United States        Off ice of         EPA-600/8-83-016B
             Environmental Protection     Research and Development    May 1983
             Agency          Washington, DC 20460
             Research and Development
&EPA      The Acidic Deposition
            Phenomenon and
            Its Effects

            Critical Assessment
            Review Papers

            Volume II Effects Sciences

            Public Review Draft

-------
    THE ACIDIC DEPOSITION  PHENOMENON AND ITS EFFECTS:
             CRITICAL ASSESSMENT REVIEW  PAPERS
Aubrey P.  Altshuller,  Editor
   Atmospheric Sciences

        Co-editors
       John S.  Nader
    Lawrence" E.  Niemeyer
Rick A.  Linthurst,  Editor
    Effects Sciences

       Co-editors
    William W.  McFee
    Dale W. Johnson
   James N. Galloway
    John J. Magnuson
     Joan P. Baker
                             Project Staff
                      Rick A. Linthurst-Director
                      Betsy A. Hood-Coordinator
                   Gary  B. Blank-Manuscript Editor
                  Clara  B. Edwards-Production Staff
                      C. Willis Williams-Crapses
                        Mike Conley-Graphics
                         Advisory Committee

                       David A. Bennett-U.S. EPA
                           Project Officer
John Bachmann-U.S. EPA
Michael  Berry-U.S. EPA
Ellis B.  Cowling-NCSU
J. Michael  Davis-U.S. EPA
Kenneth Demerjian-U.S.  EPA
  J.  H. B.  Garner-U.S.  EPA
  James L.  Regens-U.S.  EPA
  Raymond Wilhour-U.S.  EPA
     This document  has been prepared through the U.S.  EPA/NCSU Acid
Precipitation Program, a cooperative agreement between the U.S.
Environmental Protection Agency, Washington, D.C.  and North Carolina
State University, Raleigh, North Carolina.   This work was conducted
as part of the National Acid Precipitation Program and was funded by
U.S.  EPA.'

                               NOTICE

     This document  is a public review draft.  It has not been  formally
released by EPA and should not at this stage be construed to represent
Agency policy.   It  is being circulated for comment on  its technical
accuracy.
                                    U.S. Environr;
                                    Region V,  LiL.
                                    230 South D,--.
                                    Chicago, Iliinc-i:
                   ,rscy

-------
                                 Authors
 chapter
      Lawrence E.

 Chapter A.2 - Natural  and  Anthropogenic M..1... Sources

   aP E1TOr Robinson -  Washington State U.
      Jim B. Homolya -  TRW


 Chapter A-3 - Transport  Processes

      Moor V.  Gillani - Washington U.
      jac* D.  Shannon - Argonne National Lab
      David.  E.  Patterson - Kashington U.


 Chapter A-4 - Transformation Processes  .

      David F. Mfller - U.  of Nevada
      Dean />. Hegg - U. of  Washington
      Peter V. Hobbs -  U.  of Washington,
      Noor V. Gillani - Washington U.
     Michael R. Whitbeck  -  U.  of Nevada


Chapter A-5 - Atmospheric Concentrations and Distributions of Chemical
              Substances

     A. Paul Altshuller - Consultant


Chapter A-6 - Precipitation Scavenging Processes

     Jeremy M. Hales -  Battelle, Pacific Northwest Lab


Chapter A-7 - Dry Deposition Processes

     Bruce B. Hicks - National  Oceanographic and Atmospheric
                      Administration

-------
Chapter A-8 - Deposition Monitoring
          A.          .  u

Chapter A-, - Long.Range Transpon

     Chandrakant M.  Bhumralkar - National rv
            E.  ™  -  »,                              ""
Chapter E-l  -  Introduction

     Rick A. Linthurst  - North Carolina State U.


Chapter E-2  -  Effects on Soil Systems

     William W.  McFee - Purdue U.
     Fred Adams  -  Auburn U.
     Christopher S.  Cronan - U. of Maine
     Mary K. Firestone  - U. of California, Berkeley
     Charles D.  Foy  - U.S. Department of Agriculture
     Robert  D. Harter - U. of New Hampshire
     Dale W. Johnson -  Oak Ridge National Lab


Chapter E-3  -  Effects on Vegetation

     Dale W. Johnson -  Oak Ridge National Lab
     Boris I.  Chevone - Virginia Polytechnic Institute
     Patricia  M. Irving - Argonne National Lab
     Samuel  B. McLaughlin - Oak Ridge National  Lab
     Dudley  J. Raynal - Syracuse U.
     David S.  Shriner - Oak Ridge National Lab
     Lorene  L. Sigal -  Oak Ridge National Lab
     John M. Skelly  - Pennsylvania State U.
     William H.  Smith - Yale U.
     Jerome  B. Weber -  North Carolina State U.
                                  ii

-------
Chapter E-4 -  Effects  on Aquatic Chemistry

     James N.Galloway  - U.  of Virginia
     Dennis S. Anderson - U. of Maine
     M. Robbins Church - U.S. EPA
     Christopher S.  Cronan  - U. of Maine
     Ronald B. Davis - U. of Maine
     Peter J.  Dillon - Ontario Ministry of Environment
     Charles T. Driscoll -  Syracuse U.
     Steve A.  Norton - U. of Maine
     Gary C. Schafran  - Syracuse U.
Chapter E-5 -  Effects  on  Aquatic  Biology

     John J. Magnuson  - U.  of Wisconsin
     Joan P. Baker - North  Carolina State U.
     Peter G.  Daye - Daye Atlantic Salmon Corp.
     Charles T.  Driscoll  -  Syracuse U.
     Kathleen   Fischer -  Environment Canada
     Charles A.  Guthrie - N.Y.  State Dept. of Environ. Conservation
     John H. Peverly - NY State College Agric. & Life Sciences
     Frank J.  Rahel -  U.  of Wisconsin
     Gary C. Schafran  - Syracuse  U.
     Robert Singer - Colgate U.
Chapter E-6 - Indirect Effects  on Health

     Thomas W. Clarkson -  U.  of Rochester
     Joan P. Baker -  North Carolina State U.
     William E. Sharpe - Pennsylvania State U.
Chapter E-7 - Effects on Materials

     John Yocom - TRC Environ.  Consultants, Inc.
     Norbert S. Baer - New York U.
                                   iii

-------
                                 PREFACE
     The Acidic Deposition Phenomenon  and  Its  Effects:  Critical
Assessment Review Papers public  review draft,  is a technical review
document in two volumes, prepared and  released for a 90-day period of
public technical  comment.  The Environmental Protection Agency will
develop an interpretive summary, The Acidic Deposition Phenomenon and
Its Effects:   Critical  Assessment~Pocument, based upon the content of
the Review Papers and the public comments.

     The Acidic Deposition Phenomenon  and  Its  Effects:  Critical
Assessment Review Papers was requested by  the  Clean Air Scientific
Advisory Committee (CASAC) of EPA's Science Advisory Board and will be
reviewed by that committee.   The CASAC is  comprised of independent
scientists who are quite knowledgeable in  matters pertaining to
atmospheric pollution and its effects.  These  scientists will evaluate
the scientific adequacy of the Critical  Assessment Document.  As part of
this evaluation, the CASAC considers the comments and criticisms of the
general public and scientific community  as they pertain to scientific
issues and questions.  (Although the science of an issue may obviously
have implications for policy decisions,  matters of policy per se are not
in the province of the document.) This  review process is essential to
developing a scientifically unimpeachable  assessment.

     The document's original charge was  to prepare  'a comprehensive
document which lays out the state of our knowledge with regard to
precursor emissions, pollutant transformation  to acidic compounds,
pollutant transport, pollutant deposition  and  the effects (both measured
and potential) of acidic deposition.1   It  was  the decision of the
editors to provide the following guidelines to the authors writing the
Critical Assessment Review Papers to meet  this overall objective of the
document:

       1.  Contributions are written for  scientists and informed lay
          persons.

       2.  Statements are to be explained and  supported by references;
          i.e., a textbook type  of approach,  in an objective style.

       3.  Literature referenced  is to  be of high quality and not every
          reference available is to be included.

       4.  Emphasis is to be placed on  North American  systems with
          concentrated effort on U.S.  data.

       5.  Overlap between this document  and the SOX  Criteria Document
          is to be minimized.
                                   iv

-------
      6.  Potential vs known processes/effects  is to be clearly noted to
          avoid misinterpretation.

      7.  The certainty of our knowledge  should be quantified, when
          possible.

      8.  Conclusions are to be drawn  on  fact only.

      9.  Extrapolation beyond the  available data is to avoided.

     10.  Scientific knowledge is to be included without regard to
          policy implications.

     11.  Policy-related options or recommendations are beyond the scope
          of this document and are  not to be included.

The reader,  to avoid possible misinterpretation of the information
presented, is advised to consider and  understand these directives before
reading.

     Again,  the document has been designed to address our present status
of knowledge relative to the acidic deposition phenomenon and its
effects.  It is not a Criteria Document;  it is not designed to set
standards and no connections to regulations should be inferred.  The
literature is reviewed and conclusions are drawn based on the best
evidence available.  It is an authored document, and as such, the con-
clusions are those of the authors after their review of the literature.

     The success of the Critical  Assessment Review Papers has depended
on the coordinated efforts of many  individuals.  The document involved
the participation of over 54 scientists contributing material on their
special areas of expertise under the broad headings of either
atmospheric  processes or effects.   Coordination within these two areas
has been the responsibility of A. Paul Altshuller and Rick A. Linthurst,
the atmospheric and effects section editors, respectively.  Overall
coordination of the project for EPA is under David A. Bennett's
direction.  Dr. Altshuller is an atmospheric chemist, past recipient of
the American Chemical  Society Award in Pollution Control, and recently
retired director of EPA's Environmental Sciences Research Laboratory;
Dr. Linthurst is an ecologist and serves  as Program Coordinator for the
Acid Precipitation Program at North Carolina State University.  Dr.
Bennett is the Director of the Acid Deposition Assessment Staff in EPA's
Office of Research and Development  and provides liaison between the
section editors/contributors and CASAC scientific reviewers.

     The United States and Canada in 1980 signed a Memorandum of Intent
to seek agreement on transboundary  air pollution issues.  A number of
working groups are compiling technical information to support the
negotiations called for by the Memorandum.  Although the Critical
Assessment Document and the U.S.-Canada working group reports come from
different origins, and are intended for different purposes, there is
likely to be some overlap in their  areas  of coverage.

-------
     The written materials  to follow  are  contributions  from one to eight
authors per chapter,  integrated  by  the editors.  Approximately 75
scientists, with expertise  in the  fields  being addressed, have
participated in reviewing earlier drafts  of the chapters.  In addition,
200 individuals participated in  a  public  workshop held  for the review
of these materials in November of 1982.   Numerous changes resulted from
these reviews, and this document reflects those comments.  This is the
final  public review draft and comments are welcome.  However, several
guidelines and forms should be used to submit formal comments.  Please
consult the last section of the  volume for details.

          ACKNOWLEDGMENTS FROM NORTH  CAROLINA STATE UNIVERSITY

     The editorial  staff wishes  to  extend special thanks to all the
authors of this document.   They  have  been patient and tolerant of our
changes, recommendations, and deadlines,  leading to this fourth version
of the document.  These dedicated  persons are to be commended for their
efforts.

     We also wish to acknowledge our  Steering Committee, who has been
patient with our errors and deadline  delays.  These people have made
major contributions to this product,  and  actively assisted us with their
recommendations on producing this  document.  Their objectivity, concern
for technical accuracy, and support is appreciated.  Dr. J. Michael
Davis of EPA deserves special thanks, as  he directed the initial draft
of the document in December of 1981.  His concern for clarity of thought
and writing in the interest of communicating our scientific knowledge
was most helpful.  Dr. David Bennett  of EPA is specifically recognized
for his role as a scientific reviewer, and an EPA staff member who
buffered the editorial staff and the  authors from the public and policy
concerns associated with this document.   Dr. Bennett's  tolerance,
patience, and understanding are  also  appreciated.

     All the reviewers, too numerous  to list, are gratefully
acknowledged for helping us improve the quality and accuracy of this
document.  These people were from  private, State, Federal, and special-
interest organizations.  Their concern  for the truth, as we know it now,
is a compliment to all the  individuals and organizations who were
willing to deal objectively with this most important topic.  It has been
a pleasure to see all groups, independent of their personal
philosophies, work together in the interest  of producing a technically
accurate document.

     Dr. Arthur Stern is acknowledged for his contribution as  a
technical editor of the atmospheric sciences early  in the document's
preparation.  He has made an important contribution to  the final
product.

     Finally, EPA is  acknowledged for its willingness to give  the
scientists an opportunity to prepare  this document.   Its interest, as
expressed through the staff and authors,  in  having  this document be  an
authored document to assist in research  planning,  is most appreciated.
Rarely  does  a group of  scientists have  such  a  free  hand in contributing
independently to such an important issue  and  in  such  a  visible way.

-------
                      THE ACIDIC DEPOSITION PHENOMENON AND ITS EFFECTS:
                             CRITICAL ASSESSMENT REVIEW PAPERS

                                     Table of Contents

                                         Volume I
                                   Atmospheric Sciences

Note:  Comment forms and guidelines to be used by reviewers can be found at the ends  of
       Volumes I and II

                                                                                     Page

GLOSSARY  (not available)

ACRONYM LIST 	   xxiii

A-l  INTRODUCTION

     1.1  Objectives 	   1-1
     1.2  Approach—Movement from Sources to Receptor 	   1-1
          1.2.1  Chemical Substances of Interest	   1_1
          1.2.2  Natural and Anthropogenic Emissions  Sources 	   1-1
          1.2.3  Transport Processes 	   1-1
          1.2.4  Transformation Processes 	   1-1
          1.2.5  Atmospheric Concentrations and Distributions  of Chemical             1-2
                 Substances 	   1-2


A-2  NATURAL AND ANTHROPOGENIC EMISSIONS SOURCES

     2.1  Introduction	,	   2-1
     2.2  Natural Emission Sources	   2-1
          2.2.1  Sulfur Compounds 	   2-1
                 2.2.1.1  Introduction	   2-1
                 2.2.1.2  Estimates  of Natural  Sources  	   2-2
                 2.2.1.3  Biogenic Emissions of  Sulfur  Compounds	   2-5
                 2.2.1.4  Geophysical  Sources of Natural Sulfur Compounds  	   2-16
                          2.2.1.4.1   Volcanism  	   2-16
                          2.2.1.4.2   Marine  sources of  aerosol particles and
                                     gases 	   2-20
                 2.2.1.5  Scavenging  Processes and Sinks 	   2-22
                 2.2.1.6  Summary of  Natural  Sources of Sulfur Compounds 	   2-23
          2.2.2  Nitrogen Compounds  	   2-24
                 2.2.2.1  Introduction 	'....I!'.   2-24
                 2.2.2.2  Estimates of Natural Global Sources and Sinks  	   2-25
                 2.2.2.3  Biogenic Sources of NOX Compounds  	   2-29
                 2.2.2.4  Tropospheric  and Stratospheric Reactions 	   2-31
                 2.2.2.5  Formation of NOX by Lightning 	   2-32
                 2.2.2.6  Biogenic NQX  Emissions  Estimate for the United States ...  2-33
                 2.2.2.7  Biogenic Sources of Ammonia 	  2-34
                 2.2.2.8  Oceanic Source  for Ammonia 	  2-38
                 2.2.2.9  Biogenic Ammonia Emissions Estimates for the United
                          States	  2-39
                 2.2.2.10  Meteorological  and Area Variations for NOX and Ammonia
                          Emi ssions 	  2-40
                 2.2.2.11  Scavenging Processes for NOX and Ammonia 	  2-40
                 2.2.2.12 Organic Nitrogen Compounds 	  2-40
                 2.2.2.13  Summary of Natural NOX and Ammonia Emissions 	  2-41
          2.2.3   Chlorine Compounds	  2-41
                 2.2.3.1   Introduction  	\\  2-41
                 2.2.3.2  Oceanic Sources  	  2-42
                 2.2.3.3  Vol cam' sm	  2-46
                 2.2.3.4  Combustion 	  2-46
                                           VII

-------
Table of Contents (continued)

                                                                                    Page

                 2.2.3.5  Total  Natural  Chlorine  Sources	  2-47
                 2.2.3.6  Seasonal  Distributions  	  2-47
                 2.2.3.7  Environmental  Impacts of  Natural Chlorides  	  2-47
          2.2.4  Natural Sources of Aerosol Particles  	  2-49
          2.2.5  Precipitation pH in Background Conditions 	  2-50
          2.2.6  Summary 	  2-54
     2.3  Anthropogenic Emissions 	,	  2-55
          2.3.1  Origins of Anthropogenically  Emitted  Compounds and
                 Related Issues 	  2-55
          2.3.2  Historical Trends and  Current Emissions  of  Sulfur Compounds	  2-58
                 2.3.2.1  Sulfur Oxides 	  2-58
                 2.3.2.2  Primary Sulfate Emissions 	  2-66
          2.3.3  Historical Trends and  Current Emissions  of  Nitrogen  Oxides  	  2-72
          2.3.4  Historical Trends and  Current Emissions  of  Hydrochloric Acid (HC1)  2-75
          2.3.5  Historical Trends and  Current Emissions  of  Heavy Metals Emitted
                 from Fuel Combustion 	  2-79
          2.3.6  Historical Emissions Trends  in Canada 	  2-87
          2.3.7  Future Trends in Emissions  	  2-96
                 2.3.7.1  United States 	  2-96
                 2.3.7.2  Canada	  2-96
          2.3.8  Emissions Inventories  	  2-98
          2.3.9  The Potential for Neutralization of Atmospheric
                 Acidity by Suspended Fly Ash  	  2-100
     2.4  Conclusions 	  2-105
     2.5  References 	  2-109


A-3  TRANSPORT PROCESSES

     3.1  Introduction  	  3-1
          3.1.1  The Concept of Atmospheric  Residence  Times  	  3-1
     3.2  Meteorological Scales and Atmospheric Motions 	  3-3
          3.2.1  Meteorological Scales  	  3-3
          3.2.2  Atmospheric Motions 	  3-4
     3.3  Pollutant Transport Layer: Its Structure  and Dynamics  	  3-11
          3.3.1  The Planetary Boundary Laye>"  	  3-11
          3.3.2  Structure of the Transport  Layer 	  3-13
          3.3.3  Dynamics of the Transport Layer  	  3-15
          3.3.4  Effects of Mesoscale Complex  Systems  on  Transport Layer Structure
                 and Dynamics . ,	  3-28
                 3.3.4.1  Effect of Mesoscale  Convective  Precipitation Systems
                          (MCPS) 	  3-28
                 3.3.4.2  Complex Terrain Effects 	  3-32
                          3.3.4.2.1  Shoreline environment effects  	  3-32
                          3.3.4.2.2  Urban effects  	  3-35
                          3.3.4.2.3  Hilly terrain  effects 	  3-36
     3.4  Mesoscale Plume Transport and Dilution  	  3-39
          3.4.1  Elevated Point-Source  Emissions  	  3-39
          3.4.2  Broad Areal Emissions  Near  Ground  	  3-62
     3.5  Continental and Hemispheric Transport  	  3-68
     3.6  Conclusions 	  3-91
     3.7  References 	  3-94


A-4  TRANSFORMATION PROCESSES

     4.1  Introduction  	  4-1
     4.2  Homogeneous Gas-Phase Reactions 	  4-3
          4.2.1  Fundamental Reactions  	  4-3
                                           vi i i

-------
Table of Contents (continued)

                                                                                    Page

                 4.2.1.1  Reduced Sulfur Compounds  	  4-3
                 4.2.1.2  Sulfur Dioxide 	  4-4
                 4.2.1.3  Nitrogen Compounds	  4-10
                 4.2.1.4  Halogens 	  4-16
                 4.2.1.5  Organic Acids 	  4-16
          4.2.2  Laboratory Simulations of Sulfur Dioxide and Nitrogen Dioxide
                 Oxidation 	  4-18
          4.2.3  Field Studies of Gas-Phase Reactions  	  4-21
                 4.2.3.1  Urban Plumes 	  4-21
                 4.2.3.2  Power Plant Plumes  	  4-24
          4.2.4  Summary 	  4-29
     4.3  Solution Reactions 	  4-31
          4.3.1  Introduction 	  4-31
          4.3.2  Absorption of Acid 	  4-32
          4.3.3  Production of HC1  in Solution 	  4-38
          4.3.4  Production of HN03 in Solution 	  4-38
          4.3.5  Production of H2S04 in Solution 	  4-42
                 4.3.5.1  Evidence from Field Studies	  4-42
                 4.3.5.2  Homogeneous Aerobic Oxidation  of  S02-H20 to H2S04  	  4-43
                          4.3.5.2.1  Uncatalyzed 	  4-43
                          4.3.5.2.2  Catalyzed 	  4-45
                 4.3.5.3  Homogeneous Non-aerobic Oxidation of SO?'H20 to H2S04  ...  4-48
                 4.3.5.4  Heterogeneous Production  of  H2S04 in Solution  	  4-53
                 4.3.5.5  The Relative Importance of the Various H2S04
                          Production Mechanisms 	  4-54
          4.3.6  Neutralization Reactions 	  4-62
                 4.3.6.1  Neutralizati-on by NHs 	  4-62
                 4.3.6.2  Neutralization by Particle-Acid Reactions  	  4-63
          4.3.7  Summary	  4-64
     4.4  Transformation Models 	  4-64
          4.4.1  Introduction	  4-64
          4.4.2  Approaches to Transformation Modeling 	  4-67
                 4.4.2.1  The Fundamental Approach  	  4-67
                 4.4.2.2  The Empirical  Approach 	  4-70
          4.4.3  The Question of Linearity 	  4-70
          4.4.4  Some Specific Models and Their Applications 	  4-75
                 4.4.4.1  Detailed Chemical Simulations  	  4-75
                 4.4.4.2  Parameterized Models 	  4-77
          4.4.5  Summary 	  4-81
     4.5  Conclusions 	_.	  4-83
     4.6  References 	."	  4-87


A-5  ATMOSPHERIC CONCENTRATIONS AND DISTRIBUTIONS OF CHEMICAL SUBSTANCES

     5.1  Introduction 	  5-1
     5.2  Sulfur Compounds	  5-2
          5.2.1  Historical  Distribution Patterns 	  5-2
          5.2.2  Sulfur Dioxide 	  5-3
                 5.2.2.1  Urban Measurements  	  5-3
                 5.2.2.2  Nonurban Measurements	  5-4
                 5.2.2.3  Concentration Measurements at  Remote Locations 	  5-12
          5.2.3  Sulfate 	  5-13
                 5.2.3.1  Urban Concentration Measurements  	  5-13
                 5.2.3.2  Urban Composition Measurements 	  5-15
                 5.2.3.3  Nonurban Concentration Measurements 	  5-15
                 5.2.3.4  Nonurban Composition Measurements 	  5-19
                 5.2.3.5  Concentration and Composition  Measurements at Remote
                          Locations 	  5-22
          5.2.4  Particle Size Characteristics of Particulate Sulfur Compounds ....  5-23


                                           ix

-------
Table of Contents (continued)

                                                                                    Page

                 5.2.4.1   Urban Measurements  	  5-23
                 5.2.4.2   Monurban  Size  Measurements	  5-25
                 5.2.4.3   Measurements at Remote Locations  	  5-26
     5.3  Nitrogen Compounds 	  5-27
          5.3.1  Introduction 	  5-27
          5.3.2.  Nitrogen Oxides 	  5-27
                 5.3.2.1   Historical  Distribution Patterns  and Current
                          Concentrations of Nitrogen  Oxides  	  5-27
                 5.3.2.2   Measurements Techniques-Nitrogen Oxides  	  5-28
                 5.3.2.3   Urban Concentration  Measurements  	  5-28
                 5.3.2.4   Nonurban  Concentration Measurements  	  5-29
                 5.3.2.5   Measurements of Concentrations at  Remote Locations 	  5-33
          5.3.3  Nitric Acid	  5-35
                 5.3.3.1   Urban Concentration  Measurements	  5-35
                 5.3.3.2   Nonurban  Concentration Measurements  	  5-38
                 5.3.3.3   Concentration  Measurements  at Remote Locations  	  5-43
          5.3.4  Peroxyacetyl  Nitrates  	  5-44
                 5.3.4.1   Urban Concentration  Measurements  	  5-44
                 5.3.4.2   Nonurban  Concentration Measurements	  5-46
          5.3.5  Ammonia  	  5-48
                 5.3.5.1   Urban Concentration  Measurements	  5-50
                 5.3.5.2   Nonurban  Concentration Measurements  	  5-50
          5.3.6  Particulate Nitrate  	  5-51
                 5.3.6.1   Urban Concentration  Measurements	  5-53
                 5.3,6.2   Nonurban  Concentration Measurements  	  5-55
                 5.3.6.3   Concentration  Measurements  at Remote Locations	  5-55
          5.3.7  Particle Size Characteristics of Particulate Nitrogen Compounds  ..  5-56
     5.4  Ozone 	  5-58
          5.4.1  Concentration Measurements Within  the Planetary Boundary Layer
                 (PBL) 	  5-60
          5.4.2  Concentration Measurements at Higher Altitudes  	  5-63
     5.5  Hydrogen Peroxide 	  5-63
          5.5.1  Urban Concentration  Measurements  	  5-64
          5.5.2  Nonurban Concentration  Measurements  	  5-65
          5.5.3  Concentration Measurements in Rainwater	  5-65
     5.6  Chlorine Compounds 	  5-66
          5.6.1  Introduction 	  5-66
          5.6.2  Hydrogen Chloride  	  5-66
          5.6.3  Particulate Chloride 	  5-67
          5.6.4  Particle Size tharacteristies of Particulate  Chlorine Compounds  ..  5-67
     5.7  Metallic Elements 	  5-68
          5.7.1  Concentration Measurements and Particle  Sizes in Urban Areas	  5-69
          5.7.2  Concentration Measurements and Particle  Sizes In Nonurban Areas  ..  5-71
     5.8  Relationship of Light Extinction and Visual Range  Measurements  to Aerosol
          Composi tion 	  5-74
          5.8.1  Fine Particle Concentration and Light Scattering Coefficients ....  5-74
          5.8.2  Light Extinction or  Light Scattering Budgets  at Urban Locations  ..  5-75
          5.8.3  Light Extinction or  Light Scattering Budgets at Nonurban
                 Locations 	  5-77
          5.8.4  Trends in Visibility as Related to Sulfate  Concentrations 	  5-79
     5.9  Conclusions 	  5-79
     5.10 References 	  5-85


A-6  PRECIPITATION SCAVENGING PROCESSES

     6.1  Introduction	  6-1
     6.2  Steps in the Scavenging Sequence 	  6-3

-------
Table of Contents (continued)

                                                                                    Page

          6.2.1  Introduction  	  6-3
          6.2.2  Intermixing of Pollutant and Condensed Water (Step 1-2) 	  6-7
          6.2.3  Attachment of Pollutant to Condensed Water Elements (Step 2-3) ...  6-8
          6.2.4  Aqueous-Phase Reactions (Step 3-4)  	  6-15
          6.2.5  Deposition of Pollutant with Precipitation (Step 4-5) 	  6-15
          6.2.6  Combined Processes  and the Problem of Scavenging Calculations ....  6-18
     6.3  Storm Systems and Storm Climatology 	  6-18
          6.3.1  Introduction  	  6-18
          6.3.2  Frontal  Storm Systems  	  6-19
                 6.3.2.1   Warm-Front Storms	  6-20
                 6.3.2.2   Cold-Front Storms 	  6-25
                 6.3.2.3   Occluded-Front Storms	  6-25
          6.3.3  Convectlve Storm Systems  	  6-28
          6.3.4  Additional Storm Types:  Nonideal Frontal Storms, Orographlc
                 Storms and Lake-Effect Storms 	  6-28
          6.3.5  Storm and Precipitation Climatology 	  6-30
                 6.3.5.1   Precipitation Climatology  	  6-32
                 6.3.5.2   Storm Tracks  	  6-32
                 6.3.5.3   Storm Duration Statistics  	  6-35
     6.4  Summary of Precipitation-Scavenging Field  Investigation 	  6-35
     6.5  Predictive and Interpretive Models of  Scavenging 	  6-51
          6.5.1  Introduction  	  6-51
          6.5.2  Elements of a Scavenging Model  	  6-54
                 6.5.2.1   Material Balances 	  6-54
                 6.5.2.2   Energy Balances  	  6-55
                 6.5.2.3   Momentum Balances 	  6-56
          6.5.3  Definitions of Scavenging Parameters	  6-56
          6.5.4  Formulation of Scavenging Models:   Simple Examples
                 of Microscopic and  Macroscopic  Approaches 	  6-62
          6.5.5  Systematic Selection of Scavenging  Models:
                 A F1 ow Chart  Approach  	  6-65
     6.6  Practical Aspects of Scavenging Models:  Uncertainty Levels and Sources
          of Error 	  6-68
     6.7  Conclusions 	  6-72
     6.8  References 	  6-75


A-7  DRY DEPOSITION PROCESSES

     7.1  Introduction 	  7-1
     7.2  Factors Affecting Dry Deposition  	  7-1
          7.2.1  Introduction  	  7-1
          7.2.2  Aerodynamic Factors 	  7-6
          7.2.3  The Quasi-Laminar Layer	  7-9
          7.2.4  Phoretic Effects and Stefan Flow  	  7-12
          7.2.5  Surface Adhesion 	;	  7-15
          7.2.6  Surface Biological  Effects 	  7-15
          7.2.7  Deposition to Liquid Water Surfaces 	  7-16
          7.2.8  Deposition to Mineral  and Metal  Surfaces  	  7-19
          7.2.9  Fog and  Dewfall  	  7-20
          7.2.10 Resuspension  and Surface Emission 	  7-21
          7.2.11 The resistance Analog  	  7-22
     7.3  Methods for Studying Dry Deposition	  7-28
          7.3.1  Direct Measurement  	  7-28
          7.3.2  Wind Tunnel and Chamber Studies  	  7-31
          7.3.3  Micrometeorological  Measurement  Methods  	  7-33
     7.4  Field Investigations of Dry Deposition  	  7-39
          7.4.1  Gaseous  Pollutants  	  7-39
          7.4.2  Particulate Pollutants 	  7-46
          7.4.3  Routine  Handling in Networks 	  7-51

                                           XI

-------
Table of Contents (continued)

                                                                                    Page

     7.5  Micrometeorological Models of the Dry Deposition Process 	  7-52
          7.5.1  Gases 	  7-52
          7.5.2  Particles  	  7-55
     7.6  Summary 	  7-56
     7.7  Conclusions 	  7-60
     7.8  References 	  7-63


A-8  DEPOSITION MONITORING

     8.1  Introduction 	  8-1
     8.2  Wet Deposition Networks  	  8-2
          8.2.1  Introduction and  Historical Background 	  8-2
          8.2.2  Definitions 	  8-3
          8.2.3  Methods, Procedures and Equipment for Wet Deposition Networks ....  8-4
          8.2.4  Wet Deposition  Network Data Bases 	  8-7
     8.3  Monitoring Capabilities  for Dry Deposition  	  8-11
          8.3.1  Introduction 	  8-11
          8.3.2  Methods for Monitoring Dry Deposition 	  8-17
                 8.3.2.1 Direct Collection Procedures 	  8-18
                 8.3.2.2 Alternative Methods  	  8-20
          8.3.3  Evaluations of  Dry Deposition Rates  	  8-21
     8.4  Wet Deposition Network Data With Applications to Selected Problems	  8-28
          8.4.1  Spatial Patterns  	  8-28
          8.4.2  Remote Site pH  Data  	  8-50
          8.4.3  Precipitation Chemistry Variations Over Time 	  8-59
                 8.4.3.1 Nitrate  Varfation Since 1950's 	  8-59
                 8.4.3.2 pH Variation Since 1950's 	  8-61
                 8.4.3.3 Calciiro  Variation Since the I960' s 	  8-65
          8.4.4  Seasonal Variations  	  8-67
          8.4.5  Very Short Time Scale Variations 	  8-68
          8.4.6  Air Parcel Trajectory Analysis 	  8-68
     8.5  Glaciochemical Investigations as a Tool in  the Historical Delineation of
          the Acid Precipitation Problems 	  8-70
          8.5.1  Glaciochemical  Data  	  8-70
                 8.5.1.1 Sulfate  - Polar Glaciers 	  8-71
                 8.5.1.2 Nitrate  - Polar Glaciers 	  8-72
                 8.5.1.3 pH and Acidity - Polar Glaciers  	  8-72
                 8.5.1.4 Sulfate  - Alpine Glaciers 	  8-73
                 8.5.1.5 Nitrate  - Alpine Glaciers 	  8-73
                 8.5.1.6 pH and Acidity - Alpine Glaciers 	  8-73
          8.5.2  Trace Metals -  General Statement	  8-74
                 8.5.2.1 Trace  Metals - Polar Glaciers 	  8-74
                 8.5.2.2 Trace  Metals - Alpine Glaciers 	  8-76
          8.5.3  Discussion and  Future Work  	  8-76
     8.6  Conclusions 	  8-79
     8.7  References 	  8-83


A-9  LONG-RANGE TRANSPORT AND ACIDIC DEPOSITION MODELS

     9.1  Introduction 	  9-1
          9.1.1  General Principles for Formulating Pollution Transport and
                 Diffusion  Models  	  9-1
          9.1.2  Model Characteristics  	  9-3
                 9.1.2.1  Spatial  and Temporal Scales 	  9-3
                 9.1.2.2 Treatment of Turbulence 	  9-5
                 9.1.2.3 Reaction Mechanisms  	  9-5
                 9.1.2.4 Removal  Mechanisms  	  9-5
                                            xn

-------
Table of Contents (continued)

                                                                                    Page

          9.1.3  Selecting Models  for Application  	  9-6
                 9.1.3.1   General  	  9-6
                 9.1.3.2   Spatial  Range  of Application  	  9-6
                 9.1.3.3   Temporal  Range of  Application  	  9-8
     9.2  Types of LRT Models  	  9-8
          9.2.1  Eulerian Grid Models 	  9-8
          9.2.2  Lagrangian Models	  9-11
                 9.2.2.1   Lagrangian  Trajectory Models  	  9-11
                 9.2.2.2   Statistical Trajectory Models  	  9-13
          9.2.3  Hybrid Models 	  9-13
     9.3  Modules Associated with  Chemical (Transformation) Processes  	  9-14
          9.3.1  Overview 	  9-14
          9.3.2  Chemical Transformation Modeling  	  9-14
                 9.3.2.1   Simplified  Modules 	  9-15
                 9.3.2.2   Multireaction  Modules 	  9-15
          9.3.3  Modules  for NOX Transformation 	  9-16
     9.4  Modules Associated with  Wet and Dry Deposition  	  9-20
          9.4.1  Overview 	  9-20
          9.4.2  Modules  for Wet Deposition  	  9-21
                 9.4.2.1   Formulation and Mechanism  	  9-21
                 9.4.2.2   Modules  Used in Existing Models  	  9-22
                 9.4.2.3   Wet  Deposition Modules for Snow	  9-24
                 9.4.2.4   Wet  Deposition Modules for NOX  	  9-24
          9.4.3  Modules  for Dry Deposition  	  9-24
                 9.4.3.1   General  Considerations 	  9-24
                 9.4.3.2   Modules  Used in Existing Models	:	  9-26
                 9.4.3.3   Dry  Deposition Modules for NOX  	  9-26
          9.4.4  Dry Versus Wet Deposition 	  9-26
     9.5  Status of LRT Models as  Operational Tools  	  9-27
          9.5.1  Overview	  9-27
          9.5.2  Model Application 	  9-27
                 9.5.2.1   Selection Criteria 	  9-27
                 9.5.2.2   Regional  Concentration and Deposition Patterns 	  9-28
                 9.5.2.3   Use  of Matrix  Methods to Quantify Source-Receptor
                          Relationships  	  9-29
          9.5.3  Data Requirements  	  9-34
                 9.5.3.1   General  	  9-34
                 9.5.3.2   Specific  Characteristics of Data Used in Model
                          Simulations 	  9-37
                          9.5.3.2.1  Emissions 	  9-37
                          9.5.3.2.2  Meteorological Data  	  9-38
          9.5.4  Model Performance  and Uncertainties 	  9-38
                 9.5.4.1   General  	,	  9-38
                 9.5.4.2   Data Bases  Available for Evaluating Models 	  9-40
                 9.5.4.3   Performance Measures 	  9-40
                 9.5.4.4   Representivity of  Measurements  	  9-41
                 9.5.4.5   Uncertainties  	  9-41
                 9.5.4.6   Selected  Results 	  9-42
     9.6  Conclusions	  9-47
     9.7  References 	  9-49
                                           xm

-------
                      THE ACIDIC DEPOSITION  PHENOMENON  AND  ITS EFFECTS:
                             CRITICAL  ASSESSMENT  REVIEW PAPERS

                                    Table of  Contents

                                         Volume II
                                    Effects Sciences

Note:  Comment forms and guidelines to be used by reviewers can be  found at  the ends of
       Volumes I and II

                                                                                    Page

E-l  INTRODUCTION

     1.1  Objectives 	  1-1
     1.2  Approach 	  1-1
     1.3  Chapter Organization and General Content 	  1-2
          1.3.1  Effects on Soil Samples 	  1-3
          1.3.2  Effects on Vegetation 	  1-3
          1.3.3  Effects on Aquatic Chemistry  	  1-4
          1.3.4  Effects on Aquatic Biology  	  1-4
          1.3.5  Indirect Effects on Health  	  1-5
          1.3.6  Effects on Materials  	  1-5
     1.4  Acidic Deposition 	  1-5
     1.5  Linkage to Atmospheric Sciences 	  1-6
     1.6  Sensitivity 	  1-6
     1.7  Presentation Limitations 	  1-7


E-2  EFFECTS ON SOIL SYSTEMS

     2.1  Introduction	  2-1
          2.1.1  Importance of Soils to Aquatic Systems 	  2-1
                 2.1.1.1  Soils Buffer Precipitation Enroute to Aquatic Systems ...  2-1
                 2.1.1.2  Soil as a Source of  Acidity  for  Aquatic Systems  	  2-2
          2.1.2  Soil's  Importance as  a Medium for Plant Growth  	  2-2
          2.1.3  Important Soil Properties  	  2-2
                 2.1.3.1  Soil Physical Properties 	  2-3
                 2.1.3.2  Soil Chemical Properties 	  2-3
                 2.1.3.3  Soil Microbiology  	  2-3
          2.1.4  Flow of Deposited Materials Through Soil  Systems 	  2-3
     2.2  Chemistry of Acid Soils 	  2-5
          2.2.1  Development of Acid Soils  	  2-5
                 2.2.1.1  Biological Sources of H+ Ions 	  2-6
                 2.2.1.2  Acidity from Dissolved  Carbon Dioxide  	  2-6
                 2.2.1.3  Leaching of  Basic  Cations 	  2-7
          2.2.2  Soil Cation Exchange  Capacity	  2-8
                 2.2.2.1  Source of Cation Exchange Capacity in  Soils  	  2-8
                 2.2.2.2  Exchangeable Bases and  Base  Saturation  	  2-8
          2.2.3  Exchangeable and Solution Aluminum in  Soils 	  2-9
          2.2.4  Exchangeable and Solution Manganese in Soils 	  2-12
          2.2.5  Practical Effects of  Low pH 	  2-12
          2.2.6  Neutralization of Soil Acidity  .:	  2-13
          2.2.7  Measuring Soil pH 	  2-14
          2.2.8  Sulfate Adsorption 	  2-15
          2.2.9  Soil Chemistry Summary 	  2-18
     2.3  Effects of Acidic Deposition on Soil Chemistry and Plant  Nutrition	  2-19
          2.3.1  Effects on Soil pH 	  2-19
          2.3.2  Effects on Nutrient Supply  of Cultivated  Crops  	  2-24
          2.3.3  Effects on Nutrient Supply  to Forests  	  2-25
                 2.3.3.1  Effects on Cation  Nutrient Status 	  2-29
                 2.3.3.2  Effects on S and N Status 	  2-31
                 2.3.3.3  Acidification Effects on Plant Nutrition  	  2-34
                                            XIV

-------
Table of Contents (continued)

                                                                                     Page

                          2.3.3.3.1  Nutrient deficiencies 	   2-34
                          2.3.3.3.2  Metal  ion toxicities 	   2-34
                                     2.3.3.3.2.1   Aluminum toxicity 	   2-35
                                     2.3.3.3.2.2   Manganese toxicity 	   2-36
          2.3.4  Reversibility of Effects on Soil  Chemistry 	   2-36
          2.3.5  Predicting Which Soils will  be Affected Most 	   2-37
                 2.3.5.1  Soils Under Cultivation  	   2-37
                 2.3.5.2  Uncultivated, Unamended  Soils  	   2-37
                          2.3.5.2.1  Basic  cation-pH  changes in  forested soils  ....   2-40
                          2.3.5.2.2  Changes in aluminum or other  metal  concen-
                                     tration in soil  solution fn forested soils  ...   2-41
     2.4  Effects of Acidic Deposition on Soil  Biology 	   2-41
          2.4.1  Soil Biology Components and Functional  Significance	   2-41
                 2.4.1.1  Soil Animals 	   2-41
                 2.4.1.2  Algae 	   2-42
                 2.4.1.3  Fungi 	   2-42
                 2.4.1.4  Bacteria 	   2-42
          2.4.2  Direct Effects of Acidic Deposition  on  Soil  Biology 	   2-43
                 2.4.2.1  Soil Animals 	   2-43
                 2.4.2.2  Terrestrial  Algae 	   2-44
                 2.4.2.3  Fungi 	   2-44
                 2.4.2.4  Bacteria 	   2-45
                 2.4.2.5  General  Biological  Processes 	   2-45
          2.4.3  Metals--Mobilization  Effects on Soil  Biology 	   2-47
          2.4.4  Effects of Changes in Microbial Activity on Aquatic Systems  	   2-48
          2.4.5  Soil Biology Summary	   2-48
     2.5  Effects of Acidic Deposition on Organic  Matter Decomposition  	   2-49
     2.6  Effects of Soils on the Chemistry of Aquatic Ecosystems  	   2-50
     2.7  Conclusions	   2-56
     2.8  References 	   2-59


E-3  EFFECTS ON VEGETATION

     3.1  Introduction 	   3-1
          3.1.1  Overview	   3-1
          3.1.2  Background	   3-2
     3.2  Plant Response to Acidic Deposition 	   3-5
          3.2.1  Leaf Response to Acidic Deposition 	   3-5
                 3.2.1.1  Leaf Structure and  Functional  Modifications 	   3-5
                 3.2.1.2  Foliar Leaching - Throughfall  Chemistry  	   3-8
          3.2.2  Effects of Acidic Deposition on Lichens and  Mosses  	   3-10
          3.2.3  Summary 	   3-17
     3.3  Interactive Effects of Acidic Deposition with  Other Environmental
          Factors on Plants 	   3-18
          3.3.1  Interactions with Other Pollutants 	   3-18
          3.3.2  Interactions with Phytophagus  Insects 	   3-21
          3.3.3  Interactions with Pathogens  ....;	   3-21
          3.3.4  Influence on Vegetative Hosts  That Would Alter  Relationships
                 with Insect or Microbial Associate 	   3-24
          3.3.5  Effects of Acidic Deposition on Pesticides  	   3-25
          3.3.6  Summary 	   3-26
     3.4  Biomass Production 	   3-27
          3.4.1  Forests 	   3-27
                 3.4.1.1   Possible Mechanisims  of Response  	   3-28
                 3.4.1.2  Phenological  Effects  	   3-30
                          3.4.1.2.1  Seed germination  and seedling establishment ..   3-31
                          3.4.1.2.2  Mature and reproductive  stages  	   3-33
                 3.4.1.3  Growth of Seedlings and Trees  in Irrigation
                          Experiments  	  3-33
                                            XV

-------
Table of Contents (continued)

                                                                                    Page

                 3.4.1.4   Studies  of long-Term Growth of Trees	  3-34
                 3.4.1.5   Dieback  and Decline in  High Elevation Forests 	  3-37
                 3.4.1.6   Summary  	  3-41
          3.4.2  Crops 	  3-42
                 3.4.2.1   Review and Analysis of  Experimental Design	  3-42
                          3.4.2.1.1  Dose-response determination  	  3-43
                          3.4.2.1.2  Sensitivity  classification 	  3-44
                          3.4.2.1.3  Mechanisms	  3-45
                          3.4.2.1.4  Characteristics of precipitation simulant
                                    exposures  	  3-45
                          3.4.2.1.5  Yield  criteria  	  3-46
                 3.4.2.2   Experimental  Results  	  3-46
                          3.4.2.2.1  Field  studies  	  3-47
                          3.4.2.2.2  Controlled environment studies 	  3-51
                 3.4.2.3   Discussion  	  3-59
                 3.4.2.4   Summary	  3-62
     3.5  Conclusions 	  3-62
     3.6  References 	  3-65


E-4  EFFECTS ON AQUATIC CHEMISTRY

     4.1  Introduction 	  4-1
     4.2  Basic Concepts  Required  to Understand the  Effects of
          Acidic Deposition on Aquatic  Systems  	  4-1
          4.2.1  Receiving Systems 	  4-1
          4.2.2  pH, Conductivity, and  Alkalinity 	  4-4
                 4.2.2.1   pH 	  4-4
                 4.2.2.2   Conductivity	  4-4
                 4.2.2.3   Alkalinity	  4-5
          4.2.3  Acidification 	  4-6
     4.3  Sensitivity of  Aquatic  Systems  to Acidic Deposition 	  4-6
          4.3.1  Atmospheric Inputs  	  4-6
                 4.3.1.1   Components of Deposition  	  4-7
                 4.3.1.2   Loading  vs Concentration	  4-8
                 4.3.1.3   Location of the Deposition 	  4-8
                 4.3.1.4   Temporal Di stribution of Deposition 	  4-8
                 4.3.1.5   Importance of Atmospheric  Inputs to Aquatic Systems	  4-9
                          4.3.1.5.1  Nitrogen (N), phosphorus (P), and
                                    carbon (C)  	  4-9
                          4.3.1.5.2  Sulfur 	  4-9
          4.3.2  Characteristics of  Receiving Systems Relative  to  Being Able to
                 Assimilate Acidic Deposition 	  4-10
                 4.3.2.1   Canopy  	  4-10
                 4.3.2.2   Soil 	  4-12
                 4.3.2.3   Bedrock  	  4-14
                 4.3.2.4   Hydrology  	  4-15
                          4.3.2.4.1  Flow paths  	  4-15
                          4.3.2.4.2  Residence  times 	  4-17
                 4.3.2.5   Wetlands 	  4-17
                 4.3.2.6   Aquatic  	  4-18
                          4.3.2.6.1  Alkalinity  	  4-18
                          4.3.2.6.2   International production/consumption
                                    of ANC 	  4-22
                          4.3.2.6.3  Aquatic sediments	  4-24
          4.3.3  Location of Sensitive  Systems  	  4-25
          4.3.4  Summary--Sensi tivi ty  	  4-30
     4.4  Magnitude of Chemical Effects of  Acidic Deposition on
          Aquatic Ecosystems	  4-31
                                         XVI

-------
Table of Contents (continued)

                                                                                    Page

          4.4.1   Relative  Importance of HN03 vs H2S04 	  4-31
          4.4.2   Short-Term Acidification  ..'.	  4-37
          4.4.3   Long-Term Acidification  	  4-38
                 4.4.3.1   Analysis of Trends based on Historic Measurements of
                          Surface Water Quality 	  4-44
                          4.4.3.1.1  Methologlcal problems with the evaluation
                                    of historical trends 	  4-44
                                    4.4.3.1.1.1  pH 	  4-44
                                                 4.4.3.1.1.1.1  pH-early metho-
                                                                dology 	  4-44
                                                 4.4.3.1.1.1.2  pH-current metho-
                                                                dology 	  4-46
                                                 4.4.3.1.1.1.3  pH-comparability
                                                                of early and cur-
                                                                rent mesurement
                                                                methods 	  4-47
                                                 4.4.3.1.1.1.4  pH-general
                                                                problems 	  4-47
                                    4.4.3.1.1.2  Conductivity 	  4-48
                                                 4.4.3.1.1.2.1  Conductivity
                                                                methodology 	  4-48
                                                 4.4.3.1.1.2.2  Comparability of
                                                                early and current
                                                                measurement
                                                                methods 	  4-48
                                                 4.4.3.1.1.2.3  General problems..  4-48
                                    4.4.3.1.1.3  Alkalinity	  4-49
                                                 4.4.3.1.1.3.1  Early methodology.  4-49
                                                 4.4.3.1.1.3.2  Current
                                                                methodology 	  4-49
                                                 4.4.3.1.1.3.3  Comparability of
                                                                early and current
                                                                measurement
                                                                methods 	  4-50
                                    4.4.3.1.1.4  Summary of measurement
                                                 techniques 	  4-51
                          4.4.3.1.2  Analysis of trends	  4-51
                                    4.4.3.1.2.1  Introduction 	  4-51
                                    4.4.3.1.2.2  Canadian studies 	  4-53
                                    4.4.3.1.2.3  United States studies 	  4-61
                          4.4.3.1.3  Summary—trends in historic data 	  4-74
                 4.4.3.2   Assessment of Trends Based on Paleol imnological
                          Technique 	  4-77
                          4.4.3.2.1  Calibration and accuracy of paleol imnological
                                    reconstruction of pH history 	  4-78
                          4.4.3.2.2  Lake acidification determined by
                                    paleol imnological  reconstruction 	  4-78
                 4.4.3.3   Alternate Explanations -for Acidification-Land Use
                          Changes 	  4-79
                          4.4.3.3.1  Variations in the groundwater table 	  4-79
                          4.4.3.3.2  Accelerated mechanical  weathering or
                                    land scarification 	  4-79
                          4.4.3.3.3  Decomposition of organic matter 	  4-80
                          4.4.3.3.4  Long-term changes in vegetation 	  4-80
                          4.4.3.3.5  Chemical amendments 	  4-80
                          4.4.3.3.6  Summary--Effects of land use changes
                                    or acidification	  4-80
          4.4.4   Summary—Magnitude of Chemical Effects of Acidic Deposition 	  4-81
     4.5   Predictive Modeling of the Effects of Acidic Deposition
          on  Surface Waters 	  4-82
                                        xvn

-------
Table of Contents (continued)

                                                                                    Page

          4.5.1  Aimer/Dick son Relationship 	  4-83
          4.5.2  Henriksen's Predictor Nomograph  	  4-88
          4.5.3  Thompson Cation Denudation Rate  Model  (CDR)  	  4-91
          4.5.4  Summary of Predictive Modeling	  4-94
     4.6  Indirect Chemical Changes Associated with Acidification
          of Surface Waters 	  4-94
          4.6.1  Metals 	»	  4.94
                 4.6.1.1  Increased Loading of Metals From Atmospheric
                          Deposition 	  4.95
                 4.6.1.2  Mobilization of Metals  by Acidic Deposition 	  4-97
                 4.6.1.3  Secondary Effects of Metal Mobilization  	  4-98
                 4.6.1.4  Effects of Acidification  on Aqueous Metal  Speciation	  4-98
                 4.6.1.5  Indirect  Effects on  Metals 1n Surface Waters  	  4-98
          4.6.2  Aluminum Chemistry 1n Dilute  Acidic Waters  	  4-99
                 4.6.2.1  Occurrence,  Distribution,  and Sources of Aluminum	  4-99
                 4.6.2.2  Aluminum  Speciation  	  4-102
                 4.6.2.3  Aluminum  as a pH Buffer 	  4-102
                 4.6.2.4  Temporal  and Spatial Variations 1n  Aqueous
                          Aqueous Levels of Aluminum 	  4-104
                 4.6.2.5  The Role  of Aluminum In Altering Element Cycling
                          Within Acidic Waters 	  4-106
          4.6.3  Organics 	  4-108
                 4.6.3.1  Atmospheric Loading  of  Strong Acids and Associated
                          Organic Micropollutants 	  4-108
                 4.6.3.2  Organic Buffering Systems  	  4-109
                 4.6.3.3  Organo-Metallc Interactions 	  4-109
                 4.6.3.4  Photochemistry 	  4-110
                 4.6.3.5  Carbon-Phosphorus-Alumlnupi Interactions  	  4-110
                 4.6.3.6  Effects of Acidification  on Organic Decomposition
                          1n Aquatic Systems 	  4-110
     4.7  Mitigative Strategies for Improvement of  Surface Water Quality  	  4-111
          4.7.1  Base Additions	  4-111
                 4.7.1.1  Types of  Basic Materials  	  4-111
                 4.7.1.2  Direct Water Addition of  Base	  4-115
                          4.7.1.2.1  Computing base  dose requirements 	  4-115
                          4.7.1.2.2  Methods of base application 	  4-119
                 4.7.1.3  Watershed Addition of Base 	  4-123
                          4.7.1.3.1  The concept  of  watershed
                                     application  of  base 	  4-123
                          4.7.1.3.2  Experience In  watershed liming  	  4-124
                 4.7.1.4  Water Quality Response  to  Base Treatment	  4-126
                 4.7.1.5  Cost Analysis,  Conclusions and Assessment of Base
                          Addition  	  4-128
                          4.7.1.5.1  Cost analysis  	  4-128
                          4.7.1.5.2  Summary—base additions  	  4-130
          4.7.2  Surface Water Fertilization 	  4-130
                 4.7.2.1  The  Fertilization Concept  	  4-130
                 4.7.2.2  Phosphorous  Cycling  in  Acidified Water	  4-132
                 4.7.2.3  Fertilization Experience and  Water
                          Quality Response to  Fertilization	  4-133
                 4.7.2.4  Summary-Surface Water Fertilization 	  4-134
     4.8  Conclusions 	  4-134
     4.9  References 	  4-137


E-5  EFFECTS ON AQUATIC BIOLOGY

     5.1  Introduction	  5-1
     5.2  Biota of Naturally Acidic Waters 	  5-3
                                         xvm

-------
Table of Contents (continued)

                                                                                    Page

          5.2.1   Types of Naturally Acidic Waters 	  5-3
          5.2.2   Biota of Inorganic Acldotrophlc Waters 	  5-4
          5.2.3   Biota In Acidic Brownwater Habitats 	  5-6
          5.2.4   Biota In Ultra-Ollgotrophic Waters 	  5-8
          5.2.5   Summary  	  5-9
     5.3  Benthic Organisms  	  5-15
          5.3.1   Importance  of  the Benthic Community 	  5-15
          5.3.2   Effects  of  Acidification on
                 Components  of  the Benthos	  5-16
                 5.3.2.1   Mlcroblal Community 	  5-17
                 5.3.2.2   Perlphyton  	  5-18
                          5.3.2.2.1   Field surveys 	  5-18
                          5.3.2.2.2   Temporal trends 	  5-19
                          5.3.2.2.3   Experimental studies 	  5-21
                 5.3.2.3   Microlnvertebrates 	  5-22
                 5.3.2.4   Crustacea 	  5-23
                 5.3.2.5   Insecta  	  5-25
                          5.3.2.5.1   Sensitivity of different groups 	  5-25
                          5.3.2.5.2   Sensitivity of Insects from different
                                     mlcrohabitats 	  5-30
                          5.3.2.5.3   Acid sensitivity of Insects based on food
                                     sources 	  5-31
                          5.3.2.5.4   Mechanisms of effects and trophic
                                     Interactions 	  5-31
                 5.3.2,6   Mollusca 	  5-32
                 5.3.2.7   Annelida 	  5-33
                 5.3.2.8   Summary of  Effects of Acidification on Benthos 	  5-34
     5.4  Macrophytes  and Wetland Plants 	  5-39
          5.4.1   Introduction 	;	  5-39
          5.4.2   Effects  on  Acidification on Aquatic Macrophytes	  5-43
          5.4.3   Summary	  5-45
     5.5  Plankton 	  5-45
          5.5.1   Introduction 	  5-45
          5.5.2   Effects  of  Acidification on Phytoplankton 	  5-47
                 5.5.2.1   Changes in  Species Composition 	  5-47
                 5.5.2.2   Changes in  Phytopl ankton Bloroass and Productivity	  5-54
          5.5.3   Effects  of  Acidification on Zooplankton 	  5-57
          5.5.4   Explanations and Significance 	  5-70
                 5.5.4.1   Changes in  Species Composition	  5-70
                 5.5.4.2   Changes in Productivity 	  5-72
          5.5.5   Summary  	  5-75
     5.6  Fishes	  5-76
          5.6.1   Introduction 	  5-76
          5.6.2   Field Observations 	  5-77
                 5.6.2.1   Loss of Populations	  5-78
                          5.6.2.1.1  United States 	  5-78
                                     5.6.2.1.1.1   Adirondack  Region of
                                                 New York State 	  5-78
                                    5.6.2.1.1.2   Other regions of the eastern
                                                 United States 	  5-81
                         5.6.2.1.2  Canada	  5-82
                                    5.6.2.1.2.1   LaCloche Mountain Region  of
                                                 Ontario	  5-82
                                    5.6.2.1.2.2   Other areas  of Ontario 	  5-86
                                    5.6.2.1.2.3   Nova  Scotia  	  5-86
                         5.6.2.1.3  Scandinavia  and  Great Britain 	  5-91
                                    5.6.2.1.3.1   Norway 	  5-91
                                    5.6.2.1.3.2   Sweden 	  5-95
                                    5.6.2.1.3.3   Scotland 	  5-95
                                          XIX

-------
Table of Contents (continued)

                                                                                    Page

                 5.6.2.2   Popul ati on Structure  	  5-97
                 5.6.2.3   Growth  	  5-100
                 5.6.2.4   Episodic  Fish K111 s  	  5-103
                 5.6.2.5   Accumulation of  totals  in  Fish  	  5-105
          5.6.3  Field  Experiments  	  5-105
                 5.6.3.1   Experimental Acidification of Lake 223 Ontario  	  5-105
                 5.6.3.2   Experimental Acidification of Norris
                          Brook,  New Hampshi re  	  5-108
                 5.6.3.3   exposure  of Fish to Acidic Surface Waters  	  5-108
          5.6.4  Laboratory Experiments  	  5-112
                 5.6.4.1   Effects of Low pH 	  5-113
                          5.6.4.1.1  Survival  	  5-113
                          5.6.4.1.2  Reproduction 	  5-116
                          5.6.4.1.3  Growth	  5-123
                          5.6.4.1.4  Behavior  	  5-124
                          5.6.4.1.5  Physiological responses 	  5-124
                 5.6.4.2   Effects of Aluminum and Other Metals  in Acidic  Waters  ...  5-127
          5.6.5  Summary  	  5-129
                 5.6.5.1   Extent  of Impact 	  5-129
                 5.6.5.2   Mechanism of Effect  	;	  5-131
                 5.6.5.3   Relationship Between  Water Acidity and Fish
                          Population Response  	  5-133
     5.7  Other Related Biota 	  5-137
          5.7.1  Amphibians 	  5-137
          5.7.2  Birds  	  5-138
                 5.7.2.1   Food Chain Alterations  	  5-138
                 5.7.2.2   Heavy Metal Accumulation 	  5-139
          5.7.3  Mammals  	  5-140
          5.7.4  Summary	;	  5-141
     5.8  Observed and  Anticipated Ecosystem  Effects	  5-144
          5.8.1  Ecosystem Structure	  5-144
          5.8.2  Ecosystem Function 	  5-146
                 5.8.2.1   Nutrient Cycling 	  5-146
                 5.8.2.2   Energy  Cycling  	  5-146
          5.8.3  Summary  	  5-147
     5.9  Mitigative Options Relative to  Biological  Populations at  Risk 	  5-148
          5.9.1  Biological Response to  Neutralization 	  5-148
          5.9.2  Improving Fish Survival  in Acidified  Waters 	  5-150
                 5.9.2.1   Genetic Screening 	  5-150
                 5.9.2.2   Selective Breeding  	  5-151
                 5.9.2.3   Acclimation 	  5-152
                 5.9.2.4   Limitations of Techniquest to  Improve Fish Survival  	  5-153
                 5.9.2.5   Summary	  5-154
     5.10 Conclusions 	  5-154
          5.10.1  Effects of Acidification on Aquatic  Organisms 	  5-155
          5.10.2  Processes and Mechanisms by  Which  Acidification
                  Alters Aquatjc  Ecosystems 	  5-161
                  5.10.2.1  Direct Effects of Hydrogen Ions  	  5-161
                  5.10.2.2  Elevated Metal Concentrations 	  5-161
                  5.10.2.3  Altered Trophic-Level Interactions  	  5-162
                  5.10.2.4  Altered Water Clarity 	  5-162
                  5.10.2.5  Altered Decomposition of Organic Matter 	  5-162
                  5.10.2.6  Presence of  Algal  Mats 	  5-163
                  5.10.2.7  Altered Nutrient  Availability 	  5-163
                  5.10.2.8  Interaction  of Stresses  	  5-163
          5.10.3  Biological Mitigation  	  5-164
          5.10.4  Summary  	  5-164
  5.11  References  	  5-165
                                           XX

-------
 Table of Contents  (continued)

                                                                                     Page

 E-6  INDIRECT EFFECTS ON HEALTH

      6.1  Introduction  	   6-1
      6.2  Food Chain Dynamics	   6-1
           6.2.1  Introduction 	   6-1
           6.2.2  Availability and Bioaccumulation of Toxic Metals	   6-2
                 6.2.2.1  Speciation (Mercury) 	   6-2
                 6.2.2.2  Concentrations and Speciations in Water (Mercury)  	   6-5
                 6.2.2.3  Flow of Mercury In the Environment	   6-5
                          6.2.2.3.1  Global  cycles	   6-6
                          6.2.2.3.2  Biogeoc hem leal cycles of Mercury  	   6-6
           6.2.3  Accumulation in Fish 	   6-10
                 6.2.3.1  Factors Affecting  Mercury Concentrations in  Fish  	   6-11
                 6.2.3.2  Historical and Geographic Trends in Mercury  Levels  in
                          Freshwater Fish 	   6-22
           6.2.4  Dynamics and Toxicity in Humans (Mercury) 	   6-24
                 6.2.4.1  Dynamics in Man (Mercury) 	   6-24
                 6.2.4.2  Toxicity in Man 	   6-25
                 6.2.4.3  Human Exposure from Fish and Potential  for Health
                          Risks 	   6-31
      6.3  Ground Surface and Cistern Waters  as Affected by Acidic Deposition  	   6-34
           6.3.1  Water Supplies	   6-34
                 6.3.1.1  Direct Use of Precipitation (Cisterns)  	   6-35
                 6.3.1.2  Surface Water Supplies 	   6-36
                 6.3.1.3  Groundwater Supplies	   6-40
           6.3.2  Lead 	   6-43
                 6.3.2.1  Concentrations in  Noncontaminated Waters	   6-43
                 6.3.2.2  Factors Affecting  Lead Concentrations
                          in Water,  Including Effects of pH 	   6-43
                 6.3.2.3  Speciation of Lead in Natural  Water 	   6-45
                 6.3.2.4  Dynamics and Toxicity of Lead in Humans 	   6-45
                          6.3.2.4.1   Dynamics of lead 1n humans  	   6-45
                          6.3.2.4.2   Toxic effects of lead on humans 	   6-46
                          6.3.2.4.3   Intake  of lead in water and  potential for
                                     human health effects 	   6-53
          6.3.3  Aluminum 	   6-57
                 6.3.3.1  Concentrations in  Uncontaminated Waters 	   6-57
                 6.3.3.2  Factors  Affecting  Aluminum Concentrations in Water  	   6-58
                 6.3.3.3  Speciation of Aluminum  in Water 	   6-58
                 6.3.3.4  Dynamics and Toxicity in humans 	   6-58
                          6.3.3.4.1   Dynamics of  aluminum in  humans  	   6-59
                          6.3.3.4.2   Toxic effects of aluminum in man  	  6-59
                 6.3.3.5  Human  Health  Risks from Aluminum  in Water 	   6-59
    6.4  Other Metals	  6-60
    6.5  Conclusions 	   6-60
    6.6  References	  6-63


E-7  EFFECTS ON MATERIALS

     7.1  Introduction  	  7-1
          7.1.1  Long Range  and Local Effects	  7-2
          7.1.2  Inaccurate  Claims of Acid Rain Damage to Materials 	  7-5
          7.1.3  Complex Mechanisms  of Exposure and Deposition 	  7-5
          7.1.4  Laboratory  vs Field  Studies  	  7-7
          7.1.5  Measurement of Materials  Damage	  7-7
                 7.1.5.1  Metals 	  7-7
                 7.1.5.2  Coatings	  7.3
                 7.1.5.3  Masonry  	  7-8
                 7.1.5.4  Paper and Leather  	  7-8
                 7.1.5.5  Textiles and Textile Dyes 	  7-8
                                         XXI

-------
Table of Contents (continued)

                                                                                    Page

 7.2  Mechanisms of Damage  to Materials	  7-8
          7.2.1  Metals  	  7-9
          7.2.2  Stone 	  7-10
          7.2.3  Glass 	  7-12
          7.2.4  Concrete  	  7-12
          7.2.5  Organic Materials  	  7-12
          7.2.6  Deposition Velocities  	  7-13
     7.3  Damage to Materials by  Acidic  Deposition  	  7-13
          7.3.1  Metals  	k.  7-13
                 7.3.1.1  Ferrous Metals 	  7-15
                          7.3.1.1.1   Laboratory Studies  	  7-18
                         7.3.1.1.2   Field  Studies  	  7-19
                 7.3.1.2  Nonferrous  Metals  	  7-23
                         7.3.1.2.1   Aluminum  	  7-23
                         7.3.1.2.2   Copper 	  7-25
                          7.3.1.2.3   Zinc  	  7-25
          7.3.2  Masonry 	  7-26
                 7.3.2.1  Stone  	  7-26
                 7.3.2.2  Ceramics  and Glass  	  7-30
                 7.3.2.3  Concrete  	  7-30
          7.3.3  Paint 	  7-31
          7.3.4  Other Materials  	  7-35
                 7.3.4.1  Paper  	  7-35
                 7.3.4.2 Photographic Materials  	  7-35
                 7.3.4.3  Textiles  and Textile Dyes 	  7-36
                 7.3.4.4  Leather 	  7-36
          7.3.5  Cultural  Property  	  7-37
                 7.3.5.1  Architectural  Monuments  	  7-37
                 7.3.5.2  Museuns,  Librarties  and Archives  	  7-37
                 7.3.5.3  Medieval  Stained  Glass  	  7-38
                 7.3.5.4  Conservation and  Mitigation Costs  	   7-38
     7.4  Economic Implications  	  7-40
     7.5  Mitigative Measures  	  7-42
     7.6  Conclusions 	  7-43
     7.7  References 	  7-44
                                         XXI1

-------
                      Acronym  and Abbreviation List

ADI (acceptable daily intake)                               E-6
AL (Aeronomy Laboratory,  NOAA)
6-ALA (s-aminolevulinic acid)                               E-6
ANC (acid neutralizing capacity)                            E-4
ARL (Air Resources Lab, NOAA)
ARS (Agricultural  Research Service,  DOA)
BCF (biconcentration factor)                               E-6
BLM (Bureau of Land Management,  DOI)
BLMS (boundary layer models)                                A-9
BM (Bureau of Mines, DOI)
BNC (base neutralizing capacity)                            E-4
BNC aq (aqueous base neutralizing capacity)                 E-4
BOD (biologic oxygen demand)
BS (base saturation)                                       E-4
BSC (base saturation capacity)                             E-4
BUREC (Bureau of Reclamation,  DOI)
BWCA (Boundary Water Canoe Area)
CANSAP (Canadian Sampling Network for Acid Precipitation)
CB (base cation level)                                     E-4
CDR (cation denudation rate)                                E-4
CEC (cation exchange capacity)                             E-2
CEQ (Council on Environmental  Quality)
CSI (calcite saturation index)                             E-4
CSRS (Cooperative State Research Service, DOA)
                                  xxm

-------
DOA (Department of Agriculture)
DOC (dissolved organic carbon)                              E-4
DOD (Department of Defense)
DOE (Department of Energy)
DOI (Department of Interior)
DOS (Department of State)
ELA (experimental lakes area)                               E-4
ENAMAP (Eastern North America  Model  of  Air Pollutants)
EPA (Environmental Protection  Agency)
EPRI (Electric Power Research  Institute)
ERDA (Energy Research and  Development Agency  (defunct)
ESRL (Environmental Sciences Research Laboratory, EPA)
FA (fulvic acid)                                           E-4
FDA (flourescein diacetate)                                 E-2
FEP (free erythrocyte protoporphyrin)                       E-6
FGD (Flue Gas Desulfurization)
FS (Forest Service, DOA)
FWS (Fish and Wildlife Service,  DOI)
GTN (Global Trends Network)
HHS (Department of Health  and  Human  Services)
ILWAS (Integrated Lake Watershed Acidification  Study)       E-4
LAI (leaf area index)                                      A-7
LIMB (Limestone Injection/Multistage Burner)
LRTAP (Long-Range Transboundary  Air  Pollution)
LSI (Langelier Saturation  Index)                            E-6
MAP3S (Multi-State Atmospheric Power Production
       Pollution Study)
                                  xxiv

-------
 MCPS  (Mesoscale convective precipitation systems)           A-3
 MOI (Memorandum of  Intent, U.S.-Canada)
 NADP  (National Atmospheric Deposition Program)
 NASA  (National Aeronautics and Space Administration)
 NATO  (North Atlantic Treaty Organization)
 NBS (National Bureau of Standards, DOC)
 NCAR  (National Center for Atmospheric Research)
 NECRMP  (Northeast Corridor Regional Modeling Program)        A-2
 NOAA  (National Oceanic and Atmosperic Administration,  DOC)
 NPS (National Park Service, DOI)
 NSF (National Science Foundation)
 NSPS  (New Source Performance Standards)
 NTN (National Trends Network)
 NWS (National Weather Service, NOAA)
 OECD  (Organization for Economic Cooperation and
      Development)
 OMB (Office of Management and  Budget)
 ORNL  (Oak Ridge National  Laboratory)
 OSM (Office of Surface Mining, DOI)
 PAN (peroxyacetyl  nitrate)                                  E-3,  A-5
 PBCF  (practical  biconcentration factor)                     E-6
 PBL (planetary boundary layer)                              A-4
 PGF (pressure gradient force)                                A-3
 PHS (Public Health Service)
RSN (Research Support Network)
SAC ($04 adsorption capacity)                                E-4
SEAREX
                                   xxv

-------
SAES (State Agricultural  Experiment Station,  DOA)
SCS (Soil Conservation Service,  DOA)
SURE (Sulfate Regional Experiment,  EPRI)
TFE (total fixed endpoint alkalinity)                        E-4
TIP (total inflection point alkalinity)                      E-4
TVA (Tennessee Valley Authority)
USGS (United States Geological  Survey,  DOI)
VOC (Volatile Organic Compounds)
WMO (World Meteorological Organization)
                                   xxvi

-------
            THE ACIDIC DEPOSITION  PHENOMENON  AND  ITS EFFECTS
                           E-l.  INTRODUCTION

                            (R.  A.  Llnthurst)

1.1  OBJECTIVES

     The basic and applied scientific  knowledge that can be gained
through the study of the acidic  deposition  phenomenon will undoubtedly
advance our understanding of emissions, transport, scavenging, and
deposition interactions.  This knowledge  is essential for a more
complete understanding of the causes of acidic deposition and for
defining the loadings of acidic  and acidifying substances that
ultimately interact with the ecosystem.   However, it is the perception
that acidic deposition may be harming  our natural and managed
environment that has stimulated  world-wide  interest.  As a result, the
effects and/or the potential effects of acidic deposition are the
primary motivation for public concern  and research activities now
designed to learn more about this  phenomenon.

     The objectives of the effects  portion  of this document are to
define the logic behind the concerns of potential effects, present the
support, or lack of support, for these concerns and draw conclusions
relative to the effects of acidic  deposition  based on the best available
evidence.  Special attention is  given  to  quantitative information on the
magnitude and extent of effects.   However,  it will become evident that
placing statistical confidence limits  on  the  data presently available is
difficult, and in most instances,  impossible.  A lack of quantitative
cause and effect data, in itself,  defines the state of knowledge in many
of the research areas.

1.2  APPROACH

     An ecosystem approach to the  acidic  deposition effects issues has
been used.  Figure 1-1 diagramatically presents a conceptual flow of wet
and dry deposition through a forested  system.  As most of the
terrestrial landscape is covered by vegetation, most acidic inputs to a
system pass through the canopy or  down the  stems of plants, to the soil;
and finally, over or through the soil  to  aquatic  systems, lakes and/or
streams, or into the ground water  system.   At any point along this
pathway, the chemistry of precipitation can be significantly altered.
As a result, the complexities of quantifying  effects in relation to a
chemical dose becomes increasingly  difficult.

     Direct deposition of acidic and acidifying substances to soils and
aquatic systems also occurs.  The  size of the receiving system of
interest, in relation to the size  of any  other ecosystem component which
may alter the deposition chemistry  prior  to contact, becomes important.
                                  1-1

-------
                                 INPUTS
GASEOUS
 OUTPUT
   ROOT
   TURNOVER
          LEACHING
        (biological export)
                                GEOCHEMICAL EXPORT
  Figure  1-1.   Conceptual diagram  of wet and dry deposition pathways  in
                an  ecosystem context  (from Johnson et  al.  1982.  The effects
                of  acid rain in forest nutrient status.   Water Res. Research
                18(3):449-461)
                                      1-2

-------
A common example of this concept is  lake  and  watershed  interactions.
Small lakes surrounded by large  watersheds are more greatly influenced
by those waters which pass through the  terrestrial landscape prior to
entering the lake;  since most of the water received is  from the
terrestrial pathway.  Thus,  the  effect  of the terrestrial system on
precipitation/deposition chemistry becomes a  variable which ultimately
defines the chemistry of the water entering the  aquatic systems via this
path.  If a lake is large in relation to  the  area it drains, direct
deposition to the lake surface becomes  increasingly important and the
terrestrial component of the system  plays a less important role.

     Having defined a representative flow path through  a system from a
chemical perspective, one must recognize  that any part  of the system
which alters the chemistry of precipitation can  be affected.  Thus, the
vegetation, the soil, and the waters may  be altered by  incoming wet and
dry deposition.  In addition to  these direct  alterations of the system
components, indirect effects can also occur.  Soils, for example, if
chemically altered, ultimately affect vegetation responses; soils being
the medium in which plants grow. If water chemistry is affected, the
biota in those waters are then subject  to change.  Subsequently, these
changes can be of significance to human health since both vegetation and
aquatic organisms are part of the human food  chain.

     This ecosystem perspective, with all its complexities and linkages,
should be kept in mind throughout the reading of the chapters.  The
concept of acidic deposition effects can  only be fully  understood with
this perspective in mind.  However,  for convenience of  presentation,
each major ecosystem component has been somewhat artificially separated
from the others and subsequently discussed in partial isolation from the
holistic approach.

1.3  CHAPTER ORGANIZATION AND GENERAL CONTENT

     Because soils  affect both vegetation and water, the effects of
acidic deposition on soils are discussed  first.  Secondly, vegetation
effects are evaluated from a more direct  influence perspective,
capitalizing on the knowledge of soils/nutrient  cycling, i.e., the
potential indirect  effects.   Next, the  water  chemistry  component of the
system is reviewed  from a watershed  perspective, continuing to rebuild
the ecosystem perspective.  Having defined the effects  of acidic
deposition on water chemistry, a discussion of aquatic  organism
responses to changing water  chemistry follows.

     Indirect effects on human health and a discussion  of acidic
deposition on materials, man's structures of  art and shelter, are also
presented.  Although manmade structures are not  part of the 'natural
ecosystem1 concept, they are certainly  a  part of our landscape and any
effects of acidic deposition on  them are  of concern.

     The general content of  the  chapters  is presented briefly below; in
the interest of establishing a general  sense  of  what will be found in
more detail in the  chapters  to follow.


                                 1-3

-------
1.3.1  Effects on Soil  Systems

     Soils are natural  integrators of ecosystem  structure and function.
They provide a pathway  for water  delivered  to aquatic systems or for
uptake by vegetation.   Therefore,  in  this chapter, emphasis is placed on
the natural  processes that contribute to acidification, nutrient status,
and metal movement in soils.  The effects of acidic and acidifying
substances on these natural  processes is then superimposed as an
additive factor, and their contribution to  these processes is examined.
Natural and managed systems are discussed separately.  Reversibility
concepts are presented  and predictions of changes over time are made
after making several assumptions.   These sections of the chapter are
chemically oriented and some basic soil chemistry is also included.

     Nutrient cycling aspects of  acidic deposition influences on soils
is the primary emphasis of the chapter.  Both the chemical and
biological components of this process are discussed in detail.  The
importance of changing  nutrient/metal  mobilization activity in soils is
discussed as it relates to both vegetation  response and water chemistry.
The soil  organisms, their role in  nutrient  cycling, and the potential
and measured effects of acidic deposition are also discussed.

     Soils are chemically and biologically  complex systems.  The effect
that acidic deposition  will  have  on such systems is dependent on
numerous variables.  Because of this  complexity  and the expectation that
potential effects may be long-term, the definitive conclusions one can
draw are not as numerous as some  might expect.

1.3.2  Effects on Vegetation

     Most of the terrestrial  landscape is covered by vegetation.
Because vegetation collectively includes the primary producers in the
food web, its importance to man is without  question.  Thus, any change
in plant productivity,  whether it be  an increase or decrease, can have
significant implications for man's food and fiber system.

     The material presented in the vegetation section discusses a
diverse range of acidic deposition/pi ant interactions.  These include
direct effects on the smallest scale;  i.e.  physiological and cell/leaf
response, to the gross  scale of forest and  crop  productivity.  The
potential effects of acidic deposition, plant, and environmental
condition interactions, leading to quantification of plant response, are
presented.  Special attention is  given to the concept of cumulative
effects on forests over time  and  the  lack of data in this field of
acidic deposition effects at the  present time.   The effects of
vegetation on deposition chemistry, as it passes through/over vegetation
to soils, is not discussed in detail.

     Plants are subject to more environmental stress factors than any
other component of the  system.  Their fixed position in the system
causes them to be exposed regularly to changes in air quality,
precipitation chemistry, soil physiochemical characteristics, disease


                                  1-4

-------
 influence, and climate, to which their limited avoidance/tolerance
 mechanisms may or may not be able to respond.  This immobility and
 dependence on air, soil, and water regimes of high variability makes it
 difficult to isolate single causes of response, whether they be
 beneficial or detrimental.  At the present level of understanding  of
 plant response as influenced by general  stress factors, the  direct and
 indirect effects of acidic deposition that can be definitely stated are
 extremely 1 imi ted.

 1.3.3  Effects on Aquatic Chemistry

     Most of the present concern for the potential effects of acidic
 deposition, and the significance of these effects, has been  derived from
 the aquatics literature.  As already noted, lakes and streams in an
 ecosystan are not isolated units.  They  are directly subject to acidic
 deposition inputs, but they are also dependent on the terrestrial  system
 buffering, or lack of buffering, of these inputs.  Unlike the longer
 term, chronic changes in soils and vegetative productivity,  evidence
 suggests that aquatic systems are responsive to both episodic shocks of
 acidity (e.g., during snow melt) and chronic inputs of acidic and
 acidifying substances over time.

     The discussion of aquatic chemistry is designed to deal  with  the
 complexity of processes that influence water quality and the relative
 importance of these processes/events. Because considerable  emphasis has
 been placed on aquatic resources in the  study of acidic deposition,
 rather lengthy discussions of methodology and historical  trends are
 relevant to drawing conclusions regarding impacts of acidic  deposition
 and are included.  These topics have been an important source of
 controversy and are therefore dealt with in detail  in this section.
 Predictive models, sensitive regions, significance of metals, and
 mitigative strategies are also discussed extensively.

     The data base for defining historical  changes in aquatic chemistry
 as a result of acidic deposition is among the strongest for  the
 ecosystem components discussed in this document.  Like any of the  other
 system components, however, predictions  of water quality require an
 understanding of a large number of other influencing variables, e.g.,
 soils.  Unfortunately, at this time,  our ability to predict  changes
 expected from acidic deposition is limited since predictive  models have
yet to be adequately validated.

 1.3.4  Effects on Aquatic Biology

     The emphasis of the aquatic biology chapter is placed on the
 response of aquatic organisms to acidification.   For the  most part,
 these discussions do not attempt to link the acidic deposition
 phenomenon to observed biological  changes,  but rather,  define the  link
 between biological response and acidification,  whatever the  cause.

     The chapter discusses the biota  found in naturally acidic  systems,
 recognizing that such systems have and will  always exist.  Such
                                  1-5

  409-262 0-83-2

-------
information proves useful  for comparing  naturally vs artificially
acidified systems and the  biota  that are  found in both.  The components
of the food chain in oligotrophic  water  systems most susceptible to
change are discussed relative to their importance and response to
acidity.  Benthos, macrophytes,  plankton  and  fish are included.
Organisms which are dependent on aquatic  systems, for at least a portion
of their life cycle, are also discussed.   Mechanisms of response, field
and laboratory evidence for changes in aquatic biota resources, and
biological mitigation options are  also presented and evaluated.

     Although predictions  of species  survival as a  function of water
quality are feasible, the  limited  resource inventory and lack of
predictive chemistry models inhibits  quantification of the magnitude and
extent of acidic deposition impacts on aquatic resources.
Quantification of direct impacts of acidification is most likely for the
higher trophic levels, e.g., fish, especially as better resource
inventories become available.  However,  the effects of acidification on
the interactions between trophic levels  remain unclear at this time.

1.3.5  Indirect Effects on Health

     Limited data is available on  the  potential and known effects of
acidic deposition on human health. Food chain dynamics are discussed  in
a bioaccumulation context.  Particular emphasis is  placed on aquatic
organisms of importance to man,  and drinking  water  from ground, surface,
or cistern systems.  Those metals  suspected as being influenced by
acidity are highlighted.  These include  mercury, lead, and aluminum.

     Although the acidic deposition oriented  'toxicity data base1, is
somewhat limited, the authors have capitalized on the extensive toxicity
literature and research in other fields of science.  Superimposed on
these concepts is the effect of  acidification, and  conclusions are
drawn.

1.3.6  Effects on Materials

     Like the natural ecosystem, materials, both natural and manmade,
are subject to many environmental  influences.  Among them are the
effects of acidic and acidifying substances.  This  chapter of the
document reviews the rather limited data  available  on the specific topic
of acidic deposition effects, as defined  in this document, and discusses
the major building and construction materials that  might be affected by
acidic deposition.  Mechanisms of  damage,  economic  implications, and
mitigative measures are presented  and evaluated.  The importance of dry
deposition over wet deposition is  highlighted.

1.4  ACIDIC DEPOSITION

     The previous sections refer to acidic deposition without
definition.  Volume I, Chapter A-l defines this term for technical use
in the atmospheric/deposition sciences.   However, from an effects point
of view, the chemical quality of precipitation is as, if not more,
important than the pH.  Deposition, both  wet  and dry, contains both


                                  1-6

-------
essential and nonessential substances needed by ecosystems  as  part of
their natural nutrient cycle.  Therefore,  the materials  presented  In  the
effects chapters concentrate on the generic concept of acidification  and
the Importance of sulfate and nitrate loadings to  the ecosystem.
Whether these substances are deposited in  dry or wet form is not
differentiated.  Because the inputs of sulfur and  nitrogen  can be  acidic
upon delivery, or can become acidifying as they cycle through  the
system, these substances are the critical  elements for discussion.
Because the data bases were not sufficient to conclusively  define  input
limits for 'protection'  of biological  resources, there was  no  need  to
deal with a separation of wet and dry forms of deposition.  When
simulated treatments are involved, differentiation of deposition forms
is noted as necessary; e.g., in the crop productivity discussion.
Although an effort to separate the components of deposition was not
undertaken, this does not minimize the potential for differential
effects of wet vs dry deposition exposures.

     Therefore, reference to acidic deposition will  refer to total
deposition of acidic or acidifying substances.  Differentiation is  made
only as deemed appropriate by the authors  on an issue-by-issue basis.

1.5  LINKAGE TO ATMOSPHERIC SCIENCES

     Every effort to use information from  the atmospheric chapters  of
the document was made.  Reference to deposition changes  over time,
emissions levels, natural  vs anthropogenic sources of sulfur and
nitrogen, and/or sulfur and nitrogen loadings are  consistent with  those
presented in Volume I.  Any conclusion which would have  been drawn  using
data not consistent with the atmospheric/deposition chapters was
modified or removed.  Therefore, Volume I  appropriately  sets the stage
for the levels of acidity/deposition,  the  'cause', that  was considered
in the development of the effects presentations.  References to chapters
in Volume I are made, as necessary.

1.6  SENSITIVITY

     In addition to problems of interpreting the meaning of the acidic
deposition concept, other terminology  is equally subject to
misinterpretation.  In particular, the term 'sensitivity' lends itself
to varied interpretations.  Sensitivity, as used in the  effects
chapters, refers to the relative potential  for changes to occur within
an ecosystem or one of its components.  A  highly sensitive  portion  of an
ecosystem will change more noticeably, or  rapidly, to acidic inputs than
will  one that is generally classified  as having moderate, low, or no
sensitivity.  However, the reader must be  cautious in many  of  the
effects areas to be certain the reference  to sensitivity is clear.  For
example, reference to a sensitive soil is  not meaningful.   Acidic
deposition effects must be considered  with respect to a  specific
physiochemical property of the soil.  Soil-metal mobility or pH, for
example, can be classified as 'sensitive1  to change.   Likewise,
particular tree species, aquatic organisms, processes, and/or  materials
can be sensitive to change due to acidic deposition.   However,


                                  1-7

-------
developing sensitivity classifications  for  larger units of the ecosystem
can be misleading,  and comparing  dissimilar ecosystem components, e.g.,
soils and fish, is  inappropriate.  In addition, quantification of
'sensitivity1  is defined in  the aquatic chemistry chapter but only
qualitative relative usage of the word  appears in discussions of other
ecosystem components.

1.7  PRESENTATION LIMITATIONS

     A phenomenon as complex as acidic  deposition cannot be presented
with respect to every environmental  factor  that might influence
ecosystem response.  In the  discussions that follow, it is recognized
that acidic deposition is treated as if it  were isolated from other
pollutants with which it might interact.  Thus, not every possible link
between the ecosystem and influencing variables has been considered.
What is presented is the authors'/editors'  perspective of key issues.
This does not infer that other issues are unimportant.  Rather, an
absence of discussion suggests that the issue has not, as yet, been
recognized as essential to our understanding or that data to support any
relevant comments were lacking.
                                  1-8

-------
            THE ACIDIC DEPOSITION PHENOMENON AND ITS EFFECTS
                     E-2.  EFFECTS ON SOIL SYSTEMS

         (W. W. McFee, F. Adams, C. S. Cronan,  M.  K. Firestone,
              C. D. Foy, R. D. Harter, and D. W. Johnson)!

2.1  INTRODUCTION

     Soil plays a key role in ecosystems.   It is one of their most
stable components and, when combined with  climate, defines  a terrestrial
ecosystem's productivity limits.  Moreover,  because much of the  water
entering streams and lakes directly contacts soil, soil  properties also
exert important influences on aquatic systems.

     Because of soil's importance to most  ecosystems,  the impact of
acidic deposition on soils assumes prominence in our discussion.
Defining soil sensitivity to acid inputs depends on understanding soil
properties and chemistry, which are discussed early in this chapter.
Thereafter, we can locate vulnerable soils and determine expected and
potential effects on various soil  components.  Types and rates of
changes can be determined, and the effects of soil changes  on aquatic
and terrestrial ecosystems can be considered.  Specifically,  questions
concern impacts on soil fertility; nutrient, toxic substance,  and
organic acid availability; plant vitality; and  water quality.  Both
short and long-term implications must be considered in relation  to
numerous soil components, to soil-piant relationships, and  to soil-water
relationships.

2.1.1 Importance of Soils to Aquatic Systems

     Aquatic systems receive diverse outputs from terrestrial
ecosystems.  Influences of acidic deposition on transfers from
terrestrial to aquatic systems may be direct, when material  deposited
from the atmosphere passes over or through the  soil  with little
interaction, or they may be indirect, when deposited materials cause
changes in soil processes, such as weathering,  leaching, and organic
matter decomposition.  Thoroughly assessing effects of atmospheric
deposition on any element transferred from a terrestrial  to an aquatic
system requires extensive measurements of  inputs,  internal  processes,
and outflows (Gorham and McFee 1980).  These authors note that our
understanding of the processes is rather incomplete.

2.1.1.1  Soils Buffer Precipitation Enroute to  Aquatic Systems—Soil
systems are generally strongly buffered against changes  in  pH.   They  are
usually thousands of times more resistant  than  water to pH  shifts (Brady
lAll  of these authors have contributed to this  chapter.   Because of
 subsequent integration of the material,  these  authors  are  not
 identified by section.
                                 2-1

-------
1974).  Therefore,  pH of deposited precipitation  tends  to  shift  toward
that of the soil  if the water comes into intimate contact  with the  soil.
The cation exchange capacity (CEC) of the soil  and the  extent of its
saturation with basic cations (e.g.,  Ca2+,  Mg2+,  K+)  determine the
soil buffering capacity in moderately acid soils  (see Section 2.2).
Strongly acid soils may be buffered by the soil minerals.   In general,
soils with high clay content, especially smectite clays, and  with high
organic matter content are strongly buffered.   These  soils tend  to
deliver water that has come in intimate contact with  the soil matrix  to
aquatic systems at or near the soil  pH.   In areas with  alkaline,
neutral, or slightly acid soils,  the soil buffer  system removes  much  of
the acidity in acidic deposition.   Where the soils are  near the  acidity
of the incoming precipitation, they may not change the  pH  of  water  as it
passes through, especially if the soil  solution remains rather dilute.

2.1.1.2  Soil as a Source of Acidity for Aquatic  Systems--Many of the
soils in the world's humid regions have been acid for very long  periods.
Bailey (1933) pointed out that podzol soils (soil  order Spodosol) were
generally the most acidic, followed by lateritic  (Oxisols  and Ultisols)
and podzolic (Ultisols and Alfisols)  soils.  He did not consider organic
soils (Histosols), many of which  are quite acid.   For example, all  of
those designated "Dysic" at the family level of classification have a pH
less than 4.5, and some have a much lower pH (Soil Survey  Staff  1975).
Drainage waters from such acid soils may be equally acidic as the soil
and essentially control the pH of receiving lakes or  streams.  In many
cases, however, after percolating water passes  through  acid soil, it
interacts with more basic materials underneath  before reaching a stream.
Thus, a lake may be surrounded with surface soil  considerably more  acid
than the water.  Such is the case around many lakes in  the Adirondack
mountains where most of the soils are strongly  acid (Galloway et al.
1980).

2.1.2  Soil's Importance as a Medium for Plant Growth

     All of the other roles of soil fade into insignificance  when
compared to its importance as a medium for plants. Soil provides the
physical support, most of the water, nutrients, and oxygen needed by
plant roots for normal growth and development.   Well  over  95  percent  of
our food and much of our fiber come directly or indirectly from
terrestrial plants.  Soil properties set limits on the  productivity of
terrestrial ecosystems.  Even though soils tend to resist  rapid  change,
any significant reduction in their ability to support plants, such  as
the increased Al toxicity cited by Ulrich et al.  (1980) and A. H.
Johnson et al. (1981), is a serious matter.

2.1.3  Important Soil Properties

     Any changes deleterious to the soil's role as a  plant growth medium
or  that alter its output to aquatic systems are causes  for concern.
These include chemical changes, such as in acidity, nutrient  supply,
cation exchange capacity, leaching rates of nutrients,  or  mobilization
of toxic substances; physical changes, such as accelerated weathering
                                  2-2

-------
 rates  or  changes  in  aggregation; or biological changes, such as
 reductions  in nitrification or other processes.

 2.1.3.1   Soil Physical Properties--Soil physical properties are never
 independent of chemical and biological properties; however, water
 movement, water retention/storage capacity, and soil aeration are
 determined  primarily by physical properties.  Controlling water flow is
 the most  important influence of soil physical properties on interaction
 of soil with acid rain.  Soils that have high surface runoff rates, such
 as those  on  steep slopes or with low porosities, tend to transmit water
 rapidly without changing its composition.  Likewise, if the soil has
 many coarse  pores and is well drained, as are many sands and loamy
 sands, water passing through may be changed only slightly.  Therefore,
 if the primary consideration is protection of a body of water by the
 soil's buffering capacity, the two situations described are "sensitive."
 On the other hand, if changes in the soil itself are the concern, these
 soils  are not particularly sensitive from the physical standpoint.

 2.1.3.2  Soil Chemical Properties--Resistance of soil chemical
 properties to the effects of acidic deposition is measured in terms of
 the buffering capacity, initial pH, sulfate adsorption capacity, and
 amount and type of weatherable minerals.   Soils with high buffering
 capacities due to high CEC and high base status will be very slow to
 respond to acid inputs of the magnitude acidic deposition introduces.
 Weatherable minerals containing carbonates are common in lower horizons
 of the younger soils in many regions and will effectively neutralize
 acids  from all sources.  Details of these relations are discussed in
 later sections.

 2.1.3.3  Soil Microbiology—Biological  processes in soils may be
 influenced by acid precipitation and,  at the same time,  provide some of
 the means of resistance and/or recovery.   If important soil  biochemical
 processes, such as N fixation,  nitrification, organic matter decay, and
 nutrient release are changed by acid precipitation, the  impact  could be
 significant.  Studies of relationships of soil acidity to biochemical
 activity are plentiful.  However,  most have doubtful  applications to the
 acid rain problem because they were studies of natural  pH differences,
 not of shifts due to acid inputs.   A few recent studies  indicate
 alteration in microbial activity near the soil surface due to simulated
 acid rain (Strayer and Alexander 1981,  Strayer et al. 1981).  The
 capacity of most soils to buffer acid inputs as well  as  the diversity
 and adaptability of microbe in  the soil  contribute to resistance to acid
 rain effects.  A more complete discussion of soil  biology and acidic
 deposition follows in Section 2.4.

 2.1.4  Flow of deposited materials through  soil  systems

     A generalized depiction  of the flow  of deposited materials  through
a terrestrial ecosystem is  shown  in Figure  2-1.   In a forested  ecosystem
 (to a lesser degree on  cropland, also),  a major  portion  of the
precipitation is intercepted  by foliage.  The chemical properties of the
resultant throughfall  and stemflow can  be substantially  altered  from the
                                  2-3

-------
    •^v
    *r
                          INTERCEPTION
DIRECT DEPOSITION
                                                 SURFACE FLOW
                                                  =>—==—-—'

                                                  Minimum to

                                                  Moderate  soil

                                                  interaction
                                           CHANNELIZED  FLOW

                                       Minimum  soil  interaction
GROUNDWATER FLOU
                  DIFFUSION FLOW

              Maximum soil  interaction
                                  IMPERVIOUS ZONE —
                                                           - _ -
                                                        / __ i   k  ,
Figure 2-1.   Flow  paths of precipitation through a terrestrial  system.
                                  2-4

-------
incipient precipitation (see Section 3.2.1.2).   While this alternation
may be of no importance in constructing the total  system input-output
balance, it has a big impact on the nature of reactions expected at the
soil surface.

     Upon striking the surface, the water may infiltrate the soil  or
move laterally as surface flow.  In a forested ecosystem, surface flow
will usually not be visible on the forest floor but will  flow through
the surface organic layers.  This provides opportunity for water to
react chemically with surficial materials to a greater extent than does
surface flow in cultivated areas.  The amount of interaction will  be
proportional to path length and flow rate.

     In uncultivated areas, many large channels are established by
burrowing animals and decomposing roots.   These are frequently open to
the surface and provide open conduits for flow of drainage water.
These channels may carry nearly all drainage water during saturated
flow, and may be dominant conduits during all  rainfall  events.   Little
opportunity for soil interaction is provided,  and the precipitation may
be conducted through the soil  with little or no alteration.

     Water movement by unsaturated flow will  usually be through the
capillary pores where maximum opportunity exists for interaction with
the soil.  This is the major source of water to plants.   Flow through
fine pores is necessary in many deeper soil layers that have limited
macropore space.  The various flow paths are depicted in  Figure 2-1.

2.2  CHEMISTRY OF ACID SOILS

     A brief discussion of important concepts  in the chemistry of acid
soils is presented here as background for understanding the sections
that follow.  Those already familiar with these concepts  may wish  to
proceed to Section 2.3.

     Although little is known about the impact of acidic  deposition per
se on soils, much is known about acid soils in  general.   The factors
which determine the natural acidification of soils are important to the
development of an adequate comprehension of recent and/or future acidic
deposition impacts.  There are many acid soils  in the United States, and
it is appropriate to capitalize on our understanding of these systems.

2.2.1  Development of Acid Soils

     The eastern half of the United States has  a climate  in which  rain-
fall exceeds the combined losses of water by  runoff, evaporation,  and
transpiration from the soil.   The excess  water  leaches  through  the soil,
carrying with it basic cations and other soluble materials.   If leaching
removes basic cations faster than they are replenished  by natural
processes or human activities, the soil  profile becomes  increasingly
acid and impoverished of nutrients (Pearson and Adams  1967).   However,  a
prerequisite for leaching to cause soil  acidity is the  addition of H+
ions to the system (Bache 1980,  Ulrich 1980) along with mobile  anions.
The H+  ions can be donated from a variety of  sources.
                                 2-5

-------
2.2.1.1  Biological  Sources of H+ Ions— Al though  H+  ions may  be
generated by chemical  weathering of minerals  through hydrolytic
reactions, the significant sources of H+  production  in  soils  are  all
based on biological  reactions.

     Oxidation of sulfur and sul fides can be  important  natural  sources
of acidity.  Much of the sulfur in soils  is present  in  a highly reduced
state.  This includes combined S in soil  organic  matter and such  common
minerals as pyrite,  FeS2.  The release of sulfur  from organic-matter
in aerobic soils is  followed by the H+-producing  oxidation reaction

     S + 3/2 02 + H20 = $042- + 2H+.

Elemental S is sometimes used in agriculture  for  disease control  and  as
a fertilizer material.  Its contribution  to soil  acidity is readily
calculable from the  equation above, i.e., 16  kg of S per hectare  is
equivalent to one hundred cm of pH 4.0 precipitation, 1 keq H+  ha'1.

     When sul fide minerals, e.g., pyrite, are exposed to atmospheric
oxygen, oxidation of these minerals results in significant H
production, according to the reaction

     2FeS2 + 7H20 +  7 1/2 02 = 2 Fe(OH)3+ 4S042-  + 8H+.

Significant quantities of sul fide minerals are found only in  recently
exposed soil materials or those that have been maintained in  anaerobic
conditions, e.g., coastal marshes.  Therefore, their influence  is
important in only very limited areas.

     Acidity from nitrification is an important contribution  in most
soils of the humid regions.  Nitrogen is  one of the  most abundant
elements in plants and in soil  organic matter and is present  mostly in  a
highly reduced state.   It is released from organic matter as  NH3,
which hydrolyzes to  NH4+ in soil  solution.   Much  of  the NH4+  is
oxidized to nitrate  by bacteria,  according to the reaction

     NH4+ + 202 = NOa- + 2H+ + H20.
By this reaction, 9 kg NH4+ ha-1 could produce 1  keq  H+  ha-1.
The theoretical  maximum acidity from nitrification  is never realized in
soils because concurrent or subsequent reactions  involving N neutralize
a portion of the H+ produced.

2.2.1.2  Acidity from Dissolved Carbon Dioxide—Atmospheric C02
contributes some acid to soils, however,  the respiratory activities  of
plant roots and soil microbes  result in soil  air  containing considerably
more C02 than atmospheric air.   Soil air commonly contains up to 1
percent CO?. in comparison to  the 0.033 percent of  a  normal  atmosphere
(Patrick 1977). This C02 lowers the pH of pure water  according to the
equation below, which can be derived from the relationships among C02
content of air, dissolved H2C03, and H+ activity.
                                  2-6

-------
                (H+)  = [1.50 x 10-10  x  % c02]l/2.

If atmospheric C02 is 0.033 percent,  then  H+  activity of  rainwater
is 2.2 X 10-6M (moles per liter)  or pH  5.65.   If  soil air contains 1.0
percent C02, then H+  activity is  1.2  X  1Q-5M  or pH 4.91.  Thus,
biologically generated C02 is a source  of  H+  ions in soils but has
very little influence below a pH  of about  5.0.

     The dominant source of H+ in many  soils  used in nonleguminous
agricultural production in the United States  is from the  use of
ammoniacal  fertilizers, e.g., NH3, NfyNOs,  (NHa^CO, NH4H?P04,
(NH/^HPOa, and (Nfy^SO^  Because nitrogen  is often used at
rates of TOO to 200 kg ha-1, fertilizers alone may generate H+ in
soils at rates of 3.6 to 21.6 keq H+  ha-1.   it should be  noted that
the net acidification from ammoniacal N is frequently less than the
theoretical due to direct uptake  of NH4+ by  plants and H+
consuming reactions in soils.  Although these calculations are based on
fertilizer  application to agricultural  lands,  these same  relationships
are applicable for determining the acidification  impact on soils from
atmospheric N sources.  Nitrogen  additions contribute to  acidification
by increasing basic cation removal in plants  harvested and by furnishing
a mobile anion, N03~, for leaching losses.

     Acidity is also  added from soil  organic  matter.  The microbial
process by  which plant residues are converted into soil humus generates
many carboxyl ligands, RCOOH, on  the  humus.   The  protons  of such ligands
partially dissociate, adding H+ to the  soil  solution.  This source of
H+ production becomes increasingly important  when large amounts of
soil humus  are present.

     Roots  can absorb unequal amounts of anions and cations because the
uptake mechanisms are relatively  independent  of each other.  The
electroneutrality of the soil solution  is  maintained by plant release of
H+ or HC03- during the uptake process.  Plants with N-fixing
rhizobia absorb more cations than anions from the soil when N is
obtained almost entirely from ^ High yielding  legumes  may produce
H+ equivalent to several hundred  kg CaC03  per hectare (several keq
H+ ha'1).

2.2.1.3  Leaching of Basic Cations—Production of H+ resulting from
the various mechanisms does not produce acid  soils unless it is
accompanied by leaching.  In the  absence of leaching (arid and semi-arid
regions), HC03~ tends to accumulate in  soil  solution, leading to
H+ neutralization and precipitation reactions with Ca.  In the
presence of leaching, H+ in the soil  solution replaces some of the
adsorbed basic cations (Ca, Mg, K) on the  exchange surfaces of soil
particles.   As the excess soil solution moves downward through the soil
profile, it carries basic cations equivalent  to its anionic content.
Meanwhile,  the adsorbed H+ remains in place with  the soil particles,
causing the soil to become more acid.
                                  2-7

-------
2.2.2  Soil  Cation Exchange Capacity

     Many differences in the sensitivity  of soils  to  acidic  inputs can
be traced to the extent of base saturation  and  to  differences in cation
exchange capacity (CEC), the sum of the exchangeable  cations, expressed
in chemical  equivalents, in a given quantity of soil.   It  is the major
characteristic of soils that prevents them  from becoming rapidly
impoverished when leached.  This section  is presented to explain the
source of CEC and some of the variables which affect  it.

2.2.2.1  Source of Cation Exchange  Capacity in  Soi Is—To have a CEC,
soil  particles must have a net negative charge.Soil  clay particles may
have a negative charge due to isomorohous substitution of  AP+ for
Si4+ in tetrahedral  layers and of Mg*+ or Fe2+  for A13+ in
octahedral layers of the clay structure.  This  charge is termed a
"permanent charge" (Coleman and Thomas 1967).   A second mechanism is the
result of the terminal metal  atom's reaction with  water to complete its
coordination with either OH"  or h^O.   At  low pH, the  coordinating
ligand tends to be H20, which results in  a  site with  a positive
charge; at high pH,  the coordinating  ligand tends  to  be OH~, which
results in a negatively charged site.  Minerals with  this  kind of
negative charge as their primary source of  CEC  are referred  to as having
a "pH-dependent charge."  Therefore,  these  soil  particles  change CEC as
the pH changes.

     In most soils,  a significant component of  the CEC comes from
organic matter.  The major portion  of soil  humus is associated with the
clay fraction, except in extremely  sandy  soils  (Schnitzer  and Kodama
1977).  Its pH-dependent CEC is a major component  of  the CEC of surface
soils and may be almost the sole source of  CEC  in  sandy soils.  Soil
humus has many ligands from which protons dissociate,  such as carboxyl
(-COOH), phenol (-OH), and imide (-NH).  In acidic soils,  however, only
the carboxyl ligand ionizes enough  to affect pH, i.e., R-COOH ->
R-COO- + H+, creating a negatively  charged  exchange site.  The
fraction of H+ that ionizes from carboxyl ligands  increases  with
increasing pH, thereby increasing soil  CEC.

     The CEC of surface soils is determined by  their  clay  and organic
matter contents.  In the highly weathered Ultisol  soils common to the
Southeast, surface-soil clays are usually kaolinite and hydroxy-Al
intergrade vermiculite.  These soils  contain a  high percentage of sand
and low contents of clay and organic  matter, and commonly  have a CEC of
about 5 meq 100 g~*.  In soils with a more  temperate  climate in the
eastern half of the United States,  soil organic matter is  usually higher
and smectite clays are sometimes more abundant, hence the  CEC is
normally higher, about 15 meq 100 g"1 (Coleman  and Thomas  1967).

2.2.2.2  Exchangeable Bases and Base  Saturati'on--The  exchangeable
cations in acid soils consist primarily of  Ca,  Mg, K, Al,  H,  and Mn.
The basic cations are Ca, Mg, and K,  while  Al and  H are measures of soil
acidity.  The fraction of the CEC that is satisfied by basic  cations is
defined as "base saturation."  For  a  particular soil  and CEC method, a


                                  2-8

-------
well-defined, positive correlation between pH and base saturation
exists.  Unfortunately, the CEC reported in the literature is method
dependent.  The most common methods of determining CEC are (1)  sum of
exchangeable cations by neutral salt extraction, (2)  NH4+ adsorption
at pH 7.0, (3) Na+ adsorption at pH 8.2, and (4) sum  of exchangeable
cations by neutral salt extraction plus titratable acidity by
triethanolamine at pH 8.0.  The most commonly used method is probably
1.0 N NH4OAc extraction at pH 7.0, method (2) above.   For soils with
simiTar characteristics, pH can be used as a reasonable estimate of base
saturation.  For example, the "soil pH" - "base saturation" relationship
of 111tisols in Alabama is similar to the combined relationship of
Alfisols, Inceptisols, and Spodosols in New York (Figure 2-2).

     Analogous to the base-saturation concept, quantities of individual
exchangeable cations can be expressed in terms of saturation of the CEC.
This concept is particularly useful in defining the relative
availability of cations.  The cation-saturation concept is also useful
in predicting probable toxic levels of Al.   Although  Al  phytotoxicity is
a function of soil-solution Al  activity, it is more convenient to
measure exchangeable Al.

2.2.3  Exchangeable and Solution Aluminum in Soils

     Aluminum mobility is a key area of concern for both aquatic impacts
and terrestrial vegetative response relative to acidic deposition.   The
soluble Al in soils is a product of acid weathering of clay minerals.
As H+ concentration increases in soil solution, the stability of clay
minerals decreases, resulting in the release of A13+  ions from their
surface structure.  Measurable amounts of soluble Al  are found only at a
pH less than 5.5.   Only a small portion of the dissolved Al  resides in
the soil solution.  Most becomes exchangeable, since  cation-exchange
sites in soils have a strong affinity for A13+ ions.

     Even though Al saturation of strongly acid soils (pH < 5.0)  will
normally exceed 50 percent of the CEC, the concentration of Al  in soil
solution is usually < 1 ppm.  The significance of exchangeable Al  is
two-fold: (1) it is the major component of exchangeable acidity in soils
(i.e., acidity displaced by a neutral-salt solution), and (2)  it is the
source for the immediate increase of Al  into soil-solution from an acid
soil  when replaced by other cations on the exchange sites.

     Soil-solution Al  concentration is determined by  the pH dependent
solubility of Al-containing clay minerals.   For example, kaolinite
dissolves according to the reaction

     Al2Si205(OH)4 + 6H+ = 2A13+ + 2Si(OH)4 + H20.

Thus, soil-solution Al  concentration will  be determined  by the
activities of H+,  Si(OH)4, or other products of weathering
reactions.

     Aluminum oxides are common in acid soils, and  it is frequently
assumed that solution Al  is controlled by A1(OH)3  solubility.   In  that


                                  2-9

-------
    8.0



    7.0



    6.0



 2  5.0
 to
 ID
 O
 UJ
 Z3
 cr
    4.0
3.0
    2.0
    1.0
      0
       0     10     20     30    40     50     60     70     80    90    100

                        PERCENT  OF  BASE  SATURATION
Figure 2-2.  Typical  relationship of soil  pH to the percent base
             saturation.   Adapted from Lathwell and Peech (1969).
                                  2-10

-------
case, A13+ activity in soil solution is a function only of pH because
of the reaction

     A1(OH)3 + 3H+ = Al3+ + 3^0.

The equilibrium log K for this reaction, log A13+ - 3 log H+, varies
from 9.7 for the amorphous oxide to 8.0 for crystalline gibbsite.   At pH
5.0, for example, A13+ activity would vary from 20 yM for the
more-soluble amorphous oxide to 0.1 yM for gibbsite at equilibrium
with the soil solution.

     In most acid soils of the United States, clays are primarily
aluminosilicates, and solution Al is controlled by soil-solution Si  as
well as pH.  When both Al and Si are present in soil  solution,  their
activities frequently depend upon a solid-phase component with  the
general  composition of A^SigOjjfOH)^  Its solubility in acid
soils is expressed by the equation
     l/2Al2Si205(OH)4 + 3H+ = A13+ + Si(OH)4 + 1/2H20.

The equilibrium log K for this reaction,  log A13+ + i0g Si (OH)  -  3 log
H+, varies from 5.6 for amorphous halloysite to 3.25 for crystalline
kaolinite.  If Si(OH)4 in soil solution is 0.2 mM (a reasonable value
for acid soils), then Al3+ activity at pH 5.0 would range from  2  yM
for amorphous halloysite to 0.01 yM for crystalline kaolinite at
equilibrium with the soil solution.

     The relative solubilities of Al oxides and aluminosilicates  in
soils show that soil-solution Al3+ activity, at the same pH,  varies
according to the solubility of the Al -control ling mineral  as  follows:
amorphous Al oxide > amorphous halloysite > gibbsite >  kaolinite  >
smectite.  Consequently, the level of soil-solution Al ,  and its
phytotoxic effect on plants or its transport to aquatic  systems,  varies
among soils at the same pH, depending upon which mineral  is controlling
solution Al .

     Under nonagricul tural ecosystems, soils generally  contain  too
little solution phosphorus (P) to affect  soluble Al.  However,
fertilizer P is an effective agent for lowering solution Al by  forming
such insoluble precipitates as variscite,  A1(OH)2H2P04.   Dilute,
acid solutions of Al  react with sul fate to form insoluble compounds but
these compounds will  be the controlling  factor very  infrequently.   The
influence of Al  and Mn on plant nutrition is discussed  in Section
2.3.3.3.

     In the presence of organic ligands,  the solubility  of aluminum can
be greatly enhanced (Lind and Hem 1975).   Numerous reports emphasize the
importance of polyphenols and other components of soil organic  matter  in
the transport of Al  within soils (Bloomfield 1955,  Davies et  al.  1964,
Malcolm and McCracken 1968).   In many  cases,  organic-aluminum complexes
are the major form of mobile Al .   Cronan  (1980b,c)  points out the
importance of organic substances in  Al  leaching and discusses the
changes likely when strong acid anions such as sul fate are present.
                                  2-11

-------
     Inorganic aluminum is present in acid soil  solutions  primarily  as
monomeric ions, the most common ones being Al3+,  A10H2+,
A1(OH)2+, Al(OH)3o, A1S04+ and A1H9POA2+.   In  most
acid soils, A1(OH)2+ is the most abundant  solution ion.

     Since about 1920 soluble Al  has been  recognized  as an  important
factor limiting plant growth in acid soils (Adams and Pearson  1967).
Because of the pH-dependent solubility of  Al,  phytotoxic levels of
solution Al can be expected in most mineral  soils when soil  pH is <  5.0
to 5.5.  Only a fraction of a ppm is needed for  sensitive  species to
exhibit symptoms (see Section 2.3.3.3.2.1).

2.2.4  Exchangeable and Solution Manganese in  Soils

     Another result of acidification is associated with the mobility of
manganese.  Manganese occurs in soils in three valency states.  Since
divalent Mn (Mn2+) is the most soluble form, Mn  availability depends
upon the redox potential of the system.  The equilibrium between Mn
oxides and solution Mn2+ is subject to rapid  shifts in the  soil.

     In most soils with significant levels of  easily  reducible Mn, toxic
levels of Mn2+ in soil  solution can be expected  when  soil  pH is < 5.5.
The lower the pH, the more likely phytotoxicity  will  occur.  Lower redox
potentials favor Mn-oxide dissolution.  In turn,  lower redox potentials
are favored by waterlogged conditions, particularly when accompanied by
the rapid decomposition of organic matter. Consequently,  over the
short-term, toxic levels of Mn are more likely under  poorly aerated
conditions.  A long-term consequence of poor aeration, however, is the
depletion of easily reducible Mn and soluble Mn  to quite low levels
through leaching.

     It is normal for Mn and Al phytotoxic symptoms to occur
concurrently in many acid soils because the pH-dependent solubility  of
Mn oxides and the Al-containing soil  minerals  release toxic levels of Mn
and Al at about the same pH level, i.e., < pH  5.0 to  5.5.   Whereas Al
phytotoxic symtoms are not generally evident on  aerial  plant parts,
symptoms of Mn phytotoxicity are quite severe  before  plant growth is
affected significantly.

2.2.5  Practical Effects of Low pH

     Low soil  pH influences most chemical  and  biological reactions.   It
accelerates mineral weathering and the release of phytotoxic ions to the
soil solution; it affects the downward migration of clay and
organic-matter particles in the soil-profile development process, and it
affects the level and availability of most plant nutrients  in  the
soil-solution.

     The solubility of soil minerals at low pH is important to plant
growth and transport of ions to aquatic systems.   The common Al minerals
or compounds in acid soils are the aluminosilicates,  hydrated  oxides,
                                  2-12

-------
phosphates, and hydroxy-sulfates.  The relationship of low pH to Al and
Mn solubility was covered in Sections 2.2.3 and 2.2.4, and their
influence on plant nutrition is covered in Section 2.3.3.3.

     Low soil pH affects the availability of all macronutrients (N, P,
S, Ca, Mg, K) to some extent (Adams and Pearson 1967; Adams 1978;
Rorison 1980).  These effects, however, are seldom great enough to
influence plant yields.  Nitrogen availability is affected because low
pH decreases the rate at which organic matter decomposes and releases N
to the soil solution.  Phosphorus availability is affected primarily via
chemical solubilities.  At low pH (< pH 5.5), P is made increasingly
less available because of its reaction with Al and Fe.  Sulfate
availability is determined by both organic-matter decomposition and by
inorganic reactions with Al and Fe.  The result of these effects is that
sulfate becomes progressively less available as pH decreases below 6.0.

     Cation (Ca, Mg, K) availability is not readily expressed as a
function of soil pH.  The relative availability of these nutrients as a
function of pH is of no practical consequence in most cases, except that
most soils become acid only after depletion of these cations.  In
strongly acid soils, however, toxic levels of solution Al  render
vegetation less able to utilize the Ca and Mg.

     Low soil  pH affects the availability of all micronutrients (B, Cl,
Cu, Fe, Mn, Mo, Zn) except chloride {Adams and Pearson 1967; Rorison
1980).  The availability of Cu, Fe, Mn, and Zn is significantly
increased by lower soil pH in the range 6.5 to 5.0.  Boron availability
increases only slightly with decreasing pH.   Molybdenum availability
decreases with decreasing pH because of decreased solubility of
molybdate forms.  Additional  information on soil acidity and plant
nutrition is given in Section 2.3.

2.2.6  Neutralization of Soil Acidity

     In unamended soils, the natural  forces that neutralize acidity are
weathering of neutral or basic minerals, the addition of basic  materials
from the atmosphere or floods, and the deposition of basic cations by
vegetation recycling.  In humid temperate regions outside of
floodplains, the uptake of basic cations by plant roots and their
deposition on the soil  surface and weathering are the important
neutralizing forces.  These forces do not normally reverse the  natural
acidification trends, but modify the  rate and distribution of
acidification within the soil profile.

     The effectiveness with which soil  acidity can be neutralized  by
liming depends upon the purity and particle size of the lime, the  amount
of lime applied, the soil  pH, the cation exchange capacity, the
uniformity with which the lime is spread,  and the extent of soil-lime
mixing (Barber 1967).  Liming materials are restricted to the Ca and Mg
salts of carbonate, silicate, and hydroxide.   The bulk of agricultural
lime comes from ground limestone.
                                  2-13

-------
     The net reaction that causes lime to neutralize  soil  acidity  is the
result of two separate reactions.   One is the cation-exchange  reaction
that releases Al3+ and H+ to the soil  solution from exchange sites;
the other is lime dissolution and the  hydrolysis  of COs2'.  When
exchangeable Al3+ js displaced by Ca2+ from dissolving  lime, it
undergoes stepwise hydrolysis to form  a precipitate of  A1(OH)3 and
solution H+ ions.  The overall exchange-hydrolytic reaction is
expressed by the equation

     2 Al-soil  + 3 CaCOa + 3 1^0 = 3 Ca-soil  + 2  A1(OH)3 + 3 C02-

     With thorough mixing of small  lime particles with  an acid soil,  the
neutralization reaction is quite efficient in raising soil pH  to about
6.0.  Lime becomes increasingly less effective in dissolving and raising
soil pH beyond this value.

2.2.7  Measuring Soil pH

     The term "soil pH" as it is commonly used refers to the pH of the
solution in contact with the soil.  Soil  pH is one of the most useful
measurements made on soils (Adams 1978).   It is used  to predict the
likelihood of excessive toxic ions, the need for  liming a soil, a
variety of soil microbial activities,  and the relative  availability  of
several inorganic nutrients.

     The usual  method of measuring soil  pH is to  immerse a
glass-electrode, reference-electrode assembly into a  soil-water
suspension and measure the electromotive  force (emf)  of the cell.  Part
of the measured emf is due to a junction  potential at the salt-bridge,
test-solution interface.  A basic premise of soil pH  measurements  is
that the junction potential between the salt bridge and the test
solution (or soil suspension) is the same as with the standard solution.
This equality is realized only where test solutions and standard
solutions are similar in ionic compositions.   Soil suspensions hardly
meet this requirement, but they approximate it if the reference
electrode is placed in the supernatant while the  glass  electrode is
immersed in the settled suspension.

     Because soil pH is an empirical value, the method  of measurement
must be standardized.  Samples should  be  either air-dried or oven-dried
at low temperature (< 50 C); oven drying  at 105 C produces meaningless
pH values.  When soil solution is separated from  solid-phase soil,  its
pH seldom matches that of the soil  suspension. One reason for the
discrepancy is the loss or dilution of C02 in the soil  solution upon
drying of the soil sample and the subsequent addition of water.

     Soil pH is influenced by the soil-water ratio and  the salt
concentration of the water used.  There is no universal agreement  on
what the ratio should be.  Soil to water  ratios of 1:1  up to 5:1 are
commonly used.   Since most soils are highly buffered, the differences
obtained due to variations in soil:water ratio are not  of practical
importance as long as the procedure is consistent and stated with  the
resul ts.
                                  2-14

-------
     In acid soils, soil  pH generally  decreases  temporarily with  the
addition of fertilizer or other salts  and  increases with  the  dissipation
of fertilizer,  either by  crop removal  or by  leaching.   In  poorly
buffered soils, this pH change may be  as much  as 0.5 to 1.0 pH  unit for
normal  fertilizer rates.  These changes in  soil pH are  not due to  changes
in total soil  acidity but are due to shifts  of Al  and  H ions  from
exchange sites to soil solution because of cation-exchange reactions.
Some of this variation can be overcome by  use  of a 0.01M  CaCl?
solution instead of water when measuring pH.

     If soil acidity of an area is to  be monitored over years,  time of
sampling should be consistent with annual  inputs of fertilizers,  natural
vegetative cycles, and weather cycles.  The  most consistent values will
be obtained if samples are taken when  salt content is  at  a minimum.

     Spatial variation of soil  pH within a field,  both vertically and
horizontally,  requires careful  sampling to obtain a sample that
represents the area of interest.   The  area to  be represented  should be
reasonably uniform in appearance within one  soil  series and uniform in
history.  Several  identical  soil  cores should  be composited and
thoroughly mixed before a subsample of the composite for  pH measurement
is taken.

2.2.8  Sulfate Adsorption

     As pointed out in Section 2.2.1.3, the  presence of mobile  anions is
necessary for the leaching of cations  to occur.   The dominant anion in
the atmospheric deposition in North America  is sulfate ^S042~).
Therefore, the reaction of sulfate, especially its adsorption or  free
movement, is an important soil  characteristic.

     Soils containing large quantities of  amorphous Fe and Al oxides or
hydroxides have a capacity of adsorb 5042-.  Sulfate adsorption
results in the displacement of OH° or  OH2  from  iron or aluminum
hydroxide surfaces (Rajan 1978).  This results in an increased  negative
charge on the hydroxide surface which  accounts for the simultaneous
retention of sulfate and  associated cations  in soil.   Sulfate adsorption
is strongly affected by pH since deprotonization of amphoteric
adsorption sites can make them negatively-charged and  cause repulsions
of anions.  Sulfate adsorption is also affected  by the cations  present
on exchange sites, with the presence of polyvalent cations causing more
adsorption than monovalent ions.   Soil pH  is a more important factor
than cation type,  however (Chao et al. 1963).  Recently,  it was shown
that organic matter has a decidedly negative influence on  sulfate
adsorption, even when free Fe and Al oxide content is  high (Johnson et
al. 1979, 1980, Couto et  al.  1979). This  effect is thought to  be due to
the blockage of adsorption sites by organic  liquids.

     The question of reversibility of  sulfate  adsorption  is crucial to
the long-term effects of  acidic deposition on  soil leaching.  If  sulfate
is irreversibly adsorbed, sulfate adsorption can be viewed as increasing
                                  2-15

-------
the soil's capacity to accept acidic deposition before significant
leaching of cations begins.  If sulfate is reversibly adsorbed,  however,
its effects on reducing leaching are only short-term, since desorption
of sulfate will result in equivalent losses of sulfate and cations from
the soil.

     The reversibility of sulfate adsorption varies with soil  properties
and the desorbing solution used.  In some cases,  H20 recovers  all
adsorbed sulfate whereas in other cases, full  recovery is achieved only
with phosphate or acetate extractions.   Reasons for the better recovery
with phosphate or acetate include the greater affinity of these  anions
for adsorption sites and, in the case of acetate,  the increase in  pH  as
well.  Pre-treatment of soils with phosphate (such as by fertilization
in the field) is known to reduce sulfate adsorption capacity since
sulfate does not displace phosphate from adsorption sites.   However,
phosphate does not always displace all  adsorbed sulfate, as shown  by
Bornemisza and Llanos (1967) for highly-weathered tropical  soils rich in
Fe and Al oxides.

     There is evidence that "aging" or  prolonged  contact between soil
and solution reduces the recovery of sulfate (Barrow and Shaw  1977).
This effect is attributed to slow reactions and occurs with other
adsorbed anions as well.  Some soils are known to adsorb sulfate
irreversibly (against H20) under field  conditions  but not in
laboratory conditions (Johnson and Henderson 1979), a phenomenon likely
related to slow reactions.  Microbial  immobilization may be a  factor  in
the "aging" phenomenon as well.

     Sulfate adsorption is concentration-dependent, i.e., sulfate
adsorption increases with solution sulfate concentration (Chao et  al.
1963).  Thus, for any given input concentration,  sulfate will  adsorb  on
to soil sesquioxide surfaces until  the  corresponding soil  adsorbed
sulfate value is reached on the sulfate adsorption isotherm.   When that
point is reached, the soil should be in steady-state with outputs
equalling inputs.  In the case where sulfuric  acid inputs increase,
concentrations increase, thereby activating "new"  sulfate adsorption
sites and causing a net sulfate retention in the  soil.  With continued
inputs, a new steady-state condition would eventually be reached.   This
is schematically depicted in Figure 2-3 (Johnson  and Cole 1980).

     This concentration-dependent relationship will result in  a  "front"
moving downward through a sulfate adsorbing soil  when a new, higher
level of sulfate concentration is introduced,  and continually  applied to
the soil.  Soil  above (or behind) the front will  have a new higher level
of sulfate on the soil in response to the higher  solution levels.   Soil
solution samples taken behind the front might  indicate signficant
movement of cations and sulfate, while  samples at a lower depth  indicate
essentially no leaching of cations and  sulfate.   Thus, the  sulfate
adsorbing soil delays cation leaching effects  of  dilute sulfuric acid
inputs until  the adsorbing capacity (dependent on  input concentration)
is satisfied down through the soil  zones of interest.
                                  2-16

-------
     oo
     CQ
     C£.
     O
     1/1
     O
     
-------
     In the case where sulfuric  acid  inputs  decrease, sulfate will
desorb from the soil,  unless  it  is  Irreversibly adsorbed, to a point on
the isotherm at which  the equilibrium sulfate concentration equals input
concentrations.  At this  point,  inputs and outputs are equal.  Prior to
this point, outputs exceed inputs during  sulfate desorption and the
sulfate and cations previously retained during adsorption are leached
from the soil.

     Sulfate adsorption capacity of soils  is not routinely determined;
therefore, the  extent  of  soils with significant capacity to adsorb
sulfate has not been established.   Some adsorption is a common property
of many Ultisols, Oxisols, some  Alfisols, and is reported for other
soils (Singh et al. 1980). The  work  of Johnson and Todd (1983) shows
sulfate adsorption is  low in  Spodosols.   The distribution of these soil
orders within the U.S. is depicted  in Figure 2-4 (Section 2.3.5).

2.2.9  Soil Chemistry  Summary

     Acid soils are a  natural consequence of long exposure to a climate
of excess rainfall because of the leaching action of natural inputs of
acidic ions.  Unleached soils do not  become  acid.  The rate at which
leached soils become acid depends upon soil  characteristics, including
buffer capacity, and the  rate of H+ input and the accompanying anion.
Natural H+ inputs come from C02» organic  matter, nitrification, and
sulfur oxidation.  The buffer capacity of soils partially neutralizes
H+ input by reactions  with carbonates (>  pH  7.0), with exchangeable
bases (pH 5.5 to 7.0), and with  clay  minerals (< pH 5.5).  Soil-mediated
injury to vegetation from H+  inputs occurs only when pH is low enough
to cause significant dissolution of Al- or Mn-containing clay minerals
(< pH 5.0 to 5.5).

     The amount of H+  required to lower pH of an acid soil depends
upon the CEC of that soil.  For  example,  a loamy sand Ultisol with the
rather low CEC of 2.0  meq 100 g-1 requires about 1.1 meq H+ 100
g-1 to lower pH from 6.0  (65  percent  base saturated) to 4.5  (10
percent base saturated).  That would be about 22 keq H+ ha"1 to
effect the change to a depth  of 15  cm. A finer textured Ultisol  with a
CEC of 10 meq 100 g-1  requires about  five times that amount.  Soils
high in smectites (expandible clays)  or organic matter require
considerably more H+ for  a comparable pH  change.

     The weathering of alumino-silicate clays will produce strong
buffering  in soils that are already acid  (5.5 or below) such that
calculations of pH changes, based on  changes in basic cation removal by
H+ additions, grossly underestimate the amount of acid required to
cause the changes in these soils.   The presence of sulfate adsorption
capacity (see Section 2.2.8)  increases their capacity to absorb dilute
H2S04 inputs before significant change in pH or base status
occurs.
                                  2-18

-------
2.3  EFFECTS OF ACIDIC DEPOSITION ON SOIL  CHEMISTRY AND  PLANT  NUTRITION

     It is not always clear what deposition is acidic  or acidifying.
From the standpoint of the effects on neutral  to  acid  soils, the
following depositional materials could be  expected to  have  acidifying
effects:  H2S04, HN03, H2SOs,  S02, S, NHs,  NH4S04,  whereas  the
following sulfate salts are essentially neutral or slightly basic  in
effects on long-term soil  pH:   CaS04, K2S04,  Na2S04, MgS04- Carbonates
of calcium and magnesium would raise the pH.

     To alter the soil chemically, precipitation  must  bathe the soil
particles.  Runoff water will  minimally impact soil due  to  its brief
contact with soil particles.   As Tamm (1977)  has  noted,  water
percolating through soil is not necessarily at equilibrium  with the soil
solution but may move directly through old root channels, animal
burrows, and large pores at ped surfaces.   Soils  percolating similar
quantities of water may differ in the extent of their  reaction with the
water.  Under unsaturated conditions, water tends to move through  the
small pores of soil aggregates and has the best opportunity to attain
chemical equilibrium with the  soil.  During a rainfall,  the flow
velocity in the small pores within aggregates becomes  negligible
relative to that in the large  pores between aggregates.   Drainage  water,
therefore, only reacts with the soil to the extent that  dissolved
constituents diffuse between the small  and large  pores (Bolt 1979).
This effect can be demonstrated by comparing soil solution  chemistry,
obtained by porous ceramic cups, with that of free leachate water.
Using this system, Shaffer et  al. (1979) demonstrated  that  solutions
applied to a saturated soil can pass through the  soil  rapidly  and  nearly
unchanged.

2.3.1  Effects on Soil pH

     In considering the effects of acidic  deposition,  it is essential  to
realize that acids are produced naturally  within  soils (Reuss  1977,
Rosenqvist 1977, Rosenqvist et al. 1980; also see Section 2.2.1).
Atmospheric acidic inputs must be viewed as an addition  to  natural,
continual acidification and leaching processes due to  carbonic acid
formation, organic acid formation, vegetative cation uptake, and a
variety of management practices (Reuss 1977,  Johnson et  al. 1977,
Andersson et al. 1980, Soil ins et alI. 1980).   In  Table 2-1  several
values are given for potential acidifying  or neutralizing effects  of
lime, N fertilizer, acidic precipitation,  and internal acid production
in soils.  Even though most of the values  are only approximate, it is
clear that a year of rather heavy acidic deposition has  potential
acidifying effects that are small compared to common agricultural
amendments.  For that reason,  it is generally concluded  (McFee et  al.
1977, Reuss 1977) that acidic  deposition will not have a measurable
effect on the pH of soils that are under normal cultivation practices.

     The values for internal acidity production (see Table  2-5 in
Section 2.3.3.1) span a wide range.  If the lower values occur, then
acidic deposition is potentially as influential as natural  processes,


                                  2-19

-------
           TABLE 2-1.   RELATIVE ACIDIFYING AND  NEUTRALIZING  POWER  OF
                            MATERIALS  ADDED TO  SOILS
       Source
        Potential  acid or base effect
Agricultural  liming operation
  5000 kg CaCOa ha'1
Neutralizing or basic effect
  100 keq ha'1              10 eq  m"z
Nitrogen fertilization with
  reduced form of N,  such as
  urea or Nfy
  70 kg N ha'1
Acidifying effect9
  10 keq ha'1
     1  eq
Atmospheric deposition           Acidifying effect
  1 year (100 cm)  pH 4.0 rain       1  keq ha'1
                           0.1 eq  m'
                                    ,-2
  16 kg S ha-1 dry deposition    Acidifying effect
                                   1  keq ha-1
                           0.1 eq  m"2
Internal acid production in
  soils due to carbonic and
  organic acids in one year
  from Table 2.5
Acidifying effect
  .23-22.7 keq ha-1
.023-2.27 eq  nr2
 N fertilization usually has somewhat less  actual  acidifying effect.
 This is the maximum assuming complete nitrification  of the N
 fertilizer.
                                 2-20

-------
but in other cases it would be quite small  and of little consequence in
natural  ecosystems.   Unfortunately,  the  data base for including natural
acid formation in assessments of impact  on  soils is extremely limited.
Thus, current schemes, by default, often assume that atmospheric inputs
add significantly to internal acid production, an assumption that is not
universally accepted (e.g.,  Rosenqvist 1977, Rosenqvist et al. 1980).
Carbonic acid is a major leaching agent  in  some forest soils (McColl and
Cole 1968,  Nye and Greenland 1980),  yet  it  does not produce low pH
(i.e., < 5.0) solutions under normal conditions (McColl and Cole 1968;
Johnson et al. 1975, 1977).   Organic acids  may contribute substantially
to elemental leaching in forest soils undergoing podzolization (Johnson
et al. 1977) and can produce low pH  (i.e.,  < 5.0) in unpolluted natural
waters as well (Johnson et al.  1977, Rosenqvist 1977, Johnson 1981).

     Experiments that directly indicate  a change in pH due to acidic
deposition  inputs (Tamm 1977, Abrahamsen 1980b, Parrel! et al. 1980,
Wainwright 1980, Stuanes 1980,  Bjor  and  Teigen 1980) either used
accelerated application rates far exceding  natural precipitation or
applied concentrated acid.  Both create  situations unlikely to exist in
nature because they do not allow for normal influences of weathering,
and nutrient recycling.  It is also  clear that soils exposed to
concentrated acids over short periods will  undergo reactions and changes
that would never occur with more dilute  acid over longer periods.
Therefore,  the effects of acidic deposition on soil pH are often
predicted from known soil chemical relationships, using input values
similar to those measured in recent  years and without the benefit of
long-term experiments under simulated natural conditions.

     McFee et al. (1976) calculated  theoretical reductions in both soil
pH and base saturation from atmospheric  H+  inputs, assuming no
concurrent inputs of basic cations.   They concluded that most soils
resist pH change and that there is only  a "small likelihood of rapid
soil degradation due to acid precipitation."  However, they also suggest
that long-term (e.g., 100 years) soil acidification trends could have an
impact on non-agricultural soils and that these trends are very
difficult to evaluate in short-term  experiments.  Models of soil
acidification processes range from complex  ecosystem budget approaches
(Andersson et al. 1980, Soil ins et al. 1980) to process-oriented soil
leaching models (Reuss 1978).  Their quantitative applicability on a
wide range of sites has not been tested, but they can add to our
understanding of the concepts involved and  may be applied to many
terrestrial ecosystems.

     Despite uncertainties in estimating potential acidification rates,
the authors of this chapter provide  some illustrations in Table 2-2.
The data illustrate that large differences  in potential acidification
rates can be expected due to CEC alone,  even without considering such
other soil  properties as anion adsorption capacity or hydro!ogic
characteristics.  It also illustrates how the assumptions concerning
accompanying cations, H+ replacement efficiency, and weathering rates
change estimates of acidification rates.
                                  2-21

-------
     Several considerations embodied in Table 2-2 must be  understood  if
the data are to be used correctly.

1)   The input rates of acidic deposition  are considerably higher  than
     those now reported for the United States.

2)   Most natural  ecosystems within humid  regions have acid  soils.
     Soils with neutral  to slightly-acid pH  and  with  very  low CEC,  3  to
     6 meq 100 g-1, are uncommon in the humid regions.

3)   A 50 percent decrease in base  saturation in many mineral soils
     could lower pH from the slightly acid (6.6  to  6.8)  range to
     strongly acid (5.0 to 5.5) range.

4)   These estimates ignore anion adsorption capabilities  and natural
     acidifying processes.

5)   Assumptions under scenario 1 are not  realized  in nature.  Those
     under 2 and 3 are realistic for many  soils  and many deposition
     situations, but cannot be considered  universally applicable.

     If we consider a soil with a low CEC  of only 3 meq 100  g-1 and
assume a soil bulk density of 1.3 g cc"1,  this soil would  have a total
of 60 keq cation exchange capacity  per hectare in the top  15 cm (third
soil in Table 2-2).  A significant  pH change could  be accomplished by
reducing the percent base saturation by 50 percent.   This  would seem  to
be theoretically possible in 15 years:  15  yr x 2 keq  ha-1  yr-1 = 30
keq ha"1.  However, all  of the acid input  would  have  to replace and
leach an equivalence of bases (Assumption  1  in Table  2-2).  This is
highly unlikely.  Wiklander (1974)  indicates a replacement efficiency
considerably less than 1.0 in acid  soils,  pH 5.5 to 6.5.  Further,
accompanying salts of Ca, Mg, and K also reduce  the acid efficiency in
lowering pH (Assumption 2).  Such rapid change also assumes  no H+
consumption by weathering and no recycling of bases to the surface soil
whereas Abrahamsen (1980b) indicated weathering rates were keeping pace
with acid inputs in treatments with pH above 4.0.   Moreover, vegetation
may deposit significant quantities  of basic nutrient  ions  on the
surface.  A more reasonable estimate of the  years required to lower the
soil pH significantly, even in this very poorly  buffered example,  is  22
to 90 years.  If a value of 9 meq CEC or higher  is  assumed (a more
common value for most surface soils in the United States)  then the
minimum time is 67 years without weathering and  much  longer, or
infinity, with normal weathering.

     The magnitude of soil resistance to pH changes is illustrated by
the small pH changes that have resulted from natural  acid  inputs of 0.23
to 2.27 keq ha-1 yr-1 generated by  N-fixation-metabolism,  organic
matter decay and C02 from respiration (Table 2-1).  These  inputs have
not caused rapid pH changes and it  is unlikely that an additional  2 keq
ha"1 yr-1 or less from acidic deposition will  cause a significant
change in many soils.
                                  2-22

-------
  TABLE 2-2.   ESTIMATES OF  TIME  REQUIRED TO EFFECT A 50% CHANGE IN BASE
   SATURATION IN  THE  TOP 15 CM OF  SOIL.  TIME REQUIRED FOR SIGNIFICANT
  ACIDIFICATION OF UNCULTIVATED  SOILS THAT ARE SLIGHTLY ACID OR NEARLY
    NEUTRAL UNDER HIGH  RATES OF  ACIDIC DEPOSITION—100 CM OF PH 4.0
   PRECIPITATION  PLUS 16 KG S HA-1 YR"1 IN DRY DEPOSITION (TOTAL ACID
                   INPUT OF 2 KEQ  H+ HA'1 YR'1)
           Soil
                   CEC
                meq 100 g~l
     Assumption
1        2         3

       years
Midwestern Alfisol
Southeastern Ultisol
Ouartzipsamnent
15
9
3
75
45
15
110
67
22
220 oo
125 °°
45-90
  with low organic matter
Assumption 1.
Assumption 2.
Assumption 3.
All  incoming H+ exchanges for (replaces)  basic  cations
on the soil exchange complex.  There are  no accompanying
basic cations and no weathering or other  input  of  basic
cations.  This is the "worst case" situation and cannot
exist in nature.

The incoming H+ is accompanied by 0.3-0.5 keq ha~*
yr-1 of basic cations Ca, Mg, K (Cole and Rapp  1981),
and the replacing efficiency of H+ for basic cations
drops below 1.0 as the base saturation of the soil drops
(Wiklander 1975).

Same as under 2 except that acidification is further
slowed by release of basic cations from weathering 1  keq
ha'1 yr-1 (for example, 20 kg Ca ha"1 of  15 kg  Ca
plus 3 kg Mg ha~l yr"1 within range calculated  for
Hubbard Brook (Likens et al. 1977) and the cycling of
basic cations back to soil surface by plants.
                                  2-23

-------
     The evidence for acidification  of  soils  by  the present rate of
acidic deposition Is not strong.   If significant acidification is to
occur within a few decades,  it will  be  in  the limited soil areas that
combine the following characteristics:   the soil is not renewed by fresh
soil deposits; it is low in  cation exchange capacity, i.e., low in clay
and organic matter;  it is low in  sulfate adsorption capacity; it
receives high inputs of acidic deposition  without  significant basic
cation deposition; it is relatively  high in present pH (neutral to
slightly acid) and free of easily weatherable materials to one meter
depth (see Section 2.3.5.2.1).

     As Section 2.3.3.1 discusses, acid precipitation cannot leach
nutrient cations unless the  associated  sulfate or  nitrate anions in the
soil are mobile.   Evidence indicates that  sulfate  is not always mobile
(Section 2.2.8) particularly as soils become  more  acid (Johnson and Cole
1977, Johnson et al. 1979, Abrahamsen 1980b,  Singh et al. 1980).

     It is also possible for a soil  to  be  leached of cations without
concurrent acidification, if acidic  inputs stimulate the weathering of
cations from primary minerals.   Therefore, it is important to make a
distinction between cation leaching  and the process of soil
acidification.  It is unrealistic to assume either a steady-state
condition for soil exchangeable cations or a  condition where weathering
is zero and cations are depleted  from exchange sites in proportion to
H+ inputs.  These common assumptions made  in  predictive models
seriously limit the models'  applicability  to  natural systems.  Another
important factor which models do not consider is the acidification
caused by natural processes.  As  noted  in  Section  2.2.1, atmospheric
acid inputs must be viewed as an  addition  to  the natural acidification
processes of cation uptake by plants, nitrification, and soil leaching
by organic and carbonic acids (Johnson  et  al. 1977, Reuss 1977, Cronan
et al. 1978, Rosenqvist et al.  1980).

     Section 2.1.3, on leaching is closely related because long-term pH
changes require leaching of  basic cations  as  well  as acidic inputs.

2.3.2  Effects on Nutrient Supply of Cultivated  Crops

     This section deals with the significance of atmospheric additions
of S and N to crop requirements.   Few detrimental  effects of acidic
deposition are expected on nutrient  supply to cultivated crops (see
Section 2.3.1) because by comparison agricultural  practices have a
massive effect.

     Input of nutrients to plant systems from rainfall has been
documented since the mid-19th century (Way 1855).  Calculations made in
a number of U.S. regions have estimated the  seasonal atmospheric
deposition of nutrient species, particularly  sulfate and nitrate, to
agricultural and natural systems and the implications of this  deposition
on plant nutrient status. Estimates by Hoeft et al. (1972) of 30 kg S
ha~i yr~l and 20 kg N ha~l yr~l deposited  in  precipitation in
                                  2-24

-------
Wisconsin indicates the importance of atmospheric  sources of these
elements.  These values,  however,  are higher than  those usually reported
in the United States (see Chapter  A-8).  Jones et  al. (1979) reported
that atmospheric S is a major contribution  to the  agronomic and
horticultural crop needs for S as  a plant nutrient in South Carolina.

     The amount of annual S deposition at selected sites is presented in
Table 2-3.   Amounts of S recorded  for 1953-55 in rural areas along the
Gulf and southern Atlantic coasts  were usually less  than 6 kg S ha~l
yr-1.  In northern Alabama, Kentucky, Tennessee, and Virginia the
levels were much higher (10 to 30  kg ha~l yr-1)  (Jordan et al.
1959).  These can be compared with the recent NADP data for wet
deposition of S (Chapter A-8).

     These amounts of S represent  a significant  portion of that required
by crops.  The amounts of S absorbed by crops are  summarized in Table
2-4.  Terman (1978) estimates an average crop removal of 18.5 kg S
ha-1 yr-1 and concludes that if current rates of atmospheric S
deposition are greatly reduced, the need for applying fertilizer S for
satisfactory crop yield will  increase.

     The atmospheric deposition of N is usually lower than deposition of
S, but crop requirements are much  higher.   Therefore, it is generally
accepted that atmospheric N deposition plays a small or insignificant
role in nutrition of cultivated crops (see  Chapter E-3, Section 3.4.2).

     It is well known that foliar  applications of  plant nutrients can
stimulate plant growth (Garcia and Hanway 1976).   It is possible, but
unproven, that repeated exposure of plants  to small  amounts of
atmospheric deposition may be more effective in  stimulating plant growth
than a comparable amount applied to soils (see Chapter E-3, Section
3.4).

2.3.3  Effects on Nutrient Supply  to Forests

     Nutrient supply may be influenced by acidic deposition effects on
leaching of cations or by pH induced changes in mineral solubility,
microbial processes, or weathering rates in addition to the direct
influence of additions of N and S  in deposition.   Microbial processes
are discussed in Section 2.4.   Solubility (availability) and weathering
reactions are discussed in Section 2.2.

     Acid rain has created a major concern  because of the potential for
accelerated cation leaching from forest soils and  eventual losses of
productivity (Engstrom et al.  1971).  This  concern was the driving force
for numerous empirical studies of  acid rain effects  on forest nutrient
status in general and cation leaching in particular  (reviewed by Johnson
et al. 1982).

     •Perhaps because of the negative implications  of the term "acid
rain," initial speculations about  acid rain's effects on forest
                                  2-25

-------
TABLE 2-3.  AMOUNTS OF SULFUR FOUND IN PRECIPITATION IN VARIOUS STATES
State
Southern States
Al abama
Arkansas
Florida
Kentucky
Louisiana
Mississippi
North Carolina
Oklahoma
Tennessee
Texas
Virginia
Location
in state
Prattville
Muscle Shoals
Muscle Shoals
Muscle Shoals
NW and SE
Gainesville
Others
Various
Various
Various
Statesville
Others
Still water
Various
Various
Beaumont
Others
Norfolk
Various
Sites
1
19
20
23
2
1
5
6
5
7
1
15
1
7
5
1
4
1
16
Years Major source
kg
1954-55
1954
1955
1956
1954-56
1953-55
1953-55
1954-55
1954-55
1953-55
1953-55
1953-55
1927-42
1955
1971-72
1954-55
1954-55
1954-55
1953-56
General
General
Steam Plant
Steam Plant
General
Urban
General
General
General
General
Industry
General
General
General
General
Industry
General
Industry
General
Average
S ha-1 yr-1
3.7
5.4
11.9
11.0
3.7
8.8
3.2
13.1
9.0
5.0
15.5
6.0
9.7
14.2
17.1
12.1
5.7
35.2
21.4
                               2-26

-------
TABLE 2-3.  CONTINUED
State
Northern States
Indiana
Michigan
Nebraska
New York
Wisconsin
Adapted from Terman
Location Sit
in state
Gary
Others 1
Various 5
Various 7
Ithaca
Industrial Site
Urban
Rural 1
(1978). See origir
es
1
0


1
1
9
3
al
Years
1946-47
1946-47
1959-60
1953-54
1931-49
1969-71
1969-71
1969-71
for data
Major source
Industry
General
Industry
General
Urban &
Industry
Industry
Urban
General
sources.
Average
kg S ha-1 yr-1
142.2
30.0
11.3
7.2
54.9
168.0
42.0
16.0

 2-27

-------
                    TABLE 2-4.  SULFUR CONTENT OF CROPS
            Crops
 Yield    Total  S Content
tons ha-1    kg  ha"1
Grain and oil crops
Barley (Hordeum vulgare L.)
Corn (Zea Mays L.)
Grain sorghum (Sorghum bicolor L. Moench)
uats (Avena satlva L.)
Rice (Oryza sativa L.)
Wheat (Triticum aestivum L.)
Peanuts (Arachis hypogaea L.)
Soybeans (Glycine max Merr.)
Hay Crops
Alfalfa (Medicago sativa L.)
Clover-grass
Bermuda-grass (Cynodon dactylon L.)
Common
Coastal
Orchardgrass (Dactyl is glomerata L.)
Timothy (Phleum pratense L.)
5.4
11.2
9.0
3.6
7.8
5.4
4.5
4.0
17.9
13.4
9.0
22.4
13.4
9.0
22
34
43
22
13
22
24
28
45
34
17
50
39
18
Cotton and tobacco

  Cotton (lint + seed)  (Gossypium hirsutum L.)
   4.3
34
Tobacco (Nicotiana tabacum L.)
Burley
Flue-cured
Fruit, sugar, and vegetable crops
Beets
Sugar (Beta saccharifera)
Table (Beta vulgaris L.)
Cabbage (Brassica oleracea)
Irish potatoes (Solanum tuverosum L.)
uranges (Citrus sp.)
Pineapple (Ananas comosus)

4.5
3.4
67
56
78
56
52
40
21
50
50
46
72
27
31
16
Estimates by Potash/Phosphate Institute of North  America.
Terman (1978).
            Adapted from
                                  2-28

-------
productivity devoted little or no attention to concurrent sulfate and
nitrate deposition on forests deficient in S or N.   Only recently has it
been recognized that acid rain can cause increases  as well  as decreases
in forest productivity (Abrahamsen 1980b, Cowling and Dochinger  1980).
The net effect of acid rain on forest growth depends upon a number of
site-specific factors such as nutrient status and amount of atmospheric
acid input.  (See also Chapter E-3, Section 3.4.1.)

     It is also very important to consider that ions such as $042-
and N03" are already in the ecosystem and that H+ is generated
naturally by the plant community (111rich 1980).  Thus, the question is
one of relating inputs to natural  levels; e.g., does atmospheric H+
input significantly add to or exceed natural H+ production within the
soil?  Do the detrimental effects of H+ deposition  offset the benefits
of N03~ deposition in an N-deficient ecosystem or the benefits of
S042~ deposition in an S-deficient ecosystem?  In short,  the problem
of assessing the effects of acid rain on forest nutrient status  is
largely a matter of quantification and requires a nutrient cycling
approach.

2.3.3.1  Effects on Cation Nutrient Status—Cation  leaching is important
to soil properties because it may lead to a loss of plant nutrients and
depressed soil  pH.  It is important in hydrology because cations leached
from soils may be deposited in aquatic systems.

     The basic cation status of a soil depends on the net effect of
leaching and other losses versus weathering and other inputs (Abrahamsen
1980a, Ulrich et al. 1980).   Weathering is stimulated by  additional  H
input, offsetting leaching to some extent.   However,  most acid
irrigation studies (Abrahamsen 1980b)  and one study  under ambient
conditions (Ulrich et al. 1980) indicate a net decline in exchangeable
basic cations with time.   There is little doubt that acid rain can
accelerate cation leaching rates,  but the magnitudes of these increases
must be evaluated within the context of natural,  internal  leaching
processes.  The magnitude is quite variable, depending upon the  amount
of acid input,  the rate of soil leaching by natural  processes (Cole and
Johnson 1977, Cronan et al.  1978), and the degree to which  soils are
buffered against leaching (e.g., by anion adsorption;  Johnson and Cole
1977).  Furthermore, the ultimate effects of accelerated  cation  leaching
on cation nutrient status depend upon a number of variables,  most
notably exchangeable cation capital,  primary mineral  weathering  rate
(Stuanes 1980),  forest cation nutrient requirement,  and management
practices such as harvesting.

     A comparison of the effects of some of these factors on cation
nutrient status  is given  in  Table  2-5.   Various schemes for evaluating
internal acid production have been proposed (Reuss  1977,  Soil ins et al.
1980,  Ulrich 1980), but in this case,  only  the values  reported by
various investigators for soil  leaching (usually  by  carbonic acid)  are
considered.   It is obvious that atmospheric acid  inputs vary not only in
absolute magnitude, but also in their importance relative to internal
leaching processes and effects of  harvesting.


                                  2-29

  409-262 0-83-3

-------
       TABLE 2-5.  ATMOSPHERIC  H+  INPUTS VS CATION REMOVAL BY  INTERNAL H+
   PRODUCTION (CARBONIC  AND ORGANIC ACIDS) AND POTENTIAL NET  ANNUAL CATION
   REMOVAL IN BOLE ONLY  AND WHOLE-TREE HARVESTING (WTH) IN SELECTED FOREST
                ECOSYSTEMS (ADAPTED FROM EVANS ET AL. 1981)
Site
Thompson,
Washington
Soiling,
W. Germany
Jadrass,
Sweden
Findley,
Washington
H.J. Andrews,
Oregon
Precipitation
Species Age H+
(yr) input3
Pseudotsuga 42
menziesii
Fagus sylvatica 59
Pinus sylvestris
Abies amabilis, 175
Tsuga mertensiana
Pseudotsuga 450
menziesii


240<*
(4.8)
9009
1909
90h
(5.6)
289
Cation
Cation removal by
leaching by harvesting0
internal acid
production^5 Bole WTH
(eq ha"1 yr"1)
420<* 3806
(5.9)
19509 2209
2269
1410" 2726
(4.5)
227009 606

6606
3706
4606
1066
aWeighted average [H+]  times  precipitation  amount;  weighted average [H+] as
 pH appears in parenthesis where available.

bCalculated from net increase in weighted average  HC03~ or organic anion
 concentration (the latter estimated  by  anion  deficit) times water amount.
 Weighted average [H+]  as pH  for solutions  appears  in  parentheses where
 available.

CNutrient content divided by  age;  WTH =  whole  tree  harvest, removal of all
 aboveground biomass.
      Cole and Johnson (1977).

eFrom Cole and Rapp (1981).

fFrom Lindberg et al .  (1979).

9From Andersson et al . (1980).   For comparison  in  this  table,  only  H?C03
 production values are included.
                                        2-30

-------
     At the unpolluted site in Findley Lake,  it is not surprising  that
internal leaching processes and harvesting effects exceed  atmospheric
H+ inputs.  However, even in the beech stand  at Soiling, West Germany,
values for HgCOs production reported by Andersson  et al.  (1980)
exceed atmospheric H+ inputs as measured by open-bucket collectors.
In this case, the comparison is misleading, however,  since dry
deposition to the forest canopy at Soiling is known to be  exceedingly
high (Ulrich et al.  1980), and, consequently, H+ inputs to the forest
floor substantially  exceed those deposited above the canopy.   It is  also
noteworthy that Ulrich et al.  believe that while internal  H+  producing
processes are important at Soiling, acid rain is having serious,
deleterious effects  on forests there.

     Studies of basic cation leaching due to  acidic inputs sometimes
give inconsistent results.  Under ambient conditions,  Mayer and Ulrich
(1977) noted a net loss of Ca, Mg,  K, and Na  from  the  soils under  a
beech forest.  Except for Na,  however,  the loss was equal  to  or less
than nutrient accumulation in  the trees.   Roberts  et  al. (1980) reported
that acidic precipitation on Delamere forest  (pine)  of central England
may produce small  changes in litter decomposition,  but they found  no
effect on Ca, Mg,  K, or Na leaching rate.   Cole and Johnson (1977) found
no detectable effect of acid precipitation on the  soil  solution of a
Douglas-fir ecosystem.  On the other hand, Andersson  et al. (1980) noted
a net output of Ca from both a pine forest soil  in  Sweden  and  a beech
forest soil in West  Germany; both soils accumulated N  but  not sulfate.
Cronan (1980a) reported net losses  of Ca,  Mg, K, and  Na from  subalpine
soil in New Hampshire, attributing losses to  acidic precipitation.
Studies by Mollitor  and Raynal (1982) suggest that  leaching of K may be
the most serious problem of cation leaching in Adirondack  forest soils.

     Nitrate is sometimes associated with acidic deposition and differs
considerably from sulfate in that it is very  poorly adsorbed  to most
soils (Johnson and Cole 1977).  However,  biological  processes  in
N-l invited ecosystems quickly immobilize nitrate, and  since N  limitations
are common in forested regions of the world,  nitrate  is rarely mobile
(Abrahamsen 1980b).   On the other hand,  nitrogen-rich  ecosystems (where
biological immobilization of N03~ is minimal)  are  susceptible  to
leaching by HN03«

     With regard to  North American  forests, cation  deficiencies are  very
rare although they are known to occur in red  pine  (Pinus resinosa) on
some sandy soils in  New York State  (Stone  and Kszystyniak  1977, Heiberg
and White 1951,  Hart et al.  1969).   Acid rain accelerated  leaching
could, in theory,  exacerbate this situation,  but this  possibility  has
not been investigated.  It should be noted, however,  that  these
ecosystems are exceedingly conservative with  regard to  potassium (Stone
and Kszystyniak 1977), and biological cycling and  conservation may play
major roles in resisting effects of acid  rain on K+ leaching  (e.g.,
other cations may  be leached while  K+ is  conserved).

2.3.3.2  Effects on  S and N Status—Deficiencies of S  have been
indicated in forests remote from pollutant inputs  in eastern Australia


                                  2-31

-------
(Humphreys et al.  1975)  and the northwestern United States  (Youngberg
and Dyrness 1965,  Will  and Youngberg 1978).   Humphreys  et al.  (1975)
suggest that pollutant inputs from power plants benefit S-deficient
Australian forests, particularly when the soils have little  S04Z"
adsorption capacity.  In these situations continual  input of moderate
amounts of H2$04 as acid rain may be a source of fertilizer.

     At the other extreme, continual  atmospheric S  inputs may  help
alleviate sub-optimal sulfate availability in sulfate "fixing"  soils
that are rich in hydrated Fe and Al  oxides.   Although adsorbed  insoluble
sulfate is thought to be available to plants in the long run,  the
intensity or rate  of supply to the soil  solution can be less than that
required by plants, effecting an S limitation (Hasan et al.  1970).

     Research has shown that N fertilization, a practice in  some
forested regions of the world, results in rapid use of  ecosystem S
supplies, possibly leading to S limitations  (Humphreys  et al.  1975,
Turner et al. 1980).  It has been suggested  that forest N and  S status
must be evaluated because of the closely related roles  of these elements
in protein synthesis (Kelly and Lambert 1972, Turner and Lambert 1980,
Turner et al. 1980).  In relatively unpolluted regions  of the
northwestern United States, evidence indicates that lack of  growth
response to N by Douglas-fir is due to marginal  S status (Turner et al.
1977, 1979).  Thus, it seems evident that moderate  amounts of  S in
deposition could benefit forests undergoing  N fertilization.   In the
United States this currently involves a total of about  1,000,000 ha of
forest lands, primarily in the Northwest and Southeast  (Bengston 1979).

     Amounts of atmospheric S input sufficient to satisfy forest S
requirements are much smaller than many crop requirements.   In general,
S inputs of 5 kg ha-1 yr~l are sufficient to satisfy S  requirements
in most forest ecosystems (Humphreys et al.  1975, Evans et al. 1981,
Johnson et al. 1982).  Inputs of S042- in acid rain affected regions
frequently exceed  this value (often by a factor of  2 to 4),  implying
that Sis currently being deposited in excess of forest requirements
(Table 2.3).

     Several studies have shown that excess  S cycles within  vegetation
and accumulates in soils as $042- without any apparent  harm  (Kelly
and Lambert 1972,  Turner et al. 1980, Turner 1980).   The plateau between
S sufficiency and toxicity in forest ecosystems appears to be  quite
broad.  Inputs of S usually constitute a more significant increment to
the natural sulfur flux within forest ecosystems than do equivalent
inputs of H+ to the natural flux of H+.  Therefore,  it  would appear
that further emphasis ought to be given to effects  from the  $042-
component of acidic deposition.  Similarly,  further emphasis ought to be
given to the effects of N inputs, since they appear to  be increasing
(Abrahamsen 1980b) and N is commonly the limiting nutrient in  forest
ecosystems.
                                  2-32

-------
     Nitrogen deficiencies are common in forests throughout the world
 (Abrahamsen 1980b).  Inputs of N03" (as well as NH4+ and other
 forms of N) are likely to improve forest nutrient status and
 productivity in many cases.  Nearly all  forest ecosystems for which
 nutrient budgets are available appear to accumulate N03~ as well as
 other forms of N (i.e., inputs > outputs; Abrahamsen 1980b).   Since
 N03~ is very poorly adsorbed to most soils (Vitousek et al.  1979),
 this accumulation is undoubtedly due to biological  uptake.  The
 inhibiting effect of N03~ immobilization on the leaching potential
 of HN03 is the same as that of S042~ immobilization on the
 leaching potential of H2$04 even though the mechanisms of
 immobilization for those two anions are different.

     Because forest N requirements are relatively high compared to S
 requirements, instances of atmospheric N inputs in  excess of forest N
 requirements seldom occur.   An apparent exception is the Soiling site in
 West Germany, where atmospheric inputs of N, S, and H+ are high
 (111rich et al. 1980).

     If atmospheric N inputs increase to the point  where N deficiencies
 are alleviated and excess N is available in soils,  nitrification may be
 stimulated.  Nitrification  pulses are thought to be responsible for a
 large percentage of leaching at the heavily-impacted Soiling  site in
 West Germany, for example (111 rich et al.  1980).  Thus, nitrogen
 "saturation" of forest ecosystems could result in significant increases
 in cation leaching and, under extreme circumstances, soil  acidification.
 Such "saturation" would occur most readily in forests with low N demand
 (i.e., boreal coniferous forests; Cole and Rapp 1981) or in  forests with
 adequate or excessive N supplied (such as by N-fixing species).  Indeed,
 the naturally acidifying effects of red  alder,  an N-fixing species
 indigenous to the northwestern United States, have  been noted by several
 investigators.   However, there is not evidence  of widespread,  imminent
 nitrogen saturation of forests since N deficiencies are still   quite
common and most ecosystems  are still  accumulating N (Abrahamsen 1980,
Johnson et al.  1982).

     Acidic deposition may  indirectly affect N  availability  in forest
 soils.   Tamm (1976)  predicted  short-term increases  in N availability
 (due to increased decomposition and microbiological  N immobilization)
and tree growth due to acidic  precipitation.   However,  long-term
declines in both N status and  tree growth could occur due  to  net N
losses from the ecosystem.   With regard  to decomposition,  empirical
 results have been variable  (see Section  2.5).  Whether this increase in
N availability is due to changes in  microbial activity or  to  the
acid-catalyzed hydrolysis of labile soil  N is unknown.   In either event,
the results of the Norwegian studies,  in  which  both  N availability and
nitrate leaching were stimulated by  H2S04 inputs, strongly suggest
that,  contrary  to earlier predictions  that nitrification would be
inhibited by acidic  inputs  (Tamm 1976),  nitrification can  be  stimulated
by acidic  inputs.
                                  2-33

-------
2.3.3.3  Acidification Effects on Plant Nutrition—It  is  unlikely  that
many soils will  be significantly acidified  by  acid  rain at current input
levels in the United States (see Section 2.3.1).  Should  soil
acidification occur, however (e.g.,  in  restricted areas with high  acid
inputs and very  poorly buffered soils), a great deal of information is
available about  plant responses.   Also, recent results from the
heavily-impacted Soiling site in West Germany  suggest  that slight
changes in soil  pH due to the combined  effects of acid rain and  internal
processes are causing serious negative  effects on forests there  (Ulrich
et al. 1980).

2.3.3.3.1  Nutrient deficiencies.   In general, only those acidic soils
that are highly  leached (sandy and/or low CEC) are  likely to be
sufficiently low in Ca to affect growth of  higher plants.  That  is,  if
Al  and other toxic ions are not present in  excess,  most acidic soils
will  have adequate Ca for good growth of most  plants (Foy 1964,  Foy
1974a).  The evidence suggests that many, if not all, of  the Ca
deficiencies reported on acidic soils in the field  are due to Al-Ca
antagonisms rather than low Ca per se.   For a  fuller treatment of  the
Ca-deficiency Al-toxicity argument,  see earlier reviews (Kamprath  and
Foy 1972; Foy 1974a,b; Foy 1981).   Similarly,  magnesium deficiencies
observed in plants grown on acid soils  are  often due to Al-Mg
antagonisms rather than low total  soil  Mg levels.

     Phosphorus  deficiency is a common  problem in crops and forests
grown on acidic  soils because such soils are often  low in total  P  and
because native P,  as well as fertilizer P is combined with Al and  Fe in
forms that are only sparingly soluble (Adams and Pearson  1967, Kamprath
and Foy 1972, Graham 1978,  Pritchett and Smith 1972).

     Unlike other micronutrients,  Mo is less available in strongly acid
soils (Kamprath  and Foy 1972).  Molybdenum  deficiencies such as  those
reported on the  Eastern Seaboard,  in the Great Lakes states, and on the
Pacific coast of the United States generally occur  on such soils (Kubota
1978).

2.3.3.3.2  Metal ion toxicities.   Any metal  can be  toxic  if soluble in
sufficient quantities.In near-neutral  soils, heavy metals occur  as
inorganic compounds or in bound forms with  organic  matter, clays,  or
hydrous oxides of Fe, Mn, and Al.   However,  a  decrease in soil pH  can
create metal toxicity problems for vegetation.  Zinc, Cu, and Ni
toxicities have  occurred frequently in  a variety of acid  soils.  Iron
toxicity occurs  only under flooded conditions  where Fe occurs as the
reduced, soluble Fe2+ form (Foy et al.  1978).   Toxicities of Pb, Co,
Be, As, and Cd occur only under very unusual conditions.  Lead and Cd
are of particular interest because they move into the  food chain and
affect human and animal health.  For further details,  see a recent
review (Foy et al. 1978) .

     Aluminum and Mn toxicities are the most prominent growth-limiting
factors in many, if not most, acidic soils  (Foy 1973,  1974b, 1981;
                                  2-34

-------
Tanaka and Hayakawa 1974).   Hence,  this review will  emphasize  the
harmful  effects of these two elements  on plants.   The chemistry  of  Al
and Mn in soils was discussed in Sections 2.2.3 and  2.2.4.

     2.3.3.3.2.1  Al uminum  toxi city.   Because Al  is  a structural
constituent of soil clay mineral  particles,  Al  toxicity  is  theoretically
possible in most,  if not all, soils.   The primary condition required  to
produce solubility of excess Al  is  a low pH.   As Section 2.2.3 pointed
out, aluminum may  become soluble enough to be of concern when  the  soil
pH is in the range 5.0 to 5.5 or below.

     Aluminum toxicity is believed  to  be a primary factor in limiting
plant root development (depth and branching)  in many acidic subsoils  of
the southeastern United States (Foy 1981).  For example, Kokorina  (1977)
noted that acid soil  toxicity was more harmful  in dry years.  This  dry
season phenomenon  in concert with acidic deposition  may  also be  a  factor
in Ulrich's (1980) recent reports on  forest growth reduction in  West
Germany.

     On the basis  of some complex theories of ecosystem  acidification
processes on and after a decade of  monitoring at the Soiling site,
scientists at the  University of Gottingen in  West Germany state  that  the
forests of Soiling (as well as others  like it in Germany) are  being
seriously impacted by acid  rain (Ulrich 1980).   Most significantly, at
the Soiling site Al concentrations  in  soil solutions have increased
twofold (from 1-2  mg £-1 to 2-5 mg  £-1) beneath the  beech stand
and  ~ tenfold (from 1-2 mg £-1 to 15-18 mg £-1) beneath  the
spruce stand over  the last decade (Matzner and Ulrich 1981).  It is
hypothesized that Al  concentrations are reaching toxic levels, thereby
damaging or killing tree roots and  causing serious consequences  to  the
maintenance of these forest ecosystems.  An important question relative
to toxicity of Al  levels concerns the  form of Al  in  soil  solution.  It
would be important to know  the extent  of chelation by organic  materials.

     Atmospheric H+ inputs  must be  viewed as  additions to natural,
internal acid generation (Ulrich 1980).  One  very important internal
H+ generating process at Soiling is nitrification in mineral soil
layers during warm, dry years.  Nitrification during these  periods
(thought to be caused by decomposition of previously accumulated N-rich
root residues) causes a pulse of acid  production. According to  Ulrich
et al. (1980), systems that have been  acidified by acid  precipitation
are unable to withstand such pulses because their buffering capacities
are much reduced.   Thus, Al is mobilized at such times,  creating toxic
conditions for roots.

     Undoubtedly,  acid inputs to the Soiling  site are very  high.   Inputs
of H+ measured with open-bucket collectors are not themselves
excessively high,  being approximately  700 eq  ha~l yr~l;
comparatively, H+  input values of this magnitude are not uncommon  in
forests of the United States (Evans et al. 1981). However, at Soiling
                                  2-35

-------
H+ flux in throughfall  is 2  to 5  times  greater than  in  open
precipitation due to dry deposition  in  the  forest canopy.

     In contrast to results  and hypotheses  at  Gottingen,  scientists with
the Norwegian SNSF Project demonstrated the ability  of  forest  ecosystems
to tolerate acid inputs and  Al  levels exceeding those reported at
Soiling.  This ability  is shown by  results  of  an intensive series of
irrigation studies involving inputs  of  H2$04 ranging from current
background levels (approximately  0.8 keq ha-1  yr-1)  up  to
approximately 30 times  that  amount  (26  keq  ha-1 yr-1).  Although Al
concentrations in soil  solutions  and in tree foliage increased
substantially, no indications of  Al  toxicity were noted and growth
effects were small (slight growth increases occurred in some species,
slight decreases in other species,  and  no effects in some species;
Abrahamsen 1980a,b; Tveite 1980a,b).  It is also noteworthy that large
nitrification pulses occurred in  most acid  treatments (Abrahamsen
1980a).  Finally, greenhouse studies involving acid  irrigation and
liming of Norway spruce showed that  this species (which occurs also at
the Soiling site) is extremely tolerant of  high acid inputs and foliar
Al concentrations.

     Plant species and  cultivars  differ widely in their tolerances to
excess Al in the growth medium.  Published  references to  such
differences are too numerous to cite individually, but  access  to the
older literature is provided in review  papers  (Foy 1974b, 1981).
Aluminum tolerance has  been  associated  with pH changes  in root zones, Al
trapping in non-metabolic sites within  plants, P uptake efficiency, Ca
and Mg uptake and transport, root cation exchange capacity, root
phosphatase activity, internal concentrations  of Si, NH4+ - NOs"
tolerance or preference, organic  acid contents, Fe uptake efficiency and
resistance to drought.   For citations  from  the earlier  literature, see
review papers (Foy 1974b 1981, Foy  and  Fleming 1978, Foy  et al. 1978).

     2.3.3.3.2.2  Manganese toxicity.   Manganese toxicity frequently
occurs in soils with pH values of 5.5  or below, if the  soil parent
materials are sufficiently high in  easily reducible  Mn  content.
However, some soils do  not contain sufficient  total  Mn  to produce
toxicity, even at pH 5.0 or below.   Soils of the Atlantic Coastal Plain
of the United States are lower in total Mn  than those of  the Gulf
Coastal Plain (Adams and Pearson  1967).  However, within  any area, soils
vary widely in Mn contents (Sedberry et al. 1978).   In  that study, the
DTPA extractable Mn varied more with parent material and  clay  than with
pH and organic matter.   Reducing  environments  induced by  poorly aerated
conditions in soils increase Mn availability and potential for toxicity.

2.3.4  Reversibility of Effects on Soil Chemistry

     Changes in soil chemistry caused  by acidic deposition in  unmanaged
terrestrial ecosystems  must, in general, be considered  irreversible, but
there are exceptions.  Nutrients  lost  are not  readily regained.
However, exchangeable basic cations in  surface soils may  be replaced
                                  2-36

-------
gradually by weathering,  by recycling by  deep rooted species,  and  by
dust inputs if the acidic inputs are reduced.   Since basic  cation
depletion is the normal,  long-term trend  in humid  regions,  the trend
toward increased acidity  would probably not be reversed  in  such
environments even if inputs stopped.

     Since microbial activity in soils responds quickly  to  changing
environments, important soil  processes it moderates  can  be  expected to
return to former levels when the environment changes as  a result of
reductions in deposition.

     Leaching of Al  to aquatic systems in response to acidic  inputs
would likely lessen with  reduced acidic deposition.

2.3.5  Predicting Which Soils will  be Affected Most

2.3.5.1  Soils Under Cultivation—It is unlikely that acidic
precipitation will adversely affect cultivated soils.  Not  only do many
management practices result in acid production greater than that
expected to be derived from acidic deposition, but good  management also
requires controlling pH at a level  most conducive  to plant  growth  (see
Section 2.2.6).  For example, NH4+ is an  important source of
fertilizer N to soils.  This form rapidly oxidizes to N03-  in  soil,
resulting in significant  acid production  (see Section 2.2.1).   Routine
additions of N fertilizers may result in  the release of  between one and
two orders of magnitude more H+ than will  be annually derived  from
acidic deposition (McFee  et al. 1976).

2.3.5.2  Uncultivated, Unamended Soils—As indicated in  the soil
chemistry section, 2.2.1.3, arid or semi-arid region soils  that are not
normally leached do not naturally acidify, and adding acidic  deposition
will not change that nor  cause any foreseeable ill effects.

     The soils that might be affected are those of the humid  regions,
which are not normally amended with lime  and/or fertilizers.   This area
includes most of the forested land of the eastern  United States, the
Pacific Northwest and some high altitude  areas of  the west.   It is
important to identify which soils in these regions are likely  to be
adversely affected by acidic deposition.

     Various schemes for  assessing site sensitivity  to acidic  deposition
effects have been proposed.  Those directed toward aquatic  effects have
emphasized bedrock geology (Hendry et al.  1980, Norton 1980),  while
those concerned with terrestrial effects  have emphasized cation exchange
capacity and base saturation (McFee 1980, Klopatek et al. 1980).   For
the reasons previously discussed, sulfate adsorption capacity  should be
included in the sensitivity criteria for  both aquatic and terrestrial
impacts (Johnson 1980), but unfortunately, the data  base for  the latter
is limited.  In considering soil sensitivity to adverse  effects of
acidic deposition, it is  helpful to separate the effects into  two
                                  2-37

-------
           Figure 2-4.   Generalized  soil map of the United States  (SCS USDA, 1975)  showing
                        regions  dominated by suborders or groups of suborders.  The most  common
                        suborder is  named.  Many other suborders exist within  the boundaries  of
                        each  area.

                        Al f i sol s                                Mol 1 i sol s
                          Al   Aqualfs                              Ml  Aquolls
                          A2   Boralfs                              M2  Borolls
                          A3   Udalfs                              M3  Udolls
                          A4   Ustalfs                              M4  Ustolls
                          A5   Xeralfs                              M5  Xerolls

                        Aridisols                               Spodosols
                          Dl   Argids                              SI  Aquods
                          02   Orthids                              S2  Orthods

^                       Entisols                                Ultisols
i                          El   Aquents                              Ul  Aquults
oo                         E2   Orthents                             U2  Humults
                          E3   Psamments                            U3  Udults

                        Histosols                               Vertisols
                          HI   Hemists                              VI  Uderts
                          H2   Hemists and Saprists                 V2  Usterts
                          H3   Fibrists, Hemists, and Saprists

                        Inceptisols
                          II   Andepts
                          12   Aquepts
                          13   Ochrepts
                          14   Umbrepts

-------
   U. S  DEPARTMENT OFAGRICUITURE
                                                            GENERAL  SOIL  MAP OF THE  UNITED  STATES
SOIL CONSERVATION SERVICE
ro
 I
OJ
                                                                                      E3c                  Vok
                                                                                                  0     100	7uO    300    '00    500    600 M.k
                                                                                                                                                  .H3a
                                                                                                                                                             Obi' .uv

-------
categories: (1) changes related to soil pH-basic cation changes,  which
would include any direct losses of nutrients and changes in processes or
availability related to pH;  (2) changes in soil solution and/or
leachate chemistry that might affect aquatic systems or be toxic  to
plant roots, for which the primary concern is change in aluminum
concentration in solution.

     McFee (1980) has suggested that cation exchange capacity (CEC) be
used as the primary criterion for determining soil  sensitivity to acidic
deposition.  The suggested classification considers soils with CEC
greater than 15.4 meq 100 g-1,  those subject to frequent flooding, or
those with free carbonates in the upper 25 cm of the sol urn to be
insensitive.  Non-calcareous, non-alluvial soils with CEC between 6.2
and 15.4 meq 100 g-1 are classed as slightly sensitive, and those with
CEC less than 6.2 meq 100 g-1 are classified as sensitive.

     Wiklander (1974, 1980b)  proposed a more complex classification
system, which considers soil  buffering capacity as  well  as the ability
of H+ to compete for exchange sites in low pH, low  base saturated
soils.  Buffering capacity will, of course, be directly affected  by CEC
as well as by pH, base saturation, and the presence of carbonates and
ferromagnesium minerals.   Considering base saturation separately
recognizes that H+ competes best with base ions on  pH-dependent charge
sites (Snyder et al. 1969, McLean and Bittencourt 1973).   As base
saturation decreases and a larger proportion of the pH-dependent  charge
sites are filled with acidic  ions, H+ inputs become less  effective in
removing basic cations.

     Wiklander1s classification scheme still does not include all  known
factors that moderate effects of acidic deposition.  For  example,
Wiklander (1975, 1980a,b)  demonstrated that the presence  of neutral
salts, either in the precipitation or in the soil,  significantly
moderates the effect of acidic  precipitation on soil.  Sulfate
adsorption capacity  of the soil  should also be considered because  mobile
sulfate serves as a  counter ion for cation leaching (Cronan et al.  1978,
Johnson 1980).   Many acid soils have an anion retentive capacity  which
can be related to both the presence of hydrated Fe  and Al  oxides  and to
charge of the soil with decreased pH (Wiklander 1980a).   High sulfate
adsorption capacity  will  decrease soil  sensitivity  to cation removal.

     Comparisons of  above systems indicate weakness in all,  but a
tendency to agree when viewed on a national  scale.   The regions
dominated by Ultisols, Spodosols and some of the Inceptisols (Figure
2-4)  encompass most  of the areas predicted to be sensitive to acidic
deposition.  All mapping  efforts at any level  above the most detailed
(county soil  maps for example)  will  of necessity include  a wide range of
conditions within any map unit.   For that reason, all  of  the efforts
published thus far should be  used with  some caution.

2.-3.5.2.1  Basic cation-pH changes in forested soils.   Based on the
sensitivity criteria proposed by McFee (1980), Wiklander  (1980b),  and
                                  2-40

-------
Johnson (1980), it is clear that soils likely to undergo  significant
changes in basic cation content or change in pH  have  these
characteristics:

     (1)  they are not renewed by flooding or other processes;

     (2)  they are free of carbonates to considerable depth  (1.0  meter
          or more);

     (3)  they have low CEC but pH of at least 5.5  to 6.0; and

     (4)  they have a low sulfate adsorption capacity.

Because soils with low CEC (< 6.0 meq 100 g-1, McFee  1980, Klopatek  et
al.  1980)  in humid climates tend to become acid  naturally over  time,  few
soils meet criteria 3 above.   So few have, in fact, that  by  the time we
apply the other criteria, it is clear that accelerated loss  of  basic
cations and lowered soil  pH as a result of acidic deposition are
unlikely to be extensive problems.  Maps prepared by  01 sen et al.  (1982)
show areas of low CEC and moderately high pH that are extensive enough
to appear on a national map,  only in the central  portion  of  the United
States.  In that area, however, most soils do not meet criteria 2  and do
not currently receive significant acidic deposition.

2.3.5.2.2   Changes in aluminum or other metal  concentration  in  soil
solution in forested soils.  Based on the discussion  of soil  chemistry
in Section 2.2.3, it is clear that soils most likely  to have increased
Al in solution or in leachate due to acidic deposition are already acid,
(< pH 5.5), and meet criteria 1, 2, and 4 above.  Cation  exchange
capacity is not as important in this case, but effects will  be  most
pronounced where CEC is low.   In such soils, the buffer capacity  is
largely controlled by Al-mineral chemistry.  Increased acidic inputs  may
increase the rate of Al release and increase its concentration  in  soil
solution or leachate from the soil.  This is most likely  to  occur  where
total quantity of the controlling Al compounds exposed to chemical
action is  small, e.g., in a coarse-textured acid soil.

2.4   EFFECTS OF ACIDIC DEPOSITION ON SOIL BIOLOGY

2.4.1  Soil Biology Components and Functional  Significance

     The biological component of soil  is of primary importance  in  the
functioning of the complete ecosystem.   In this  section,  the soil  biota
will be briefly described in  terms of functional  significance.  For
general reference, see Alexander (1980a), Richards  (1974), or Gray and
Williams (1971).

2.4.1.1  Soil Animals—The most significant roles played  by  the
invertebrate soil fauna pertain to turnover of organic material and  soil
physical characteristics.  Many members of this  group, such  as
earthworms, mites, ants,  and termites are involved  in mixing the  organic
                                  2-41

-------
and inorganic soil constituents.   The quantity of organic material
actually assimilated by these organisms is small, generally less than 10
percent, but the relatively large quantity of material  consumed is
frequently altered chemically by  enzymes or microorganisms present  in
the animal's gut.   Thus, by maceration and mixing, these organisms  play
an important role  in the conversion of plant material  to soil  humus.

2.4.1.2  Algae--Chlorophyta (green algae), Cyanobacteria (blue-green
algae) and Chrysophyta (diatoms)  are common inhabitants of the soil
surface.  Since algae are dominantly photoautotrophic  organisms (using
light as an energy source and C02 as a carbon source)  they can
colonize environments lacking the organic carbon required by many life
forms.  In areas where higher life forms are largely absent, such as
fresh volcanic deposits, beach sands, eroded areas, and freshly burned
areas, algae commonly appear as the pioneering species, frequently
supplying the organic material  required for subsequent colonization by
other life forms.   Some blue-green algae (bacteria) can convert
atmospheric N2 to  organic compounds.  In many environments, such as
flooded paddy fields, this ability to fix nitrogen provides a critical
input of nitrogen  to the system.   Lichens, an intimate association
between certain algae and fungi,  are also important pioneering species,
and some have the  ability to fix  nitrogen.  Ubiquitous on rock surfaces
and other extremely harsh environments, lichens are instrumental in  the
long-term breakdown and dissolution of rocks ultimately to form soil.

2.4.1.3  Fungi—Soil fungi are involved in degrading a wide range of
organic compounds, from simple sugars to complex organic polymers.  Many
members of this group possess the enzymatic capacity to attack the  major
plant constituents, such as cellulose, hemicellulose,  and lignin.  Fungi
are normally the dominant initial colonizers of plant  debris and are
ultimately responsible for many of the steps occurring during the
conversion of plant material to soil organic matter.  The complex
network of fungal  hyphae which totally permeates the fabric of soil
constitutes a major portion of the soil biomass as well as binding
together soil particles to form aggregates.  Products  of fungal
metabolism in soil, such as carbohydrates, can act as  glues for primary
soil particles.

     Certain types of soil fungi  can play direct roles in nutrient
availability to plants by forming mycorrhizal associations with plant
roots.  The fungal hyphae greatly expand the volume of soil from which
plant roots can effectively draw  nutrients.  In deficient soils, the
fungal partner can substantially  improve phosphorus, copper, zinc,  and
possibly nitrogen   (ammonium) availability to plants.  In addition,  the
mycorrhizal association may enhance water availability, increase salt
tolerance, enhance heavy metal  resistance, and affect plant growth  via
hormone production.  Although relationships are not yet well understood,
each of these effects is currently under investigation.

2.4.1.4  Bacteria--The procaryotic microflora of soils are also
extremely important in the decomposition of plant litter and the
                                  2-42

-------
synthesis and breakdown of soil  organic matter.   Bacteria are primarily
responsible for making organic forms of N,  S,  and P  available to  plants
by mineralizing organic matter.   For substantial  plant uptake to  occur,
S must be as S042- and N as either N03~ or  NH4+.   Oxidation
of NH4+ to N03~ (nitrification)  is dominantly  catalyzed by
autotrophic soil bacteria.  Nitrogen is lost from the soil through
anaerobic bacterial reduction of N(h- to the gaseous species N2
and N20 (denitrification).  Most nitrogen enters  ecosystems  through
bacterial reduction of atmospheric N£ to NHA+  (No-fixation).
Fixation by bacteria living symbiotically with plants can contribute
significant amounts of nitrogen to both agricultural and forest  systems.
Nitrogen nutrition of many leguminous plants is enhanced through
N2~-fixation by bacteria of the genus Rhizobium.   Fixation by
actinomycetes, such as Frankia.  in association with  woody species may
contribute critical amounts of nitrogen to  some forest systems.   The
oxidation and reduction of S roughly parallel  that of N.  In addition to
bearing primary responsibility for the availability  of N and S to
plants, soil microbes also strongly influence  the availability of
phosphorus, iron, and manganese through organic mineralizations  and
redox reactions.

     The distribution of microbial activity in soil  generally reflects
the fact that many of these microbes are heterotrophs, that  is,  they
require preformed organic compounds.  Soil  microbial activity is
generally greatest in regions of high organic  carbon availability.
While most types of microbial activity do occur to some extent
throughout the soil profile, recognizing that maximal activity commonly
occurs in somewhat discrete areas of the soil  is  important to
understanding potential effects of acidic deposition.  Microbial  attack
on plant debris takes place largely in the  surface litter layer.
Production and breakdown of soil humus occur dominantly in the upper
portion of the soil profile, reflecting the site  of  initial  leaf, stem,
and root material deposition.  Heterotrophic microbial activity  is also
high in soil near plant roots, where root-derived material provides
carbon for soil bacteria and fungi.

2.4.2  Direct Effects of Acidic Deposition  on  Soil Biology

     The effects of acidic deposition should be expected to  vary
tremendously, depending on the type of organism and  the characteristics
of the soil which it inhabits.  While soil  acidification does affect
many biological processes, it is often impossible to distinguish  direct
effects of acidification from secondary effects resulting from
acid-induced changes in the soil solution.   The following section
documents some effects which have been attributed to soil acidification
resulting from acid inputs.

2.4.2.1  Soil AnimalIs—Many classes of soil animals, such as earthworms
(Lumbricidae).millipedes  (Myriapoda), and nematodes (Nematoda),  are
known to be less abundant in acid soils than in neutral soils.  However,
high populations of other soil animals, such as springtails  (Collembola)
                                  2-43

-------
and potworms (Enchytraeidae),  are common in acid soils  high  in  organic
matter (Richards 1974).

     Effects of simulated acid precipitation on  soil  fauna vary markedly
according to the species observed.   Studies by Baath  et al.  (1980),  in
which soils were treated with  50 or 150 kg ha~l  H2S04 for 6  years,
showed that the numbers  of Collembola increased, Enchytraeidae
decreased, but mites (Acarina) were generally unaffected by  both
application rates.   In a two-year exposure to simulated rain of pH 2.5
to 6.0 (25 or 50 mm per  month),  Collembola, Acarina,  and Enchytraeidae
were generally unaffected or increased in number with the acid
treatments.  However,  a  few species of Acarina and the  dominant
Enchytraeid were significantly reduced by the more extreme acidification
(Hagvar 1978, Abrahamsen et al.  1980).  It should be  noted that the
soils studied by these two groups were naturally very acidic; hence  the
indigenous soil fauna  may have been relatively acid tolerant.   In less
acid deciduous woodland  soils  (Kilham and Wainwright  1981),  the native
population of soil  animals appeared to be much more sensitive to acid
rain (pH 3.0) localized  near a coking works, but these  results  also
reflect the presence of  substantial dry deposition on the litter.

2.4.2.2  Terrestrial Algae—While green algae (Chlorophyta)  readily
colonize Telatively acid soils,  blue-green algae (Cyanobacteria) have
been reported to be particularly sensitive to soil  acidity (Dooley and
Houghton 1973, Wilson and Alexander 1979).  While there is little
experimental verification in soil systems, the general  sensitivity of
free-living Cyanobacteria to acidity suggests they may  be susceptible to
acidic deposition.   The  sensitivity of blue-green algae to acid
precipitation has been demonstrated in a lichen symbiosis.   Simulated
acidic deposition of pH  4.0 or less substantially reduced ^-fixation
by the dominant No-fixing lichen in a deciduous forest  (Denison et al.
1977).

2.4.2.3  Fungi—Fungi  become increasingly important in  acid  soils as
compared to neutral-alkaline soils (Gray and Williams 1971). The
commonly observed dominance of fungi over bacteria in acid soils may, in
part, result from a greater sensitivity of heterotrophic bacteria to
H+ concentration and the consequent reduction in competition
(Alexander 1980a).

     The relative tolerance of fungi to acid precipitation was
demonstrated by Wainwright (1979),  who isolated fewer heterotrophic
bacteria but more fungi  from soils exposed to acid rain and  heavy
atmospheric pollution than from similar but unaffected  soils.   The
presence of nitrifying fungi  in acid soils lacking autotrophic
nitrifiers (Remacle 1977, Johnsrud 1978) also appears to indicate the
relative resistance of fungi  to soil acidity.

     Most investigations of the effects of acidic deposition on soil
fungi, however, have used traditional  plate count methods, which do  not
                                  2-44

-------
necessarily reflect viable fungal  biomass.   Baath et al.  (1980)  found
that FDA (fluorescein diacetate)  active fungal  biomass  decreased
significantly under the two acid  regimes described earlier (Section
2.4.2.1) while total  fungal mycelia (the sum of viable  and non-viable
hyphae) increased.

     To date, little information  available concerns the response of
mycorrhizal associations to acidic deposition.   Sobotka (1974)  reported
a reduction in the fungal  mantle  of spruce mycorrhizae  receiving heavy
atmospheric pollution, including  acid rain.   In a short-term experiment,
Haines and Best (1975) found no visible damage  to endomycorrhizae of
sweetgum exposed to pH 3.0 treatments.   To explain deviations in
nutrient flux data, these researchers suggested that cation carriers of
mycorrhizal roots may be more susceptible to inhibition by H+ than are
non-mycorrhizal roots.

2.4.2.4  Bacteria—The discussion in this section pertains largely to
soil bacteria.  In  many soil  microbial  processes, however, it is
impossible or meaningless to isolate bacterial  functions  from soil
fungal and faunal  processes with  which  they  are inherently integrated.
For example, leaf litter decomposition  requires fungal,  bacterial, and
faunal attack.

     Bacteria are generally considered  to be less acid  tolerant than
fungi.  Some bacteria, however, are extremely acid tolerant.   For
example, species of the chemoautotrophic thiobacilli  can  survive at  pH
0.6 and thrive at pH 2.0 (Butlin  and Postgate 1954).

     Acidic deposition may affect heterotrophic bacteria  in soil  by
causing changes in  total  numbers  and/or species composition.   Francis et
al. (1980) reported that the total number of bacteria and actinomycetes
generally declined  in soil  acidified from pH 4.6 to 3.0 with an  addition
of H2$04, although  the magnitude  of these effects was not reported.
In soils transferred to a site receiving pH  3.0 rain  and  dry deposition,
Wainwright (1980)  found that over a one-year period bacterial  numbers
did not change significantly, even though the soil  pH fell  from 4.2  to
3.7.  Baath et al.  (1980)  noted a shift towards spore-forming bacteria
in soils receiving  H2S04 inputs for six years as compared to control
soils, suggesting a response to adverse conditions.   In the same
experimental series,  total  bacterial  numbers (by plate  counts)  did not
change, but bacterial biomass and FDA-active bacteria did decrease with
increasing severity of treatment  (Baath et al.  1979,  1980).

2.4.2.5  General Biological Processes—Net heterotrophic  activity
(bacterial, fungal, and faunal) and the rate of organic matter
decomposition are commonly determined by measuring C02  evolution.  The
rate of glucose mineralization was reduced in surface soils receiving
100 cm of simulated rain (pH 3.2  and 4.1),  continually  or
intermittently, over a 7-week period (Strayer and Alexander 1981).
However, the 7-week treatments caused less significant  effects  than  did
the continuous exposure,  and the  reductions  were less severe in soils of
                                  2-45

-------
greater natural  acidity.   The authors therefore suggested that  some
microbial  adaptation was  occurring over time.

     Respiration in soils transferred to a site receiving pH  3.0  rain
was reduced by 50 percent after a one-year exposure (Wainwright 1980).
Similar effects  of simulated acid precipitation have  also been  reported
by Tamm et al. (1977).   Observed effects of simulated acid precipitation
on litter decomposition are summarized in Section  2.5.

     Several  reports now indicate that acid inputs can  slightly
accelerate mineralization of organic nitrogen  (Wainwright 1980, Strayer
et al.  1981)   Tamm et al. (1977) similarly found increased accumulation
of NH4+ in acid-treated humus samples, but they interpreted this  to
mean that immobilization  was more retarded than mineralization  (a
hypothesis for which no substantiating data existed).  Conversely,
Francis et al. (1980) found lower NH4+ production  in  a  soil  that  had
received an addition of ^04.   For all of this work, the treatment
periods were  relatively short (from 1 hour to  1 year);  longer exposures
may yield more consistent results.  The data,  however,  are compatible
with the fact that "natural" soil  acidity does not have a uniform effect
on N-mineralization (Alexander 1980b).

     Since nitrification is generally believed to  be  catalyzed  by
relatively autotrophic  nitrifiers (known to be acid-sensitive on  labora-
tory media),  researchers have predicted that this  process should  be  one
of the microbial processes most sensitive to acid  precipitation (Tamm
1976, Alexander 1980b).  While evidence indicates  that acid inputs to
soil inhibit  autotrophic  nitrification, the overall effects on  NH4
oxidation to  N0o~ are neither uniform nor easily interpreted.
Francis et al. (1980) could detect little nitrifying  activity in  the
naturally acid forest soil studies (pH 4.6) or in  the soil sample that
had received  an  addition  of ^$04, but they concluded that further
acidification of an acid forest soil would lead to a  significant
reduction in  nitrification.  Wainwright (1980) found  essentially  no
effect on nitrifying activity in a soil exposed to acid rain (pH  3.0)
from a coking works.  Strayer et al. (1981) examined  the effects  of
acute acidification on nitrification in surface soil  from soil  columns
and found interesting but somewhat complex results.   When high  NH4
amendments (100 ppm-N)  were added to the nitrification assay, all acid
treatments tested (pH 3.3 to 4.1) caused substantial  reductions in
nitrification rates.  However,  when NH4+ was not added to the soil,
the acid treatments caused no detectable effect, or in  some cases,
caused a slight stimulation in N03~ production.  Since forest
soils would be expected to have relatively low natural  concentrations of
NH4+, the authors conclude that short-term exposures  to acid rain
should not substantially affect nitrification  in forest soils.  The
results reported by Strayer et al. (1981) are consistent with the
occurrence of heterotrophic nitrifying organisms in naturally acidic
forest soils; these heterotrophic nitrifiers are considered much  less
sensitive to  acidity than are autotrophic nitrifiers  (Remade 1977,
Johnsrud 1978).
                                  2-46

-------
     Little published data concern effects of acidic deposition on soil
denitrification.   While slight soil  acidification may not alter the
overall rate of this process,  it should be expected to increase NgO
production relative to N£ (Firestone et al.  1980).

     A substantial  amount of work on the sensitivity of N2-fi*ation by
legume-Rhizobium associations to soil acidity has been published.   In
some cases, the bacterial symbiont appears to be sensitive to acidity
(Bromfield and Jones 1980, Lowendorf et al.  1981);  in other cases, the
nodule formation or activity are affected (Evans et al. 1980, Munns et
al. 1981).  However, work on the effects of acidic deposition on No-
fixation by legumes is scant.   Shriner and Johnston (1981) reported that
simulated rain of pH 3.2 applied for 1 to 9 weeks caused decreased
nodulation in kidney beans.  The authors suggest that similar effects
would be unlikely to occur under normal agricultural management
practices but might be expected to occur in natural, unmanaged
ecosystems (Shriner and Johnston 1981).  No data are available
concerning effects of acid rain on the associations of actinomycetes
with woody plants.

2.4.3  Metals—Mobilization Effects on Soil  Biology

     Two questions concerning mobilization of metals and effects on soil
biology must be addressed.  First, the input of acidity to soil  can
cause mobilization of Al and Mn from mineral forms indigenous to the
soil.  Can mobilization of Al  and Mn by acid inputs be expected to  have
toxic effects on the soil biota?  Second, acidic deposition is sometimes
accompanied by atmospheric deposition of various heavy metals. Does the
acidity of the rain increase the potential toxicity of these metals?
While few data available directly or realistically address these
potential effects of acidic deposition, a small body of pertinent
background literature exists.

     The toxicity of available Al to soil microbial activity has been
reported by Mutatkar and Pritchett (1966), who found that additions of
Al to soils with pH maintained below 4.0 significantly reduced the rate
of soil respiration.  Ko and Hora (1972) have identified Al3+ ions as
being fungitoxic in acid soil  extracts.  These workers found germination
of ascospores to be totally inhibited by aqueous solutions (pH 4.8)
containing as little as 0.65 ppm Al.  They did not identify Mn as  toxic
to the fungi tested, but the concentrations of this metal in the soil
extracts examined were low compared to Al concentrations.  In studies
dealing with the growth of the Rhizobiurn-bean symbiosis in acid tropical
soils, Dobereiner (1966) found that additions of 40 ppm Mn to acid soils
reduced either N2-fixation efficiency or nodule numbers.  Since
preliminary evidence suggests that the threshold concentrations for
toxicity of mobilized aluminum are relatively low, such an indirect
consequence of acid input to soil may be a possibility.  However,  acid
rain, within current pH limits, has not been shown to mobilize these
metals in quantities toxic to soil biota.
                                  2-47

-------
     Soils in the vicinity of metal-smelting and coal-burning are likely
to be subject to atmospheric deposition of heavy metals (Little and
Martin 1972, Freedman and Hutchinson 1980) in addition to acidic
deposition.  The input of heavy metals to these soils is significant
because metal solubilization and biological  toxicity are pH dependent.
Numerous pure culture studies demonstrate increasing metal  toxicity with
decreasing pH of solution (e.g., Babich and Stotzky 1979).   However,
many of these studies should not be extrapolated to soils because of the
complexity of the metal cation interactions with soil constituents.
Babich and Stotzky (1977) found that Cd toxicity to microbes in soil was
a function of soil pH; however, this may  have been an anomaly, since
toxicity increased with increasing soil  pH.

     Metals vary in potential toxicity;  work by Somers (1961)  indicated
that the microbial toxicity of heavy metals is highly correlated with
the electronegativity of the metal.  When attempting to assess the
potential effects of acidic deposition in association with metal
deposition, one must consider several  factors:   1) the toxicity
potential of the metal, 2) the quantities and speciation of metals
deposited and degree of association with  acid inputs, and 3) the pH
dependence of metal toxicity in the recipient soil environment.
Mobilization of metal ions in soils receiving acid inputs,  and
subsequent toxicity of these metals, may  be a mechanism by which acidic
deposition affects soil biological  activity, but experimental  evidence
is lacking.

     Apparently certain plant-microbial  associations are able to
protect plants from metal toxicity.  Bradley et al. (1981)  found that
mycorrhizal infection of an ericaceous,  Calluna species reduced heavy
metal uptake by the plant.  The authors suggested that protection by the
fungal symbiont allowed this species to colonize heathland soils in
which the low pH increases availability of metal  cations to levels which
are toxic to many non-ericaceous species.

2.4.4  Effects of Changes in Microbial Activity on Aquatic Systems

     Because our current understanding of the effects of acidic
deposition on microbial activity in terrestrial ecosystems is limited,
extrapolations to possible secondary effects on aquatic systems are
tenuous at best.  It is important to recognize, however, that even a
small change in microbial activity in soil may cause profound changes in
aquatic systems, into which much of the soil water will  ultimately
drain.

2.4.5  Soil Biology Summary

     The following statements represent simplifications of complex and
sometimes contradictory trends in the existing data.  They reflect both
the complexity of microbial processes and the variability in
experimental protocols.  The extreme variability in pH and ionic
                                  2-48

-------
 composition of simulated rain, as well as differences in important soil
 characteristics, makes comparing data difficult.  Treatment durations in
 the experiments reported ranged from 1 hour to 6 years.  Short-term
 "accelerated" treatments may not only overlook potential long-term
 effects, but also may yield misleading predictions.  The shortcomings of
 long-duration experiments involving infrequent sampling should also be
 recognized.  Acid rain rarely occurs in isolation, rather it occurs in
 association with other pollutants such as heavy metals and the gaseous
 precursors of acid species.  The potential synergisms among these
 pollutants should not be overlooked.  The following statements summarize
 or interpret the limited data available and should be read with the
 above-mentioned limitations in mind.

     Acidic deposition will not substantially affect soil  biological
 activity in cultivated soils because of the much greater influence of
 soil amendments.

     The following statements pertain to uncultivated soil  systems:

    °   The effects of acidic deposition on animals in strongly acid
        soilsare not significant.   In less acid soils, pH  3.0 simulated
        rain significant changes in litter animals.

    o   Certain types of soil  microbial  activity are more  sensitive to
        soil acidity than are others.  Soil  fungi are probably the
        components of the soil  biota least sensitive to acid inputs;
        but little is known about effects on mycorrhizal symbionts.

    0   Preliminary evidence indicates that No-fixation by  lichens
        is inhibited by rain of pH less  than 4.0.  The evidence for
        acidic deposition influences on  Rhizobium or actinomycete
        symbiotic N-fixation is insufficient for a conclusion.

    o   Autotrophic nitrification  in surface soils is reduced by
        artificial  acid inputs; however,  no evidence exists to prove
        that acidic deposition  at  the rates  currently common in the
        United States will  cause such a  decrease.  Net nitrification
        may not be similarly decreased because of the acid  tolerance
        of heterotrophic nitrifiers.

    0   Slight increases and decreases in N-mineralization  rates
        result from treatments  of  short  duration, but little direct
        evidence  concerning long-term responses to realistic inputs
        exists.

2.5   EFFECTS OF ACIDIC  DEPOSITION  ON  ORGANIC  MATTER DECOMPOSITION

     One of the long-standing hypotheses  regarding the environmental
effects of acidic deposition has been that increased acid loading to
                                  2-49

-------
forest soils will  result in decreased decomposition  rates  for  organic
matter.  This hypothesis has been  addressed by  a  number  of investigators
(Tamm et al. 1977;  Abrahamsen et al.  1976,  1980;  Abrahamsen and Dollard
1978; Alexander 1980a,b; Baath et  al.  1980; Cronan 1980a,b;  Hovland  et
al. 1980; Francis et al. 1980; Lohm  1980;  Roberts et al. 1980; Hovland
1981; Kilham and Wainwright 1981;  Strayer  and Alexander  1981;  and
Strayer et al. 1981).  Unfortunately  the results  from these studies  have
appeared mixed and inconsistent (Table 2-6).  However,  if  one  screens
the published studies and selectively excludes  the results from those
investigations that represent extremely acute treatments,  then the
following summary statements emerge.

     (1)   Most decomposition studies related to  acidic  deposition  have
           been conducted with coniferous  litter  materials.

     (2)   Results suggest that it is important to interpret data  from
           decomposition studies in  relation to H+ loading and
           not simply with respect to the  pH of the  artificial rain
           treatments.

     (3)   It is important to distinguish  between the physical-chemical
           and the biological components of organic  decomposition.
           Based upon shorter-term studies (2 to 4 months  or less),  it
           has been shown that increased H+ loading  generally  will
           increase leaching of cations and organic  constituents  from
           forest litter.  This response may help to explain why  acidic
           precipitation treatments increase the initial rate  of  weight
           loss in some experiments.   Over the  longer term (>  4
           months), it appears that the biologically-mediated
           mineralization of organic matter in  forest soils will  be
           only slightly inhibited by acidic deposition (< 1 to 2
           percentdecrease in decomposition rate).

     (4)   Overall, unless average precipitation inputs were to drop to
           pH 3.0 or below, one would not expect significant impacts of
           acidic deposition on litter decomposition.

2.6  EFFECTS OF SOILS ON THE CHEMISTRY OF AQUATIC ECOSYSTEMS

     Much of the evidence for atmospheric depositions'  contribution to
surface  water acidification, while convincing in many cases (e.g.,  N.  M.
Johnson  1979), is circumstantial.   Only recently have efforts  been made
to establish the mechanisms by which atmospheric acid inputs are
transferred to aquatic ecosystems (Abrahamsen et al. 1979, Seip  1980,  N.
M. Johnson et al. 1981).  If acidic precipitation passes through  soil
prior  to entering an aquatic ecosystem, it will usually be strongly
influenced by the chemical nature of the soil.   Even barren rock  has
some influence on the chemistry of runoff water (Abrahamsen et al.
1979).   The pH of water  leaving the soil is not necessarily the  same as
the  soil solution pH in  intimate contact with the soil.
                                  2-50

-------
        TABLE  2-6.   REVIEW  OF  STUDIES  CONCERNED  WITH THE  IMPACT  OF ACIDIC  DEPOSITION  ON  ORGANIC DECOMPOSITION
                    Author
Soil  Type
 Duration
    of
Experiment
Treatments
Results
          1.  Abrahamsen et al. 1980   Lodgepole pine needles     75-90 days
                                      Norway spruce needles
                    3-9 mos.
ro
 i
en
                                      Raw coniferous humus   •  unspecified
           2.  Abrahamsen and Dollard 1978   General Review
          3.  Abrahamsen et al.  1976
                                           Lodgepole pine needles   90 days
             Needles from field experiments
               at pH 5.6 and 3.0 were  incu-
               bated in moist condition and
               weighed.

             Spruce needles in lysimeters
               were watered 2x weekly  with
               pH 5.6, 3, or 2 water at a
               rate of 100 mm mo.'1  or 200
                                                                             nun mo.
                                                                                   -1
                                Raw humus in litterbags ex-
                                 posed to pH 5.3, 4.3, and
                                 3.5 treatments.
                                   Needles moistened with di-
                                     lute H2$04 solutions.
                                           Cellulose/Wood
                                                                  Unspecified
                                                                Acid treatment increased  decom-
                                                                  position-29% greater at pH  3
                                                                  than 5.6
                       Relatively small  effects  from
                         acid treatments.   No  signifi-
                         cance at 100 mm mo.~l.   At
                         200 mm mo."1, the  pH  3  and 2
                         treatments decreased  decompo-
                         sition by < 5%

                       Increased leaching of K,  Mq, Mn,
                         Ca.

                       pH 4.3 treatment caused 8% de-
                         crease in decomposition  rate,
                         while pH 3.5 caused 10%
                         decrease.

                       Decomposition of organic  matter
                         in acidic coniferous  forest
                         soils  is apparently  only
                         slightly sensitive to acidifi-
                         cation.  Decomposition  of
                         fresh  litter and cellulose  is
                         influenced only at pH _< 3.

                       Decomposition was depressed at
                         pH 1.8 as compared to 3.5.  No
                         difference between pH 3.5 and
                         4.0
                                   Unspecified acid treatments     No consistent trends.

-------
                                                                TABLE 2-6.   CONTINUED
                      Author
 Soil  Type
 Duration
    of
Experiment
                                                                                           Treatments
                                                                                                                             Results
            4.  Alexander 1980a
                Strayer and Alexander
                   1981
Honeoye silt  loam  (pH 7.1)   2+ wk.
           Soils  were exposed  to  pH  4.1  and
             3.2  acid rain  treatments and
             were incubated with  C^
             glucose.
                                pH 4.1 treatment had no
                                  effect on glucose
                                  mi neralization
                                  pH 3.2 treatnent decreased
                                  glucose mineralization rate
                                  by 30-66%.
            5.  Alexander 1980b
Spodosols  from the           14-61
  central  Adirondacks        days
           Soils  were exposed  to  100  cm  of
             pH 3.5 and  5.6  artificial rain
             for  14 consecutive or  35
             intermittent  days.
ro
tn
ro
                                In 14 day consecutive rain,
                                  the rate of total  organic
                                  carbon (TOC) leaching was
                                  initially greater at pH 3.5
                                  than 5.6.  This later re-
                                  versed.

                                In 35 day intermittent treat-
                                  ment:  pH 3.5 leached more
                                  TOC than pH 5.6.
              6.  Baath et al.  1980.
  Coniferous iron Podzol
   12 mo.
              7.  Cronan 1980a.
  Coniferous forest floor    4 mo.
Litterbags  were placed  in
  field  plots  exposed to
  H2S04  treatments  P  50
  and 150 kg ha'1.
            Forest floor microcosms were
              exposed to pH 5.7, 4.0 and
              3.5 artificial rains.
C02 evolution response variecf
  with soil  pH -- inhibition  in
  more acid  soils,  but
  stimulation by  pH 3.5 rain  in
  less acid  soil.

No significant difference com-
  pared to controls for Scots
  pine needle litter.

Root litter  exposed to 150 kg
  ha~l had 21* decrease in
  decomposition rate.

Increased rainfall  acidity
  caused increased  leaching of
  Ca, Mq, K,  and  NH4 +.
  Compared to the pH  4
  treatment,  the  pH 3.5 rain
  caused 50-150%  more  K, Ca,  and
  Mg leaching.

-------
                                                               TABLE  2-6.   CONTINUED
                       Author
                                                   Soil  Type
                           Duration
                              of
                          Experiment
         Treatments
                                           Results
             8.  Cronan 1980b
                                              Coniferous and hardwood
                                                forest floors
                            3 mo.
             9.  Hovland 1981
IX)
 i
en
CO
                                              Norway spruce needle
                                                litter
                            5 yr.
 Forest  floor microcosms were
   subjected to weekly 3.5 cm
   simulated rains at pH 5.7
   and 4.0
 Field  plots were exposed to
   pH 6.1, 4.0, 3.0, and 2.5
   rains over  5 yr.  Litter
   collected from these plots
   was  assayed.
 Hardwood  forest  floors showed
   60% more Ca  leachinq and 65%
   more Mq leachinq at pH 4.0.
   Coniferous forest floors
   showed  40% more Ca and 25%
   more Mg leaching at pH 4
   compared to  pH 5.7.  In
   general, cation fluxes from
   the hardwood litter were much
   greater than from coniferous
   litter.

 Acid rain' treatments produced
   very little  effect on biolo-
   gical activity in litter as
   measured by  respiration and
   cellulose activity.
            10.  Hovland et al. 1980
Norway  spruce needles    16-38 wk.
Lysimeters containing spruce
  needles were exposed to pH
  5.6, 3.0 and 2.0 solutions
  at 100 and 200 mm mo"^.
Small  effects on decomposition.
  Treatments at pH 3 and 2 ini-
  tially increased the decompo-
  sition rate at 100 mm mo"1.
  After 38 wk., decomposition
  had  decreased relative to
  controls in pH 3 and 2 treat-
  ments at 200 mm mo'l.
                                                                                                                Effect of acid treatments on
                                                                                                                  monosaccharide content was not
                                                                                                                  consistent.  However, there
                                                                                                                  was an indication of reduced
                                                                                                                  lignin decomposition at 200
                                                                                                                  mm mo~l for pH 3 and 2.

                                                                                                                Acid treatments caused increased
                                                                                                                  leaching of Mg, Mn, and Ca.

                                                                                                                Initially, acid rains decreased
                                                                                                                  P leaching; later, this
                                                                                                                  reversed.

-------
                                                                   TABLE  2-6.   CONTINUED

Author

Duration
Soil Type of Treatments
Experiment

Results

ro
 i
en
             11.  Francis et al.  1980
             12.  Lohm 1980
13.   Roberts  et  al.  1980
             14.  Tamm et al. 1976
                                 Oak-pine sandy  loam  (pH 4.6)   5 mo.
                                 Coniferous  iron  Podzol
Coniferous  Podzol
                                 Coniferous  Porizol
                              6 yr.
5 mo.
                             5-6 yr.
Soils were adjusted  with
  acid or base to give  a
  soil pH of 3.0 or  7.0,
  and were then incubated
  with controls.

Plots were exposed to 0,
  50, and 150 kg ha"1
  H2$04 per yr.
  Litter bags were exposed
  for 2 yr.

Field plots were subjected
  to biweekly 5 mm appli-
  cations of pH 3.1  and
  2.7 acid rain.
        Field  plots  received 0,
          50,  and  100  kg  ha'1
          yr~l  applications
          of H2S04.
                                                                   The acidified soil showed 6-52%
                                                                     less C02 production, depend-
                                                                     ing upon amendments.
                                    Acid treatments lowered the
                                      decomposition rate by 5-7%.
No significant  effect  of acid
  treatments on respirtion.
  Litterbaqs showed  significant
  increase in weight loss (15%)
  with increased  acidity.

Found decreased C02
  respiration with  increased
  H2S04.

-------
     Rosenqvist (1977, 1978, Rosenqvist et al.  1980)  has argued that the
influence of soil  and bedrock on the chemistry  of waters is overwhelming
and that the pH of runoff water would be the same whether snowmelt was
acid or neutralized by a suitable base.  Seip et al.  (1980) carried out
an experiment to test Rosenqvist1s hypothesis by applying NaOH to one of
the mini-catchment watersheds in Norway; results showed that,  indeed,
the neutralization of snow with NaOH had little effect on runoff pH.
The investigators attributed the lack of effect, to  differences in
weather conditions, and Na content of the deposition.

     Seip (1980) presented a hypothesis for surface  water acidification
which has met with agreement among soil scientists as to its mechanism
but not necessarily to its magnitude.  This has been termed the "mobile
anion mechanism."   In essence, it states that the introduction  of a
mobile anion into an acid soil will cause the pH of  a soil  solution to
drop.  This is because of the requirement for cation-anion  balance in
solution and because most exchangeable cations  in acid soils are H+
and Al3+.  Thus, due to cation exchange processes and the requirement
for cation-anion balance, increased anion concentration in  an acid soil
solution causes increased H+ and Al3+ concentrations, regardless of
whether the anion is introduced as a salt or an acid.  This mechanism
has been known to soil scientists for decades as the "salt  effect,"
wherein soil pH is usually more acid in CaCl2 solutions than in H20
(Yuan 1963).  Field studies have confirmed that this  mechanism is valid
(Abrahamsen et al. 1979; Seip et al. 1979a,b,  1980;  Abrahamsen and
Stuanes 1980).  However, doubt remains as to whether  the magnitude of pH
change this mechanism can produce could cause the pH  changes reported
for acidified surface waters (Abrahamsen and Stuanes  1980;  Johnson 1981;
Rosenqvist 1981, pers. comm.).  It is clear, however, that  neutral  salts
can, when added to an acid soil, cause a flux of Al  in a low-pH solution
to streams.

    Natural acid production, changes in land use patterns,  and
management practices such as harvesting, burning, and fertilizing are
suggested alternative sources for surface water acidification
(Rosenqvist 1977,  1978;  Patrick et al.  1981).   These  possibilities have
been explored to some extent in southern Norway, but  we have no concrete
evidence that changes due to harvesting and land use  have caused surface
water acidification (Drabltfs et al. 1980)  although the debate
continues.   Evidence suggests, however, that fish kills associated with
acidic pulses have been occurring in at least one place in  southern
Norway (Roynelandsvann)  since the 1890's (Torgenson  1934).   In this
instance, liming was successful  as a mitigative measure for short-term
effects on fish populations (Abrahamsen, pers.  comm.).   The causes of
these acid pulses  are unknown, but presumably acid rain effects were
much smaller nearly a century ago.

     Some attention has  been given to neutralization  processes affecting
acid rain as it passes through terrestrial  to aquatic ecosystems.   N.  M.
Johnson et al. (1981) found a two-stage process operative in the Hubbard
                                  2-55

-------
Brook, NH ecosystem in which H+ in acid rain is initially neutralized
by dissolution of reactive alumina in the soil  before both H+ and
A13+ are neutralized by chemical  weathering of alkali and alkaline
earth minerals in bedrock.  Because stage 2 proceeds more slowly than
stage 1, first- and second-order streams may contain H+ and A13+,
but neutralization is usually complete before surface waters reach
third-order streams.

     Kilham (1982) reports a case in which deposition appears to have
caused an increase in lake alkalinity.  Alkalinity in Weber Lake,
Michigan, has increased twofold over the last thirty years, and
theoretical  considerations of acid-base budgets lead to the hypothesis
that this alkalization has resulted from plant nitrate uptake,  bacterial
sulfate reduction, and carbonate mineral  weathering, all  enhanced  by
acid precipitation.  This effect, while no more desirable than
acidification, contradicts the assumption that acid rain  always causes
surface water acidification and is ample testimony to the complexity of
terrestrial-aquatic interactions,  Kilham (1982)  indicates that
alkalization is likely only in lakes of high alkalinity with abundant
carbonates in the watershed.

     In view of the lack of understanding of terrestrial-aquatic
transport processes, assigning "sensitivity" ratings to acid deposition
on a regional scale is premature.  Nonetheless, agencies  alarmed by
reports of ecological effects of acid precipitation insist upon knowing
something about the geographical  magnitude of the acid rain "problem,"
and scientists must make their best guesses as to appropriate criteria,
even though the mechanisms are not completely understood.  This
situation reflects a gap in understanding and a critical  research need
that encompasses not only soil and bedrock chemical  reactions but also
hydro!ogical processes.  Recent studies have shown the important
contribution of variable source areas (i.e., portions of watershed
landscapes that contribute to streamflow during storm events) to surface
waters and their chemical  composition during stormflow (Henderson  et al.
1977, Huff et al. 1977, Johnson and Henderson 1979).

     Similarly, water flow through soil macropores (See Figure  2-1) can
be a very important component of soil water flux during periods of
saturated flow (Luxmoore 1981).  Both variable source areas and
macropore flow reduce the amount of contact between soils or bedrock and
waters passing through terrestrial ecosystems.   Integrated studies of
terrestrial-aquatic transport processes involving both hydrological and
chemical components are essential to an understanding of the effects of
acid rain on aquatic ecosystems.

2.7  CONCLUSIONS

     Effects of acidic deposition related to soils are in these general
categories:   soil acidification,  nutrient supply, metal mobility,  and
rnicrobial activity.  The following conclusions, relative to these
general categories, can be drawn from Chapter E-2:
                                  2-56

-------
Soils amended in agricultural  practice will  not be harmed by
acidic deposition (Section 2.3.5).

Soil acidification is a natural  process in humid regions.
It is obvious that acidic deposition contributes to this
process; however, at current levels, it is a minor
contribution (Section 2.3.5).

Most soils that were easily acidified are already acid;
therefore, soils likely to become perceptibly more acid  due
to deposition are limited. They are the soils that have  low
buffering capacity,  a relatively high pH (slightly acid, pH
5.5 to 6.5), low sulfate adsorption capacity, no carbonates,
and no basic inputs  (Section 2.3.5).

The availability of  sulfur and nitrogen to plants will be
enhanced by their presence in  the deposition.  Because
nitrogen limitations are so common  and cation limitations are
so rare in forests of the United States, it seems likely that
HN03 inputs generally will be  beneficial.  Exceptions  may
occur on sites with  adequate or excessive N  supplies. Benefits
of H2S04 deposition  are probably minimal,  because S
deficiencies are rare and probably  easily satisfied with
moderate atmospheric S inputs  (Section 2.3.2).

The long-term effect (i.e., over decades or  centuries) of
acidic deposition can be expected to remove  cations from
forest soils, but it is not clear whether this will  reduce
available cations and enhance  acidification  of soils.  For
example, cation leaching rates,  although increased by acid
precipitation, may remain insignificant relative to total
soil supplies and forest growth  requirements; furthermore,
exchangeable cations may be replaced by weathering from
primary minerals at  rates sufficient to maintain their current
status partially as  a result of acid precipitation inputs
(Section 2.3.3).

Assessing acidic deposition effects on forest nutrient status
involves quantifying amounts of inputs involved and the  S, N,
and cation nutrient  status of  specific sites.  It cannot be
stated that forest ecosystems, in general, respond to acidic
deposition in a single predictable  way.  Indeed, the
contrasting behavior of Norway spruce in Germany and in  Norway
exemplifies the variable response that can be expected from
different sites (Section 2.3.3).

The most likely damage to forest productivity, and the one for
which some evidence  exists, would result from Al toxicity.
This may occur on already acid soils where acidic deposition
plus natural acidifying processes increase acidity enough to
cause a significant rise in Al availability.  If soil pH is
low enough (< pH 5.0 to 5.5) in  mineral soils to cause the
                        2-57

-------
dissolution of Al- and Mn-containing clay  minerals,  any  H+
input will  increase solution  of Al  and  Mn  concentration
(Section 2.3.3).

The increased mobility of Al  in uncultivated,  acid  soils is
probably the most significant effect of acidic deposition on
soils as they influence terrestrial  plant  growth  and aquatic
systems (Section  2.3.3).

Based upon shorter term studies,  we can expect that increased
H+ loading will generally cause increased  loss of cation and
organic components from forest litter.   Over the  longer  term,
it appears that the biologically-mediated  mineralization of
organic matter in forest soils will  be  only slightly inhibited
by acidic deposition (< 1 to  2 percent  decrease in
decomposition rate).  In general,  experimental  data  suggest
that decomposition processes  are relatively unaffected by
simulated precipitation pH's  above 3.0. Thus, unless average
precipitation inputs were to  drop to pH 3.0 or below,
significant impacts of acidic deposition or litter
decomposition in  natural  systems are not expected (Section
2.3.3).

Soil microbial activity may be significantly influenced  near
the surface if inputs are great enough  to  affect  pH or
nutrient availability.  Evidence for effects of acidic
deposition Rhizpbium or actinomycete is reduced by  artificial
acid inputs, but  no evidence  exists that current  rates common
in the United States will cause such a  decrease.  Slight
decreases and increases in N  mineralization rates result from
short-term acid inputs, but long-term responses are not
documented.  Possible effects of acidic deposition  as soil
microbial activity in natural systems have not been ruled out,
but important effects under field conditions have not been
clearly demonstrated (Section 2.4).
                        2-58

-------
2.8  REFERENCES

Abrahamsen, G.  1980a.  Impact of atmospheric sulfur deposition on
forest ecosystems, pp. 397-416.  ^Atmospheric Sulfur:   Environmental
Impact and Health Effects.  D. S. Shriner,  C. R. Richmond,  and S.  E.
Lindberg, eds.  Ann Arbor Science.

Abrahamsen, G.  1980b.  Acid precipitation, plant nutrients,  and forest
growth, pp. 58-63.  lr± Ecological Impact of Acid Precipitation. D.
Drabltfs, and A. Tollan, eds.  Proc.  of an International  Conference,
The Norwegian Interdisciplinary Research Programme Acid  Precipitation-
Effects on forest and Fish. Sandfjord, Norway,  March 11-14,  1980.

Abrahamsen, G. and G. J. Dollard.  1978.  Effects of acidic
precipitation on forest vegetation and soil.   SNSF FA 32/78,
SNSF-project, Oslo-As, Norway.

Abrahamsen, G., and A. 0.  Stuanes.  1980.  Effects of simulated rain  on
the effluent from lysimeters with acid, shallow soil  rich in  organic
matter, pp. 152-153.  J^n Ecological  Impact  of Acid Precipitation.  D.
Drablj6s, and A. Tollan, eds.  Johs.  Grefslie Tyrkkeri, A/S,  Mysen,
Norway.

Abrahamsen, G., K. Bjor, R. Horntvedt, and  B. Tveite.  1976.   Effects of
acid precipitation on coniferous forest, pp.  36-63.  lr±  Impact of  Acid
Precipitation on Forest and Freshwater Ecosystems in Norway.   F. H.
Braekke, ed.  SNSF FR 6/76, SNSF-project, Oslo-As, Norway, pp.
36-63.

Abrahamsen, G., J. Hovland, and S. Hagvar.   1980.  Effects of artificial
acid rain and liming on soil organisms and  the decomposition  of organic
matter, pp. 341-362.  In Effects of Acid Precipitation on Terrestrial
Ecosystems. T. C. HutcTrTnson and M.  Havas,  eds.   Plenum  Press,  New York.

Abrahamsen, G., A. 0. Stuanes, and K.  Bjor.  1979.  Interaction between
simulated rain and barren  rock surface.  Water,  Air,  and Soil  Pollut.
11:191-200.

Adams, F.  1978.  Liming and fertilization  of Ultisols and Oxisols.  j£
Mineral Nutrition of Legumes in Tropical and Subtropical  Soils.  C. S.
Andrew and E. J. Kamprath.  CSIRO.  Australia,   pp. 377-394.

Adams, F. and R. W. Pearson.  1967.   Crop response to lime  in the
southern United States and Puerto Rico.  Agronomy 12:161-206.

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.

Alexander, M.  1980a.  Effects of acidity on  microorganisms and
microbial processes in soil, pp. 363-374.  j£ T. C. Hutchinson  and M.
Havas, eds.  Effects of Acid Precipitation  on Terrestrial Ecosystems.
Plenum Press, New York.

                                  2-59

-------
Alexander, M.  1980b.  Effects of acid precipitation on biochemical
activities in soil, pp. 47-52.  ^_n D.  Drabltfs and A. Tollan,  eds.
Proc. Int. Conf. Ecological  Impact of Acid Precipitation.   SNSF-Project,
Sandefjord, Norway.

Andersson, F., T. Fagerstrom, and S.  I.  Nilsson.   1980.  Forest
ecosystem responses to acid deposition - Hydrogen ion budget  and
nitrogen/tree growth model  approaches, pp. 319-334.   Iji effect  of  acid
precipitation on terrestrial ecosystems.  C. Hutchinson and M.  Havas,
eds.  Plenum Press, New York.

Baath, E., B. Berg, U. Lohm, B. Lundgren,  H. Lundkvist,  T.  Rosswall, B.
Sonderstrom, and A. Wiren.   1980.  Soil  organisms and litter  decompo-
sition in a Scots pine forest - effects  of experimental  acidification.
In Effects of Acid Precipitation on Terrestrial  Ecosystems.   T.  C.
Wtchinson and M. Havas, eds.  Plenum Press, New York.   pp. 375-380.

Baath, E., B. Lundgren, and B. Soderstrom.  1979.  Effects  of artificial
acid rain on microbial activity and biomass.  Bull.  Environ.  Contam.
Toxicol. 23:737-740.

Babich, H. and G. Stotzky.   1977.  Sensitivity of various bacteria,
including actinomycetes, and fungi to cadmium and the influence of pH on
sensitivity.  Appl. Environ. Microbiol.  33:681-695.

Babich, H. and G. Stotzky.    1979.  Abiotic factors  affecting the
toxicity of lead to fungi.   Appl. Environ. Microbiol. 38:506-513.

Bache, B. W.  1980.  The acidification of soils.   In Effects  of Acid
Precipitation on Terrestrial Ecosystems.  T. C.  Hutchinson  and  M.  Havas,
eds.  Plenum Press, New York.  pp. 163-202.

Bailey, E. H.  1933.  The correlation of the pH values of the soil
profiles with the great soil groups.   Amer. Soil  Survey Assn. Bull. 14,
pp. 93-95.

Barber, S. A.  1967.  Liming materials and practices.  Agronomy
12:125-160.

Barrow, N. J., and T. C. Shaw.  1977.   The slow reactions between  soil
and anions:  7.  Effects of time and temperature of contact between an
adsorbing soil and sulfate.  Soil Sci. 124:237-240.

Bengston, G. W.  1979.  Forest fertilization in the United  States.
Progress and Outlook.  J. For. 78:222-229.

Bjor, K. and 0. Teigen.  1980.  Effects of acid precipitation on soil
and forest.  6. Lysimeter experiment in greenhouse,  pp.  200-201.  J^n
Ecological Impact of Acid Precipitation.  D. Drablj6s and A. Tollan,
eds.  Proc. of International Conference, Sandefjord, Norway.
                                  2-60

-------
Blcornfield, C.  1955.  Leaf leachates as a factor in pedogenesis.   J.
Sci. Food Agr. 6:641-651.

Bolt, G. H.  1979.  Movement of solutes in soil:   Principals of
adsorption/exchange chromatography.  ^n Soil  Chemistry B.
Physico-chemical Models.  G. H. Bolt, ed.   Elsevier Scientific Pub.  Co.,
Amsterdam.

Bornemisza, E., and R. Llanos 1967.  Sulfate  movement, adsorption  and
desorption in three Costa Rican soils.  Soil  Sci. Soc. Amer. Proc.  31:
356.

Bradley, R., A. J. Burt, and D. J. Read.  1981.   Mycorrhizal infection
and resistance to heavy metal toxicity in  Calluna vulgarls.   Nature
292:335-337.

Brady, N. C.  1974.  The Nature and Properties of Soils.   Macmillan, New
York, NY.

Bromfield, E. S. P. and D. Gareth Jones.  1980.   Studies on  acid
tolerance of Rhi'zobium trifolii in culture and soil.  J. Applied
Bacteriol. 487Z57:ZH37

Butlin, K. R. and J. R. Postgate.  1954.  Economic importance of
autotrophs.  lr± Fourth Symposium of the Society for General
Microbiology, Cambridge.  B. A. Fry and J. L. Peel, eds.

Chao, T. T., M. E. Harward, and .S. C. Fung 1963.   Cationic effects  on
sulfate adsorption by soils.  Soil Sci. Soc.  Amer. Proc. 27:35-41.

Cole, D. E. and D. W. Johnson.  1977.  Atmospheric sulfate additions and
cation leaching in a Douglas Fir ecosystem.  Water Resources Research
13:313-317.

Cole, D. W. and M. Rapp.  1981.  Elemental cycling in forest ecosystems.
In Dynamic Properties of Forest Ecosystems.  D.  E. Reichle,  ed.
Cambridge Univ. Press, MA.

Coleman, N. T. and G. W. Thomas.  1967.  The  basic chemistry of soil
acidity.  Agronomy 12:1-42.

Couto, W., D. J. Lathwell, and D. R. Bouldin.  1979.  Sulfate sorption
by two Oxisols and an Alfisol of the tropics.  Soil Sci. 127:108.

Cowling, E. B. and L. S. Dochinger.  1980.  Effects of acidic
precipitation on health and productivity of forests, pp. 165-173.   Ir±
Effects of Air Pollutants on Mediterranean and Temperate Forest
Ecosystems.  P. R. Miller, Technical Coordinator.  U.S. Forest Service
Gen. Tech. Rep. PSW-43, Upper Darby, PA.

Cronan, C. S.  1980a.  Controls on leaching from coniferous  forest floor
microcosms.  Plant Soil 56:301-322.
                                  2-61


  409-262 0 - 83 - it

-------
Cronan, C. S.  1980b.   Consequences  of sulfuric  acid  inputs  to a  forest
soil, pp. 335-343.   In Atmospheric Sulfur  Deposition-Environmental
Impact and Health Effects.   D.  S.  Shriner,  C.  R.  Richmond, and S. E.
Lindberg, eds.   Ann Arbor Science  Publishers,  Ann Arbor, MI.

Cronan, C. S.  1980c.   Solution chemistry  of a New  Hampshire subalpine
ecosystem:  A biogeochemical  analysis.   Oikos  34:272-281.

Cronan, C. S.,  W. A. Reiners,  R. L.  Reynolds,  and G.  E. Lang.  1978.
Forest floor leaching:  Contributions  from mineral, organic, and
carbonic acids  in New Hampshire subalpine  forests.  Science  200:309-311.

Davies, R. I.,  Coulson, C.  B.,  and Lewis D.  A.  1964.  Polyphenols  in
plant, humus, and soil.  IV.  Factors leading to  increase in  biosynthesis
of polyphenols  in leaves and their relationship  to  mull and mor
formation.  J.  Soil Sci. 15:310-317.

Denison, R., B. Caldwell, B.  Bormann,  L. Eldred,  C. Swanberg, and S.
Anderson.  1977.  The effects of acid  rain on  nitrogen fixation in
western Washingtron coniferous  forests.  Water,  Air,  and Soil Pollut.
7:409-416.

Dobereiner, J.   1966.   Manganese toxicity  effects on  modulation and
N-fixation of beans (Phaseolus  vulgaris L.)  in acid soils.   Plant Soil
24:153-166.

Dooley, F. and J. A. Houghton.   1973.   The nitrogen-fixing capacities
and the occurrence of blue-green algae in  peat soils.  Brit. Phycol. J.
8:289-300.

Drablj6s, D., I. Sevaldrud,  andJ.  A. Timberlid.   1980.  Historical
land-use changes related to fish status development in different  areas
in southern Norway, pp. 367-389.  In Ecological  Impact of Acid
Precipitation.   D.  Drabltfs  and  A. Tollan,  eds.  Proc. of International
Conference, Sandefjord, Norway.

Engstrom, A., G. Backstrand,  and H.  Stenram.  1971.   Air pollution
across natural  boundaries - The impact on  the  environment of sulfur in
air and precipitation.  Sweden's Case  Study for  the United Nations
Conference on the Human Environment, Royal  Ministry for Foreign Affairs,
Royal Ministry  of Agriculture,  Sockholm, Sweden.

Evans, L. S., G. R. Hendrey,  G. J.  Stensland,  D.  W. Johnson, and  A. J.
Francis.  1981.  Acidic precipitation:  Consideration  for an  air quality
standard.  Water, Air, and Soil Pollut.  16:468-509.

Evans, L. S., R. F. Lewin,  and P.  A. Vella.   1980.  Effect of nutrient
medium pH on symbiotic nitrogen fixation by Rhizobium leguminosarum and
Pisum sativum.   Plant Soil  56:71-80.
                                  2-62

-------
Farrell, E. P., I.  Nilsson,  C.  0.  Tamm,  and G.  Wiklander.   1980.
Effects of artificial  acidification  with sulfuric  acid on  soil chemistry
in a Scots pine forest,  pp.  186-187. ^n Ecological  Impacts of Acid
Preciptiation.  D.  Drablds and  A.  Tollan,  eds.   Jons. Grifslie
Trykkeri A/S, Mysen,  Norway.

Firestone, M. K., R.  B.  Firestone,  and J.  M. Tiedje.  1980.   Nitrous
oxide from soil denitrification:   Factors  controlling its  biological
production.  Science  208:749-751.

Foy, C. D.  1964.  Toxic factors  in  acid soils  of  the southeastern
United States as related to  the response of alfalfa  to lime.  USDA Prod.
Res. Rep. 80.

Foy, C. D.  1973.  Manganese and  plants, pp. 51-76.   In  Manganese.   Nat.
Acad. Sci. Nat. Res.  Council, 2101  Constitution Ave.,~HW,  Washington,
D.C.

Foy, C. D.  1974a.   Effects  of  soil  calcium on  plant growth,  pp.
135-151.  The Plant Root and Its  Environment.   E.  W.  Carson,  ed.
University Press of Virginia, Charlottesville,  VA.

Foy, C. D.  1974b.   Effects  of  aluminum  on plant growth, pp.  601-642.  In
The plant root and  its environment.   E.  W. Carson, ed.   Univ. Press  of
Virginia, Charlottesville.

Foy, C. D.  1981.  Effect of nutrient deficiencies and toxicities in
plants:  Acid soil  toxicity. Manuscript for Handbook of Nutrition and
Food.  CRC Press, 200 NW 24th Street, Boca Raton,  FL.

Foy, C. D. and A. L.  Fleming.   1978. The  physiology  of  plant tolerance
to excess aluminum  and manganese  in  acid soils, pp.  301-328.  In Crop
Tolerance to Suboptimal  Land Conditions.  ASA Spec.  Publ.  No. 3~2~, ASA,
CSSA, SSSA, Madison,  WI.

Foy, C. D., R. L. Channey, and  M.  C. White.   1978.   The  physiology of
metal toxicity in plants.  Ann. Rev. Plant Physiol.  29:511-566.

Francis, A. J., D.  Olson,  and R.  Bernatsky.   1980.   Effect of acidity  on
microbial processes in a forest soil, pp.  166-167.  Jji Ecological Impact
of Acid Precipitation.  Proceedings  of an  International  Conference,
Sandefjord, Norway, March  11-14,  1980, D.  Drabltfs  and A. Tollan, eds.,
SNSF-project, Oslo-As, Norway.

Freedman, B. and T. C. Hutchinson.   1980.   Pollution  inputs from the
atmosphere and accumulations in soils and  vegetation near  a
nickel-copper smelter at Sudbury,  Ontario, Canada.   Can. J.  Bot.
58:108-132.
                                  2-63

-------
Galloway, J. N., C. L. Schofield,  G.  R.  Hendrey,  E.  R.  Altwicker,  and  D.
E. Troutman.  1980.  An analysis of lake acidification  using  annual
budgets.  ^n_ Ecological Impact of Acid Precipitation.   D.  Drablj6s  and
A. Tollan, eds.  Proc. Intern. Conf.,  Sandefjord,  Norway.   SNSF  Project,
Oslo.

Garcia, R. L. and J. J. Hanway.  1976.  Foliar fertilization  of  soybeans
during the seed filling period.  Agron.  J.  68:653-657.

Gorham, E. and W. W. McFee.   1980.   Effects of acidic deposition upon
outputs from terrestrial  to aquatic ecosystems, pp.  465-480.  JJT_ Effect
of Acid Precipitation on  Terrestrial  Ecosystems.   T. C.  Hutchinisbn and
M. Havas, eds.  Plenum Press,  New York.

Graham, P. H.  1978.  Some problems and  potentials of field beans
(Phaseolus vulgaris L.) in Latin America.   Field  Crops  Res. 1:295-317.

Gray, T. R. G. and S. T.  Williams.   1971.   Soil Microorganisms.
Longman, New York.

Hagvar, S.  1978.  Effects of acidification and liming  on  collembola and
acarina.  SNSF Project IR 36/78, Oslo, Norway.

Haines, B. and G. R. Best.  1975.   The influence  of  an  endomycorrhizal
symbiosis on nitrogen movement through soil columns  under  regimes  of
artificial throughfall and artificial  acid  rain,   ^ji Proc.  1st Int.
Symp. on Acid Precipitation and the Forest  Ecosystem.   L.  S.  Dochinger
and T. A. Seliga, eds.  USDA For.  Serv.  Gen.  Tech. Rep.  NE-23:951-961.

Hart, J. B., A. L. Leaf,  and S. J.  Stutzbach.   1969.  Variation  in
potassium available to trees within an outwash soil.  Soil  Sci.  Soc. Am.
Proc. 33:950-954.

Hasan, S. M., R. L. Fox,  and C. C.  Boyd.  1970.   Solubility and
availability of sorbed sulfate in  Hawaiian  soils.  Soil  Sci.  Soc.  Am.
Proc. 34:897-901.

Heiberg, S. 0. and D. P.  White.  1951.  Potassium  deficiency  of
reforested pine and spruce stands  in  northern New  York.  Soil Sci. Am.
Proc. 15:369-376.

Henderson, G. S., A. Huntley,  and  W.  Selvidge.  1977.   Nutrient
discharge from Walker Branch Watershed,  pp. 307-320.  In Watershed
Research in North America.  D. L.  Correll,  ed.  Smithsonian Institution,
Edgewater, MD.

Hendry, G. R., J. N. Galloway, S.  A.  Norton,  C. L. Schofield, D. A.
Burns, and P. W. Schaffer.  1980.   Sensitivity  of  the eastern United
States to acid precipitation impacts  on  surface waters,  pp. 216-217.  Jjn
Ecological Impacts of Acid Precipitation.   D.  Drabl0s,  and  A. Tollan,
eds.  Jons. Grefslie Trykkeri  A/S,  Mysen, Norway.
                                  2-64

-------
Hoeft, R. G., D.  R.  Kenney,  and L.  M.  Walsh.   1972.  Nitrogen  and  sulfur
in precipitation  and sulfur  dioxide in the  atmosphere in Wisconsin.  J.
Environ. Qual. 1:203-208.

Hovland, J.  1981.   The effect of artificial  acid  rain  on  respiration
and cellulase activity in  Norway spruce needle litter.  Soil Biol.
Biochem. 13:23-26.

Hovland, J., G. Abrahamsen,  and G.  Ogner.   1980.   Effects  of artificial
acid rain on decomposition of spruce needles  and on mobilization and
leaching elements.   Plant  Soil 56:365-378.

Huff, D. D., R. J.  Luxmoore, J. B.  Mankin,  and 0.  L. Begovich.  1977.
TEHM: A Terrestrial  Ecosystem Hydrology Model. ORNL/NSF/EATO-27,  Oak
Ridge National Laboratory, Oak Ridge,  TN.   147 p.

Humphreys, F. R., M. J. Lambert, and J. Kelly.  1975.   The occurrence  of
sulfur deficiency in forests, pp. 154-162.   In Sulfur  in Australasian
Agriculture.  K.  D.  McLachlan, ed.   Sidney  LfnTv. Press, Sidney,
Australia.

Johnson, A. H., T.  G. Siccama, D. Wang, T.  H. Barringer, and R. S.
Turner.  1980.  Decreases  in stream pH and  tree growth  rates in the New
Jersey pinelands:  A possible link  to acid  rain.   Agronomy Abstr., 1980,
p. 202.

Johnson, A. H., T.  G. Siccama, D. Wang, R.  S. Turner and T. H.
Barringer.  1981.  Recent  changes in patterns of tree  growth rate  in the
New Jersey pinelands:  A possible effect of acid rain.  J. Environ.
Qual. 10(4):427-430.

Johnson, D. W.  1980.  Site susceptibility  to leaching  by  H2S04 in
acid  rainfall, pp.  525-536.   In Effects of  Acid Precipitation  on
Terrestrial Ecosystems.  T.  CTHutchinson and M. Havas, eds.   Plenum
Press,  New York.

Johnson, D. W.  1981.  The natural  acidity  of some unpolluted  waters  in
southeastern Alaska and potential impacts of acid  rain.  Water, Air, and
Soil  Pollut.   16:243-252.

Johnson, D. W. and D. W. Cole.  1977.  Sulfate mobility in an  outwash
soil  in  western Washington.   Water, Air, and Soil  Pollut.  7:489-495.

Johnson, D. W. and D. W. Cole.  1980.  Anion mobility  in  soils-
Relevance  to  nutrient transport from  forest ecosystems.  Environ.
Internat.  3:79-80.

Johnson, D. W. and G. S. Henderson.   1979.   Sulfate adsorption and
sulfur  fractions in a highly-weathered soil under a mixed deciduous
forest.  Soil  Sci.  128:34-40.
                                  2-65

-------
Johnson, D. W., 0. E. Todd.   1983.   Some relationships  among Fe, Al,  C
and $04 in a variety of forest soils.   Soil  Sci.  Soc. Amer. 0. 47:(in
press).

Johnson, D. W., D. W. Cole,  and S.  P.  Gessel.   1975.  Processes of
nutrient transfer in a tropical  rain forest.   Biotropica  7:208-215.

Johnson, D. W., D. W. Cole,  S. P. Gessel,  M. J.  Singer, and R. V.
Minden.  1977.  Carbonic acid leaching in  a  tropical, temperate,
subalpine and northern forest soil.   Arctic  Alpine  Res.   9:329-343.

Johnson, D. W., D. W. Cole,  and S.  P.  Gessel.   1979.  Acid precipitation
and soil sulfate adsorption  in a tropical  and  in a  temperate forest
soil.  Biotropica 11:38-42.

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. 507-520.  In  Atmospheric Sulfur
Deposition: Environmental  Impact and HealthfTffects.  D.  S. Shriner,  C.
R. Richmond, and S. E. Lindberg, eds.   Ann Arbor Science, Ann Arbor,  MI.

Johnson, D. W., J. Turner, and J. M. Kelly.  1982.  The effects of acid
rain on forest nutrient status.   Water Res.  Res.  18(3):449-461.

Johnson, N. M.  1979.  Acidification of headwater streams in the New
Jersey pine barrens.  J. Environ. Qual.  8:383-386.

Johnson, N. M., C. T. Driscoll,  J.  S.  Eaton, G.  R.  Likens, and W.  P.
McDowell.  1981.  Acid rain, dissolved aluminum and chemical weathering
at the Hubbard Brock Experimental Forest,  New  Hampshire.  Geochimica  et
Cosmochimica Acta 45:1421-1437.

Johnsrud, S. C.  1978.  Heterotrophic nitrification in  acid forest
soils.  Holarctic Ecol. 1:27-30.

Jones, U. S., M. G. Hamilton, and J. B.  Pitner.   1979.  Atmospheric
sulfur as related to fertility of Ultisols and Entisols in South
Carolina.  Soil Sci. Soc.  Amer.  J.  43:1169-1171.

Jordan, H. V., C. E. Bardsley, Jr.,  L.  E.  Ensminger, and  J, A. Lutz.
1959.  Sulfur content of rainwater  and atmosphere in southern states.
USDA Tech. Bull. 1196.  16 p.

Kamprath, E. J. and C. D.  Foy.  1972.   Lime-fertilizer-pi ant
interactions in acid soils,  pp.  105-151.  Ir± R.  A.  Olsen  et al., eds.
Soil Sci. Soc. Amer., Madison, VII.

Kelly, J. and M. J. Lambert.  1972.   The relationship between sulphur
and nitrogen in the foliage  of Pinus radiata.   Plant Soil 37:395-408.

Kilham, P.  1982.  Acid precipitation:   Its  role in the alkalization  of
a lake in Michigan.  Limnology and  Oceanography 27(5):856-867.
                                  2-66

-------
Killham, K. and M. Wainwright.  1981.  Deciduous leaf litter and
cellulose decomposition in soil exposed to heavy atmospheric pollution.
Environ. Pollut. Series A 26:79-85.

Klopatek, J. M., W. F. Harris, and R. J. Olson. 1980.  A regional
ecological assessment approach to atmospheric deposition:  Effects on
soil  systems, pp. 559-564.  lr± Atmospheric Sulfur Deposition-
Environmental Impact and Health Effects.  D. S. Shriner, C. R. Richmond,
and S. E. Lindberg, eds.  Ann Arbor Science, Ann Arbor, MI.

Ko, W. H. and F. K. Hora.  1972.  Identification of an Al ion as a soil
fungi toxin.  Soil Sci. 113:42-45.

Kokorina, A. L.  1977.  Effect of soil acidity on yield and chemical
composition of herbage under different soil moisture regimes (Ru, e).
Rationyi Zhurnal 5:55:363:88-91.

Kubota, J.  1978.  Molybdenum status of United States soils and plants.
j£ The Biology of Molybdenum. W. R. Chappell and K. Petersen, eds.  Vol.
I.  Chap. 6, Marcel Dekker, New York.

Lathwell, D. J. and M. Peech.  1969.  Interpretation of chemical  soil
tests. .Cornell Univ. Agr. Exp. Sta. Bull. No. 995.  Cornell  Univ.,
Ithaca, NY.  39 p.

Likens, G. E., F. H. Borman,  R. S. Pierce, J. S. Eaton, and N. M.
Johnson.  1977.  Biogeochemistry of a forested ecosystem.  Springer-
Verlag, New York.

Lind, C. J. and J. D. Hem.  1975.   Effects of organic solutes on
chemical reactions of aluminum.  USGS water suppl.  paper 1827-6.   83 pp.

Lindberg, S. E., R. C. Harris,  R. R. Turner, D. S. Shriner,  and D. D.
Huff.  1979.  Mechanism and rates of atmospheric deposition of selected
trace elements and sulfate to a deciduous forest watershed.
ORNL/TM-6674.   Oak Ridge Nat. Lab.,  Oak Ridge, TN.   514 pp.

Little, P. and H. Martin.   1972.  A survey of zinc, lead, and cadmium in
soil and natural vegetation around a smelting complex.   Env.  Poll.
3:245-254.

Lohm, U.  1980.   Effects of experimental  acidification  on soil  organism
populations and decomposition pp.  178-179.  Jji^ Ecological Impact of Acid
Precipitation.   Proceedings of an  International  Conference, Sandefjord,
Norway, March 11-14. 1980, D.  Drablj6s and A. Tollan,  eds.,
SNSF-project,  Oslo-As, Norway.

Lowendorf, H.  S.-, A. M.  Baya,  and  M.  Alexander.   1981.   Survival  of
Rhizobium in acid soils.   Appl.  Environ.  Microbiol. 42:951-957.

Luxmoore, R. J.   1981.  Micro-,  meso-, and macroporosity of soil.   Soil
Sci. Soc. Am.  J. 45:671-672.
                                  2-67

-------
Malcolm, R. L. and R. J. McCracken.  1968.  Canopy Drip:   A source of
mobile soil organic matter for mobilization of iron and aluminum.   Soil
Sci. Soc. Amer. Proc. 32:834-838.

Matzner, E. and B. Ulrich.  1981.  Effect of acid precipitation on soil,
pp. 555-564.  Di Beyond the Energy Crisis—Opportunity and Challenge.
R. A. Fazzolare and C. B. Smith, eds.  Pergamon Press, New York.

Mayer, R. and B. Ulrich.  1977.  Acidity of precipitation as influenced
by the filtering of atmospheric sulphur and nitrogen compounds  - its
role in the elemental balance and effect on soil.  Water, Air,  Soil  and
Pollut. 7:409:416.

McColl, J. G. and D. W. Cole.  1968.   A mechanism of cation transport in
a forest soil.  Northwest Sci. 42:134-140.

McFee, W. W.  1980.  Sensitivity of soil regions to long-term acid
precipitation, pp. 495-506.  In Atmospheric Sulfur Deposition:
Environmental Impact and HealTTT Effects.  D. S. Shriner,  C. R.  Richmond,
and S. E. Lindberg, eds.  Ann Arbor Science, Ann Arbor, MI.

McFee, W. W., J. M. Kelly, and R. H.  Beck.  1976.  Acid precipitation
effects on soil pH and base saturation of exchange sites, pp. 725-736.
In Proceedings of the First International  Symposium on Acid
"Precipitation and the Forest Ecosystem.  L. S. Dochinger and T. A.
Seliga, eds.  USDA Forest Service Gen. Tech. Rep. NE-23.

McFee, W. W., J. M. Kelly, and R. H.  Beck.  1977.  Acid precipitation
effects on soil pH and base saturation of exchange sites.  Water,
Air, Soil Pollut. 7:401-408.

McLean, E. 0. and V. C. Bittencourt.  1973.  Complementary ion effects on
potassium, sodium, and calcium displacement from bi-ionjc bentonite and
illite systems as affected by pH-dependent charges.   Soil Sci.  Soc.
Amer. Proc. 37:375-379.

Mollitor, A. V. and D. J. Raynal.  1982.  Acid precipitation and ionic
movements in Adirondack forest soils.  Soil Sci. Soc.  J.  46:137-141.

Munns, D. N., J. S. Holtenberg, T. L. Tighetti, and D. J. Lauter.   1981.
Soil acidity tolerance of symbiotic and nitrogen-fertilized soybeans.
Agron. J. 73:407-410.

Mutatkar, V. K. and W. L. Pritchett.   1966.  Influence of added aluminum
on carbon dioxide production in tropical soils.  Soil  Sci.  Soc. Am.
Proc. 30:343-346.

Norton, S. A.  1980.  Geologic factors controlling the sensitivity of
aquatic ecosystems to acidic precipitation, pp. 521-532.   Ir^ Atmospheric
Sulfur Deposition:  Environmental Impact and Health Effects.  D.  S.
Shriner, C. R. Richmond, and S. E. Lindberg, eds.  Ann Arbor Science,
Ann Arbor, MI.
                                  2-68

-------
Nye, P. H. and D.  J.  Greenland.   1980.   The soil  under  shifting
cultivation.   Commonwealth Bureau of Soils  Tech.  Comm.  No. 51,
Commonwealth Agricultural  Bureau, Farnham Royal,  Bucks.

Olsen, R. J., D.  W. Johnson,  and D.  S.  Shriner.   1982.   Regional  assess-
ment of potential  sensitivity of soils  in the Eastern United  States  to
acid precipitation.  ORNL/TM, Oak Ridge National  Lab.   No. 8374.

Patrick, R., V. P. Binetti, and S. G. Halterman.   1981.   Acid lakes  form
natural and anthropogenic  causes.  Science  211:446-448.

Patrick, W. H., Jr.  1977.  Oxygen content  of soil  air  by a field
method. Soil Sci.  Soc. Am. J. 41:651-652.

Pearson, R. W. and F. Adams.   1967.   Soil acidity and liming.  Am. Soc.
Agron., Madison,  WI.

Pritchett, W. L.  and  W. H. Smith.  1972. Fertilizer responses in young
pine plantations.   Soil Sci.  Soc. Amer. Proc. 36:660-663.

Rajan, S. S. S.  1978.  Sulfate adsorbed on hydrous alumina,  ligands
displaced, and changes in  surface charge.   Soil  Sci. Soc. Amer. J. 42:
39.

Remade, J.  1977.  The role of heterotrophic nitrification  in acid
forest soils - preliminary results.   Ecol.  Bull.  (Stockholm)  25:560-561.

Reuss, J. 0.  1977.  Chemical and biological  relationships relevant  to
the effect of acid rainfall on the soil-plant system.   Water,  Air, and
Soil Pollut. 7:461-478.

Reuss, J. 0.  1978.  Simulation of nutrient loss  from soils  due to
rainfall acidity.  Research Report EPA-600/3-78-053, U.S. Environmental
Protection Agency, Corvallis, Oregon.  45 p.

Richards, B. N.  1974.  Introduction to the Soil  Ecosystem.   Longman,
New York.

Roberts, T. M., T. A. Clarke, P. Ineson, and T.  R. Gray.  1980.   Effect
of sulfur deposition on litter decomposition and nutrient leaching in
coniferous forest  soils, pp. 381-393.  In Effects of Acid Precipitation
on Terrestrial Ecosystems.  T. C. HutchTnson and M. Havas, eds.   Plenum
Press, New York.

Rorison,  I. H.  1980.  The effects of soil  acidity on  nutrient
availability and plant response, pp. 283-304. J^n Effects of Acid
Precipitation on Terrestrial Ecosystems.  T. C.  Hutchinson and M. Havas,
eds.   Plenum Press, New York.

Rosenqvist,  I. Th.  1977.   Sur Jord/Surt Vann.,  Ingeniorforlaget, Oslo,
Norway.   123 p.
                                  2-69

-------
Rosenqvist, I.  Th.  1978.   Alternative  sources  for  acidification of
river water in  Norway.   The Science  of  the  Total  Environment 10:39-49.

Rosenqvist, I.  Th.,  P.  Jorgensen,  and H.  Rueslatten.   1980.  The
importance of natural  H+ production  for acidity in  soil  and water. Jjn
Ecological Impact of Acid  Precipitation.  Drablos,  D., and A. Tollan,
Eds.  Johs. Grefslie Trykkeri  A/S, Mysen, Norway.

Schnitzer, M. and H. Kodama.  1977.   Reactions  of minerals with soil
humic substance, pp. 741-770.  ^Minerals  in Soil  Environments.  Soil
Sci. Soc. Amer., Madison,  WI.   Chapter  21.

Sedberry, J. E., Jr.,  Hatim Karim, and  B. J. Nitler.   1978.  The profile
distribution of total  and  DTPA extractable  manganese  in  selected soils
from the major mineral  soil areas  in Louisiana.   Comm. Soil Sci. PI.
Anal. 9(5):389-401.

Seip, H. M.  1980.  Acidification  of freshwater - Sources and mechanism,
pp. 358-366. ^n_ Ecological Impacts  of  Acid Precipitation.  D.
Drabltfs, and A. Tollan, eds.  Johs.  Grefslie Trykkeri  A/S, Mysen,
Norway.

Seip, H. M., G. Abrahamsen, E. T.  Gjessing, and A.  0.  Stuanes.  1979a.
Studies of soil-, precipitation-,  and run-off chemistry  in six small
natural plots ("mini-catchments").  SNSF-Project IP 46/79.

Seip, H. M., E. T. Gjessing, and H.  Kamben. 1979b.   Importance of the
composition of precipitation for the pH in  runoff - Experiments with
artificial precipitation on partly-soil-covered mini-catchments.
SNSF-Project IR 47/79.

Seip, H. M., S. Anderson,  and  B. Halsvik.   1980.  Snowmelt studied in a
mini-catchment with  neutralized  snow.   SNSF-Project IR 65/80.

Shaffer, K. A., D. D.  Fritton, and D. E.  Baker.   1979.   Drainage water
sampling in a wet, dual-pore soil  system.   J. Environ. Qua!. 8: 241-246.

Shriner, D. S.  and J.  W. Johnston.  1981.   Effects  of  simulated,
acidified rain on nodulation of leguminous  plants by Rhizobium spp.
Environ. Exp. Bot. 21:199-109.

Singh, B. R., G. Abrahamsen, and A.  Stuanes.  1980.   Effects of
simulated acid rain  on sulfate movement in  acid forest soils.  Soil Sci.
Soc. -Amer. J. 44:75-80.

Snyder, G. H.,  E. 0. McLean, and R.  E.  Franklin.  1969.  Interactions of
pH-dependent and permanent charges of clay: I.  Calcium  and Rb bonding
to bentonite and illite.  Soil Sci.  Soc.  Amer.  Proc. 33:392-396.

Sobotka, A.  1974.  Einfluxx von Immissionen auf die Wurzelnbildung mit
begrantem Wuchs bei  der Fichte.  I.  U.  F. R. 0.  IXC.   International
Tagung uber die Luftverunreinigung und  Fortwirtschaft.   15-18 Oktober.
Marianske Lazne, Tschechoslowakei, Tagungsbericht.
                                  2-70

-------
Soil Survey Staff.   1975.   Soil  Taxonomy.   USDA,  SCS.   Agric.  Hdbk. 436.
Govt. Printing Office,  Washington,  D.C.

Soil ins, P., C. C.  Grier,  F.  M.  McCorison,  K.  Cromack,  Jr.,  R. Fogel,
and R. L. Fredriksen.   1980.   The internal  element cycles  of an  old-
growth Douglas-fir  ecosystem  in  Western  Oregon.   Ecol.  Mon.  50:261-285.

Somers, E.  1961.  The  fungitoxicity of  metal  ions.   Ann.  Appl.  Biol.
49:246-253.

Stone, E. L. and R. Kszystyniak.  1977.   Conservation of potassium  in
the Pinus resinosa  ecosystem.  Science 198:192-193.

Strayer, R. F. and  M.  Alexander.  1981.   Effect of simulated acid rain
on glucose mineralization and some physicochemical properties of forest
soils.  J. Environ. Qual.  10:460-465.

Strayer, R. F., C.  J.  Lin, and M. Alexander.   1981.   Effect  of simulated
acid rain on nitrification and nitrogen  mineralization in  forest soils.
J. Environ. Qual. 10:547-551.

Stuanes, A. 0.  1980.   Effects of acid precipitation on soil  and forest.
5.  Release and loss of nutrients from a Norwegian forest  soil due  to
artificial rain of varying acidity, pp.  198-199.   In Ecological  Impact
of Acid Precipitation.   D. Drablj6s and A. Tollan, "eels.  Proc. of
International Conference, Sandefjord, Norway.

Tamm, C. 0.  1976.   Acid precipitation:   Biological  effects  in soil  and
forest vegetation.   Ambio 5:235-238.

Tamm, C. 0.  1977.  Acid precipitation and forest soils.  Water, Air,
and Soil Pollut. 7:367-369.

Tamm, C. 0., G. Wiklander, and B. Popovic.  1976.  Effects of
application of sulfuric acid to poor pine forests, pp. 1011-1024.   _In
Proceedings of the First International Symposium on Acid Precipitation
and the Forest Ecosystem.  L. S. Dochinger and T. A. Seliga, eds.   USDA
Forest Service General  Technical Report NE-23, Upper Darby,  PA.

Tamm, C. 0., G. Wiklander, and B. Popovic.  1977.  Effects of
application of sulfuric acid on poor pine forests.  Water, Air,  and Soil
Pollut. 8:75-87.

Tanaka, A. and Y. Hayakawa.  1974.  Comparative studies on plant
nutrition.  Crop tolerance to soil acidity.  1. Tolerance to low pH.   2.
Tolerance  to high levels of Al and Mn.  3. Tolerance to soil acidity
(J).  J.  Soil  Sci. PI. Nutr. 21: 305-308, 1974, 1975, 45,  46, 122.
561-570.   19-25, 26-32.

Terman,  G.  L.  1978.  Atmospheric Sulfur—The Agronomic Aspects.  Tech.
Bull. 23.   The Sulfur Institute, Washington, D.C.
                                  2-71

-------
Torgensen, H.   1934.   Forsok  med aparbeidelse av  sure,  fisketomme  vann.
Staugflskeren.   pp.  38-47.

Turner, J.  1980.   Interactions of sulphur with nitrogen  in  forest
stands.  Ir± Forest Fertilization Conference.   R.  Kennedy,  ed.   College
of Forest Resources,  University of Washington, Seattle.

Turner, J. and M.  J.  Lambert.   1980.   Sulfur  nutrition  of  forests,  pp.
321-333.  J_n Atmospheric Sulfur Deposition: Environmental  Impact and
Health Effects.  D.  S. Shriner, C. R.  Richmond, and  S.  E.  Lindberg, eds.
Ann Arbor Science, Ann Arbor,  MI.

Turner, J., D.  W.  Johnson,  and M.  J.  Lambert. 1980.  Sulphur  cycling  in
a Douglas-fir forest and its  modification  by  nitrogen applications.
Aecol. Plant 15:27-35.

Turner, J., M.  J.  Lambert,  and S.  P.  Gessel.   1977.  Use  of  foliage
sulphate concentration to predict  response to area application by
Douglas-fir.  Can. J. For.  Res. 7:476-480.

Turner, J., M. J.  Lambert,  and S.  P.  Gessell. 1979.  Sulfur
requirements of nitrogen fertilized Douglas-fir.   For.  Sci.  25:461-467.

Tveite, B.  1980a.  Effects of acid precipitation on soil  and  forest.
8. Foliar nutrient concentrations  in field experiments,  pp.  204-205.   Ir±
Ecological Impact of Acid Precipitation.   D.  Drabltfs and  A.  Toll an,
eds.  Norwegian Interdisciplinary  Research Programme -  Acid
Precipitation - Effects on  Forest  and Fish.  Proc. of International
Conference, Sandefjord, Norway.

Tveite, B.  1980b.  Effects of acid precipitation on soil  and  forest.
9. Tree growth in field experiments,  pp.  206-207. Jji Ecological  Impact
of Acid Precipitation.  D.  Drablj6s and A.  Tollan, eds.   Norwegian
Interdisciplinary Research  Programme - Acid Precipitation -  Effects on
Forest and Fish.  Proc. of an International Conference,  Sandefjord,
Norway, March 11-14, 1980.

Ulrich, B.  1980.   Production and consumption of  hydrogen ions in  the
ecosphere, pp. 255-282.  _In_ Effects of Acid Precipitation on Terrestrial
Ecosystems.  T. C. Hutchinson and M.  Havas, eds.   Plenum Press, New
York.

Ulrich, B., R. Mayer, and P.  K. Khanna.  1980.  Chemical  changes  due  to
acid  precipitation in a loess-derived soil in central  Europe.   Soil Sci.
130:193-199.

Vitousek, P. M., J. R. Gosz,  C. C. Grier,  J.  M.  Melillo,  W.  A. Reiners,
and R. L. Todd.  1979.  Nitrate losses from disturbed  ecosystems.
Science 204:469-474.

Wainwright, M.  1979.  Microbial S-oxidation in  soils  exposed  to  heavy
atmospheric pollution.  Soil  Biol. Biochem. 11:95-98.


                                  2-72

-------
Wainwright, M.  1980.  Effect of exposure to atmospheric pollution on
microbial  activity in soil.   PI. Soil  55:199-204.

Way, T.  1855.  The atmosphere as a source of nitrogen  to plants.   Royal
Agr. Soc.  J. 16:249-267.

Wiklander, L.  1974.  The acidification of soil  by acid precipitation.
Grundforbattring 26:155-164.

Wiklander, L.  1975.  The role of neutral salts  in the  ion exchange
between acid precipitation and soil.  Geoderma 14:93-105.

Wiklander, L.  1980a.  Interaction between cations and  anions
influencing adsorption and leaching, pp. 239-254.   In Effects of Acid
Precipitation on Terrestrial  Ecosystems.  T. C.  HutcEinson and M.  Havos,
eds.  Plenum Press, New York.

Wiklander, L.  1980b.  The sensitivity of soils  to acid precipitation,  .
pp. 553-567.  Jji Effects of Acid Precipitation on  Terrestrial
Ecosystems.  T. C. Hutchinson and M. Havas, eds.   Plenum Press, New
York.

Will, G. M. and T. C. Youngberg.  1978.  Sulfur status  of some central
Oregon pumice soils.  Soil Sci. Soc. Amer J. 42:132-134.

Wilson, J. T. and M. Alexander.  1979.  Effect of  soil  nutrient status
and pH on nitrogen-fixing algae in flooded soils.   Soil Sci.  Soc.  Am. J.
43:936-939.

Youngberg, C. T. and C. T. Dyrness.  1965.  Biological  assay  of pumice
soil fertility.  Soil Sci. Soc. Am. Proc. 29:182-187.

Yuan, T. L.  1963.  Some relationships among hydrogen,  aluminum, and pH
in solution and soil systems.  Soil Sci. 95:155-163.
                                  2-73

-------
            THE ACIDIC DEPOSITION  PHENOMENON AND  ITS EFFECTS
                      E-3.   EFFECTS  ON  VEGETATION

3.1  INTRODUCTION

3.1.1  Overview

     This chapter examines  diverse plant-pollutant  relationships to
assess potential  and recognized  effects of acidic deposition as
described in the extant literature.   Vegetation responses discussed
include morphological  and physiological responses,  species/varieties and
life-stage susceptibilities,  disease and  insect stresses, indirect
effects of nutrient cycle alterations,  and crop and forest productivity.

     Since the close relationship  between soils and plants bears
examination in terms of how soil acidification affects productivity.  It
is important to recall from the  previous  chapter  the following points:

     o   soils amended in agricultural  practice will not likely be
         negatively impacted by  acidic  deposition;

     o   soil acidification is a natural  process  in humid regions, so
         most soils that are easily  acidified are already acid; and

     o   soils with low buffering  capacity, relatively high pH, low
         sulfate adsorption capacity, no  carbonates, and no basic inputs
         are susceptible to increased acidification rates from
         atmospheric inputs of acidic and acidifying substances.

     Keeping these points in mind, Chapter E-3 will deal more with the
direct effects of acidic deposition  on  plant response, and to
interactive effects of acidic deposition  with other factors, such as
other pollutants, insects,  pathogens, and pesticides.

     Given the uncertainty  still surrounding effects on plant
productivity, however, this document does not attempt to make economic
assessments of recognized or potential  damage to  vegetation; nor does it
consider mitigative measures to  counter acidic deposition inputs to
plant systems.  Discussions of nutrient cycling and forest productivity
are included in both this chapter  and the soils chapter, from slightly
different perspectives.  Both chapters  should be  read carefully to gain
a more complete understanding of the issues.
                                  3-1

-------
3.1.2  Background (P.  M.  Irving  and  S.  B. Mclaughlin)

     The observation that both gaseous  and  rain-borne pollutants affect
vegetative growth is not  limited to  recent  years.  Robert Angus Smith
(1872) in his manuscript,  "Air and Rain:  The Beginnings of a Chemical
Climatology," included a  section on  "Effect of Acid Gases on Vegetation
and Capability of Plants  to  Resist Acid Fumes."  As early as 1866 the
Norwegian playwrite Ibsen  (1866)  referred to the phenomenon in the drama
"Brand":
          A sickening fog  of  smoke  from British coal
     Drops in a grimy pool  upon the land,
     Befouls the vernal  green and chokes  to death
     Each lovely shoot,
Of course the fog of smoke  referred to by  Ibsen was from imported
British coal  and not from the  long-range transport of pollutant gases.
An  intensive effort to study  the effect of acidic deposition was not
initiated until  the Norwegian  SNSF (Sur Nedbj6rs Virkning Pa Skog Og
Fisk--"Acid Rain Effects on Forests and Fish") Project was established
in 1972.  The phenomenon was first widely  recognized in North America at
the First International Symposium on  Acid  Precipitation and the Forest
Ecosystem in Ohio (USDA 1976), and at the  NATO Conference on Effects of
Acid Precipitation on Vegetation and  Soils (Toronto 1978).  At the Ohio
conference, Tamm and Cowling (1976) speculated upon the potential
effects of acidic deposition,  but few existing studies directly
supported their hypotheses  of  damaging effects.

     As the acid rain phenomenon gained increasing attention and its
occurrence was reported over large areas of North America, economic
damage to vegetation was predicted and a number of research programs to
investigate the effects were initiated in  the mid-1970's.

     Anthropogenic and natural  air contaminants are usually inventoried
on a separate basis (e.g.,  chemical speciation) when information is
sought as to sources, dispersion, or  induced effects (see Chapters A-2
and A-5).  Categorically, the  National Ambient Air Quality Standards
(NAAQS) for criteria pollutants (ozone, sulfur oxides, hydrocarbons,
nitrogen dioxides, carbon monoxide, and particulate matter) have been
established to protect human health and welfare.  Comprehensive
documents that describe vegetation effects of the major phototoxic air
pollutants are available (U.S. EPA 1978, 1982).  As distances from
pollutant sources increase, chances for combinations to occur also
increase, or, as in the case of large metropolitan/industrial areas,
pollutant combinations are  the rule rather than the exception.

     The wet deposition of  acidic pollutants may consist of a number of
variables affecting vegetation (i.e., hydrogen, sulfur, and nitrogen
doses).  The influence of predominant gaseous pollutants that may be
present within the defined  isopaths of acidic precipitation must also be
                                  3-2

-------
taken into account.   If  results of  such interaction studies are not
available or understood, effects may be attributed to acidic depositions
but instead be due to gaseous  pollutants alone, or as combined with the
influence of acidic  depositions.  Because of the potential for
interactions with biotic and abiotic entities, factorial research
designs and multivariate analyses may be necessary to gain a more
complete understanding of  vegetative response to acidic deposition.

     In the United States,  the eastern half of the country is the
geographical area of major concern  for impacts of air pollution (both
gaseous pollutants and acid rainfall) on crop and forest productivity.
The combination of a high  density of fossil-fuel combustion plants, a
high frequency of air stagnation episodes, and elevated levels of both
photochemical oxidants and rainfall acidity over widespread areas of the
eastern United States have resulted in exposure of large acreages of
forests to increased deposition of  atmospheric pollutants (Mclaughlin
1981).  An overlay of isopleths of  air stagnation frequency (a measure
of the potential of pollutants to accumulate during periods of limited
atmospheric dispersion), isopleths  of rainfall acidity, and forest zones
of the United States is  shown  in Figure 3-1.  This overlay highlights
this juxtaposition of stress potential and forest types.  While air
stagnation episodes are  not in themselves a measure of air pollution
stress, they do provide  an indication of the potential for pollutants
from multiple sources to be concentrated within regional air masses.
The eastern half of the  United States, with approximately 80 percent of
the total fossil-fueled  electric power plants, thus has both the
emissions and the atmospheric  conditions to create regional scale
elevation of air pollutants (see Chapter A-2).  Comparable conditions
also appear to exist in  coastal California, where severe air stagnation
has led to very high levels of phdtochemical oxidants.  The acidity of
rainfall in much of the  Northeast  quadrant of  the United States (Figure
3-1) averages about pH 4.1 to  4.3  annually--about 30 to 40 times as acid
as the hypothetical  carbonate-equilibrated natural rainfall with a pH of
5.6 (Likens and Butler 1981).  Vegetation in the high-altitude boreal
forests of New England experiences even greater inputs, being exposed to
hundreds of hours during the growing season to clouds with pH values in
the range of 3.5 to 3.7  (Johnson  and Siccama 1983).  Photochemical
oxidants, principally ozone, which are formed  both naturally in
reactions involving ultraviolet  radiation and  from biogenic and
anthropogenic hydrocarbon and  nitrogen oxide precursors, occur at
potentially phytotoxic levels  over the entire  eastern region (Westburg
et al. 1976). Forest productivity  losses from  this pollutant have not
been quantified except in southern California, where extreme urban
pollution from the Los Angeles Basin and poor  air dispersion have
combined to produce the highest oxidant concentrations  in the United
States and widespread mortality  and decline of forests  in the nearby San
Bernadino Mountains (Miller et al.  1977).
                                  3-3

-------
     1-BOREAL FOREST ECOSYSTEM
     J-UAICE STATES FOREST ECOSYSTEM
     3-EASTERN DECIDUOUS FOREST ECOSYSTEM
     4-SOUTH EASTERN PINE FOREST ECOSYSTEM
     6-TROPICAL FOREST ECOSYfEM
     6-WESTERN MONTANE FOREST ECOSYSTEM
     7-SUBALPINE FOREST ECOSYSTEM
     8-PACIFIC COAST FOREST ECOSYSTEM
     9-CALIFORNIA WOODLAND
     10-SOUTHWESTERN WOODLAND
Figure  3-1.
Distribution of frequency isopleths for total  number of
forecast days with high  meteorological  potential  for air
pollution over  a 5-year  period.   Isopleths are shown in
relation to major forest types  of the  United  States
(adapted from Miller  and McBride 1975}  and in relation  to
mean  annual hydrogen  ion (kg  ha'1 yr-1) deposition in
precipitation  (adapted from Henderson  et al.  1981).
                                      3-4

-------
3.2  PLANT RESPONSE TO ACIDIC DEPOSITION

3.2.1  Leaf Response to Addle Deposition  (D.  S.  Shriner)

     Any discussion of foliar effects of acidic  deposition must  be
prefaced by a recognition that our knowledge of  the  potential  effects
are drawn from experimental  observations with simulated rain  solutions
rarely typical of ambient events.   As a result,  in the  absence of field
observation of effects due to ambient precipitation  events, it is
important to recognize that these  experimental observations are  most
useful for understanding mechanisms of effect, and less so for
extrapolation to field-scale impacts.

     Most of the terrestrial landscape being impacted by acidic
deposition is covered by a minimum of one  layer  of vegetation.   As  a
result, a large proportion of the  incident precipitation ultimately
affecting soils and surface water  chemistry has  previously contacted
vegetation surfaces.  The fact that vegetation surfaces are perhaps the
most probable primary receptors of deposited pollutants raises two
important issues regarding the interactions between  water droplet and
receptor surface:

     1)  effects of incident precipitation chemistry on the receptor
         surface structure and function; and

     2)  effects of the receptor surface on incident precipitation
         chemistry.

3.2.1.1  Leaf Structure and Functional Modifications—Based on experi-
mental evidence with simulated rain, a wide range of plant species  1s
believed to be sensitive to direct Injury  from some  elevated  level  of
wet acidic deposition (Evans et al. 1981b, Shriner 1981; see  also
Section 3.4).  Other species have  been noted to  be tolerant of equally
elevated levels (to pH 2.5 for up  to 10 hours total  exposure)  without
visible injury (Haines et al. 1980).  These results  suggest that
generalizations about sensitivity  to Injury may  be difficult,  and some
understanding of the mechanisms by which  injury  may  occur is  necessary.
The sensitivity of an individual species of vegetation  appears to be
influenced by structural features  of the vegetation, which 1)  influences
the foliage wettability; 2)  makes  the foliage more vulnerable to injury
(e.g., through differential  permeability of the  cuticle); or  3)  retains
rainwater due to leaf size,  shape, or attachment angle.  In those
instances where one or more of the above conditions  renders a plant
potentially sensitive to acidic deposition, effects  may be manifested in
alterations of leaf structure or function.

     Injury to foliage by simulated acidic precipitation largely depends
on the effective dose to which sensitive tissues are exposed.  The
effective dose, that concentration and amount of hydrogen 1on, and  time
period responsible for necrosis of an epidermal  cell, for example,  are
influenced by the contact time of  an individual  water droplet or film on
the foliage surface (Evans et al.  1981b, Shriner 1981).  Contact time,
                                3-5

-------
in turn, can be regulated by the wettability of the leaf,  or by  leaf
morphological features that prevent rapid runoff of water  from the
surface.  Physical characteristics of the leaf surface (e.g., roughness,
pubescence, waxiness) or the chemical  composition of the cutin and
epicuticular waxes determine the wettability of most leaves (Martin and
Juniper 1970).

     For injury to occur at the cellular level, the ions responsible
must penetrate these protective physical and chemical  barriers or enter
through stomata (Evans et al.  1981b).   Crafts (1961a)  has  postulated
that cuticle penetration occurs through micropores.   Evidence indicates
that these micropores are most frequent in areas such  as at the  bases  of
trichomes and other specialized epidermal  cells (Schnepf 1965).  However,
the occurrence of such micropores is not well  documented for all  plant
cuticles (Martin and Juniper 1970).  Hull  (1974) demonstrated that basal
portions of trichomes are more permeable than adjacent areas; cuticles
of guard cells and subsidiary  cells are preferred absorption sites
(Dybing and Currier 1961, Sargent and  Blackman 1962).   In  addition,
Linskens (1950) and Leonard (1958)  found that the cuticle  near veins is
apparently a preferential site for absorption of water-soluble
materials.

     Perhaps as important as the greater density of micropores
associated with these specialized cells is Rentschler's (1973) evidence
that, at least in certain species,  epicuticular wax is less frequently
present on certain of these specialized epidermal  cells. Such an absence
of wax, in combination with increased  cuticular penetration at those
sites,  would tend to maximize  the sensitivity of those sites. Evans et
al. (1977a,b; 1978) have determined that approximately 95  percent of the
foliar lesions occurring on those plant species observed by them
occurred near the bases of such specialized epidermal  cells as
trichomes, stomatal guard and  subsidiary cells, and along  veins.
Stomatal penetration by precipitation,  on the other hand,  is thought to
be infrequent (Adam 1948; Gustafson 1956,  1957; Sargent and Blackman
1962) and is considered a relatively insignificant route of entry of
leaf surface solutions (Evans  et al. 1981b).

     Solution pH has also been shown to influence the  rate of cuticular
penetration in studies with isolated cuticles (Orgell  and  Weintraub
1957, McFarlane and Berry 1974).   The  rate of penetration  of acidic
substances increased with a decrease in pH, while the  rate of
penetration of basic substances increased with an increase in pH (Evans
et al.  1981b).

     Preliminary work by Shriner (1974) suggested that,  in addition to
the physical abrasion of superficial wax structure by  raindrops,  leaves
exposed to rainfall of pH 3.2  appeared to weather more rapidly than did
leaves  of pH 5.6 control treatment plants.  However,  it was impossible
to determine from those experiments whether chemical  processes at the
wax surface were responsible for the differences or whether the  acidic
rain induced physiological  changes that retarded regeneration of the
                                  3-6

-------
 waxes and recovery  from mechanical damage.  The latter explanation may
 be the most  tenable because the waxes would be expected to resist
 chemical  reaction with dilute  strong acids (Evans et al. 1981b), and
 because numerous reports of physiological imbalance resulting from
 acidic precipitation exposure  exist (Shriner 1981).  Hoffman et al.
 (1980) proposed a mechanism by which precipitation acidity can act as a
 chemical  factor in  weathering  epicuticular waxes.  They pointed out that
 the wax composition, as polymeric structures of condensed long-chain
 hydroxy carboxylic  acids, may  result in an "imperfect" wax matrix in
 which the uncondensed sites containing hydroxy functional groups are
 more readily weathered.  Strong acid inputs to such a system would
 oxidize and  release a wide range of carbon chain acids from the basic
 waxy matrix, conceivably yielding the type of change in weathering rate
 Shriner observed.

      Rentschler (1973) and, more recently, Fowler et al. (1980) have
 shown relationships between the superficial wax layer of plants and
 plant response to gaseous air pollution.  The work of Fowler et al.
 compared  the rate of epicuticular wax degradation of Scots pine needles
 from "polluted" and unpolluted sites in the field.  These "polluted"
 sites included exposure to both dry deposition of gaseous pollutants and
 wet deposition as acid rain, making it impossible to distinguish between
 relative  effects of the two forms of deposition.   Needles at the
 polluted  site showed greater epicuticular wax structure degradation
 during the first eight months of needle expansion.  Determing the
 quantity  of wax per unit leaf area showed very small  differences between
 polluted  and clean  air sites.   Fowler et al.  concluded that observed
 differences  (by scanning electron microscopy)  were "due more to changes
 in  form than gross  loss of wax."  Since the fine structure of the wax
 layer is  controlled largely by the chemical  composition of the wax
 (Jeffree  et al. 1975), the observed changes may also reflect
 stress-induced changes In wax synthesis.   Fowler et al. estimated that
 increased water loss due to accelerated breakdown of cuticular
 resistance would only influence trees  if water were a limiting factor.
 They concluded that "the extra water loss may reduce the period (or
 degree) of stomatal  opening" and that  the magnitude of the effect on dry
 matter productivity would not be greater than 5 percent at their
 polluted  site.   Because study sites used by  Fowler et al.  were exposed
 to  gaseous sulfur dioxide as well  as to acidic precipitation,  their work
 does not allow identification  of a single causative factor.

     Histological  studies of foliar injury caused by  acidic
 precipitation have revealed evidence of modification  of leaf structure
 associated with plant exposure to acidic  precipitation (Evans  and Curry
 1979).  Quercus palustris,  Tradescantia sp.,  and  Populus  sp. exposed to
 simulated acidic precipitation experienced abnormal cell  proliferation
 and cell  enlargement.   In Quercus  (oak)  and Populus (poplar) leaves,
 prolonged exposure to treatment at pH  2.5 produced hypertrophic  and
 hyperplastic  responses in mesophyll  cells.  Lesions developed,  followed
by enlargement  and proliferation of adjacent cells, resulting  in
 formation of  a  gall  on adaxial  leaf surfaces.   In  poplar  test  plants,
                                  3-7

-------
this response involved both palisade and spongy mesophyll parenchyma
cells, while in oak test plants,  only spongy mesophyll cells were
affected (Evans and Curry 1979).   Because other similar histological
studies have not been reported,  it is impossible to evaluate how
frequent or widespread such structural modification may be.  Because
species that have been reported  to show  hyperplastic and hypertrophic
response of leaf tissues were consistently injured less than species
that did not show these responses, gall  formation may be linked to
characteristics common to species tolerant of  acidic precipitation
exposure.

     Several studies have reported modification of various physiological
functions of the leaf as a result of exposure  to simulated acidic
precipitation.  Sheridan and Rosenstreter (1973), Ferenbaugh (1976),
Hindawi et al. (1980), and Jaakhola et al.  (1980) reported reduced
chlorophyll  content as a result of tissue exposure to acidic solutions.
Ferenbaugh,  however, observed that significant reduction in chlorophyll
content did not occur at pH 2.0,  and that chlorophyll content slightly
increased at pH 3.0.  Irving (1979)  also reported higher chlorophyll
content of leaves exposed to simulated precipitation at pH 3.1.  Hindawi
et al. observed a steady reduction in chlorophyll content in the range
between pH 3.0 to 2.0, and found  no change in  the ratio of chlorophyll
a:b.

     Ferenbaugh (1976) determined photosynthesis and respiration rates
of test bean plants exposed to simulated acidic precipitation.
Respiration and photosynthesis were significantly increased at pH 2.0.
Ferenbaugh concluded that because growth of the plants was significantly
reduced, photophosphorylation was uncoupled by the treatments.  Irving
(1979) reported increased photosynthetic rates in some soybean
treatments,  attributing them to  increased nutrition from sulfur and
nitrogen components of the rain  simulant, which overcame any negative
effect of the pH 3.1 treatment.   Jacobson et al. (1980) reported a shift
in photosynthate allocation from  vegetative to reproductive organs as a
result of acidic rain treatments  of pH 2.8 and 3.4, also suggesting that
the primary effect was not on the photosynthetic process itself.

3.2.1.2  Foliar Leaching - Throughfall Chemistry—Rain, fog, dew, and
other forms  of wet deposition play important roles as sources of
nutrients for vegetation and as mechanisms of  removal from vegetation of
inorganic nutrients and a variety of organic substances:  carbohydrates,
amino acids, and growth regulators (Kozel  and  Tukey 1968, Lee and Tukey
1972, Hemphill and Tukey 1973, Tukey 1975).  Tukey (1970, 1975, 1980)
and Tukey and Morgan (1963) have  extensively reviewed the leaching of
substances from plants as the result of  water  films on plant surfaces.

     During periods between precipitation events, the vegetation canopy
serves as a sink, or collection surface,  upon  which dry particulate
matter, aerosols, and gaseous pollutants accumulate by gravitational
sedimentation, impaction, and absorption.  Throughfall can be defined as
that portion of the gross, or incident,  precipitation that reaches the
                                  3-8

-------
forest floor through openings In the forest canopy  and by  dripping  off
leaves, branches, and stems (Patterson 1975).   Throughfall  generally
amounts to between 70 and 90 percent of gross  rainfall, with  the  balance
divided between stemflow and interception loss to the  canopy.

     Chemical  enrichment of throughfall  has been well  documented  for  a
broad variety of forest species (Tamm 1951, Madgwick and Ovington 1959,
Nihlgard 1970, Patterson 1975, Lindberg and Harriss 1981).  This
enrichment has three potential sources:   1) reactions  on the  leaf
surface in which cations on exchange sites of  the cuticle  are  exchanged
with hydrogen from rainfall; 2) movement of cations directly  from the
translocation stream within the leaf into the  surface  film of  rainwater,
dew, or fog by diffusion and mass flow through areas devoid of cuticle
(Tukey 1980);  and/or 3) washoff of atmospheric particulate matter that
has been deposited on the plant surfaces (Patterson 1975,  Parker  et al.
1980, Lindberg and Harriss 1981).

     The exchange of hydrogen ions in precipitation for cations on  the
cuticle exchange matrix can result in significant  scavenging  of hydrogen
ions by a plant canopy.  Eaton et al. (1973),  for example,  found  the
forest canopy to retain 90 percent of the incident  hydrogen ions  from pH
4.0 rain (growing season average), resulting in less acidic (~ pH 5.0)
solutions reaching the forest floor.  The removal of H+ by exchange
processes in the forest canopy does not eliminate the  effects  of  H+
deposition on the forest ecosystem, however.  Cations  leached  from  the
foliage may eventually be leached from the ecosystem if the anion
associated with H+ inputs ($042- or NOa") is mobile (see
Figure 2-1, Chapter E-2).  Plant response to this may  be 1) accelerated
uptake to compensate for foliar cation losses, or 2) reduced  foliar
cation concentrations, if H+ inputs and foliar exchange are of
significant magnitude and duration.  In either event,  the  introduction
of H+ with a mobile anion will cause the net loss of cations  from the
ecosystem, whether the H+ cation exchange  occurs 1n the forest canopy
or the soil.  Further aspects of cation leaching are discussed in
Chapter E-2 (soils).

     An example of the second case has recently been hypothesized by
Rehfuess et al. (1982) for Norway spruce in high elevation forests  of
eastern Bavaria.  Trees experiencing symptoms  of decline and  dieback
were paired with non-symptomatic trees in the  same  stands  and  site
conditions.  Large differences were noted in foliar content,
particularly of older leaves, of Ca and Mg, with declining trees
consistently showing lower levels of Ca and Mg content than healthy
trees.  The Mg contents were characterized by  the authors  as  in "extreme
deficiency," with calcium in "poor supply." The authors further
speculated that since these nutrient deficiences occurred  on  soils
varying considerably in content of both elements, that soil depletion
was probably not the dominant contributing factor,  but rather  that  the
deficiency is mainly a consequence of enhanced leaching of Ca  and Mg
from the foliage as a result of acidic deposition of strong acids.  The
authors further speculated that Ca and Mg uptake from  soil  pools  may  be
                                  3-9

-------
inadequate to replace this foliar leaching.   Such  nutritional  disorders
have been reported to subsequently make  foliage more  susceptible  to
additional leaching (Tukey 1970).

     Separating relative contributions of internal  (leached)  and
external (washoff) fractions of throughfall  enrichment  is  difficult  and
has been attempted infrequently. Parker et al.  (1980) have reviewed
those attempts to estimate the importance of dry sulfur deposition to
throughfall enrichment by sulfate-sulfur (Table 3-1).   For those  studies
that have attempted such an analysis,  the estimated percentage
contribution of dry deposition to throughfall  enrichment ranged from 13
to 100 percent, or from 0.3 to 14.4 kg ha~l  yr-1.   Parker  et  al.
concluded that for temperate hardwood  forests in industrialized regions,
40 to 60 percent of annual net throughfall  (throughfall  enrichment)  for
sul fate is due to washoff of dry deposition, with  30  to 50 percent being
typical for conifers of the same regions.  For hardwoods and  conifers  in
regions typified by low background levels of dry sulfur deposition,
washoff may range from 0 to 20 percent of throughfall enrichment.
Similar data have been developed for several trace elements (Lindberg
and Harriss 1981).

     Through the application of simulated rainfall  in controlled
experiments, precipitation acidity has been  studied as  a variable
influencing the leaching rate of various cations and  organic  carbon  from
foliage (Wood and Bormann 1974, Fairfax  and  Lepp 1975,  Abrahamsen et al.
1977).  Foliar losses of potassium, magnesium, and calcium from bean and
maple seedlings were found to increase as the acidity of simulated rain
increased.  Tissue injury occurred below pH  3.0, but  significant
increases in leaching rates occurred as  high as pH 4.0  (Wood  and  Bormann
1974).  Phaseplus yulgaris L.  foliage  exposed by Evans  et  al.  (1981a)  to
citrate-phosphate buffer solutions with  a range in acidity from pH 5.7
to pH 2.7 also demonstrated that greater acidity of these  solutions
preferentially leached greater amounts of calcium,  nitrate, and sulfate,
while less acidic solutions leached greater  amounts of  potassium  and
chloride.  Abrahamsen and Dollard (1979) observed  that  Norway spruce
(Picea abies (L.) Karst) lost greater  quantities of nutrients  under
their most acidic treatments,  but no related change in  foliar cation
content occurred, in contrast to the observations  of  Rehfeuss  et  al.
(1982) discussed above.  Wood and Bormann (1977) noted  results similar
to those of Abrahamsen and Dollard (1979) for eastern white pine  (Pinus
strobus L.).

3.2.2  Effects of Acidic Deposition on Lichens and Mosses  (L.  L.  Si gal)

     The objective of this section is  to review the literature on the
effects of acidic deposition on lichens  and  mosses and  also to review
the literature that describes the effects of realistic,  low levels of
gaseous sulfur dioxide (S02) on lower  plants (Grennfelt et al. 1980).
Several researchers (Skye 1968, Turk and Wirth 1975)  have  concluded  that
S02 toxicity and pH effects are not independent factors.
                                  3-10

-------
    TABLE 3-1.   REPORTED VALUES FOR SULFATE-SULFUR DEPOSITION RATES FOR THROUGHFALL  AND INCIDENT

                                    PRECIPITATION IN WORLD FORESTS
co
i
Forest system
Subalpine balsam fir,
New Hampshire
Hemlock,
British Columbia
Conifers,
southern Norway
Conifers,
southern Norway
Conifers,
southern Norway
Beech,
central Germany
Spruce,
central Germany
Hemlock-spruce,
southeastern
Alaska
Tropical rain forest,
Costa Rica
Douglas fir,
Reference
Cronan 1978
Feller 1977
Haughbotn 1973
Haughbotn 1973
Haughbotn 1973
Heinrichs and Mayer
1977
Heinrichs and Mayer
1977
Johnson 1975
Johnson 1975
Johnson 1975
S deposition
Incident
24.4
11. Qa
32. 3b
17.7
10.0
24. ld
24.1
0
12.5
4.0
kg ha"1 yr"1
Throughfall
46.4
40.0
111.2
69.1
21.1
47.6
80.0
16.4
23.3
5.2
Precipitation
amount
(cm)
203d
245C
77
77
77
106
106
270
390
165
        Washington

-------
TABLE 3-1.  CONTINUED
Forest system
Subalpine silver fir,
Washington
Hardwoods,
Amazonian Venezuela
Hardwoods,
Amazonian Venezuela
Hard beech,
New Zealand
CO
jL Beech,
^ Southern Sweden
Spruce,
Southern Sweden
Oak,
Southern France
Loblolly pine,
North Carolina
Chestnut oak,
Tennessee
Mixed oak, Tennessee
Mixed oak, Tennessee
Reference
Johnson 1975
Jordan et al. 1980
Jordan et al . 1980
Miller 1963
Nihlgard 1970
Nihlgard 1970
Rapp 1973
Wells et al. 1975
Lindberg et al . 1979
Kelly 1979
Kelly 1979
S deposition
Incident
16. 8f
44.5
46.6
8.4
7.9<1
7.9"
16.4
7.93
13.2b»e
8.73
11.3a»b
kg ha"1 yr'1
Through fall
5.3
16.7
19.6
10.4
18.5
54.2
22.6
9.9
32.0
15.0
14.0
Precipitation
amount
(cm)
300
391
412
135
95
95
NA
NA
143
154
75

-------
                                               TABLE 3-1.  CONTINUED
      aScaled up from a subannual  estimate.
       In  vicinity of factory  or power plant.
      °Mean of extreme estimates.
       Includes stem flow.
      eSeveral  years data.
      fLittle throughfall.
CO
I

-------
     Lichens and mosses are considered by some researchers  (Nieboer  et
al. 1976) to be among the most pollution sensitive plants,  and  by  others
to be more sensitive and better indicators of chronic  pollution than
vascular plants (Hawksworth 1971,  Nash 1976,  Guderian  1977,  Winner et
al. 1978).  In addition to their roles in the ecosystem,  they are  also
valuable as biomonitors of air quality.   However,  it must be noted that
lichens and mosses integrate the effects of all  ambient pollutants,  and
in most cases, their use as bioindicators is  only  an index  of general
air pollution.

     Lichens are sensitive to air pollutants  such  as sulfur dioxide,
(Ferry et al. 1973), ozone and peroxyacetyl  nirate (PAN)  (Nash  and Sigal
1979, Sigal and Taylor 1979), fluorine (Nash  1971, Roberts  and  Thompson
1980), and metals (Rao et al. 1977;  lead, Lawrey and Hale 1981;  nickel,
Nieboer et al. 1972; mercury, Steinnes and Krog  1977;  zinc,  Nash 1975;
and chromium, Schutte 1977).   Scientists in many countries  have
demonstrated that it is possible to  correlate the  distribution  of
lichens around air pollution  sources with mean levels  of  air pollutants.
Laboratory and transplant studies  have'corroborated the data from  field
investigations.  However, the importance of peak concentrations of
pollutants relative to long-term average levels  has not been
established.  Excellent summaries  on the theory  and application of
lichens in pollution studies  have  been published by Ferry et al. (1973),
Gilbert (1974), Hawksworth and Rose  (1976), Le Blanc and  Rao (1975),
Richardson and Nieboer (1981), Skye  (1968, 1979),  and  Saunders  (1970).
In addition, the air pollution literature is  regularly indexed  in  the
British journal "The Lichenologist"  (1974-81).

     Moss species are also sensitive to  air pollution  (Gilbert  1968,
1970; Nash 1970; Nash and Nash 1974; Stringer and  Stringer  1974; Turk
and Wirth 1975; Winner and Bewley  1978a,b).   However,  less  attention has
been given to mosses in air pollution research.  Laboratory  studies  with
mosses have shown that 1) photosynthesis decreases in  relation  to  a
decrease in pH of sulfuric acid solutions (Sheridan and Rosenstreter
1973), 2)  sulfite and bisulfite solutions reduce photosynthesis  (Ing!is
and Hill 1974, Ferguson and Lee 1979), and 3)  growth of four species of
Sphagnum moss was reduced when they  were fumigated for several  months
with mean S02 concentration of 130 yg nr3 (Ferguson et al.  1978).
It has been suggested that sulfate at "feasible" atmospheric
concentrations has no effects upon photosynthesis  in mosses; however,
the fall in pH that accompanies the  oxidation of atmospheric SOg to
504 is capable of reducing photosynthesis (Ferguson and Lee  1979).
The phytotoxic effect of S02  for both mosses  and lichens  is  known  to
be greater at low pH (Gilbert 1968,  Puckett et al.  1973,  Inglis  and  Hill
1974, Hallgren and Huss 1975).

     The generally accepted mechanisms of injury are disruption  of cell
and chloroplast membranes (Wellburn  et al.  1972, Puckett  et  al.  1974,
Malhotra 1976, Ferguson and Lee 1979), and destruction of chlorophyll
(Rao and Le Blanc 1966, Nash  1973, Puckett et al.  1973).  Susceptibility
to S02 injury .is greatest when lichens are in  a  moistened or saturated
                                  3-14

-------
condition (Rao and Le Blanc 1966;  Nash  1973,  1976;  Turk  et al.  1974).
In an air-dried state, lichens  have been  shown  to be  relatively
insensitive to S02 (Showman 1972,  Nash  1973,  Turk et  al.  1974,  Marsh
and Nash 1979).

     The sensitivity of lichens to air  pollutants is  due to  a number of
factors:  (1) they rapidly absorb  moisture in different  forms (e.g.,
rain, fog, dew) and most toxic  substances dissolved in the water
(Richardson and Nieboer 1981);  (2) they are long-lived,  and  accumulated
sulfur metabolites, metals, etc. are not  eliminated seasonally  (Nash
1976); (3) they lack a vascular system  with which to  eliminate
pollutants through translocation (Nieboer et al. 1976);  (4)  they lack
structures such as epidermis and stomata  to exclude pollutants
(Sundstrom and Hallgren 1973);  (5) they probably have less buffering
capacity than vascular plants (Nieboer  et al. 1976);  and (6)  the
relationship of the alga and the fungus is delicately balanced; air
pollution probably disrupts that balance, resulting in disassociation
and destruction of the plant (Neiboer et  al.  1976).

     The ecology of lichens can be drastically  changed by air
pollutants.  As a result, ecosystems are  affected because lichens  are
integral parts of many relationships and  processes.   As  pioneer species
in disturbed areas (Treub 1888),  lichens  initiate soil formation (Ascaso
and Galvin 1976) and stabilize  soil (Rychert and Skujins 1974,  Drouet
1937).  They fix an estimated 10 to 50  percent  of the newly-fixed
nitrogen in old growth forests  in  the United States (Denison 1973,
Becker 1980, Rhoades 1981).  They  act as  sinks  for  air pollutants  and
contribute to the cleansing of  the atmosphere (A. C.  Hill 1971).

     Many invertebrates (mites, caterpillars, earwigs, snails,  slugs,
etc.) as well as vertebrates (caribou,  reindeer, squirrels,  woodrats,
voles) feed partly or wholly on lichens (Llano  1948,  Richardson 1975,
Gerson and Seaward 1977, Richardson and Young 1977).  Other  animals have
adaptive camouflage that resembles lichen-covered trees  or rocks
(Richardson and Young 1977).  The  interrelations among birds and lichens
and insects are multifaceted.  Birds use  lichens for  nest-building,
camouflage, and feeding behavior (Kettlewell  1973,  Ewald 1982), while
many insects have co-evolved with  lichens to escape predation from birds
(Cott 1940).

     Reports of injury to lichens  at low  levels of  S02 are found in
several recent studies.  Showman (1975) found that  Parmelia  caperata and
P_. rudecta were absent in regions  around  a coal-fired power  plant  when
the annual S02 average exceeded 50 yg nr3.  Will-Wolf (1980)  found
that Parmelia caperata and P. bolliana  showed morphological  alterations
in areas where maximum S02 Tevels  were  389 yg m-3,  and annual
averages were 5 to 9 yg m-3.  Eversman  (1978) found decreased
respiration rates in Usnea hirta after  field fumigations with S02  at
about 47 yg m-3 for 96 days, and plasmolysis of algal cells  in  both
U^ hirta and Parmelia chlorochroa  after 31 days of  S02 at the same
concentration!Le Blanc and Rao (1975) concluded that long-range
                                  3-15

-------
average concentrations for S02  between  16  to 79 yg m-3  (0.006 to
0.03 ppm) cause chronic injury  to  epiphytes.

     In the Ohio River Valley,  maximum  annual  averages  of $03 ranged
from about 50 to 80 yg m-3 in 1977 and  1978.   Maximum 1-hr averages
ranged from 300 to 500 yg m-3 (Mueller  et  al.  1980).  At the same
sites (Rockport and Duncan Falls), mean rainfall  pH  for August 1978 to
March 1980 were 4.12 and 4.36,  with ranges of  3.17 to 5.64 and 3.24 to
6.03, respectively [digital  (9  track  tape)  or  hard copy (printout)
versions of these data are available  upon  request directly from Peter K.
Mueller at EPRI].  Recent experimental  evidence  shows that photosyn-
thesis was reduced by 40 percent in the lichen Cladina  stellaris by
field fumigations with fluctuating S02  concentrations of less than 655
vg m-3 (0.25 ppm; Moser et al.  1980).   Laboratory exposures of the
same lichen species wetted by artifical precipitation having a pH = 4.0
and a sulfate concentration = 10.00 mg  £-1 reduced photosynthesis by
27 percent (Lechowicz 1981). From these and succeeding data, it appears
that at least some of the mechanisms  of injury for S02  and acid
precipitation are similar and that existing, long-term  low levels of the
pollutants are influencing lichen  distribution on a  regional scale.

     The effect of direct acidic deposition on lichens  is a new area of
research and therefore has produced few published results other than
those of Lechowicz (1981).  Evidence  from  previous laboratory studies of
the effects of pH on lichens is indirect and based generally on aqueous
solutions of sulfur compounds.   Puckett et al. (1973, 1974) found that
low pH enhanced aqueous sulfur  dioxide  toxicity  in buffered solutions
even when the exposure times were  brief.  D. J.  Hill  (1971) found that
sulfite in buffered solutions was  toxic at pH  4.0 and below but not
toxic at pH 5.0 and above.  Turk and  Wirth (1975) found that damage to
lichens exposed to sulfur dioxide  and subsequently submersed in buffer
solutions from pH 8.0 to pH 2.0 increased  with increasing acidity.
Baddeley et al. (1971) studied  the effect  of pH  in buffered solutions on
the respiration of several lichen  species  found  in eastern North
America.  Exposure times were short,  about 15  minutes,  but respiration
was clearly pH-dependent, and there were definite pH optima for each
species, mostly acidic (pH 4.0).  Repeated exposures might show
different patterns of respiration.

     Little is known about the effects  of  acidic deposition on nitrogen
fixation by lichens.  Denison et al.  (1977) reported a  trend toward
decreased nitrogen fixation in  the lichens Lobaria pulmonaria and L_^
pregana as a function of decreasing pH  of  the  water  in  which the lichens
were soaked.  These results must be considered preliminary, and
additional work in this area is needed  because lichens  can be important
contributors of fixed nitrogen  in  forest ecosystems  (Forman 1975, Pike
1978, Becker 1980, Rhoades 1981),  in tundra and  grasslands  (Alexander
1974), and in deserts (Shields  et al.  1957, Rychert  and Skujins 1974).

     Evidence from the few existing field  studies of acid precipitation
effects on lichens (Robitaille et al.  1977, Plummer  1980) is
                                  3-16

-------
inconclusive because separating pH  effects from  potential  ambient  SO?
(or other gaseous pollutant)  toxicity is  impossible.   Few  of  the studies
that suggest a pH response In lichens (Brodo 1974)  actually include  the
measurement of pH of the aqueous solutions in which the  lichens are
bathed.  Several field studies suggest that acidification  of  lichen
substrates may prevent establishment and  development  of  lichen
propagules (Barkman 1958, Skye 1968, Gilbert 1970,  Grodzinska 1979).
Other studies (Abrahamsen et  al. 1979, Dahl  et al.  1979) show that
lichens alter the chemistry of "rainwater" flowing  over  granite surfaces
partly covered with lichens.   Pyatt (1970) notes that lichens are
capable, to some extent, of exerting a modifying influence upon the
environment.  According to Gilbert, the pH and buffer capacity of  the
lichen thallus and substrate  are Important for the  survival and
regeneration of lichens 1n polluted areas because pH  and buffer capacity
control the distribution and  proportions  of toxic compounds in solution
and the rates of breakdown of these compounds.  Under conditions of  acid
precipitation and reduced buffer capacity, heavy metal absorption  by
lichens is increased (Rao et  al. 1977).

3.2.3  Summary (D. S. Shriner and L. L. Sigal)

     Leaf structure may play  two roles in the sensitivity  of  foliar
tissues to acidic precipitation:  1) leaf morphology  may selectively
enhance (broad-leaved species) or minimize  (needle or laminar-leaved
species) the surface retention of incident precipitation;  and 2)
specific cells of the epidermal  surface,  by virtue  of a  more  permeable
cuticle or the absence of waxes, may be initial  sites of foliar injury.
Once such a lesion occurs, further  development of local  lesions appears
to be enhanced by water collected in the  depression formed by the
necrotic tissue.

     Information on the effects of  acidic deposition  on  the accelerated
weathering of epicuticular wax of plants  is very preliminary  and at
present must be considered no more  than a "testable hypothesis."   Should
further research support the  hypothesis,  virtually  all of  the important
functions of the wax layer could be subject to alteration  due to acidic
deposition.

     Chlorophyll degradation  may occur following prolonged exposure  to
acidic precipitation.  Conclusive linkage to decreased photosynthetic
rates 1s currently missing, but premature senescence  resulting from
chlorophyll degradation may reduce  overall photosynthetic  capacity of
plants affected in this manner.   Further  study is needed before
photosynthetic rate, chlorophyll content, and premature  senescence can
be causally linked to acidic  deposition exposure.  Because simulated
acid precipitation experiments have been  conducted  at extreme ranges,
more attention must be paid to pH values  commonly observed in nature.

     Acid deposition is frequently  partially neutralized by cation
exchange and other reactions  on leaf surfaces.  These reactions reduce
the direct inputs of H+ to soils, but they do not prevent  cation
losses from the ecosystem.  If the  anion  associated with acidic
                                  3-17

-------
deposition is mobile,  cation  losses will occur whether H+ is exchanged
in the canopy or soils.

     Information on which  to  assess the effects of acidic deposition on
lichens is inadequate.   Studies  should investigate the direct effects of
H+ concentration and the other acidic deposition components (S, N) on
lichens.  A comparison of  process-level physiological mechanisms of
response to acidic deposition is necessary, followed by an analysis of
the resulting effects,  if  any, on the overall growth, yield, or
ecosystem function of  lichens.   In addition, the relevance of laboratory
studies to field observations must be established.  Given the
sensitivity of lichens to  related stress agents, they are probably
sensitive to acidic deposition.  In certain ecosystems (e.g., boreal
forests) lichens are a major  system component, and potential effects
should be regarded as  a serious  concern for long-term ecosystem
stability.

3.3  INTERACTIVE EFFECTS OF ACIDIC DEPOSITION WITH OTHER ENVIRONMENTAL
     FACTORS ON PLANTS

     Several important,  but often overlooked, indirect effects of acidic
deposition are potential interactions with other pollutants, alterations
of host-insect interactions,  host-parasite interactions, and symbiotic
associations (Figure 3-2).  These relationships could involve a direct
influence of acidic deposition on a host plant; a direct influence of
acidic deposition on an insect,  microbial pathogen, or microbial
symbiont; or a direct  influence  of acidic deposition on the interactive
process of plant and agent, i.e., infestation, disease, or symbiosis
(Figure 3-2).

3.3.1  Interactions with Other Pollutants (J. M. Skelly and B. I. Chevone)

     The available literature concerning interactive effects of acidic
precipitation and gaseous  air pollutants on terrestrial vegetatation
consists of only three separate  studies.  Shriner (1978b) examined the
intreaction of acidic  precipitation and sulfur dioxide or ozone on red
kidney bean (Phaseolus vulgaris) under greenhouse conditions.
Treatments with simulated  rain at pH 4.0 and multiple 03 exposures
resulted in a significant  reduction in foliage dry weight.  Simulated
precipitation and sulfur dioxide in combination did not affect
photosynthesis or biomass  production.  Troiano et al. (1981) exposed two
cultivars of soybean to ambient  photochemical oxidant and simulated rain
at pH 4.0, 3.4, and 2.8 in a  field chamber system.  The interactive
effects of oxidant and acidic precipitation were inconclusive, with seed
germination greater in plants grown in the absence of oxidant at each
acidity level.  Irving and Miller (1981) also examined the response of
field-grown soybeans to simulated acidic rain at pH 5.3 and 3.1 in
combination with sulfur dioxide  and ambient ozone concentrations.  No
interactive effects on soybean yield occurred from acid treatments with
sulfur dioxide.  Sulfur dioxide  alone, however, resulted in substantial
yield reductions.
                                  3-18

-------
                           ACID  DEPOSITION
INSECT
FOLIAGE
FEEDER
                                                              MICROBE
                                                              FOLIAR
                                                              PATHOGEN
                                             MICROBE
                                             STEM PATHOGEN
               INSECT
               BARK BEETLE
                                                             MICROBE
                                                             ROOT  PATHOGEN
                                                             SYMBIONT
INSECT
SOIL ARTHROPOD
 Figure 3-2.   Acid deposition  may  influence  insects,  pathogens, and
              symbionts  associated with  forest  trees  by direct influence
              (solid arrows) or indirect influence  via host  alteration
              (dashed arrows).   Direct influence  on soil  inhabiting
              insects and microbes is judged less likely  than direct
              influence  on aboveground organisms.   Alterations of soil pH
              or chemistry by  acid deposition may indirectly impact soil
              organisms.
                                   3-19
 t09-262 0-83-5

-------
     With information from only  three studies,  current  assessment  of  the
potential detrimental  Interactive  effects  of gaseous  air  pollutants and
acidic rain on terrestrial  plants  can be considered only  preliminary.
No studies have been conducted with  non-agricultural  vegetation which,
because of potential soil  impacts,  is considered more sensitive to the
indirect effects of acidic precipitation.

     Research efforts at present have addressed the indirect  interaction
of acidic precipitation and gaseous  pollutant stress  to plants.  Plants
have been exposed to pollutants  individually so that  any  interactive
effects are mediated through the plant response, whether  directly  or
indirectly to each pollutant.  With  this exposure  regime,  each pollutant
may predispose the plant to additional  injury and  elicit  a more
sensitive response to the second pollutant.   It is advantageous, under
these conditions, to use experimental systems that are  most sensitive to
both acidic inputs and gaseous pollutant stress.   Due to  crop management
practices, agronomic systems are probably  least sensitive to  Increased
acidic input and alterations in  soil  physiochemical properties.
Additional research in which both  acidic precipitation  and gaseous
pollutants can exert their individual effects on the  various  components
of an ecosystem is required.

     Effects of acidic deposition  on soil  chemistry and nutrient
recycling are unlikely to occur  rapidly (Chapter E-2, Section 2.3).
After more than a decade of research in Scandinavia,  the  observed
changes in forest soil chemical  properties that can be  attributed  to
acidic precipitation still remain  undetermined (Overrein  et al. 1980).
It is, therefore, unlikely that  interactive effects of  acidic deposition
and gaseous pollutants on plants,  which may be expressed  through changes
in soil properties, will become  evident within a single growing season.
Because only annual plants have  been used  in interactive  studies,  the
effect of acidic rain in combination with  other air pollutants stressing
perennial plant species on a yearly  basis  for several years is unknown.
Also, research efforts have not  addressed  the temporal  relationship
between precipitation events and the occurrence of other  gaseous air
pollutants in the ambient atmosphere.

     No information exists on the  interaction of a gaseous air pollutant
with a wet leaf surface.  Such direct interactions can  occur  only  with
the same frequency as precipitation events, but liquid  phase  reactions,
especially with S02, can alter the chemical  form of the pollutant
species.  Sulfur dioxide in water  can exist as the hydrated sulfur
dioxide molecule, the bisulfite  ion, or the sulfite ion,  depending upon
the pH of the solution  (Gravenhorst et al. 1978).  At pH  greater than
3.5, hydrated sulfur dioxide dissociates almost completely into hydrogen
ions and bisulfate ions.  Increased solubility of  sulfur  dioxide can
occur  if the bisulfite  ion is oxidized irreversibly to  the sulfate ion.
This oxidation process can be catalyzed by metal cations, specifically
iron (Fuzzi 1978) and manganese (Penkett et al. 1979).  Particulate
deposits on the leaf surface, containing either iron  or manganese, may
act as sources of these catalysts.  Depending upon the  rate of this
oxidation and the mecham'sm(s)  involved, increased dissolution of
                                  3-20

-------
gaseous sulfur dioxide will occur 1n leaf surface water,  generating
additional hydrogen ions.  Whether such reactions do occur at the  leaf
surface, the extent to which they occur, and their importance in
pollutant stress to plants are unknown.

3.3.2  Interactions with Phytophagus Insects (W.  H.  Smith)

     The damaging Influence of high population densities  of certain
insects can be very visible and cause widespread forest destruction;
however, substantial evidence supports the hypothesis that forest
insects, even those that cause massive destruction in the short run, may
play essential and beneficial roles in forest ecosystems  in a long-term
context. These roles may involve regulating tree species  competition,
species composition and succession, primary production, and nutrient
cycling (Huffaker 1974, Mattson and Addy 1975).  As  a result, assessing
Interrelationships between acidic deposition and phytophagous Insects is
Important.

     Air pollutants may directly affect insects by influencing growth
rates, mutation rates, dispersal, fecundity, mate finding, host finding,
and mortality.  Indirect effects may occur through changes in host age
structure, distribution, vigor, and acceptance.  Few researchers have
Investigated the effects of acidic deposition on Insects.   Some studies
relative to acidity effects on aquatic insects are available (e.g.,
Borstrum and Hendrey 1976).  Terrestrial arthropods, on the other  hand,
have been the subject of very few studies.  Hagvar et al.  (1976) have
concluded that acidic precipitation from western and central  Europe
increases the susceptibility of Scots pine forests to the  pine bud moth
(Exotelela dodecella).

     Various studies have presented data indicating  that  species
composition or population densities of Insect groups are  altered In
areas of high air pollution stress, for example,  roadside  (Przybylski
1979) or Industrial (S1erp1nsk1 1967, Novaskova 1969, Lebrun 1976)
environments.  Further specific information 1s available  on the general
influence of polluted atmospheres on population characteristics of
forest insects (Tempiin 1962; Schnaider and Sierpinski  1967;  Sierplnski
1970, 1971, 1972a,b; Boullard 1973; Wiackowski and Dochinger 1973; Hay
1975; Charles and Villemant 1977; Sierpinski and Chlodny  1977;  Dahlsten
and Rowney 1980).   Johnson (1950, 1969) has reviewed much  of the
literature dealing with air pollutants and insect pests of conifers.
One of the most comprehensive literature reviews  available concerning
forest insects and air contaminants has been presented by  Villemant
(1979).  Recently, Alstad et al. (1982) provided  an  excellent overview
of the effects of air pollutants on Insect populations.

3.3.3  Interactions with Pathogens (W. H.  Smith)

     Abnormal physiology, or disease, in woody plants follows infection
and subsequent development of an extremely large  number and diverse
group of microorganisms within or on the surface  of  tree  parts.  All
stages of tree life cycles and all  tree tissues and  organs are subject,
                                  3-21

-------
under appropriate environmental  conditions,  to impact by a heterogeneous
group of microblal pathogens including vlroids,  viruses, mycoplasmas,
bacteria, fungi, and nematodes.   As with insect interactions,  microbes
and the diseases they cause play important roles in  succession,  species
composition, density, composition,  and productivity.  In the short term,
the effects of microbial  pathogens  may conflict with forest management
objectives and assume a considerable economic or managerial  as well  as
ecologlc significance (Smith 1970).

     The interaction between air pollutants  and microorganisms in
general is highly variable and complex.  Babich and  Stotzky (1974)  have
provided a comprehensive overview of the relationships between air
contaminants and microorganisms. A specific air pollutant,  at a given
dose, may be stimulatory, neutral,  or inimical  to the growth and
development of a particular virus,  bacterium, or fungus.  In fungi,
fruiting body formation,  spore production, and spore germination may be
stimulated or inhibited.

     Microorganisms that normally develop in plant surface habitats  may
be especially subject to air pollutant influence. These microbes have
received considerable research attention and have been the subject of
review (Saunders 1971, 1973, 1975;  Smith 1976).   Numerous comprehensive
reviews have summarized the interactions between air contaminants and
plant diseases (Laurence 1981).   Heagle (1973)  summarized nearly 100
references and found that sulfur dioxide, ozone, or  fluoride had been
reported to Increase the incidence  of 21 diseases and decrease the
occurrence of nine diseases 1n a variety of nonwoody and woody hosts.
Treshow (1975) has provided a detailed review concerning the influence
of sulfur dioxide, ozone, fluoride, and partlculates on a variety of
plant pathogens and the diseases they cause.  Treshow lamented the fact
that most of the data available deal  with in vitro or laboratory
accounts of microbe-air pollutant interactions,  while only a few
investigations have examined the Influence of air pollutants on  disease
development under field conditions.

     A review provided by Manning (1975) pointed out that most research
attention has been directed to fungal pathogen-air pollutant
interactions.  Greater research  perspective  is needed concerning air
pollution influence on viruses,  bacteria, nematodes, and the diseases
they cause.  Macroscopic  agents  of  disease,  most importantly true- and
dwarf-mistletoes, must also be examined relative to  air pollution
impact, especially in the western part of North  America, where the
latter are extremely important agents of coniferous  disease.

     Forest trees, because of their large size,  extended lifetimes,  and
widespread geographic distribution  are subject to multiple
m1crob1ally-1nduced diseases frequently acting concurrently  or
sequentially.  The reviews of Heagle (1973), Treshow (1975),  and Manning
(1975) considered a variety of pollutant-woody plant pathogen
interactions but were not specifically concerned with forest tree
disease.  In their review of the impact of air pollutants on fungal
pathogens of forest trees of Poland,  Grzywacz and Wazny (1973) cited
                                  3-22

-------
literature indicating that air pollution stimulated the activities  of at
least 12 fungal tree pathogens while restricting the activities  of  at
least 10 others.

     Our understanding of the influence of acidic deposition  on
pathogens and the diseases they cause is meager.  Shriner (1974,  1975,
1977) has provided us with some valuable perspectives in this important
but understudied area.  Falling precipitation and the precipitation
wetting of vegetative surfaces (see Section 3.2.1), play an enormously
important role in the life cycles of many plant pathogens.  Recognizing
this, Shriner (1974, 1975, 1977) has examined the effects of  simulated
rain acidified with sulfuric acid on several  host-parasite systems  under
greenhouse and field conditions.  The simulated precipitation he
employed had a pH of 3.2 and 6.0, approximating the common range of
ambient precipitation pH.

     Applying simulated precipitation of pH 3.2 resulted in (1)  an  86
percent restriction of telia production by Cronartiurn fusiforme  (fungus)
on willow oak, (2) a 66 percent inhibition of Meloidogyne ha^Ta
(root-knot nematode) on kidney bean, (3) a 29 percent decrease in
percentage of leaf area of kidney bean affected by Uromyces phaseoli
(fungus), and (4) both stimulated and inhibited development of halo
blight of kidney bean caused by Pseudomonas phaseolicola (bacterium).
In the latter case, the influence of acidic precipitation varied and
depended on the particular stage of the disease cycle when the exposure
to acidic precipitation occurred.  Simulated  sulfuric acid rain  applied
to plants prior to inoculation stimulated the halo blight disease by  42
percent.  Suspension of inoculum in acidic precipitation decreased
inoculum potential by 100 percent, while acidic precipitation applied to
plants after infection occurred inhibited disease development by 22
percent.

     Examining willow oak and bean leaves with a scanning electron
microscope revealed distinct erosion of the leaf surface by rain of pH
3.2 (see Section 3.2).  This may suggest that altered disease incidence
may be due to some change in the structure or function of the cuticle
(see Section 3.2.1.1).  Shriner has also proposed that the low pH rain
may have increased the physiological  a§e of exposed leaves.   Shriner
(1978a)  concluded his initial  experiments by  suggesting that  he  had not
established threshold pH levels at which significant biological
ramifications to pathogens occur from acidic  precipitation.   He  did
suggest, however, that artificial precipitation of extremely  low pH
probably alters infection and disease development of a variety of
microbial pathogens.

     In  recent years, a very serious disease  of hard pines caused by  a
twig and leaf pathogen called Gremmeniena abietina has increased in
importance in the northeastern United StatelTThe disease, termed
Scleroderris canker, was first reported on red pine in New York  in  1959.
Currently, G. abietina is causing significant large tree mortality  in
Vermont and~New York.  Because it may be more than coincidence that this
region is included within the highest acidic  precipitation zone  of  North


                                  3-23

-------
America, Paul D. Manlon, SUNY,  Syracuse,  Initiated an  acidic  rain
Scleroderrls research project.   The laboratory and field studies
reported to date Indicate the disease may be affected  by precipitation
pH, but there was no indication that abnormally high acidified  rain
Increased disease incidence.   In fact, the opposite may be true.   That
1s, acidic rain may reduce the  Importance of the canker disease (Raynal
et al. 1980, Bragg 1982, Manlon and Bragg 1982).

     Armillarlella me!1ea is  an extremely important forest tree root
pathogen throughout the temperate zone.   The fungus 1s geographically
very wide-spread, has an extremely broad host range, and Is especially
significant in causing disease  In trees  under stress.   Shields  and Hobbs
(1979) have Indicated that soil pH is related to disease development
caused by A. mellea.  If acidic deposition influences  soil  pH (see
Chapter E-7) or tree vigor, it may indirectly impact tree susceptibility
to A_. mellea Infection.   In the northeast, spruce decline 1n  high
elevation forests has been a  recent concern.  A_. mellea Is associated
with spruce trees exhibiting  dieback and decline symptoms in  northern
New England and may play an Important role In the morbidity and
mortality of this species. The habitats of soil pathogens such as A.
roellea are buffered relative  to plant-surface habitats, so for  acidTc
deposition to Influence these pathogens  an alteration  of soil pH or
chemistry or host susceptibility would have to occur.

     Fusiform rust caused by  £. fuslforme Is the most  important disease
of managed pines In the southeast!Bruck et al. (1981) applied
simulated rain of various pH  levels to loblolly pine at the time of
inoculation with rust basidlospores.  Significantly fewer galls formed
on trees treated with simulated rain at pH 4.0 or less than formed on
trees treated with rain at pH 5.6.

     Various bacterial species  are important components of tree leaf
microfloras.  Lacy et al. (1981) observed that populations of Erwinia
herbicol a and Pseudomonas syringae were  reduced on soybean 1 eaves  when
host plants were treated with water acidified to pH 3.4 relative to
leaves exposed to distilled water (pH 5.7).

3.3.4  Influence on Vegetative  Hofjts That Would Alter  Relationships  with
       Insect or Microbial Associate (W. H. Smith)

     As Section 3.2 discussed,  exposure to acidic deposition  may lead  to
acidification of plant surfaces, leaf cuticle erosion, and foliar
lesions.  Foliar lesions could release plant volatHes attractive  or
repulsive to insect pests or  may serve as infection courts for  microbial
disease agents.

     The Influence of acidic  deposition leached chemicals on  Insects
infesting tree leaves or bark could prove attractive,  repulsive, or
provide chemical orientation.  In the case of surface  microbes, leached
compounds may inhibit vegetative growth  or spore germination  (alkaloids,
phenolic  substances) or stimulate vegetative growth  (as nutrients)  or
spore germination (as Inducers or nutrients—sugars,  amino acids,


                                  3-24

-------
 vitamins).  Leaching of toxic radioelements from plant surfaces could
 have  a  restrictive  impact on plant surface biota (Myttenaere et al.
 1980).

      Plant  growth and yield may be stimulated or inhibited by acidic
 deposition.  If growth is either stimulated or suppressed, it is
 probable  that differential influence on insects and pathogens would
 follow.   In the case of some host-pathogen and host-insect
 relationships, a tree under stress is more vulnerable to infestation or
 Infection.  Bark beetles and root-infecting or canker forming fungi are
 generally more successful in less vigorous Individuals.  Trees
 exhibiting vigorous growth, on the other hand, may be predisposed to
 more  serious Impact from certain rust fungi and other disease agents.

 3.3.5   Effects of Acidic Deposition on Pesticides (J. B. Weber)

     Pesticides are used annually to manage pests in terrestrial and
 aquatic systems.  The majority of these materials are organic chemicals
 that  selectively control unwanted and injurious Insects, pathogens,  or
 weeds.  They are applied directly to animals,  vegetation, soils, and/or
 inland waters, but  ultimately they end up in soils and/or waters.   The
 behavior and fate of pesticides in the environment depend upon the
 following:

      (1)  method of application of the pesticide;

      (2)  chemical  properties of the pesticide;

      (3)  edaphic properties of the system;

      (4)  dissipation routes of the pesticide; and

     (5)  climatic  conditions.

     No studies on effects of acidic deposition on pesticides were found
 in the literature; however,  pH  changes have been reported to  affect
 factors 1 through 5 listed above.

     Foliar absorption and injury from herbicides applied directly to
 vegetation have been reported to be greatly enhanced by lowering the pH
 for both phenoxyacetic acid (Crafts 1961b)  and dinitrophenol  (Crafts and
 Reiber 1945) type compounds.  Acidic  conditions promote formation  of the
 un-ion1zed species that more readily  penetrate and injure vegetative
membranes than do Ionized species.   Thus,  acidic deposition could
conceivably result in enhanced  injury to weeds and/or crops In certain
instances.  The most likely  possibility of this occurring would  be in
herbicide applications to forests,  pastures, minimum-till age  crop
production systems,  or aquatic  systems  where the foliage  has  had ample
time to accumulate acidic deposition.

     Chapter E-4  (Section 4.4)  reports  that acidic  deposition causes
significant lowering of the  pH  of  inland waters  in  certain instances.
                                  3-25

-------
This would have a substantial  effect on the direct biological  activity
and longevity of herbicides used 1n  aquatic weed and  algae  control.  One
would expect a significant Increase  1n the herblddal  activity of  the
phenoxyacetlc add compounds.   Aquatic herbicides such as slmazlne would
perform less satisfactorily under acidic conditions since many
Investigators (Armstrong et al.  1967,  Jordan et al. 1972) have reported
that chloro-s-tr1az1nes decompose at a much faster rate under  acidic
conditions. ~T"h1s would make 1t necessary to Increase the rates of
chloro-j;-tr1az1ne herbicides and to  make more frequent applications for
satisfactory aquatic weed control  1n waters where the pH levels were
below normal levels.

     Organic pesticides are categorized Into five major types  depending
on Ionizing characteristics.   Examples of the five types are:

     (1)  catlonlc (dlquat, paraquat);

     (2)  basic (atrazlne, slmazlne, prometryn);

     (3)  acidic (2,4-D, fenac,  plcloram)

     (4)  phosphates and arsenates (glyphosate, DSMA); and

     (5)  nonlonlc (alachlor,  carbaryl, methomyl).

     Changes In pH levels In waters  or soil solutions affect the
Ionizing properties of basic and acidic properties to the greatest
extent.  At lowered pH levels acidic and basic pesticides tend to  be
more readily adsorbed by soil  parti oil ate matter, hence less
biologically active and less mobile  (Weber 1972, Weber and  Weed 1974).
Under such circumstances, higher rates of these pesticides  would be
required to provide satisfactory performance, and the longevity of the
chemicals may be affected, depending on their modes of decomposition.

     Pesticides degraded biologically would be affected by  changes In
mlcroblal populations.  Captan, dlcamba, amltrole, vernolate,
chloramben, crotoxyphos (Hamaker 1972), metrlbuzln (Ladlle  et  al 1976),
2,4-D and MCPA (Torstensson 1975), and prometryn (Best and  Weber 1974)
were reported to persist longer under acidic conditions than under
neutral conditions.  Conversely, dlazlnon and dlazoxon (Hamaker 1972)
were degraded more readily at lower pH levels.

     Pesticides degraded chemically  are directly affected by soil  pH
levels.  Malathlon and parathlon (Edwards 1972) persisted much longer In
acidic soils than 1n neutral soils,  while atrazlne (Best and Weber 1974)
and slmazlne were degraded much more rapidly under acidic conditions
than under neutral conditions.

3.3.6  Summary (W. H. Smith)

     A review of the evidence on the Interaction of acidic  deposition
with other pollutants, Insect and mlcroblal pests, does not allow


                                  3-26

-------
 generalized statements concerning stimulation or restriction of blotlc
 stress agents, or their activities, by acidic deposition.  Certain
 studies report stimulation of pest activities associated with acidic
 deposition treatment, while other studies report restriction of pest
 activities following treatment.  No studies report significant
 Interactive effects between acidic deposition and other pollutants,
 although the potential for such effects 1s very real.

     Future research must combine field and controlled environment
 studies.  Mechanisms for addle deposition Impact on predisposition/
 protection of forest trees to/from disease caused by mlcroblal
 pathogens, and Infestation caused by phytophagous Insects must be
 examined.  Evidence available comes from laboratory and controlled
 environment studies, but no evidence on this topic from studies
 employing large trees under field conditions exists.

     We cannot, however, rule out the possibility of Indirect,  subtle
 interaction of acidic deposition with other pollutants, phytophagous
 insects, and mlcroblal pathogens.

     No known studies demonstrate that acidic deposition on plant
 surfaces directly affects the biological  activity of pesticides.
 However, ample evidence shows that pH of aqueous solutions of certain
 herbicides greatly affects herbicidal activity, and observed effects
 were greatest between pH levels of 6.0 and 3.0.  These occurrences have
 been reported for herbicides applied to terrestial  and aquatic  weeds.

     No studies show indirect effects of acidic deposition on pesticide
 inactivation, mobility, and decomposition in soils; however, ample
 evidence shows that soil  pH greatly affects all of these processes.  It
 is likely that if acidic deposition is found to affect soil  and water
 pH, then pesticide behavior and fate will  likewise be affected.

 3.4  BIOMASS PRODUCTION

 3.4.1  Forests (S. B. Mclaughlin, D. J. Raynal, and A. H.  Johnson)

     Changing levels and patterns of emissions  of atmospheric pollutants
 in recent decades have resulted in Increased exposure of extensive
 forests in Europe and North America to both gaseous pollutants  and acid
 precipitation.   Reports of decreased growth and increased mortality of
 forest trees in areas receiving high rates of atmospheric  pollutant
 deposition have stressed the need to quantify the rates of changes in
 forest productivity and identify the causes of  such changes.  The
 complex chemical  nature of combined pollutant exposures and the fact
 that these pollutants may have both direct effects  to vegetation and
 indirect (possibly beneficial)  effects makes quantification of  such
effects particularly challenging.   The complexity of forest growth and
 succession and the sensitivity of forest  trees  to natural  environmental
 stresses add further to the challenge of  quantifying effects of
 anthropogenic pollutants on forest productivity.
                                  3-27

-------
     Such quantification requires that several  critical  tasks be
addressed:   (1)  definition  of the chemical  nature of the present and
past air quality within the regions of principal concern,  (2)
documentation of the basis  for assuming that  detectable  effects may be
occurring within those regions, and (3) identification of  the types of
effects that might be produced under present  and likely  future exposure
regimes.

     A critical  need in evaluating stress effects on perennial forest
systems is documenting the  magnitude, rate, and point of inception of
historical changes in air quality.  Unfortunately,  the maximum period of
record for the present National Atmospheric Deposition Program (NADP)
network is four years, while ozone monitoring data  have  not  been
collected by standardized methods in network  fashion beyond  1975.  The
most recently published estimates of historical changes  in isopleths of
precipitation acidity (Likens and Butler 1981)  suggest that  initial
intensification of acidity of northeastern precipitation may have  begun
in the 1950's.  However, because of the limited data points  and the
uncertain chemical techniques used, the validity  of these  earliest data
has been questioned (see Chapter A-8).  Other sources of information
currently being developed include emissions inventories  coupled with
regional air dispersion modeling, evaluation  of historical stream  and
lake chemistry data, historical reconstruction of weathering rates of
marble monuments, and analysis of changes in  elemental composition of
annually-formed lake sediments and tree rings. Collectively,  these
techniques offer possibilities for documenting the  period  of
intensification of atmospheric deposition of anthropogenic pollutants.
(Further discussion of such documentation can be  found  in  Chapter  A-8).

3.4.1.1  Possible Mechanisms of Response—A wide  variety of  potential
direct and  indirect responses of forest trees to  acid  deposition  have
been hypothesized based on fundamental responses  of biological  systems
to acidity  and other stresses  (Tamm and Cowling 1976).  Included  among
these are increased leaching of nutrients from foliage,  accelerated
weathering  of leaf cuticular surfaces, increased  permeability  of leaf
surfaces to toxic materials, water, and disease agents,  altered
reproductive processes, and altered  root-rhizosphere relations.   In
addition to the direct effects of acidity from contact with  foliage,
roots,  and  rhizosphere organisms, a major area of interest is  the
indirect effects of increased  acidity on soil nutrient availability  to
vegetation  and the consequences of  soil leaching  losses to aquatic
systems  (SMA 1982).  Many of the key processes to be considered in
evaluating  the effects of acidic deposition on forest systems  are
identified  schematically in Figure 3-3. The diversity  of these processes
illustrates the complexity of  potential interactions of acidic
deposition  with forest systems and the need for better understanding of
system  level  integration of potential  effects  on multiple processes.

     Forest responses must be  examined both from the perspective of
today's  mature forests which have been produced over the last 50 to 100
years  (a  period of  significant changes in  atmospheric emissions)  as well
as with  respect to the  forests of the  future,  which by contrast are


                                  3-28

-------
                                  WET
                               DEPOSITION
       ACID DEPOSITION

       -ACID RAIN  (WET)
       -POLUTANT GASES (DRY)

         	i	
         1DIRECT EFFECTSI
              GROWTH
              VIGOR
           REPRODUCTION
I
                               THROUGHFALL
                                    I
                                    I
         {INDIRECT EFFECTSI
       NUTRIENT AVAILABILITY
           TOXIC  EFFECTS
         	31	
         MICRQBIAL PROCESSES
           NUTRIFICATION
          DENITRIFICATION
           IMMOBILIZATION
              RELEASE
            MYCORRHIZAE
         I  FOREST   |
         PRODUCTIVITY
Figure 3-3.   Key components and processes to be considered in evaluating
             effects of acidic deposition on forested ecosystems.

                                   3-29

-------
growing under atmospheric stresses that will  likely span their entire
life cycle.  Thus, productivity  of these  forests may be more influenced
by alteration of the potentially more  sensitive life stages including
reproduction, seedling establishment,  and growth.

     Seedling emergence,  establishment, and early growth phases are
considered to be potentially  among the most susceptible stages affected
(Abrahamsen et al. 1976,  Likens  1976,  Lee and Weber 1979, Raynal et al.
1980).  Additionally, reproductive phases of  growth may be the most
sensitive to acidic deposition (Likens 1976,  Cowling, 1978, Jacobson
1980.  Various controlled field  and laboratory studies in Scandinavia
and in the United States  have been conducted  to quantify possible
effects of simulated acid rain on  seed germination, seedling
establishment, and growth of  trees in  field plots.

3.4.1.2  Phenological Effects—Plants  may respond to the deposition of
acidic substances in a manner which depends on genetic characteristics
of the species;  sensitivity of individual developmental stages; timing,
duration, frequency, and  severity  of deposition events; and nature of
meteorological and microenvironmental  conditions (Cowling 1978).  Thus,
a complete assessment of  the  influences of acidic deposition on plants
must include consideration of phenology--changes in life cycle stages as
affected by environment and season. Seed germination and seedling
emergence and establishment are  early  growth  phases potentially
susceptible to acidic deposition (Abrahamsen  et al. 1976; Lee and Weber
1979; Raynal et al. 1982a,b). As  well, mature and reproductive phases
of growth may be sensitive to acidic deposition (Likens 1976, Cowling
1978, Jacobson 1980, Evans 1982).   However, differences in the
sensitivity of vegetation to  acidic deposition are not documented from
natural field studies.

     Plant growth, development,  and reproduction may be affected by
acidic deposition both positively  and  negatively.  Response depends upon
species sensitivity, plant life  cycle  phase,  and the nature of exposure
acidity.  Considerable variation in plant species susceptibilty exists,
and at the individual level the  effect of acidification on different
plant organs or tissues,  including marketable crops, may vary widely.
Controlled environment studies indicate that  the deposition of acidic
and acidifying substances from the atmosphere may have stimulatory,
detrimental, or no apparent effects on plant  growth, development, and
reproduction.  Both stimulatory  and detrimental effects may
simultaneously occur, making  determination of both acute and chronic
effects quite difficult.   For example,  tree seedling growth may be
enhanced by deposition of nitrate  and  possible sulfate when soils are
deficient in these while  concomitantly foliar injury may occur due to
hydrogen ion deposition.   Because  many biotic and abiotic factors
interact to influence plant performance,  plant dieback or reduction in
growth or yield must be evaluated  in terms of physiological stress, soil
toxicity and nutrient deficiency problems, plant disease, and direct and
indirect effects of acidic precipitation, if  chronic effects of
deposition of acidic substances  are to be fully characterized.
                                  3-30

-------
3.4.1.2.1  Seed germination and seedling establishment.  Laboratory
studies indicate that a wide range of sensitivity of seed germination to
acidic substrate conditions exists (Abrahamsen et al. 1976, Lee and
Weber 1979, Raynal et al. 1982a).  Studies focused on woody plants
reveal that seed germination of some species, including yellow birch and
red maple, is inhibited, but other species, such as sugar maple, are not
affected when exposed to substrate acidity of pH 3.0 or less (Raynal et
al. 1982a).  In some coniferous species such as white pine and white
spruce, substrate acidity of pH 3.0 may promote seed germination, but it
produces no effect in other species such as eastern hemlock.  Figure 3-4
illustrates the contrasting response of seed germination of three tree
species to different substrate acidity levels.

     Interaction of substrate solution reaction (pH) and osmotic
potential may be significant, and the effect of acidity may vary due to
differences in ionic characteristics of the germination medium (Chou and
Young 1974, Abougendia and Redman 1979).  Leaching of various substances
from the seed or fruit coat by acidic solutions may also occur,
subsequently causing neutralization.  The necessity of continually
adjusting the pH of in vitro solutions to maintain constant acidity
levels in germination studies suggests that seed tissues may effectively
buffer the germination medium, thus reducing potential  detrimental
effects of acidic deposition {Raynal et al. 1982a).  Under natural  field
conditions, vegetation canopy, litter, organic matter,  and mineral  soils
may further buffer emerging seedlings from highly acidic deposition
(Raynal et al. 1982b, Monitor and Raynal 1982).  Thus, seeds are often
protected from direct influence by acidic deposition and seed
germination typically may be minimally affected, as indicated by much of
the research to date.

     Emergence and establishment of the seedling have been shown to be
more sensitive to low substrate pH than is seed germination itself
(Abrahamsen et al. 1976, Lee and Weber 1979, Raynal  et al. 1982b).
Certain species, such as sugar maple, show no detrimental  effect of
acidity on seed germination at pH 3.0 but may be inhibited at the
establishment phase, as shown in studies of effects of  simulated acidic
precipitation on juvenile growth {Raynal et al. 1982a,b).   Injury to the
emerging seedling radicle and hypocotyl  may be direct,  due to hydrogen
ion concentration, and/or indirect,  resulting from increased
susceptibility to microbial  pathogens that tolerate acidic conditions
and changing nutrient levels (Raynal et al. 1982b).   Seedling growth
studies in which young plants are exposed to simulated  acidic
precipitation have shown that juvenile plants may exhibit reduced or
stimulated growth, depending on the species (Wood and Bormann 1974,
Raynal et al.  1982b).

     Possible changes in soil nutrient status associated with acidic
deposition must be considered in evaluating plant growth response to
acidification (see Section 2.3).  Some workers (Benzian 1965, Abrahamsen
et al. 1976,  Abrahamsen 1980) have demonstrated that optimal  height
growth of coniferous seedlings (including species of pine, spruce,  and
fir) occurs in soils having  a pH between 4.0 and 5.0.   Whether hydrogen
                                  3-31

-------



g
GERHINAT]
•j*



100
90
80
70
60
50
40
30
20
10
0
• * *
(a) ,fi|ft
:
*
.* SUGAR MAPLE
B
a*
I A pH 5.6
| o pH 4.0
I • pH 3.0
••' 1 1 1 1 1
                            10     20     30    40

                           DAYS SINCE START OF GERMINATION
                                                       50
3
M
70


60


50


40


30


20


10
                                                                WHITE PINE
     J	I
                              pH 5.6
                              pH 4.0

                          ,..., pH 3.0


                              J_
                                I
     2   4   6   8   10   12   14  16

      DAYS SINCE START OF GERMINATION
                                                          5      10     15     20

                                                        DAYS SINCE START OF GERMINATION
 Figure  3-4.   Mean cumulative percent  germination of sugar maple, yellow
               birch, and white pine  seeds  subjected to different substrate
               acidity levels.  Arrows  indicate point at which differences
               in response become  significant (p < 0.05) determined  by
               Tukey's test for mean  separation following analysis of
               variance.  Data show contrasting responses of species to
               increasing acidiy:   (a)  no significant difference at  pH 3.0,
               4.0, and 5.6 for sugar maple, (b) decreased germination in
               yellow birch at pH  3.0 compared with that at pH 4.0 and 5.6
               (no significant difference between pH 4.0 and 5.6), and
               (c) increased  germination in white pine at pH 2.4 and 3.0
               compared with  that  at  pH 4.0 and 5.6 (no significant  differ-
               ence between 2.4 and 3.0 or 4.0 and 5.6).  Adapted  from
               Raynal et al.  (1982a).

                                    3-32

-------
ion deposition directly influences seedling growth or whether it,  in
association with the deposition of other cations and anions, causes
variation in soil nutrient characteristics affecting growth is not fully
known (Abrahamsen 1980).  When nutrients in soil are not limiting  for
plant growth, detrimental effects of deposition of acidifying substances
are not likely to occur.  However, at low fertility levels, simulated
acidified canopy throughfall of pH 3.0 or less has been found to promote
seedling growth in some species (Raynal  et al. 1982b).   Such a benefical
response could result from deposition of nitrate or other nutrients.
(See Chapter E-2 for detailed discussions of forest nutrient effects.)

     Even where growth is stimulated by acidic deposition, however,
foliar injury may simultaneously occur in some species (Raynal et  al.
1980, 1982b).  Thus, competitive promotive and inhibitory effects  of
acidic deposition may concomitantly affect seedling growth and
development.  It is, therefore, not surprising that studies of the
effects of simulated acidic precipitation or forest canopy throughfall
on plant growth have produced variable results, ranging from no apparent
effects, stimulation of growth, and inhibition of growth (Wood and
Bormann 1974, Matiziris and Nakos 1977,  Raynal et al.  1980).

3.4.1.2.2  Mature and reproductive stages.  Studies of interference of
acidic deposition on flower or cone development in flowering plants and
conifers have not been made.  Should highly acidic precipitation events
coincide with floral or gamete development, pollination, or fruit  or
seed set, effects on plant populations and regeneration processes  could
possibly be altered.  Numerous studies reveal  that various air
pollutants, including sulfur dioxide and ozone, may cause reductions in
cone size and weight (Smith 1981).  Studies of air pollutant effects on
pollen germination and pollen tube-elongation  suggest  that pollen
function may be altered because of acidification of floral  tissues,
including stigmas (Karnosky and Stairs 1974).   Findings that red and
white pine pollen germination and tube elongation were  greater in  a
relatively unpolluted site compared with one of high pollution incidence
provide circumstantial evidence that pollen gametogenesis and
development may be altered by acidic deposition (Houston and Dochinger
1977).  Evaluating acidic precipitation  effects on plant reproduction
demands that the coupling of effects of  air pollution and acidification
be understood.

     Controlled-environment studies of effects of simulated acidic
precipitation on agricultural  crops indicate differential  sensitivity of
species and contrasting effects on different plant parts (Evans and
Lewin 1981, Lee et al. 1981; see also Section  3.4.2).   Research by Evans
and Lewin (1981) indicates that yield of pinto beans may be reduced by
simulated acid  rain because of a decrease in the number of seeds per
pod.  In contrast, under similar exposure conditions,  soybean yield
showed an increase due to a larger dry mass per seed.

3.4.1.3  Growth of Seedlings and Trees in Irrigation Experiments--
Abrahamsen (1982)  has reviewed field experiments in Sweden  and Norway
designed to determine the effects of artificial  acidification on growth


                                  3-33

-------
of forest trees and tree seedlings.   In Swedish experiments (Tamrn and
Wlklander 1980), young (18-yr-old) Scots pine were Irrigated below the
canopy with dilute sulfurlc  add  (annual application, 50 to 150 kg
ha"1 HeSOd In one application per year) 1n both with and without
prior addition of fertilizer.  After  6 years of application a negative
correlation between treatment acidity and basal area growth was found on
the fertilized plots (<_ 10 percent decrease at highst acidity) whereas
growth responded positively  (approximately + 30 percent increase at
highest acidity level)  on the unfertilized plots.  Increased nitrogen
uptake was considered a probable  cause of positive responses.  Results
of these studies were complicated by  changes 1n nutrient availability in
the soil and associated with the  effects of high acidity on soil fungi,
bacteria, and competing understory vegetation (Tamm and Wlklander 1980).

     In Norwegian experiments (Abrahamsen et al. 1976, Tveite and
Abrahamsen 1980) a variety of combinations of acidified groundwater
treatment (pH values between 6.0  and  2.0 by ^$04 addition),
treatment volume (25 to 50 mm per month) application technique (below or
above canopy), lime application 500 to 4500 kg CaO ha'1), and tree
species (lodgepole pine, Norway spruce, silver birch, and Scots pine)
were used.  The overall effects of these experiments were small where
treatment effects were found after 4  to 7 years of treatment application
(Tveite 1980a).  In studies  with  Scots pine, positive growth effects
were found at pH levels of 3.0, 2.5,  and 2.0 after 4 years of treatment,
followed by significant growth reduction by pH 2.0 1n the 5th year.
Norway spruce showed reduced diameter growth at all acid treatment
levels on the year after 6 years  of prior treatment.  Height growth of
silver birch was stimulated  by rainfall acidity.  Lime application had
little or no effect on observed responses.  Effects of add irrigation
on foliar nutrient levels were also generally small (Tveite 1980b).

     In evaluating the results of the Scandinavian Irrigation
experiments Abrahamsen (1980) concluded that the data give "no
substantial evidence of effects on tree growth at acidity levels
presently found In precipitation."  However, he cautions that add
effects produced particularly at  highest acidity levels may be partly
attributable to soil effects that were artifacts of the highly acid
treatment levels and hence not representative of longer-term responses
to be expected under actual  field conditions.

     Such results corroborate findings of researchers in the United
States who have demonstrated differential effects of simulated acidic
precipitation on plant growth (Wood and Bormann 1977; Raynal et al.
1980, 1982b).  Conclusions regarding  plant growth response from
experiments where vegetation and  soils have been subjected to
accelerated acidic deposition rates or concentrated acidic Inputs must
be viewed with caution, however,  for  reasons discussed 1n Chapter E-2,
Section 2.3.1.

3.4.1.4  Studies of Long-Term Growth  of Forest Trees—The evidence for
effects of regional-scale anthropogenic pollutants on productivity of
forests comes from a limited number of studies in the united States and
                                  3-34

-------
Europe in which long-term growth trends determined from tree rings  have
been analyzed.  In Scandinavia,  where acid precipitation was first
recognized and studied as an environmental problem,  research on changing
patterns of tree growth based on tree-ring chronologies have provided
circumstantial evidence of growth declines that occurred at about the
time acidity of rainfall  is thought to have intensified.  In Norway,
research by Abrahamsen et al. (1976)  and Strand (1980)  showed in Norway
spruce and Scots pine a decrease in growth (generally less than 2.3
percent per year) that became evident around 1950, primarily in the
eastern third of the country.  These responses could not be clearly
associated with the geographical patterns of most acid  rainfall, which
occurred in the southern (pH average = 4.3) rather than the eastern (pH
average = 4.5) part of the country.  Some drawbacks of  these studies,
however, were that individual sites were not characterized with respect
to soil chemical characteristics, and neither the Influences of climate
nor aging trends were removed from the data.

     Preliminary analysis of differences in responses between sites of
differing productivity class (high vs low) In southern  Norway showed no
differences 1n response to acidic precipitation (Abrahamsen et al.
1976).  On the other hand, studies in Sweden by Jonsson (1975)  and
Jonsson and Sundberg (1972) Involving Scots pine and Norway spruce
showed similar temporal trends in growth reduction beginning around
1950, and these effects were most pronounced in areas of greatest
expected susceptibility to acidic deposition.  Site susceptibility  was
estimated based on the average pH of precipitation and  pH and ion
content of lakes and rivers In 1965 and 1970 and the distribution of
soil types.  Jonsson (1975) concluded from these studies that
"acidification cannot be excluded as a possible cause of poorer growth
development, but may be suspected to have had an unfavorable effect on
growth within the more susceptible regions."  Differences in growth
reductions between susceptible and non-susceptible regions were
estimated to be in the range of 0.3 to 0.6 percent per  year.

     Since this original  study,  a second study has been initiated
covering an additional 9 years,  1965-1974, since the first survey was
completed (Jonsson and Svensson 1983).  These data confirmed the earlier
downward trend beginning In 1950 but showed a period of Increased
productivity beginning in the mid- to late 1960's.  For sites of
relatively poor quality,  growth of both pine and spruce in the 1970's
had increased substantially since Its minimum in the mld-60's but was
still substantially less than that attained up to 1940.  The overall
trend was still downward over the Interval 1910-1974.  By contrast,
growth of these species on good sites showed an upswing in the 1965-74
interval which resulted in a growth rate equal to or above that attained
during the preceding 50 years.  In explaining these trends and
summarizing the results of their own and the Norwegian  SNSF project the
Swedes make the following statements (SMA 1982).

     "A conceivable explanation  of these changes is  that the
     mathematical model that was used has not compensated for or
     caught those effects in the ground that are the results of more


                                  3-35

-------
     long-term cyclical  changes  in climate.  These changes may, for
     example,  affect the supply  of nitrogen in the ground that is
     available to plants.   It must also be noted that the Swedish
     forests have to take increased quantities of nitrogen that are
     deposited along with precipitation.  This gives a fertilizing
     effect.  There are  at the present time no clear signs or
     evidence of either  increased or reduced forest production
     resulting from the  effects  of acid precipitation on
     Scandinavian forest!and and its fertility."

     The final report on the Norwegian SNSF project makes the point
that:

     "decreases in forest growth due to acid deposits have not been
     demonstrated.  The  increased nitrogen supply often associated
     with acid precipitation may have a positive growth effect.
     This does not exclude, however, the possibility that adverse
     influences may be developing over time in the more susceptible
     forest ecosystems.   The most serious consequence for
     terrestrial  ecosystems of regional acidification at levels
     currently observed  in Norway may be the increased rate of
     leaching of major elements  and trace metals from forest soils
     and vegetation.  This also  has a bearing on the aquatic systems
     receiving these effluents.  From an ecological point of view it
     is difficult to forecast the ultimate results of the
     atmospheric acidification and related air pollutants on
     terrestrial systems and to  judge the rate and even the
     direction of changes. In the more susceptible areas it seems,
     however,  to be a question of proportion and time required
     rather than whether any ecological effects appear or not."

     In examining the Scandinavian work it is important to note that the
character of their atmospheric emissions and the chemistry of their
rainfall have changed dramatically in recent years, resulting in
substantial increases in nitrogen inputs from the atmosphere.  Emission
of S02 in Sweden increased 85 percent (from 240 to 445 thousands of
tons of S yr-1) during the interval 1950 to 1970, but had decreased
back to 240 tons yr"1 by 1978.   Sulfate in precipitation showed a
substantial (65 percent) increase (from 55 to 90 microequivalents per
liter) during the interal 1955 to 1964, but then remained constant
through 1974.  By contrast, nitrate levels increased by 33 percent (15
to 20 meq 2,"1) from 1955 to 1964 and by 1974 had reached 35 meq
jf1, a level 133 percent above that in 1955 (SMA 1982).  Thus, while
it will be difficult to  interpret the Scandinavian tree-ring studies
until both climatic and  age-related trends are removed from the data,
the most recent analysis suggests the possibility that relatively recent
significant increases in atmospheric inputs of nitrogen (coupled with
the trends in atmospheric chemistry) may be an important factor in most
recent changes in growth patterns.

     In the United States, Cogbill (1976) examined growth of beech,
birch, and maple in the  White Mountains of New Hampshire and red spruce


                                 3-36

-------
1n the Smoky Mountains of Tennessee.   From analysis of tree-ring
chronologies, he concluded that no synchronized regional decrease in
radial growth had occurred.  The ring chronologies presented for all of
the species he studied, however, showed  evidence of a decreasing growth
trend from around 1960 until  1970. More recent studies  In New York by
Raynal (1980) with red spruce and white  pine and by Johnson et al.
(1981) 1n the New Jersey pine barrens with pitch, shortleaf, and
loblolly pine have shown patterns of  decline among most  of these species
during the past 26 years.

      In New Jersey, a strong statistical  relationship between annual
variation in stream pH and growth rates  suggested that acidic
precipitation may have been a growth-limiting  factor for the past two
decades (Johnson et al. 1981).  Stream pH, In  this poorly buffered soil
system, was closely correlated with precipitation pH during a 36-month
period of concurrent records.  Of the trees examined, approximately
one-third showed normal growth, one-third showed noticeable abnormal
compression of annual Increments during  the past 20 to 25 years, and the
remainder showed dramatic reduction in annual  growth over this time
Interval.  This effect was evident in trees of different species and at
different sites and occurred regardless  of age or whether trees were
planted or native.  An interesting response of both these trees and the
four species examined by Puckett (1982)  in southeastern  New York was a
change in the Influence of climate on tree growth over the past 25 to 30
years.  Increased sensitivity of trees in these studies  to climatic
variables suggests the possibility that  changes in the physiological
relationship of these trees to their  growing environment may have
occurred during recent decades.

     Of the above studies, only that  of  the pine barrens by Johnson et
al. (1981) examined the possible Influences of gaseous pollutants on
observed growth trends.  In those studies, growth reponses were
demonstrably unrelated to 03  levels.   Although uncertain, we might
anticipate that gaseous air pollutants would also have played only a
minor Influence on growth trends observed in Scandinavia where the
density of gaseous pollutant sources  is  rather low and concentrated in
coastal areas (SMA 1982).  In central Europe where Ulrlch et al. (1980)
have reported dieback and decline of  Norway spruce and beech and 1n
inland areas of the eastern United States, contributions of gaseous
pollutants, primarily 03 and  $03, can be considered to have changed
over the same time spans as has add  precipitation and thus should be
considered 1n any study of long-term  growth effects.

3.4.1.5  Dieback and Decline  in High  Elevation Forests--Within the
United States, the forests presently  receiving the highest levels of
acidic deposition are those at high elevations in the northeast.
Forests characterized by varying proportions of spruce,  fir, and white
birch occur at the high elevations of the  Appalachian Mountains from
eastern Canada to North Carolina.  The northern boreal forests of New
York, Vermont, and New Hampshire have received considerable attention
with respect to the potential for acidic deposition Impacts.  Although
the mountain summits are remote from  large point sources of sulfur, they


                                  3-37

-------
receive extraordinarily high rates of H+,  sulfur,  and  heavy metal
deposition (Lovett et al.  1982,  Fried!and  et al. 1983).   In addition,
the vegetation is subjected to very acid cloud  moisture  for a
considerable portion of the year (Johnson  et al. 1983).   Typically,
cloud moisture pH is in the range 3.5 to 3.7, whereas  ambient
precipitation is about pH  4.1 to 4.3.  Another  cause for attention stems
from the quantitative documentation of a red spruce decline in the Green
Mountains of Vermont, the  causes of which  are obscure  at present
(Siccama et al. 1982).

     The northern boreal forests are characterized by  red spruce  (Picea
rubens), balsam fir (Abies balsamea)  and white  birch (Betula papyrifera
var. cordifolia) in the canopy,  mountain ash (Pyrus americana) and
mountain maple (Acer spicatum) as important understory trees, and an
herb layer dominated by ferns (Dry op ten's  sp.)  and pxalis montana
(Siccama 1974).  The lowermost elevation to which  the  boreal forests
extend varies from 250 m above sea level in Maine  and  Nova Scotia to 750
m in New Hampshire and Vermont,  900 to 1000 m in the Adirondack and
Catskill Mountains of New  York,  and about  1500  m in North Carolina
(Costing 1956, Siccama 1974). The presence of boreal vegetation is
believed to be related to  the incidence of cloud moisture, with the
boreal  vegetation occupying the  often cloud-capped upper slopes, and
hardwoods holding the lower elevation sites (Nichols 1918, Davis 1966,
Vogelmann et al. 1968, Siccama 1974).  In  the Green Mountains of
Vermont, the boreal forests are above cloud base for 800 to 2000 hours
per year, depending on elevation (Johnson  et al. 1983).

     Although there is considerable interest in cloud  moisture pH and
there are several ongoing  studies in the mountains of  the Northeast (H.
Vogelmann, University of Vermont; F.  H. Bormann, T. G. Siccama, Yale
School  of Forestry; G. E.  Likens, J.  Eaton, Cornell University; V.
Mohnen. J. Kadlecek, State University of New York, Albany; C. V.
Cogbill, Center for Northern Studies), there are few published data.
Data from especially designed cloud moisture collectors  at Mt.
Moosilauke, NH, indicate that growing season cloud moisture pH is
generally in the mid-3 range (Lovett et al. 1982). The  few reported
cloud pH measurements obtained from airplane flights suggest that
growing season cloud moisture pH is distinctly  lower than moisture
precipitated from the cloud, and that clouds are most  acid near cloud
base (Scott and Laulainen  1979).  The current indication is that cloud
moisture pH is approximately 0.5 pH units  lower than ambient rain or
snow pH, but considerably  more data are needed  to  characterize the
nature of cloud acidity.  The implication  is that  boreal  forest
vegetation is exposed to moisture with pH  of 3.0 to 4.0  frequently and
for a total of 30 to 80 days per year.

     In the mountainous areas of New England, precipitation increases
with altitude.  Lovett (1981, in Cronan 1983) estimates  precipitation
rates of 240 cm yr-1 in the balsam fir forests  of  New  Hampshire.
Low-elevation precipitation in New England ranges  from about 100 to 150
cm yr-1.  Siccama (1974) determined that growing season  throughfall
increased by 2.9 cm per 100 m~2  in the Green Mountains of Vermont due
                                  3-38

-------
to increased rainfall and an increase in the cloud moisture intercepted
by vegetation.  Vogelmann et al. (1968)  report that at 1087 m in the
Green Mountains, open collectors fitted with screens to intercept cloud
moisture collected 66.8 percent more water than control collectors
without screens.  Throughfall collectors placed under balsam fir at 1250
and 1300 m in the White Mountains collected 8 percent more water than
precipitation collectors placed in the open at the same elevation, and
36 percent more water than precipitation collectors located at 520 and
640 m.  Thus, high precipitation rates coupled with intercepted cloud
moisture probably produce H+ deposition rates far in excess of the
regional rates reported by precipitation collection networks based on
samples collected at lower elevation.

     Cronan (1983) estimated H+ input to the canopy at 77 to 100 meq
m-2 for the 6 month period May through October, 1978 in the high
elevation fir stands.  The hardwood canopy at 520 and 640 m received 50
to 62 meq H+ m-2 during this period.  Based on Cronan1s data,  it
appears that the boreal  forest canopy is not effective at neutralizing
atmospherically deposited H+ as throughfall collectors indicated that
the H+ input to the forest floor under fir was 98 mg nr2 for the
growing season.   Probably the best estimate of H+ deposition has been
made by Lovett et al. (1982), who used field collection of cloud
moisture samples and modeling of cloud droplet interception to estimate
H+ deposition in the subalpine zone of the White Mountains to be ~
340 meq m-2 yr-1.

     As a result of the substantial input and the inferred low
neutralization capacity of the canopy (Cronan 1983), the potential  for
accelerated leaching of bases is high, but to date,  no quantitative data
from high elevation forests indicate if  the rate has actually  increased
over the past few decades.  Changes in soil pH are not expected to be
rapid, as the forest floor of the boreal  zone soils  is naturally
extremely acid.  Siccama (1974) reported soil pH in  HgO of 3.4 to 3.7
in the forest floor (0 horizons)  at Camels Hump, Vermont in the
mid-1960's.  Johnson et al. (1983)  found that at the same sites, pH was
slightly but not significantly higher in 1980.

     Estimates of dry deposition have not been made  for high-elevation
forests, but as wind velocities increase with altitude (Siccama 1974)
and as conifers have a high surface area and have foliage all  year, dry
deposition may add substantially to the  quantity of atmospherically
deposited H+ processed.

     A decline of red spruce (but not fir or white birch) has  been
quantitatively documented in the Green Mountains of Vermont (Siccama et
al. 1982)  and observed in New York  and Mew Hampshire (Johnson  et al.
1983).  An overall.reduction of approximately 50 percent in basal  area
and density was observed in the Green Mountains between 1965 and 1979.
Trees in all  size classes were affected.   The primary cause is presently
unknown, but it is not likely to be successional  dynamics,  climatic
changes, insect damage,  or primary  pathogens (Hadfield 1968, Roman and
Raynal 1980,  Siccama et al. 1982).   Studies of pathogens in declining


                                  3-39

-------
spruce Indicate the presence of secondary  fungal pathogens, with
Arm-Mi aria me! 1 ea,  Fomes pini,  and Cytospora kunzi'1 most prominent
(Hadfleld 1968).  Hadfleld (1968) speculated that the Infected trees had
been weakened by the drought of the early  1960's prior to Invasion by
the fungi.  Using the framework of Manlon  (1981), the spruce decline has
the characteristics of a complex blotlc-ablotlc disease related to
environmental stress.  Currently, there  are no data which Implicate
acidic deposition as a contributing stress, nor are there data which
rule out all of the possible pathways by which acidic deposition could
affect forest trees.

     At present, serious dleback of spruce (Plcea ables) and fir (Abies
alba) Is under study 1n Germany. From long-term, Intensive, ecosystem-
Tevel studies, UlHch (Ulrlch et al.  1980; Ulrlch 1981a,b, 1982)
suggested that acidic deposition has  contributed to changes 1n H+
generation and consumption which have caused soil acidification,
mobilization of Al, mortality of fine roots, and ultimately, dleback and
decline 1n spruce, fir, and beech (Fagus sylvatica).  That contention 1s
based on careful documentation  of changes  In soil solution chemistry, a
nearly parallel decrease In fine root blomass and Increase 1n soil
solution Al concentrations during the growing season, and nutrient
solution studies  which Indicated that the ratio of uncomplexed Al
(I.e., Al3+) to Ca found In the soil  solution was sufficient to cause
abnormal root growth and development. While those findings suggest the
possibility of Al toxldty, they are  not definitive.  Bauch (1983)
determined that the roots of declining spruce and fir were Ca deficient,
but had the same levels of Al as healthy spruce and fir.  Rehfuess
(1981) has observed declining fir on  calcareous soils which would seem
to preclude Al toxlclty or Ca deficiency In  those cases.  More recently,
however, Rehfuess et al. (1982) noted Mg and possible Ca deficiencies by
foliar analysis even In base-rich  soils.  They speculate that
accelerated foliar leaching may be  reponsible (see Section 3.2.1.2).
Rehfuess points out that the parallel change in soil solution Al and
fine root blomass noted by Ulrlch was not  synchronized 1n that marked
decreases in fine root blomass  preceded  the  increase in soil solution
Al.  Rehfuess cites several studies  (Goettsche 1972, Deans 1979, Persson
1980) in support of his contention  that  late simmer declines 1n fine
root blomass are naturally controlled, and need not be related to Al
levels.  Ulrich's extrapolation of  nutrient  solution Al:Ca levels to the
field situation are also questionable, since the soil matrix may alter
the availability of those and other plant-essential or phytotoxic
elements.

     The hypothesis of Ulrich appears to have limited applicability to
the North American spruce decline,  where dieback and decline Is most
prominent in the high elevations where soils are Borofolists or
Cryofolists which have  ~ 80 percent organic  matter by weight
(Friedland et al. 1983), and Al toxicity would likely be masked by
complexation with organic matter (Ulrich 1982).  Data on spruce root
chemistry from Camels Hump, VT, indicate that CarAl ratios increase with
increasing elevation.  As mortality increases with elevation, it 1s
                                  3-40

-------
not likely that imbalances of Al  and Ca in root tissue are  the major
cause of spruce decline (Lord 1982,  Johnson et al.  1983).

     Whether the red spruce decline  is related to  acidic deposition has
been the focus of considerable speculation.  The deline is  widespread,
easily discerned, dramatic, and of unknown origin.  It has occurred in an
environment that receives very high  annual input of H+ from the
atmosphere and where trees are frequently subject  to intensely acid
cloud moisture; hence, it is logical that research on acidic deposition
effects in high-elevation forests has been initiated.

     At present, there are few testable hypotheses regarding how  acidic
deposition could have contributed to spruce mortality. The Al toxicity
proposed by Ulrich (1981a,b; 1982) is not supported by the  data
collected to date.  The foliar leaching hypothesis of Rehfuess et al.
(1982) remains untested as yet, however.

     The spruce decline appears to be a stress-related disease, where
the trees are probably predisposed to stress by the site conditions
where some short-term stress, possibly the drought of the early 1960's
triggered a loss of vigor, and where biotic stress imposed  by fungal
attack is sufficient to cause widespread mortality.   Acidic deposition
could act to intensify the predisposing stresses,  exacerbate the  effects
of the triggering stress, or increase the susceptibility to fungal
attack, and those possibilities warrant research in the future.

3.4.1.6  Summary—At present there is no proof that acidic  deposition is
currently limiting growth of forests in either Europe or the United
States.  From field studies of mature forests trees it is apparent that
altered growth patterns of principally coniferous  species examined to
date have occurred in recent decades in many areas of the northeastern
U.S. and in some areas of Europe with high atmospheric deposition
levels.  Recent increases in mortality of red spruce in the northeastern
U.S. and Norway spruce and beech in  Europe add further to the concern
that forests are undergoing significant adverse change, however,  no
clear link has been established between these changes and anthropogenic
pollutants, particularly acidic rainfall.  This must be presently viewed
from the perspective of two possible hypotheses:   (1)  recent changes are
purely circumstantial and not in an  way linked to  acid precipitation, or
(2) we have not yet adequately studied a  very complex association in
which multiple and interactive factors may be involved and  responses may
be subtle and chronic.

     It is too early to conclude that acidic deposition has not nor will
not affect forest productivity.  Irrigation studies with seedlings and
young trees provide no indication of immediate alarm yet are difficult
to interpret because of potential artifacts of experimental  protocols.
Detecting responses of mature forest trees is made difficult by the
complexities of competition, climate, and site factors, the potential
interactions between acid precipitation,  gaseous pollutants,  and  trace
metals, and the lack of control or unattended sites with which acid
precipitation impacted sites can  be  compared.   Although the task


                                  3-41

-------
of assessing potential  impacts of forest productivity will  assuredly be
difficult, the potential  economic and ecological  consequences of even
subtle changes in forest growth over large regions  dictates that it
should be attempted.

     To address these problems it will  be necessary to  evaluate the
long-term dynamics of forest systems over a broad enough  range of
environmental  conditions to document both whether systematic changes
have occurred and the extent to which such changes  are  linked to
variables such as levels of deposition of anthopogenic  pollutants, soil
fertility, moisture status, species composition,  and stand  stocking.   A
combination of approaches will be needed:  dendroecological  studies to
document past growth patterns of trees in a broad range of  conditions,
permanent long-term growth plots to study changes in stand  dynamics, and
forest growth models to examine the potential  long-term significance of
changing growth rates to forest growth and compensation.  The above
approaches will be correlative in nature and should be  used to focus on
the range of conditions in which responses have occurred.   However, they
must also be coupled with mechanistic studies aimed at  specific
mechanisms of effect before acid precipitation effects  on forest
productivity are ever conclusively established or refuted.

3.4.2  Crops (P. M. Irving)

     A considerable number of studies on the vegetative effects of
acidic precipitation have been published in the last 5  years.  However,
because of limitations in research design, few of these studies can be
used to estimate crop loss realistically.  Among the large  scale field
studies which are most potentially useful for estimating  yield effects,
differences in methodologies make intercomparisons  difficult and results
appear to be inconsistent.  The following is a discussion of the
approach used in acid precipitation effects studies, an analysis of the
design limitations of those studies, and a comparison of  their
methodologies and results.

3.4.2.1  Review and Analysis of Experimental Design—The  most widely
used method for making crop loss assessments in the past  has been  field
surveys in which observers estimate vegetation injury from  visible
symptoms under ambient conditions and subjectively  relate leaf damage  to
yield loss.  Since visible injury to crops has never been reported as
the result of ambient acid precipitation, experiments  using simulated
rain in field or controlled environment (i.e., greenhouse,  growth
chamber, laboratory) studies have been used to determine  the threshold
acidity levels that produce visible injury.

     Three general approaches have been used to determine impacts  on
plants from acidic deposition:  (1) Determination of a  dose-response
function for a specific species in a defined environment; (2)
classification of relative sensitivity based on morphological,
physiological, or genetic characteristics; (3) determination of
mechanisms of action.  Both field and controlled-environment
methodologies with simulated rain have been used in these approaches.


                                  3-42

-------
Only dose-response studies provide quantitative data  to estimate  growth
and yield effects.

3.4.2.1.1  Dose-response determination.   Current methods for determining
whether crop yield losses are occurring  from add rain exposure Include
dose-response studies to mathematically  relate yield  to pollutant dose.
The term 'dose-response' suggests a unlvarlate relationship; however,  a
number of potentially Important variables comprise 'acid rain dose1  (see
next section).  Complex, factorial designs and multlvarlate analyses may
be necessary to describe the relationships adequately.   Dose-response
studies of pollutant effects on crops fall Into two basic categories:
(1) field studies and (2) control!ed-env1ronment studies.  Each type of
study has Its advantages and limitations.

     Field studies are often a more realistic means of estimating actual
effects because the experimental  plants  can be grown  under normal
environmental conditions, especially 1f  common agricultural  practices
are used.  Because different environmental conditions related to
geography (I.e., temperature, soil type, and water availability)  may
lead to different responses, field studies are useful  1n estimating
regional Impacts of pollutants when similar experiments are performed  In
various regions and then compared.  Field research, however, demands
considerable time and labor and Is thus  expensive. Adding to the
expense Is the need for either a high degree of replication so that  the
sometimes subtle treatment effects can be observed above the differences
caused by environmental variability or a large number of treatment plots
for response surface analyses.  Reliable dose-response predictions
cannot usually be made without at least  2 to 3 years  of replicate
studies conducted using normal agronomic practices.

     A lack of comparable unpolluted (control) plots  1s also a problem
for field studies In most regions.  This has led to the use of such
devices as open-top chambers for the elimination of gaseous pollutants
from field plots and to the use of rain  exclusion shelters.   Experiments
using these devices must be designed properly for valid comparisons  to
be made.  For example, In a study by Kratky et al. (1974), plots  of
tomato plants were placed Inside and outside plastic  rain shelters 1n
the Kona district of Hawaii during a volcanic eruption period. The
plants growing outside the ralnshelter received rain  with a pH of 4.0
and produced no salable yield, while plants under the shelter averaged 5
kg per plant of salable fruit.  However, an explanation other than acid
rain should be considered for the Kratky study because of a possible
shelter effect.  Dry deposited materials from the volcanic eruption,
possibly acidic, may have been dissolved by rainfall  on leaf surfaces
outside the shelter but remained In the  non-reactive  dry form Inside the
shelter.  Thus rainfall, acidic or not,  would have had an effect  by
acting as a wetting agent.  The problem  of separating the effects of dry
deposition when 1t occurs In conjunction with wet deposition 1s one
facing all field researchers.

     Controlled-envlronment studies are  useful  Indicators of potential
effects and may suggest subtle changes not measureable 1n an
                                  3-43

-------
uncontrolled situation.   Controlled studies  also allow the investigator
to reduce the dimensionality or number of  variables in the experiment.
These types of studies,  for example,  may be  necessary to determine which
characteristics of rain  (i.e.,  intensity,  droplet size, ionic
composition) must be simulated  in field studies.  Their use is limited,
however, because plants  may be  more sensitive  to stress when grown under
short photoperiod, low light intensity, medium temperature, and adequate
soil moisture (Leung et  al. 1978),  conditions  which frequently occur in
a growth chamber or greenhouse  as compared to  the field.  Since
controlled-environment studies  may  overestimate acid rain stress because
of greater plant sensitivity, they  should  be used with caution when
assessing potential  damage.  For example,  Lee  and Neely (1981) found
chamber-grown radish and mustard greens to be  more sensitive to
simulated acidic rain than  were field-grown  plants.  Troiano et al.
(1982) observed that greenhouse-grown plants developed foliar injury
more readily from acid rain simulants than did field-grown plants.
Since light intensity and wind  speed affect  cuticular development
(Juniper and Bradley 1958), which in turn  affects leaf wettability,
greenhouse-grown plants  may be  affected more by acidic deposition than
field plants because of  decreased wax development (see Section 3.2.1.1).
On the other hand, under some conditions plants may be more stressed in
controlled environments  (due to restricted root growth or lower
photosynthetic rate) and thus less susceptible to treatment stress
because of lower metabolic  rates and thus  lower pollutant uptake.

     Soil factors, nutrition and cultural  practices (i.e., application
of fertilizer, pesticides and other chemicals, irrigation, planting
schedules) may all affect the sensitivity  of a plant to pollution and
therefore should be recorded in experimental methods and, for greater
accuracy, should reflect common agricultural conditions as closely as
possible.  To determine  the interaction of these factors with pollutant
effects, controlled environment studies are  necessary.

     Pollutants rarely occur alone, and since  pollutant combinations
have been found to cause more-than-additive  or less-than-additive
effects (Ashenden and Mansfield 1978, Jacobson et al. 1980), the
concentrations of other pollutants should  be monitored and reported in
conjunction with acid precipitation studies.   Exposures of various
pollutant combinations in controlled studies are necessary to determine
interactive effects.

3.4.2.1.2  Sensitivity classification. There  may be considerable
variability in sensitivity  to pollutant stress between plant
communities, species within communities, cultivars within species, and
growth stages of cultivars  (Heggestadt and Heck 1971; see also Section
3.4.1.2).  Gaseous pollutants (i.e., ozone,  sulfur dioxide) have been
found to affect certain  crop cultivars more  than others, and limited
information indicates that  this is also true for cultivar response to
acidic precipitation (see following section).   Because it would be
prohibitively expensive and time-consuming to  perform dose-response
studies on all crop cultivars,  some experimental studies are aimed at
identifying plant characteristics that can be  used to indicate a plant's


                                  3-44

-------
 relative  sensitivity or resistance to acidic deposition.  For example,
 leaf wettabillty, which 1s related to surface morphology, has been
 suggested as a parameter that may Indicate sensitivity to acidic
 precipitation (Evans et al. 1977a).

     It has been suggested that crop classes can be grouped according to
 their sensitivity to add precipitation.  Based on a study of 28
 different crops, Lee et al. (1981) reported that Inhibition of
 marketable yield was observed only 1n the dicotyledons that were
 studied,  and within this group root crops, leaf crops, cole crops, tuber
 crops, legumes and fruit crops were ranked 1n decreasing order of
 sensitivity.  But the data are contradicted by other studies.  For
 example,  Evans et al. (1982) In a study of two root crops found radishes
 to be resistant and garden beets to be sensitive to simulated acidic
 precipitation.

     Plant response may also be related to stage of development when
 exposure  occurs.  The possibility that a particular life stage may be
 more susceptible to an acid precipitation event than other stages must
 be considered when researchers Investigate and report acid precipitation
 effects.

 3.4.2.1.3  Mechanisms.  Studies having the objective of determining
 mechanisms of action of an air pollutant (mechanistic) can provide
 Information to explain the basis of an observed plant growth response.
 In studies of this type, measurements are made to determine effects on
 basic processes such as photosynthesis, respiration, transpiration, and
 metabolism.  Examples of such measurements Include  C02 uptake and
 emission,  leaf diffusive resistance, metabolite pools, and enzyme
 activities.  This Information may then be Interpreted and applied
 through the use of plant growth models to predict total plant response.
 Physiological  measurements may also be used to support and explain plant
 yield response.   For example,  Irving and Miller (1980), using a
 J^cog assimilation technique In the field, reported that S02
 exposures reduced both photosynthesis and yield of soybeans but that
 acid rain treatments had apparently stimulated the photosynthetlc rates
 with no effect on soybean yield.   Usually physiological determinations
 alone are 1n adequate to estimate the economic damage of pollutants to
 crops.

 3.4.2.1.4  Characteristics of precipitation simulant exposures.   The
 effects of a pollutant on crop yield may be defined by correlating yield
 variations with variations In pollutant dose.   Acidic precipitation,
 however, consists of a number of variables that may have an effect on
 crop yield.  For example,  the sulfate and nitrate concentrations,  which
 are frequently correlated with the hydrogen ion concentration of the
 rain,  may be more Important in affecting plant response than the pH of
 the rain (Irving and Sowlnski  1980).   Lee and  Neely (1980)  found that
 simulated rain acidified with sulfuric acid resulted in a different
effect on the growth of mustard green,  onion,  fescue,  radish,  lettuce,
 and orchard grass than simulated rain at the same pH,  acidified with
 sulfuric and nitric  acids  (2:1 equivalent weight ratio;  refer to Tables


                                  3-45

-------
3-2 and 3-3 in Section 3.4.2.2).   Acid  rain  dose should therefore be
described by concentrations  of  sulfate, nitrate, and other important
ions (e.g., NH4+,  Ca2+, Mg2+, etc.),  as well  as hydrogen ion
(pH).  For a complete analysis,  it may be  necessary to determine the
effect of each individual  ion as well as their combination so that all
important ions are simulated at levels  found in polluted and unpolluted
rain.

     Plant injury  responses  are a function of pollutant concentration
and exposure time  or quantity  (i.e.,  acid  rain dose = [H+ x cm rain] +
[$042- x cm] + [N03~ x cm].  Response to a given dose of gaseous
pollutant is frequently greater if deposited in a shorter exposure time.
Response to acid rain, however,  may be  positively correlated with the
amount of time the leaf is wet.   When comparing experimental results,
one must compare concentration  and duration  of exposure to understand
the response in terms of dose and rate.  In  the case of acid rain,
reporting the pH of applied  precipitation  is inadequate without total
dose or deposition of important ions  (i.e.,  kg ha~l of SCtyZ-,
N03-, and H+), rate or intensity (i.e., cm hr-1), duration, and
frequency.  Physiological  systems can be quite resilient due to
activation of defense and repair systems during periods of stress.
Therefore, time between stress  events may  be important for repair
functions.  It has been reported that the  "recovery" period between
gaseous pollutant  exposures  may affect  the total plant response.
Similarly, the number of "dry"  days between  precipitation events may
influence the net  response of a plant to acidic deposition.  Because of
differences in leaf wettability,  plants may  respond differently to a
rain or mist; thus droplet size is yet  another important characteristic
(see Section 3.2.1.1).

3.4.2.1.5  Yield criteria.  Because crop production is measured in terms
of the yield of a  marketable product, it is  useful to express pollutant
injury in terms of the economically valuable portion of the crop.
However, this is not easily  applied uniformly in experimental studies.
Leaf injury estimates have been commonly used to assess pollution
damage, but economic loss is not always closely related to leaf damage
(Brandt and Heck 1968).  Assessing loss based on visible injury may
overestimate or underestimate  the economic loss.  For example, in a
study of defoliation effects on yield,  Jones et al. (1955) found no
reduction in root  yield or sugar content of  sugar beets after removal of
50 percent of the leaves.  Irving and Sowinski (1981) reported increased
yield of greenhouse-grown soybeans that had  also exhibited necrosis as a
result of acid rain exposures.   Increased  yield was also reported by Lee
et al. (1980) for  alfalfa that  exhibited foliar injury from acidic rain.
Conversely, chlorosis or necrosis of  leaves  could result in
considerable economic loss of  a crop  such  as lettuce or mustard greens
without causing measurable changes in leaf weight.

3.4.2.2  Experimental Results—To allow comparisons of acid pre-
cipitation effects research  by  investigators using various techniques,
it is necessary (although perhaps not sufficient) to describe the
experimental conditions, the dose, and  the responses for each
                                  3-46

-------
Investigation in comparable units.   Accordingly, calculations were made,
based on information in the literature  or  by  personal communication, to
describe each investigation in comparative terms.  These changes 1n
units were made only for comparison  purposes.  None of the experimental
results described below have been changed  from those of the orglnal
author.  Given the experimental  design  limitations discussed in the
previous section, conclusions based  on  the following research results
must be made cautiously.

3.4.2.2.1  Field studies.  The studies  described 1n Table 3-2 were
performed in the field, using accepted  agricultural practices to the
extent experimental design would permit.   Because hydrogen, sulfate, and
nitrate ions are those components of precipitation that are believed to
most likely affect the growth and yield of crops, they were used in
describing the precipitation dose.   In  all  experiments, simulated rain
was applied at regular intervals during the Hfe cycle of the crop and,
except for 'Beeson' and 'Williams' soybeans,  was applied in addition to
ambient precipitation.  Thus, total  deposition received by the crop is
the sum of simulant plus ambient loadings.

     Among the 14 crop cultlvars (9  species)  studied, only one exhibited
a consistently negative yield effect at all acidity levels used (garden
beet), three were negatively affected by at least one of the acidity
levels used 1n the study ('So. Giant Curled'  mustard green,  'Pioneer
3992' field corn, and 'Amsoy1 soybean), and six had higher yields from
at least one acidity level ('Champion'  and 'Cherry Belle' radish,
'Vernal' alfalfa, 'Alta' fescue, 'Beeson1  soybean, and  'Williams'
soybean).  The most frequent response reported to result from simulated
acidic rain was "no effect" ('Red Kidney'  kidney bean,  'Davis1 and
'Wells' soybean, 'Cherry Belle'  radish, 'So.  Giant Curled' mustard
green, 'Improved Thick Leaf spinach, and  'Vernal' alfalfa).  Some
experiments demonstrated both positive  and negative response to acid
rain, depending on the H+ concentratipn.   There 1s little evidence for
a linear response function, however, since no effect frequently occurred
at doses greater than those producing positive or negative response.
Except for garden beet, this was true for  each study that reported a
negative response to at least one level of acidic deposition.  For
example, a 9 percent decrease in the yield of corn resulted from
treatments with 42 times the ambient H+ deposition (six times ambient
H+ concentration), but no effect occurred  at  132 and 187 times (pH
4.0, 3.5, 3.0, respectively).  In the garden  beet study, the yield
decrease from acid rain was not the  result of lower beet root weights
but because of fewer number of marketable  roots per plot.  Perhaps the
acid rain treatments affected germination  or  seedling establishment.
The ratio of sulfate:nitrate ions 1n the precipitation  simulant also
affected the response of some plants (i.e., alfalfa, fescue, mustard
green; Table 3-2), independent of pH.

     A comparison of studies on five different cultlvars of soybeans by
four different investigators appears to indicate that the 'Amsoy'
cultivar may be more susceptible to  acidic deposition than 'Beeson',
                                  3-47

-------
                TABLE 3-2.   FIELD  RESEARCH  ON CROP  GROWTH  AND YIELD AS  AFFECTED BY  ACID  PRECIPITATION
GO
CO
Total deposition
kg ha"1
(simulant ••• ambient)
H+ S042- N03-
Simulant concentration
mg l"1
H+ S042- N03- S(
Rate
cm hr'
J4^~:N03~
Events
1 f
hr
events-1
Droplet
size
PH
Effect*

Alfalfa, 'Vernal', Hedlcago saliva L. (Lee and Neely 1980)
0.017
0.171
0.833
2.611
0.011
0.017
0.271
0.833
2.611
0.011
2.13
13.31
38.89
120.84
0.75
2.13
9.07
30.44
89.15
0.75
(Garden) Beet,
0.077

0.078

0.082

0.090

0.077
Corn, '
0.028
0.594

1.847
5.814
0.014
Fescue
0.017
0.271
0.833
2.611
0.011
0.017
0.271
0.833
2.611
0.011
Kidney

1.14










2»2o
2.26
2.26
2.26
0.30
2.26
7.89
19.64
60.85
0.30
0.'OQ25 0.53 0.753
0.10 4.83 0.753
0.316 14.67 0.753
1.00 46.19 0.753
0.016 1.07 0.434
0.0025 0.53 0.753
0.10 3.20 2.92
0.316 11.42 7.44
1.00 34.00 23.29
0.016 1.07 0.434
'Perfected Detroit V-9041, Beta vulgarls









'Pioneer 3992' Zea
4.03
19.51

67.20
198.16
0.96
(Tall),
2.13
13.31
38.89
120.84
0.75
2.13
9.07
30.44
89.15
0.75
4T7F
17.33

43.54
135.47
0.39
O.OOZ 1.26 3TD4

0.010 5.47 3.04

0.079 37.07 3.04

1.995 106.6 3.04

0.087
mays L. (Lee and Neely 1980)
JJ.0025 0.53 0.753
0.10 3.20 2.92

0.316 11.42 7.44
1.00 34.00 23.29
0.016 1.07 0.434
0.7
6.4
19.5
61.3
2.5
0.7
1.1
1.5
1.5
2.5
L. (Evans
0.4

1.8

12.2

35.1



0.7
1.1

1.5
1.5
2.5
'Alta1, Festuca elatlor L. var. arundlnacea Schreb.
2.26
2.26
2.26
2.26
0.30
2.26
2.26
2.26
2.26
0.30
Bean, 'Red Kidney'
13.02
98.05
12.99
5.57
5.57
5.37
0.0025 0.53 0.753
0.10 4.83 0.753
0.316 14.67 0.753
1.00 46.19 0.753
0.016 1.07 0.434
0.0025 0.30 0.753
0.10 3.20 2.92
0.316 11.42 7.44
1.00 34.00 23.29
0.016 1.07 0.434
0.7
6.4
19.5
61.3
2.5
0.7
1.1
1.5
1.5

0.67
0.67
0.67
0.67

0.67
0.67
0.67
0.67

et al
35.0

35.0

35.0

35.0



0.67
0.67

0.67
0.67
(Lee
0.67
0.67
0.67
0.67

0.67
0.67
0.67
0.67

26
26
26
26

26
26
26
26

. 1982)
19

19

19

19



58
58

58
58
and Neely
26
26
26
26

26
26
26
26

1.5
1.5
1.5
1.5

1.5
1.5
1.5
1.5


0.001

0.001

0.001

0.001



1.5
1.5

1.5
1.5
1980)
1.5
1.5
1.5
1.5

1.5
1.5
1.5
1.5

1200
1200
1200
1200

1200
1200
1200
1200


353

353

353

353



1200
1200

1200
1200

1200
1200
1200
1200

1200
1200
1200
1200

5.6
4.0
3.5
3.0
4.8
5.6
4.0
3.5
3.0
4.8

5.7

4.0

3.1

2.7

4.1

5.6
4.0

3.5
3.0
4.8

5.6
4.0
3.5
3.0
4.8
5.6
4.0
3.5
3.0
4.8
Control
9« greater yield than pH 5.6
No effect on yield compared to pH 5.6
No effect on yield compared to pH 5.6
Ambient
Control
No effect on yield compared to pH 5.6
No effect on yield compared to pH 5.6
No effect on yield compared to pH 5.6
Ambient

101 greater shoot growth, 16% greater root
yield than ambient
Lower number of marketable roots per plot
than ambient or pH 5.7
Lower number of marketable roots per plot
than ambient or pH 5.7
Lower number of marketable roots per plot
than ambient or pH 5.7
Ambient

Control
9% lower yield; no effect on growth compared
pH 5.6
No effect on growth or yield compared to pH 5
No effect on growth or yield compared to pH 5
Ambient

Control
24% greater yield than pH 5.6
19% greater yield than pH 5.6
No effect on yield compared to pH 5.6
Ambient
Control
No effect on yield compared to pH 5.6
No effect on yield compared to pH 5.6
No effect on yield compared to pH 5.6
Ambient






















to

.6
.6










, Phaseolus vulgarls L. (Shrlner and Johnston 1981)
0.001 O.OZ 0.12
0.2
0.631 50.0 0.12 417
2.30 0.95
2.4
3.0
3.0

27
27

0.17
0.17

900
900

6.0
3.2

Control
No effect on growth or yield compared to pH 6
Ambient

.0

         aEffects are reported when statistical significance 1s < 0.05 level.

-------
                                                           TABLE  3-2.   CONTINUED
co
Total deposition Simulant
concentration
kg ha'1 mg ir1
(simulant + ambient)
H+ S042- N03- H+ S042-
Hustard
0.033
0.189
0.535
1.629
0.029
0.033
0.189
0.535
1.629
0.029
Radish,
0.106

0.130
0.231
0.733
0.105
0.139

0.169
0.243
0.915
0.138
Radish,









Radish,
0.018

0.081
0.090
0.129
0.081
Green, 'So. Giant Curled' , Brassica
2.78 1.95 0.0025
9.66 1.95 0.10
25.40 1.95 0.316
75.83 1.95 1.00
1.93 0.78 0.016
2.78 1.95 0.0025
7.05 5,45 0.10
20.20 12.68 0.316
56.33 38.04 1.00
1.93 0.78 0.016
'Champion', Raphanus satlvls
0.0025

0.06
0.32
1.585
0.17
0.0025

O.OS
0.32
1.58
0.16
0.53
4.83
14.67
46.19
1.07
0.53
3.20
11.42
34.00
1.07
N03-
Rate
cm hr'l
S042~:N03~
Events Droplet
1 hr
events-1
japonlca Hort. (Lee and Neely 1980)
0.753
0.753
0.753
0.753
0.434
0.753
2.92
7.44
23.29
0.434
L. (Trolano et
0.72 0.31

2.9
11.7
55.6

0.72

2.90
11.70
55.60

'Cherry Belle', Raphanus satlvus L.
0.0025
0.10
0.316
1.00
0.026
0.0025
0.10
0.316
1.00
0.026
0.53
4.83
14.67
46.17
0.96
0.53
3.20
11.42
34.00
0.96
'Cherry Belle', Raphanus satlvus L.
0.002

0.010
0.079
1.995
0.087
1.26

5.47
37.07
106.6


1.4
5.8
27.6

0.31

1.40
5.80
27.6

0.7
6.4
19.5
61.3
2.5
0.7
1.1
1.5
1.5
2.5
al. 1982)
2.3

2.1
2.0
2.0

2.3

2.1
2.0
2.0

(Lee and Neely 1980)
0.753 0.7
0.753
0.753
0.753
0.471
0.753
2.92
7.44
23.29
0.471
(Evans
3.04

3.04
3.04
3.04

6.4
19.5
61.3
2.5
0.7
1.1
1.5
1.5
2.5
et al. 1982)
0.4

1.8
12.2
35.1

0.67
0.67
0.67
0.67

0.67
0.67
0.67
0.67

1.0

1.0
1.0
1.0

1.0

1.0
1.0
1.0

0.67
0.67
0.67
0.67

0.'67
0.67
0.67
0.67

35.0

35.0
35.0
35.0

16
16
16
16

16
16
16
16

5

5
5
5
9
6

6
6
6
11
12
12
12
12

12
12
12
12

9

9
9
9


1.5
1.5
1.5
1.5

1.5
1.5
1.5
1.5

1

1
1
1

1

1
1
1
1
1.5
1.5
1.5
1.5

1.5
1.5
1.5
1.5

0.001

0.001
0.001
0.001

size
Mm

1200
1200
1200
1200

1200
1200
1200
1200

1900

1900
1900
1900

1900

1900
1900
1900

1200
1200
1200
1200

1200
1200
1200
1200

353

353
353
353

pH

5.6
4.0
3.5
3.0
4.8
5.6
4.0
3.5
3.0
4.8
5.6

4.2
3.5
2.8
3.8
5.6

4.2
3.5
2.8
3.8
5.6
4.0
3.5
3.0
5.6
5.6
4.0
3.5
3.0
5.6
5.7

4.0
3.1
2.7
4.1
Effect*

Control
No effect on growth or yield compared to pH 5.
No effect on growth or yield compared to pH 5.
No effect on growth or yield compared to pH 5.
Ambient


6
6
6

Control
31% lower yield; 29% lower root wt than pH 5.6
No effect on yield or growth compared to pH 5.6
33% lower yield; 24% lower root wt than pH 5.6
Ambient

No effect on yield but 5% higher shoot wt than
ambient
7% higher root wt (yield) than pH 5.6
7% higher root wt (yield) than pH 5.6
13% higher root wt (yield) than pH 5.6
Ambient
•12% lower root wt (yield), 71 higher shoot wt
than ambient
31 higher root wt (yield) than pH 5.6
11% higher root wt (yield) than pH 5.6
17% higher root wt (yield) than pH 5.6
Ambient
Control
No effect on growth or yield compared to pH 5
25% greater yield than pH 5.6
No effect on growth or yield compared to pH 5
Ambient
Control
No effect on growth or yield compared to pH 5
No effect on growth or yield compared to pH 5
No effect on growth or yield compared to pH 5
Ambient
No effect on growth or yield compared to pH
4.06 (ambient)
No effect on growth or yield compared to pH 5
No effect on growth or yield compared to pH 5
No effect on growth or yield compared to pH 5
Ambient












.6

.6


.6
.6
.6



.7
.7
.7

         aEffects are reported when statistical significance Is £ 0.05 level.

-------
                                                          TABLE 3-2.   CONTINUED
tn
O
Total deposition Simulant concentration
kg ha-1
(simulant + ambient)
H+ SO*2' N03- H*
Soybean. 'Amsoy',







Soybean,
0.229
0.916

3.262

0.218
Soybean,
0.198
0.496
1.976
4.965

0.216
0.431
1.717
10.834

Soybean,
0.077

0.464

0.076
Soybean,
0.229
0.916

3.262

0.218
Spinach,
0.033
0.134
0.503
1.529
0.029
0.033
0.134
0.503
1.529
0.029







,b 'Bee son
2.88
10.21

39.97

10.51
, 'Davis1,
6.19
7.25
93.19
256.31

11.13
25.25
127.33
683.91

'Wells',
9.02

18.72

8.90
Glyclne max (L.)
0.10
0.794
1.995

10.0

0.79
' , Glyclne max (L

-------
 'Davis1, 'Williams1 or 'Wells'; however, the experimental  conditions
 such as soil type and characteristics of the rain simulant 1n the
 'Amsoy' study were markedly different than the other studies and may be
 responsible for the observed effect.  Figure 3-5 Indicates the location
 and results of the four soybean field studies 1n relation  to the
 principal production regions and soil types.  The one cultlvar that
 responded negatively to add rain treatments ('Amsoy') was grown 1n an
 area with a sandy soil, unlike the other studies.  The simulated rain
 used in this study was applied more frequently and also had high
 concentrations of heavy metals (i.e., 20 ppb Cd, 50 ppb Pb, 100 ppb F;
 Evans et al. 1977) that were not present in the rain simulants used by
 other Investigators.  The 'Beeson' and 'Williams' cultivars, which  were
 studied in a location near the 'Amsoy1 study, but with more structured
 soil, reponded positively to the acid rain treatments when ambient  ozone
 was removed.  The 'Davis'  and 'Wells' cultivars were studied in major
 soybean-growing areas with highly buffered soils and had no response to
 acid rain treatments as much as ten times more acidic than ambient.
 This comparison suggests that the region may be an Important component
 of the response because of differences 1n major soil types, cultivars
 grown, climatic conditions, and ozone concentrations.  Perhaps of equal
 Importance is the presence of toxic heavy metals which become more
 soluble with Increases 1n acidity of simulated rain.

     In the five separate studies of radish (two cultivars), a positive
 linear correlation between yield and acidity was observed  in two studies
 (Troiano et al. 1982), a non-linear positive correlation was observed in
 another study (Lee and Neely 1980), and no effect was reported in two
 studies (Lee and Neely 1980, Evans et al. 1982).  The differences In
 results could be due to factors such as cultivar differences,
 environmental  variability, or differences in total  deposition of H+,
 $042-, N03~, or $042- to NOa- ratios.
     Experimental results from some of these studies also demonstrate
that the response of unharvested biomass is not a  reliable predictor of
yield response.  Effects on marketable yield will  not necessarily be
reflected in changes In shoot or root growth.   For example,  field corn
(Table 3-2) exhibited lower grain yield at pH  4.0  but no  effect on shoot
growth.  The results from these studies are Inadequate to indicate and
whether the average concentration or total  deposition of  H+,  $042-,
N03~ is important in determining yield response.

3.4.2.2.2  Controlled environment studies.   As with the field studies,
experimental conditions, dose, and response In all  controlled
environment studies are expressed in comparable units,  based on
calculations from published and private communications (Table 3-3).  To
compare total  deposition In Tables 3-2 and 3-3 multiply g m-2 (Table
3-3) by 10 to obtain kg ha-1 (Table 3-2).   A comparison of effects on
the same species grown 1n a controlled environment as opposed to in the
field indicates a similar response 1n most species (alfalfa,  spinach,
mustard green, soybean) although radishes  exhibited a negative effect in
a controlled environment and a positive effect In  the field.  In
                                  3-51


   409-262 0-83-6

-------
                                                                SOYBEANS
                                                  Crop yield - kg ha~! (harvested)
                                                     1978 Census  of Agriculture
                                                 0  - 1500
                         1500 -  2000
                                     > 2000
oo
en
ro
     ANL-Argonne National Laboratory
         Soil:  silt  loam 'Martinton1
         Cultivar:   'Wells'
         Acidity Effect:  None
         Irving and Miller 1981
     BNL-Brookhaven National  Laboratory
         Soil:  loamy sand 'Plymouth'
         Cultivar:   'Amsoy'
         Acidity Effect:  Negative
         Evans et al.   1981d
     BTI-Boyce Thompson Institute
         Soil:  sandy loam
         Cultivar:   'Beeson1, 'Williams'
         Acidity Effect:  Positive
         Troiano et al. 1983
     NCS-North Carolina State University
         Soil:  sandy clay loam  'Appling'
         Cultivar:   'Davis'
         Acidity Effect:  None
         Heagle et al.  1983
                                                                                                                       NCS
Figure 3-5.
Location of  four  soybean  field  studies,  indicating
production regions and soil  types.

-------
       TABLE 3-3.   CONTROLLED ENVIRONMENT STUDIES  ON  CROP  GROWTH AND YIELD AS  AFFECTED  BY ACID PRECIPITATION
01
oo
Total deposition
Simulant concentration
g m"z ( i events)
H* S042- N03- H+
Alfalfa,
0.001
0.056

0.177

0.56
Barley,
0.001
0.045
0.142
0.45
'Vernal', Mod 1c ago
0.300 0.417
2.744 0.417

17.35 0.417

54.92 0.417
'Steptoe', Hordeum
0.238 0.335
2.205 0.335
13.945 0.335
44.13 0.335
satlva
0.0025
0.10

0.316

1.00
vulgare
0.0025
0.10
0.316
1.00
mg r1
L. (Cohen et al.
0.53 0.74
4.90 0.74

30.99 0.74

98.07 0.74
L. (Cohen et al.
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74
S042-:N03-
Rate
cm hr-1
Events Droplet Fertilizer
1
hr ,
events"1
size
vm
N-P-K
PH
Effect*
1981, Lee et al. 1981)
0.7
6.6

41.8

132.5
0.67
0.67

0.67

0.67
56
56

56

56
1.5
1.5

1.5

1.5
1200
1200

1200

1200
67-252-252b
67-252-252

67-252-252

67-252-252
5.6
4.0

3.5

3.0
Control
No effect market yield, increased
shoot wt
31% greater market yield, Increased
shoot/root wt
No effect growth or market yield
1981, Lee et al. 1981)
0.7
6.6
41.8
132.5
Beet, 'Detroit Dark Red'. Beta vulgarls L. (Cohen et al. 1981.
0.001
0.026
0.082
0.26

0.140 0.193
1.274 0.193
8.057 0.193
25.50 0.193

Bibb lettuce, 'Limestone',
0.0002
0.009
0.028
0.09

0.048 0.067
0.441 0.067
2.789 0.067
8.826 0.067

Bluegrass, 'Newport', Ppa p
0.002
0.072
0.227
0.720
Broccol 1
0.0006
0.022
0.070
0.22
0.382 0.472
3.528 0.472
22.31 0.472
70.61 0.472
0.0025
0.10
0.316
1.00

Lactuca
0.0025
0.10
0.316
1.00

0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74

satlva L. (Cohen
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74

iratensls L. (Cohen et al
0.0025
0.10
0.316
1.00
, 'Italian Green Sprouting
0.117 0.164
1.078 0.164
6.319 0.164
21.58 0.164
0.0025
0.10
0.316
1.00
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74
0.7
6.6
41.8
132.5

et al . 1981
0.7
6.6
41.8
132.5

. 1981. Lee
0.7
6.6
41.8
132.5
', Brassica oleracea L. var.
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74
0.7
6.6
41.8
132.5
0.67
0.67
0.67
0.67
45
45
45
45
1.5
1.5
1.5
1.5
1200
1200
1200
1200
112-224-224b
112-224-224
112-224-224
112-224-224
5.6
4.0
3.5
3.0
Control
No effect growth or market yield
No effect growth or market yield
No effect growth or market yield
Lee et al. 1981)
0.67
0.67
0.67
0.67

, Lee et al.
0.67
0.67
0.67
0.67

et al. 1981)
0.67
0.67
0.67
0.67
Botrytls L.
0.67
0.67
0.67
0.67
26
26
26
26

1981)
9
9
9
9


72
72
72
72
(Cohen
22
22
22
22
1.5
1.5
1.5
1.5


1.5
1.5
1.5
1.5


1.5
1.5
1.5
1.5
et al. 1981
1.5
1.5
1.5
1.5
1200
1200
1200
1200


1200
1200
1200
1200


1200
1200
1200
1200
, Lee
1200
1200
1200
1200
112-224-2240
112-224-224
112-224-224
112-224-224


112-224-2240
112-224-224
112-224-224
112-224-224


224-448-443b
224-448-448
224-448-448
224-448-448
et al. 1981)
168-224-2240
168-224-224
168-224-224
168-224-224
5.6
4.0
3.5
3.0


5.6
4.0
3.5
3.0


5.6
4.0
3.5
3.0

5.6
4.0
3.5
3.0
Control
No effect growth or market yield
No effect growth or market yield
43% decrease market yield; decrease
root/shoot growth

Control
No effect growth or market yield
No effect growth or market yield
No effect growth or yield; decrease
root growth

Control
No effect market yield or growth
No effect market yield or growth
No effect market yield or growth

Control
No effect market yield or growth
No effect market yield or growth
25% lower market yield
      'Effects are reported when statistical
      OFerttltzer as kg ha"1 of N-P»05-K20.
      fertilizer as percentage of N-PjOs-l^f
significance Is < 0.05 level.

-------
                                                           TABLE  3-3.    CONTINUED
CO

-fi
Total deposition Simulant concentration
g m-z ( r events)
H* 304?- NOa- H*
mg l-1
S042- HOa- $04*':NO
Bush bean, 'Blue Lake 274', Phaseolus vulgarls L.
0.000004
0.00017

0.000004
0.00017
0.00083


Cabbage,
0.001
0.051
0.067
0.51
Carrot,
0.001
0.044
0.139

0.44

0.001 0.001 0.0025
0.008 0.001 0.10

0.001 0.001 0.0025
0.007 0.001 0.10
0.041 0.001 0.631


0.60
5.33

0.60
5.33
30.70


'Golden Acre', Brassfca oleracea
0.270 0.379 0.0025
2.499 0.379 0.10
15.80 0.379 0.316
50.02 0.379 1.00
0.53
4.90
30.99
98.07
0.83
0.70

0.83
0.70
0.75


L. var.
0.74
0.74
0.74
0.74
Rate
cm hr-1
3"

1
Events
hr ,
events"1
Droplet Fertilizer
size N-P-K
ym
PH
Effect*
(Johnston et al. 1982)
0.7
7.6

.7
7.6
40.9


Capltata L.
0.7
6.6
41.8
132.5
'Danvers Half Long', Daucus carota L. var. Sativa DC
0.230 0.327 0.0025
2.156 0.327 0.10
13.636 0.327 0.316

43.15 0.327 1.00

0.53
4.90
30.99

98.07

-0.74
0.74
0.74

0.74

Cauliflower, 'Early Snowball', Brasslca oleracea
0.0006
0.023
0.073
0.23
0.122 0.171 0.0025
1.127 0.171 0.10
7.128 0.171 0.316
22.56 0.171 1.00
Corn, 'Golden Midget', Zea mays L.
0.0005
0.020
0.063
0.20
Fescue,
0.001
0.059

0.186
0.59

0.11 0.149 0.0025
0.980 0.149 0.10
6.198 0.149 0.316
19.61 0.149 1.00
'Alta', Festuca elatlor L.
0.31 0.439 0.0025
2.891 0.439 0.10

18.20 0.439 0.316
57.86 0.439 1.00

0.53
4.90
30.99
98.07
(Cohen
0.53
4.90
30.99
98.07
0.74
0.74
0.74
0.74
et al.
0.74
0.74
0.74
0.74
0.7
6.6
41.8

132.5

L. var. Botr
0.7
6.6
41.8
132.5
1981, Lee et
0.7
6.6
41.8
132.5
.64
.64

.64
.64
.64


(Cohen et al
0.67
0.67
0.67
0.67
(Cohen et al.
0.67
0.67
0.67

0.67

18
18

16
16
16


. 1981,
51
51
51
51
1981,
44
44
44

44

ytls L. (Cohen et al
0.67
0.67
0.67
0.67
al. 1981)
0.67
0.67
0.67
0.67
var. arundlnacea Schreb. (Cohen et al.
0.53
4.90

30.99
98.07

0.74
0.74

0.74
0.74

0.7
6.6

41.8
132.5

0.67
0.67

0.67
0.67

23
23
23
23

20
20
20
20
0.67
0.67

0.67
0.67
0.67


Lee et al
1.5
1.5
1.5
1.5
Lee et al.
1.5
1.5
1.5

1.5

900
900

900
900
900


. 1981)
1200
1200
1200
1200
1981)
1200
1200
1200

1200

. 1981, Lee et al.
1.5
1.5
1.5
1.5

1.5
1.5
1.5
1.5
1981, Lee et al.
59
59

59
59

1.5
1.5

1.5
1.5

1200
1200
1200
1200

1200
1200
1200
1200
1981)
1200
1200

1200
1200

0-20-OC
0-20-0

0-20-OC
0-20-0
0-20-0



224-224-224b
224-224-224
224-224-224
224-224-224

224-224-224b
224-224-224
224-224-224

224-224-224

1981)
224-224-224D
224-224-224
224-224-224
224-224-224

168-336-336b
168-336-336
168-336-336
168-336-336

168-336-336l>
168-336-336

168-336-336
168-336-336

5.6
4.0

5.6
4.0
3.2



5.6
4.0
3.5
3.0

5.6
4.0
3.5

3.0


5.6
4.0
3.5
3.0

5.6
4.0
3.5
3.0

5.6
4.0

3.5
3.0

Control
No effect yield; older leaves age
-------
                                                         TABLE 3-3.    CONTINUED
Total deposition
g m~z ( i events)
H* S042- N03-
Simulant concentration
H*
Green pea, "Marvel '. PI sum satlvum
0.001
0.028
0.088
0.28
0.150 0.208
1.372 0.208
8.677 0.208
27.46 0.208
Green pepper, 'California
0.001
0.038
0.128

0.380

0.20 0.283
1.86 0.283
11.78 0.283

37.27 0.283

Kidney bean. 'Red Kidney'.
0.0004
Y* 0.029
Oi
CJi 0.095

0.105

0.107

0.134

0.200

0.229

Lettuce.
0.00096
0.03024
0.03024

0.03024

Mustard
0.0004
0.014
0.044
0.14
0.007 0.043
2.274 0.043

7.564 0.043

8.32 0.043

9.831 0.043

10.59 0.043

15.88 0.043

18.14 0.043

0.0025
0.10
0.316
1.00
Wonder' ,
0.0025
0.10
0.316

1.00

mg fl
S042- N03-
S042-:N03-
Rate
cm hi-1
Events
1
Droplet
hr , size
events*1 u»
Fertilizer
N-P-K
PH
Effect*
L. (Cohen et al, 1981, Lee et al. 1981)
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74
Capsicum annuum L.
0.53 0.74
4.90 0.74
30.99 0.74

98.07 0.74

0.7
6.6
41.8
132.5
(Cohen et
0.7
6.6
41.8

132.5

0.67
0.67
0.67
0.67
al. 1981.
0.67
0.67
0.67

0.67

28
28
28
28
Lee et al
38
38
38

38

1.5
1.5
1.5
l.S
. 1981)
1.5
1.5
1.5

1.5

1200
1200
1200
1200

1200
1200
1200

1200

67-224-2240
67-224-224
67-224-224
67-224-224

224-448-448*
224-448-448
224-448-448

224-448-448

5.6
4.0
3.5
3.0

5.6
4.0
3.5

3.0

Control
No effect growth or market yield
No effect growth or market yield
No effect growth or market yield

Control
No effect market yield or growth
20% greater market yield; Increased
shoot growth
No effect market yield; decreased
shoot growth
Phaseolus vulgaMs 1. (Shrlner 1978a)
0.001
*

*

*

*

*

*

0.631

'Oakland'. Lactuca satlva
0.02304 0.02976
0.13824 1.7856
0.78384 0.95232

1.38336 0.11904

green, 'So. Giant
0.074 0.104
0.687 0.104
4.339 0.104
13.73 0.104
0.002
0.63
0.63

0.63

Curled.'
0.0025
0.10
0.316
1.00
0.02 0.12
0.02/50 0.12

0.02/50 0.12

0.02/50 0.12

0.02/50 0.12

0.02/50 . 0.12

0.02/50 0.12

50 0.12

0.2/416
0.2/416

0.2/416

0.2/416

0.2/416

0.2/416

0.2/416

416.67

3.0
3.0

3.0

3.0

3.0

3.0

3.0

3.0

24
24

24

24

24

24

24

24

0.17
0.17

0.17

0.17

0.17

0.17

0.17

0.17

900
900

900

900

900

900

900

900


6.0/3.

3.2/6.

6.0/6.

3.2/3.

6.0/3.

3.2/6.



6.0
2/6 .Od

0/6.0

0/3.2

2/6.0

2/3.2

,0/3.2

3.2

Control
751 Increased pod number; greater
shoot and root wt
SOT lower pod number; greater
shoot wt
501 lower pod number; greater
shoot wt
No effect pod numher; lower shoot/
root wt
751 greater pod number; greater
root wt
No effect pod number; lower shoot/
root wt
50% greater pod number; lower
shoot wt; greater root wt
L. (Jacobson et al. 1980)
0.48 0.62
2.88 37.20
16.33 19.84

28.82 2.48

Brasslca japonlca
0.53 0.74
4.90 0.74
30.99 0.74
98.07 0.74
0.77
0.08
0.82

11.6

O.RO
0.80
0.80

0.80

Hort. (Cohen et al .
0.7
6.6
41.8
132.5
0.67
0.67
0.67
0.67
3
3
3

3

1981, Lee
14
14
14
14
2.0
2.0
2.0

2.0

et al.
1.5
1.5
1.5
1.5
900
900
900

900

1981)
1200
1200
1200
1200
Half-strength 5.7
Hoaglands





112-224-2240
112-224-224
112-224-224
112-224-224
3.2
3.2

3.2


5.6
4.0
3.5
3.0
Control
No effect growth or yield
7% Increase root wt; 24% Increase
apical leaf wt
10% Increase root wt; 29% Increase
apical leaf wt

Control
14% lower market yield
No significant effect
31% lower market yield
*0.001/0.631.
'Effects are reported when statistical significance Is £ 0.05 level.

dpH sequence Is:  10 events prior to'Halo blight Infection/3 events during Infection period/11 events post Infection.

-------
                                                           TABLE  3-3.   CONTINUED
CO

9l
Total deposition
g «rz ( i events)
H* »42- H03-
Simulant concentration
mgt-
1
N03-
»,'-,
Rate
cm hr-
Events
1 » hr i
events"1
Droplet Fertilizer pH
size N-P-K
Effect*
Oats. 'Cay use', Avena satlva L. (Cohen et al. 1981, Lee et al. 1981)
0.001
0.048
0.152

0.48
Onion, '
0.002
0.065
0.205
0.65

0.254 0.357
2.354 0.357
14.87 0.357

47.07 0.357
Sweet Spanish'.
0.34 0.484
3.185 0.484
20.14 0.484
63.75 0.484

Orcnardgrass, 'Potomac'
0.001
0.035

0.111
0.35

0.19 0.260
1.715 0.260

10.85 0.260
34.32 0.260

0.0025 0.53
0.10 4.90
0.316 30.99

1.00 98.07
0.74
0.74
0.74

0.74
Allllum cepa L. (Cohen et al.
0.0025 0.53
0.10 4.90
0.316 30.99
1.00 98.07

, Dactyl Is glomerata
0.0025 0.53
0.10 4.90

0.316 30.99
1.00 98.07

0.74
0.74
0.74
0.74

L. (Cohen
0.74
0.74

0.74
0.74

Pinto bean, 'Univ. Idaho 111'. Phaseolus vulgar Is L.
0.003
1.355
2.149
3.405
5.396
Potato,
m

0.164

0.52
Radish.
0.0003
0.012
0.033
0.12

8.533 1.365
64.033 1.365
102.16 1.365
162.49 1.365
258.16 1.365
0.002 5.0
0.794 37.52
1.259 59.86
1.995 95.21
3.162 151.27
'White Rose', Solanum tuberosu* L.
litt 8-J887-

16.11 0.387

51.00 0.387
'Cherry Belle',
0.064 0.089
0.588 0.089
3.719 0.089
11.77 0.089

Red clover, 'Kenland' ,
0.001
0.056
0.177
0.56
0.300 0.417
2.744 0.417
17.35 0.417
54.92 0.417
8:?825 l-M

0.316 30.99

1.00 98.07
Raphanus satlvus L.
0.0025 0.53
0.10 4.90
0.316 30.99
1.00 98.07

Trl folium pratense L
0.0025 0.53
0.10 4.90
0.316 30.99
1.00 98.07
0.80
0.80
0.80
0.80
0.80
(Cohen et
Sift

0.74

0.74
0.7
6.6
41.8

132.5
1981,
0.7
6.6
41.8
132.5

et al
0.7
6.6

41.8
132.5

(Evans
6.2
46.9
74.8
119.0
189.1
0.67
0.67
0.67

0.67
Lee et al.
0.67
0.67
0.67
0.67

. 1981, Lee
0.67
0.67

0.67
0.67

and Lewln
0.72
0.72
0.72
0.72
0.72
al. 1981, Lee et
u

41.8

132.5
8:l77

0.67

0.67
48
48
48

48
1981)
65
65
65
65

et al. 1981)
35
35

35
35

1981)
45
45
45
45
45
al. 1981)
if

52

52
1.5
1.5
1.5

1.5

1.5
1.5
1.5
1.5


1.5
1.5

1.5
1.5


0.33
0.33
0.33
0.33
0.33

hi

1.5

1.5
1200
1200
1200

1200

1200
1200
1200
1200


1200
1200

1200
1200


ass
353
353
353
353

ifoo

1200

1200
112-224-2240
112-224-224
112-224-224

112-224-224

336-336-3360
336-336-336
336-336-336
336-336-336


112-224-2240
112-224-224

112-224-224
112-224-224


manure and
1 Imestone
added



247-224-2240
224-224-224

224-224-224

224-224-224
5.6
4.0
3.5

3.0

5.6
4.0
3.5
3.0


5.6
4.0

3.5
3.0


5.7
3.1
2.9
2.7
2.5

5.6
4.0

3.5

3.0
Control
No effect market yield or growth
No effect market yield; Increased
root growth
No effect market yield or growth

Control
No effect market yield or growth.
No effect market yield or growth
No effect market yield; Increased
shoot growth

Control
No effect market yield; decreased
root growth
No effect market yield or growth
231 greater market yield; Increased
root growth

Control
No effect yield
281 lower seed yield
29% lower seed yield
39% lower seed yield

Control
No effect yield; Increased shoot
growth
lit greater market yield; Increased
shoot growth
8% lower market yield
(Cohen et al. 1981, Lee et al. 1981)
0.74
0.74
0.74
0.74

0.7
6.6
41.8
132.5

. (Cohen et al.
0.74
0.74
0.74
0.74
0.7
6.6
41.8
132.5
0.67
0.67
0.67
0.67

12
12
12
12

1.5
1.5
1.5
1.5

1200
1200
1200
1200

112-224-224»
112-224-224
112-224-224
112-224-224

5.6
4.0
3.5
3.0

Control
No effect growth or market yield
Lower market yield
Lower market yield; decreased shoot
growth
1981, Lee et al. 1981)
0.67
0.67
0.67
0.67
56
56
56
56
1.5
1.5
1.5
1.5
1200
1200
1200
1200
67-336-3360
67-336-336
67-336-336
67-336-336
5.6
4.0
3.5
3.0
Control
No effect growth or market yield
No effect growth or market yield
No effect growth or market yield
      •Effects are reported when statistical  significance Is £0.05 level.
      '"Fertilizer as kg ha-1 of N-P

-------
                                                         TABLE  3-3.   CONTINUED
co
Total
g m-2
H*
deposition Simulant concentration
( i events) mg i-l
5042- N03- H+ S042-
Ryegrass, "L1nn", Loll urn 'perenne L.
0.001
0.055
0.183
0.58
Spinach,
0.0004
0.014
0.044
0.14
Soybean,
0.004
1.549

6.169

Soybean ,
0.002
0.002
0.002
0.105

0.105

0.105

0.700

0.700

0.700

0.31 0.432 0.0025
2.842 0.432 0.10
17.97 0.432 0.316
56.88 0.432 1.00
' Improved
0.074
0.687
4.339
13.73
'Amsoy 71
9.755
73.20

295.12

'Wells',
0.669
0.980
1.113
3.738

2.485

5.600

20.55

25.52

29.40

c
N03- S042":N03~
(Cohen et al. 1981, Lee et al
0.53
4.90
30.99
98.07
0.74'
0.74
0.74
0.74
Thick Leaf, Sp1nac1a oleracea L.
0.104 0.0025
0.104 0.10
0.104 0.316
0.104 1.00
', Glydne max (L
1.561 0.002
1.561 0.794

1.561 3.162

Glydne max (L.)
0.721 0.0025
0.490 0.0025
0.371 0.0025
3.731 0.15

5.530 0.15

3.619 0.15

18.55 1.0

12.82 1.0

9.800 1.0

0.53
4.90
30.99
98.07
.) Merr.
5.0
37.52

151.27

0.74
0.74
0.74
0.74
(Evans et
0.80
0.80

0.80

Merr. (Irving and
0.96
1.40
1.6
5.34

7.09

8.00

29.36

36.45

42.00

Strawberry, 'Qulnalt', Fragarla chlloensls
0.002
0.080

0.253

0.800

0.42
3.920

24.79

78.46

0.595 0.0025
0.595 0.10

0.595 0.316

0.595 1.00

0.53
4.90

30.99

98.07

1.03
0.70
0.53
5.33

3.55

5.17

26.50

18.32

14.00

0.7
6.6
41.8
132.5
Rate
:m hr-1
. 1981)
0.67
0.67
0.67
0.67
Events Droplet Fertilizer
1

58
58
58
58
hr .
events"1

1.5
1.5
1.5
1.5
size

1200
1200
1200
1200
N-P-K

112-224-224b
112-224-224
112-224-224
112-224-224
pH Effect*
•

5.6 Control
4.0 No effect market yield; decreased
root growth
3.5 No effect market yield; decreased
root growth
3.0 No effect market yield; decrease!
root growth
(Cohen et al. 1981, Lee et al. 1981)
0.7
6.6
41.8
132.5
al. 19810)
6.2
46.9

189.1

0.67
0.67
0.67
0.67

0.72
0.72

0.72

14
14
14
14

78
78

78

1.5
1.5
1.5
1.5

0.17
0.17

0.17

1200
1200
1200
1200

353
353

353

112-224-224°
112-224-224
112-224-224
112-224-224

manure and
1 Imestone

added

5.6 Control
4.0 No effect growth or yield
3.5 No effect'growth or yield
3.0 No effect growth or yield

5.7 Control
3.1 11% greater seed yield; decreased
shoot growth
2.5 11% lower seed yield; decreased
shoot growth
Sowlnskl 1980)
1.0
2.0
3.0
1.0

2.0

1.5

1.0

2.0

3.0

21.2
21.2
21.2
21.2

21.2

21.2

21.2

21.2

21.2

Duchesne var. ananassa (Cohen
0.74
0.74

0.74

0.74

.7
6.6

41.8

132.5

0.67
0.67

0.67

0.67

10
10
10
10

10

10

10

10

10

et al.
80
80

80

80

0.33
0.33
0.33
0.33

0.33

0.33

0.33

0.33

0.33

1981, Lee et
1.5
1.5

1.5

1.5

2300
2300
2300
2300

2300

2300

2300

2300

2300

al.
1200
1200

1200

1200

15-30-15C
15-30-15
15-30-15
15-30-15

15-30-15

15-30-15

15-30-15C

15-30-15

15-30-15

1981)
224-336-3360
224-336-336

224-336-336

224-336-336

5.6 1:1 S04:N03; control
5.6 2:1 S04:N03; control
5.6 3:1 S04:N03; control
3.8 No effect growth or yield compared
to 1:1 control
3.8 No effect growth or yield compared
to 2:1 control
3.8 Lower root nodule wt compared to 3:1
control; no effect yield
3.0 No effect growth or yield compared
to control
3.0 No effect growth or yield compared
to control
3.0 25% greater yield than 1:1 control,
19% greater than 2:1 control

5.6 Control
4.0 51% greater market yield; Increased
shoot growth
3.5 72. greater market yield; Increased
shoot/root growth
3.0 72% greater market yield; Increased
shoot/root growth
     'Effects are  reported when statistical significance Is £0.05 level.
     fertilizer as kg ha"1 of N-P

-------
                                                   TABLE 3-3.   CONTINUED
Total deposition Simulant concentration
g m-2 ( £ events) mg
H+ S042- M03- H* S042-
l-l
N03-
Swiss chard, 'Lucullus', Beta vulgar Is var. clcla
0.001
0.032
0.101
0.32
Timothy
0.001
0.033
0.104
0.33
V Tobacco
en
CD o.OOl
0.024
0.076
0.24
Tom to,
0.001
0.051

0.161

0.510

Wheat,
0.001
0.046

0.145

0.46
0,17
1.568
9.92
31.38
. 'Climax'
0.17
1.617
10.23
32.36
, 'Burley

0.127
1.176
7.438
23.537
'Patio',
0.27
2.50

15.80

50.02

0.238 0.0025 0.53
0.238 0.10 4.90
0.238 0.316 30.99
0.238 1.00 98.07
0.74
0.74
0.74
0.74
, Phleum pratense L. (Cohen et al
0.256 0.0025 0.53
0.256 0.10 4.90
0.256 0.316 30.99
0.256 1.00 98.07
21' . Nlcotlana tabacun L

0.179 0.0025 0.53
0.179 0.10 4.90
0.179 0.316 30.99
0.179 1.00 98.07
Lycoperslcon esculentum
0.379 0.0025 0.53
0.379 0.10 4.90

0.379 0.316 30.99

0.379 1.00 98.07

0.74
0.74
0.74
0.74
. (Cohen

0.74
0.74
0.74
0.74
S042":MO
Rate
cm hr-1
3~
L. (Cohen et al. 1981, Lee
0.7
6.6
4V. 8
132.5
. 1981, Lee
0.7
6.6
41.8
132.5
et al. 1981

0.7
6.6
41.8
132.5
Mill. (Cohen et al.
0.74
0.74

0.74

0.74

'F1el
-------
 general,  total deposition of H+, $042-, and N0s~ applied was
 greater 1n the controlled environment studies than In the field studies
 because of a higher deposition rate or greater number of exposures.

     There were 34 crop varieties (28 species) studied in
 controlled-environment experiments; six exhibited a negative response
 from acid rain exposure (pinto bean, mustard green, broccoli, radish,
 beet and carrot), eight exhibited a positive response (alfalfa, tomato,
 green pepper, strawberry, corn, orchard grass, timothy, and 'Oakland
 lettuce1), 17 showed no effect (bush bean, 'Wells' soybean, spinach,
 'Limestone' lettuce, cabbage, cauliflower, onion, fescue, bluegrass,
 ryegrass, swiss chard, oats, wheat, barley, tobacco, green pea, and  red
 clover), and three species showed both positive and negative yield
 response depending on the H ion concentration (potato, 'Amsoy 71'
 soybean), or conditions of exposure (kidney bean).

 3.4.2.3  Discuss1on--Interpreting and comparing results of experiments
 on the effects of acidic deposition on crop plants must include
 considering the exposure conditions, simulant characteristics, dose
 rate, and total  dose of important ions (H+, S042~, and N03")«
 Unexplained inconsistencies among experimental results could be due  to
 differences 1n experimental  design or exposure conditions.  For example,
 in all field studies except those of 'Champion*  radish and 'Beeson*  and
 'Williams1 soybeans, the ratio of sulfate to nitrate in the rain
 simulant differed among treatments and was usually much higher than  the
 sulfaternitrate ratio in ambient rain.  Rain chemistry data from the
 National Atmospheric Deposition Program (NADP) Indicate that weekly
 precipitation pH values can vary widely for a particular area (i.e., an
 range of pH 3.7 to 6.8 for New York) while the 50*2- to N03~
 ration appears to be Independent of pH (Figure 3-6).  Because
 preliminary evidence Indicates that plants are affected by the
 sulfate:nitrate ratio 1n rain (Irving and Sowlnski 1980, Lee et al.
 1980), the differences reported among treatments in these Investigations
 may be the result of this ratio rather than the hydrogen 1on deposition.
 All published experiments used treatments having the same chemistry  from
 event to event although the chemistry of ambient rain can fluctuate
 greatly from one event to another (Figure 3-6).   Some crops may be
 affected by peak concentrations of acidity while others may respond  to
 the total deposition of ions.  No experiments separating the peak  versus
 total loading response have yet been reported, although Irving et  al.
 (1981) found that rain with a chemistry that varied from event to  event
 had a different effect on plant growth than did a constant rain
 chemistry with the same mean pH.

     The majority of the 14 crop cultivars studied in the field and  the
 34 studied in controlled environments exhibited no effect on growth  or
yield as a result of exposure to simulated rain more acidic (usually up
 to 10 times more acidic) than ambient.  The growth and yield of some
 crops, however,  were negatively affected by acidic rain while others
 exhibited a positive response.  The 9 percent reduction in the yield of
 field corn exposed to pH 4.0 rain (0.594 kg ha-1 depositon of H+)  1s
 an alarming result; however, treatments with greater acidity levels


                                  3-59

-------
   13


   12 -


   11 -


   10


    9


    8
 co  7
o   '
%r  6
 to

     5


     4


     3


     2


     1
                 N
                       N
                             0
                                                                     N
  N  N 0  0 0   0        N
         N     0    N      N N
   OP 0 0 ONOPNN    N
    NPPOPNN  POO  P N         °
N PONONOOOO NNPN OOON
 P   PON    N PP       P
 N   ON        P        N  N
   N   P
     N

      I	I	1	
                                                SYMBOL  IS LETTER OF STATE
                                                	i	I	
     3;5
              4.0
              4.5
5.0
5.5
6.0
6.5
                                                                        7.0
                                      pH
   Figure 3-6.  Ratio of
                              to  N0o~  versus  pH of precipitation in New
               York  (N),  Pennsylvania  (P), and Ohio  (0) during the
               growing  season.   Data are  from the NADP network,  1979.
                                   3-60

-------
produced no effect on the corn yield.   The experiment was repeated the
following year with similar (although  not statistically  significant)
results and for a third year with no negative effects observed (J.  J.
Lee, pers. comm.).  The reduction in the yield of one ("Amsoy")  of the
five cultivars of soybeans that have been studied suggests that genetic
or soil factors or the presence of heavy metals 1n the rain simulant may
control plant response to acidic rain.  If these possibilities are
substantiated, ramifications of the negative effects of  acid rain
observed in the above two studies could be considerable  since soybeans
and field corn are two of the most economically important agricultural
crops 1n the United States.  For reasons discussed in this review,
however, these studies do not offer definitive proof that ambient acidic
precipitation is damaging corn and soybean production.

     The positive response of some crops to acidic rain  suggests a
fertilizer response to the sulfur and  nitrogen components of the rain.
The net response of a plant to acid rain appears to result from the
interaction between the positive effects of sulfur and nitrogen
nutrition and the negative effects of acidity.  Input of nutrients to
plant systems from rainfall has been documented since the mid-19th
century (Way 1855).  Calculations made in a number of regions in the
United States estimate the seasonal  atmospheric deposition of nutrient
species, particularly sulfate and nitrate, to agricultural  and natural
systems and the Implications of this deposition on plant nutrient
status.

     Estimates by Hoeft et al. (1972)  of 30 kg S ha-1 per year and 20
kg N ha'1 per year deposited in precipitation in Wisconsin indicated
the importance of atmospheric sources of these elements, although N
requirements certainly could not be completely satisfied in this way.
Jones et al. (1979) reported that atmospheric S is a major contribution
to the agronomic and horticultural crop needs for S as a plant nutrient
in South Carolina.  Although the amount of S and N in a  single rain
event is small compared to a fertilizer application, it  is known that
foliar applications of plant nutrients may stimulate plant growth and
yield (Garcia and Hanway 1976).  The repeated exposure of plants to
rain, especially during the critical reproductive stage, suggests that
nutritional benefits from rain may be significant, even  in comparison  to
a one-time fertilizer application.

     Reports of most acid rain field studies contain little or no
characterization of the soil conditions.  Soil fertility may determine
whether a plant responds positively or negatively to acidic
precipitation.  Long-term effects of acidic deposition on poorly
managed, unamended agricultural soils  may have negative  effects on crop
productivity through the leaching of soil nutrients or mobilization of
toxic metals.  This effect has more potential for becoming significant
in those soils with low cation exchange capacity (low in clay and
organic matter), low sulfate retention capacity, and high permeability
(sands).  Although such an effect may  not become measurable for decades
or more, it will be most Important in forage crops that  are not usually
highly managed.  Some speculation exists that agricultural  management
practices may be modified as a result of acidic deposition but

                                  3-61

-------
agricultural soil scientists generally  accept that the  Influence of
acidic deposition on the need for additional  fertilizer and Hrne
application 1s probably mlnlscule.

     Another consideration that may be  Important 1n controlling the
Impact of ambient add precipitation, 1s  that crop cultlvar
recommendations are based on productivities obtained under ambient
conditions of acidic deposition.   Therefore,  crops currently being grown
may have been selected, Indirectly, for their adaptations to rainfall
acidity and the presence of other pollutants.

3.4.2.4  Summary--

     1)  Because of limitations In research design, differences In
         methodologies and Inconsistent results, 1t 1s  difficult to
         compare research results directly or arrive at an overall
         conclusion regarding crop response to acidic deposition without
         a thorough description and comparison of experimental methods.

     2)  Complex factorial research designs and multlvarlate analyses
         may be necessary to describe adequately the relationship
         between acid rain dose and plant response rather than the
         simple unlvarlate approach (treatment pH vs yield) used In  the
         past.

     3)  Given the above limitations to making generalizations about
         past research, analysis of experimental results  from  field  and
         controlled environment experiments Indicates that the majority
         of crop species exhibited no effect on growth  or yield as a
         result of exposure to simulated  acidic rain (acidity  treatments
         had pH values of 4.2 or less). Growth and yield  of a  few crops
         1n some studies, however, were negatively affected by acidic
         rain, while other crops exhibited a positive response.

     4)  Interpretation of available research results suggests that  the
         net response of a crop to acidic deposition 1s the result of
         the Interaction between the positive effects of  sulfur and
         nitrogen fertilization, the negative effects of  acidity, and
         the Interaction between these  factors and other  environmental
         conditions such as soil type and presence of other pollutants.
         Available experimental results appear to Indicate  that the
         effects of acidic precipitation  on crops are minimal  and that
         when a response occurs 1t may  be positive or negative.
         However, many crops and agricultural systems have  not been
         adequately studied.

3.5  CONCLUSIONS

     Chapter E-3 has examined vegetative  response to acidic deposition,
reviewing literature from studies that  shed light on divers
plant-pollutant relationships.  Documented experiments  concern widely
varying situations, from controlled-envlronment studies to  field


                                  3-62

-------
studies, and from intensively managed agricultural  systems  to natural
plant communities.  Control1ed-env1ronment studies  are  useful Indicators
of potential effects and may suggest subtle changes not easily
measurable in an uncontrolled situation.   Field  studies,  however, are a
more realistic means of estimating actual  effects because in these
studies experimental plants are grown under normal  agricultural
conditions.

     The following statements summarize Chapter  E-3:

     0   Leaf structure may play two roles 1n the sensitivity of foliar
         tissues to acidic deposition:  1)  leaf morphology may
         selectively enhance or minimize  surface retention  of incident
         precipitation, and 2)  specific cells of the epidermal surface
         may be Initial sites of foliar injury.  Information on the
         effects of acidic deposition on  the accelerated  weathering of
         epicuticular wax of plant leaves Is very preliminary.
         Chlorophyll degradation may occur following prolonged exposure
         to acidic precipitation (Section 3.2.1).

     0   Leaching mechanisms are major factors in nutrient  cycling in
         terrestrial ecosystems and are critical to the redistribution
         of nutrients within these cycles.  If the  rate of  leaching
         exceeds the rate of mineral  nutrient uptake, plant growth and
         yield reductions are likely (Section 3.2.1.).

     0   Under laboratory conditions, gaseous pollutant combinations and
         Integration have well  defined effects.  However, ozone is the
         single most important gas pollutant to  plant life  located at
         great distances from the* Industrial  and urban  origin of
         nitrogen oxides and hydrocarbon  precursors.  Direct effects due
         to ozone Include foliar Injury and growth  and  yield reductions
         in numerous agronomic and forest species (Section  3.3.1).

     0   A review of the evidence on the  interaction of forest trees,
         Insect and microbial  pests,  and  acidic  deposition  does not
         allow generalized statements concerning stimulation or
         restriction of biotic  stress agents,  or their  activities, by
         acidic deposition.   Certain studies report stimulation of pest
         activities associated  with  acidic deposition treatment, while
         other studies report restriction  of pest activities following
         treatment.  Further research must combine  field  and controlled
         environment studies.   Available  evidence suggests  that the
         threshold of ambient pH capable of Influencing certain Insect
         and microbial pests lies within  the range  of pH  3.0 to 4.0
         (Sections 3.3.2, 3.3.3, and  3.3.4).

     »   Performance and longevity (persistence) of certain pesticides
         depend on the pH of the systems  to which these pesticides are
         applied or in which  they ultimately reside; thus,  it is likely
         that acidic deposition will  have  significant but limited
         effects (Section 3.3.5).
                                 3-63

-------
At present we have no direct evidence that acidic  deposition
currently limits forest growth in either North America or
Europe, but we do have indications that tree  growth  reductions
are occurring, principally in coniferous species that have been
examined to date, that these reductions are rather widespread,
and that they occur in regions where rainfall acidity is
generally quite high, or pH Is low (- pH 4.3) for  an annual
average (Section 3.4.1).

Controlled environment studies Indicate that  the deposition of
acidic and acidifying substances may have stimulatory,
detrimental, or no apparent effects on pi ant  growth  and
development.  Response depends upon species sensitivity, plant
life cycle stage, and the nature of exposure  to acidity.  Some
simulation studies have Indicated that acidic deposition may
result in simultaneous stimulation of growth  and the occurrence
of visible foliar Injury (Section 3.4.1).

The majority of crop species studies in field and  controlled-
environment experiments exhibited no effect on growth or yield
as a result of exposure to simulated acidic precipitation (pH
3.0).  In a few studies, though, growth and yield  of certain
crops were negatively affected by acidic deposition, while
others exhibited positive responses (Section  3.4.2).

A crop's net response to acidic deposition results from a
combination of the positive effects of sulfur and  nitrogen
fertilization, the negative effects of acidity, and  the
interaction between these factors and other environmental
conditions such as soil type and presence of  other pollutants
(Section 3.4.2).

Available experimental results do not appear  to Indicate  that
the negative effects of acidic precipitation  outweight  the
positive effects, however, many crops and agricultural  systems
have not been properly or adequately studied  (Section 3.4.2).
                         3-64

-------
3.6  REFERENCES

Abougendia,  Z. M. and R.  Redman.   1979.  Germination and early seedling
growth of four conifers on  acidic  and alkaline substrates.  Forest
Science 25:350-360.

Abrahamsen,  G.  1980.  Acid precipitation  effects on forest and fish.  .In
Ecological  Impact of Acid Precipitation.   D. Drablos, and A. Toll an, eds.
Proc. of an  International Conference, The  Norwegian Interdisciplinary
Research Programme Acid Precipitation-Effects on forest and Fish.
Sandfjord, Norway, March  11-14, 1980.

Abrahamsen,  G.  1980.  Acid precipitation, plant nutrients, and forest
growth, pp.  58-63.  In Ecological  Impact of Acid Precipitation.  D.
Drablj6s and  A. TollaTHT eds.  Proceedings of an International
Conference,  Sandefjord, Norway.

Abrahamsen,  G. and G. J.  Dollard.   1979.   Effects of acidic precipitation
on forest vegetation and  soil.  In Ecological Effects of Acid
Precipitation.  G. Howells, ed. TPRI SOA77-403.  Electric Power Research
Institute,  La Jolla, CA.

Abrahamsen,  G., K. Bjor,  R. Horntvedt, and B. Tveite.  1976.  Effects of
acid precipitation on coniferous forest,   pp. 37-63.  In Impact of Acid
Precipitation on Forest and Fresh  Water Ecosystems in "Norway.  F. H.
Braekke, ed.  Research Report 6, SNSF-Project.

Abrahamsen,  G., H. Horntvedt, and  Bjorn Tveite.  1977.  Impacts of acid
precipitation on coniferous forest ecosystems.  Water, Air, and  Soil
Pollut. 8:57-73.

Abrahamsen,  G., A. Stuanes, and K. Bjor.   1979.  Interaction between
simulated rain and barren rock  surface.  Water, Air, and Soil Pollut.
11:191-200.

Adam, N. K.   1948.  Principles  of  penetration of liquids into solids.
Discuss. Faraday Soc. 3:5-11.

Alexander,  V.  1974.  A synthesis  of the IBP tundra blome drcumpolar
study of nitrogen fixation, pp. 109-121.   Ir^ Soil Organisms and
Decomposition in Tundra.  A. J. Holding, 0. W. Heal, S. F. Maclean, and
P. W. Flanagan, eds.  Tundra Biome Steering Committee, Stockholm.

Alstad, D.  N., G. F. Edmunds, Jr., and L.  H. Weinstein.  1982.  Effects
of air pollutants on insect populations.   Ann. Rev. Entomol. 27:369-384.

Armstrong,  D. E., G. Chesters,  and R. F. Harris.  1967.  Atrazine
hydrolysis  in soil.   Soil Sci.  Soc. Am. Proc. 31:61-66.

Ascaso, C. and J. Galvin.   1976.   Studies  on the pedogenic action of
lichen acids.  Pedobiologia 16:321-331.
                                  3-65

-------
Ashenden, T. W. and T.  A.  Mansfield.  1978.  Extreme pollution
sensitivity of grasses  when SO? and N02 are present in the atmosphere
together.  Nature 273:142-143.

Babich, H. and G. Stotzky. 1974.  Air pollution and microbial ecology.
Crit. Rev. Environ. Cont.  4:353-420.

Baddeley, M. S., B. W.  Ferry, and E. J. Finnegan.  1971.  A new method of
measuring lichen respiration:  Response of selected species to
temperature, pH, and sulphur dioxide.  Lichenologist 5:18-25.

Barkman, J. J.  1958.   Phytosociology and Ecology of Cryptogamic
Epiphytes.  Assen: The  Netherlands.

Bauch, J.  1983.  Biological alterations in the stem and root of fir and
spruce due to pollution influences.  j£ Workshop on the Effects of
Accumulation of Air Pollutants in Forest Ecosystems.  D. Reidel Co.
Hingham, Mass, (in press).

Becker, V. E.  1980. Nitrogen fixing lichens in forests of the southern
Appalachian mountains of North Carolina.  The Bryologist 83:29-39.

Becker, V. E., J. Reeder,  and R. Stetler.  1977.  Biomass and habitat of
nitrogen-fixing lichens in an oak forest in the North Carolina piedmont.
Bryologist 80:93-99.

Benzian, B.  1965.  Forestry Commission Bulletin.  London 1(37):1-251.

Best, J. A. and J. B. Weber.  1974.  Disappearance of sL-triaz1nes as
affected by soil pH using a balance-sheet approach.  Weed Sci.
22:364-373.

Borstrom, F. and G. R.  Hendrey.  1976.  pH tolerance of the first larva
stages of Lepidurus arcticys  (Pallas) and adult Gammarus lacustris.  G.
0. Sars. Report, Zoological Museum, Oslo Univ., Oslo, Norway, 37 pp.

Boullard, B.  1973.  Interactions entre les pollutants
atmospheriques et certains parasites des essences forestiees.
(Champignons et insects).  For.  Prive e 94:31,33,35,36.

Bragg, R. J.  1982.  Interactions of Schleroderrfs canker disease of red
pine with acid rain. M.S. Thesis.  State University of New York, College
of Environmental Science and  Forestry, Syracuse, NY.  61 pp.

Brandt, C. S. and W. W. Heck.   1968.  Effects of air pollutants on
vegetation.   In Air Pollution.   A.  C. Stern, ed.  2nd ed. Vol  I.  pp
401-442.  Academic Press, New York.

Brodo,  I. M.  1974.  Substrate  ecology, pp. 401-441.  In The Lichens.  V.
Ahmadj1an and M. E. Hale, eds.   Academic Press, New Yoflf.
                                  3-66

-------
Bruck, R. I., S.  R.  Shafer,  and A. S. Heagle.  1981.  Effects of
simulated acid rain  on  the development of fusiform rust on loblolly pine.
Phytopathology 71:864.

Charles, P.  J. and C. Vlllemant.  1977.  Modifications des nivereaux de
population d'lnsectes dans les jeunes plantations de pins sylvestres de
la foret de Roumare  (Seine-Maritime) soumlses a  la pollution
atmospherlque. C. R. Acad.  Agrlc. Fr., 63:502-510.

Chou, C. H.  and C. C. Young.  1974.  Effects of osmotic concentration and
pH on plant growth.  Talwanla 19:157-165.

Cogblll, C.  V.  1976.   The history and character of add precipitation In
eastern North America,  pp. 363-370.  In Proceedings of the First
International Symposium on Add Precl'pTtatlon and the Forst Ecosystem.
L. S. DocMnger and  T.  A. Sellga, eds.  USDA Forst Service General
Technical Report  NE-23, Upper Darby, PA.

Cohen, C. J., L.  C.  Grothaus and S. C. Perrlgan.  1981.  Effects of
simulated sulfurlc acid rain on crop plants.  Agricultural Experiment
Station, Oregon State University, Special Report 619.

Cott, H. B.   1940.  Adaptive Coloration In Animals.  Methuen & Co. Ltd.,
London.

Cowling, E.  B.  1978.   Effects of add precipitation and atmospheric
deposition on terrestrial vegetation, pp. 47-63.  J^n A National Program
for Assessing the Problem of Atmospheric Deposition (Acid Rain).
Galloway et al.,  eds.   A Report to the Council on Environmental Quality,

Crafts, A. S.  1961a.   The Chemistry and Mode of Action of Herbicides.
Intersdence, New York.

Crafts, A. S.  1961b.   Translocatlon 1n Plants.  Holt, Rlnehart and
Winston, New York.  182 pp.

Crafts, A. S. and H. G. Relber.   1945.  Studies of the activation of
herbicides.   HUgardla  16:487-500.

Cronan, C. S.  1978.  Solution chemistry of a New Hampshire subalplne
ecosystem: blogeochemlcal patterns and processes.  Ph.D. thesis,
Dartmouth College.

Cronan, C. S.  1983.  Blogeochemlcal responses of forest canopies to add
precipitation, jji The  Effects of Acidic Deposition on Vegetation.  Am.
Chem. Soc. Meeting,  Los Vegas, NV.  Ann Arbor Science, Ann Arbor, MI. (In
press).

Dahl, J. B., C. Qvenlld, 0.  Toll an, N. ChHstophersen, and H. M. Selp.
1979.  Methodology of studies of chemical processes 1n water runoff from
rock and shallow  soil  cover  using radioactive tracers.  Water, A1r, and
Soil Pollut. 11:179-190.
                                  3-67

-------
Dahlsten, D. L. and D.  L.  Rowney.   1980.   Influence of air pollution on
population dynamics of  forest insects and  on tree mortality, pp. 125-130.
In Effects of Air Pollutants on Mediterranean and Temperate Forest
"Ecosystems.  P. R. Miller, ed.   U.S.D.A. Forest Service, Genl. Tech. Rep.
No. PSW-43, Berkeley, CA.

Davis, R. B.  1966.  Spruce-fir forests of the coast of Maine.  Ecol.
Monographs 36:79-94.

Deans, J. D.  1979.  Fluctuations  of the soil environment and fine root
growth in a young Sitka spruce plantation.  Plant Soil 52:195-208.

Denison, R., B. Caldwell,  B. Bormann, L. Eldred, C. Swanberg, and S.
Anderson.  1977.  The effects of acid rain on nitrogen fixation in
western Washington coniferous forests.  Water, Air, and Soil Pollut.
8:21-34.

Denison, W. C.  1973.  Life in tall  trees.  Sci. Amer. 228(6):74-80.

Drouet, F.  1937.  The  Brazilian Myxophyceae I.  Amer. Jour. Bot.
24:598-608.

Dybing, C. D. and H. B. Currier.  1961.  Foliar penetration by chemicals.
PI. Physiol. 36:169-174.

Eaton, J. S., G. E. Likens, and F. H. Bormann.  1973.  Throughfall and
stemflow chemistry in a northern hardwood  forest.  J. Ecology
61:495-508.

Edwards, C. A.  1972.  Insecticides', pp. 513-568.  In Organic Chemicals
in the Soil Environment, C. A. Goring and  J. W. HamaTer, eds.  Marcel
Dekker, Inc., New York.

Electric Power Research Institute  (EPRI).   1981.  Computer output of data
from  sampling stations.  Personal  communication from C. S. Hakkarinen to
L. L. Sigal, Oak Ridge National Laboratory.

Evans, L. S.  1982.  Biological effects of acidity on vegetation:  A
review.  Environ, and Expt. Bot. 22:155-169.

Evans, L. S. and T. M. Curry.  1979. Differential responses of plant
foliage to  simulated acid rain.  Amer. J.  Bot. 66:953-962.

Evans, L. S. and K. F.  Lewin.  1981. Growth, development, and yield
responses of pinto beans and soybeans to  hydrogen ion concentrations of
simulated acidic rain.   Environmental and  Experimental Botany 21:103-114.

Evans, L. S., G. R. Hendrey, G. J. Stensland, D. W. Johnson, and A. J.
Francis.  1981b.  Acid precipitation:   Considerations for an air quality
standard.   Water, Air, and Soil Pollut. 16:469-509.
                                  3-68

-------
Evans, L. S., K.  F.  Lewin, and E. A. Cunningham.  1982.  Effects of
simualted acid rain  on yields of field-grown radishes and garden beets.
Agriculture and the  Environment 7:285-298.

Evans, L. S., N.  F.  Gmur, and F. DaCosta.  1977a.  Leaf surface and
hlstologlcal  perturbations of leaves of Phaseolus vulgaris and He!1anthus
annuus after exposure to simulated acid rain.A¥. J. Bot. 64:903-913.

Evans, L. S., N.  F.  Gmur and J. J. Kelsch.  1977b.  Perturbations of
upper leaf surface structures by acid rain.  Environ. Exp. Bot.
17:145-149.

Evans, L. S., N.  F.  Gmur and F. DaCosta.  1978.  Foliar response of six
clones of hybrid  poplar  to simulated acid rain.  Phytopathology
68:847-856.

Evans, L. S,, T.  M.  Curry, and K. Lewin.  1981a.  Responses of leaves of
Phaseolus vulgaris L. to simulated acid rain.  New Phytol. 88:403-420.

Evans, L. S., K.  F.  Lewin and E. A. Cunningham.  1981c.  Effects of
simulated acid rain  on yields of field-grown radishes and garden beets.
Prepr., Am. Chem. Soc.,  Div. Environ. Chem. 21:82-87.

Evans, L. S., K.  F.  Lewin, C. A. Conway and M. J. Patti.  1981d.  Seed
yields (quantity  and quality) of field-grown soybeans exposed to
simulated acidic  rain.   New Phytol. 89:459-470.

Eversman, S.  1978.   Effects of low-level S02 on Usnea hirta and
Parmelia chorochroa. Bryologist 81:368-377.

Ewald, P. W.  1982.   Hummingbirds:  The nectar connection.  National
Geographic 161:222-227.

Fairfax, J. A. W. and N. W. Lepp.  1975.  Effect of simulated 'Acid Rain1
on cation loss from  leaves.  Nature 255:324-325.

Feller, M. C. 1977.   Nutrient movement through western hemlock-western
redceder ecosystems  in  southwestern British Columbia. Ecology
58:1269-1283.

Ferenbaugh, R. W.  1976. Effects of simulated acid rain on Phaseolus
vulgaris L. (Fabaceae).   Amer. J. Bot. 63:293-288.

Ferguson, P. and  J.  A.  Lee.  1979.  The effects of bisulphite and sulfate
upon photosynthesis  In  Sphagnum.  New Phytologist 82:703-712.

Ferguson, P., J.  A.  Lee, and J. N. B. Bell.  1978.  Effects of sulfur
pollutants on the growth of Sphagnum species.  Environmental Pollution
16:151-162.

Ferry, B. W., M.  S.  Baddeley and D. L. Hawksworth, eds.  1973.  Air
Pollution and Lichens.   University of London Athlone Press, London.
                                  3-69

-------
Forman, R. T. T.  1975.   Canopy lichens with blue-green algae:  A
nitrogen source 1n a Columbian rain forest.  Ecology 56:1176-1184.

Fowler, D., J. N.  Cape,  I. A. Nicholson, J. W. K1nna1rd, and I. S.
Paterson.  1980.  The Influence of a polluted atmosphere on cuticle
degradation In Scots pine (Plnus  sylyestrls).  In Ecological Impact of
Acid Precipitation.  D.  Drablls and A. Toll an, e3s.  Proceedings of an
International Conference, Sandefjord, Norway.

Frledland, A. J.,  A. H.  Johnson and T. G. Slccama.  1983.  Accumulation
of trace metals 1n the forest floor In Vermont:  Spatial and temporal
patterns.  Water A1r and Soil Pollution. (1n press).

Fuzzl, S.  1978.  Study  of Iron (III) catalysed sulfur dioxide oxidation
1n aqueous solution over a wide range of pH.  Atmos. Envlr.
12:1439-1442.

Garcia, R. L. and J. J.  Hanway.   1976.  Foliar fertilization of soybeans
during the seed-filling period.   AGron. J. 68:653-657.

Gerson, U. and M. R. D.  Seaward.  1977.  Lichen-Invertebrate
associations,  pp. 69-119.  In M. R. D. Seaward (ed.), Lichen Ecology.
Academic Press, New York. GlTEert, 0. L.  1968.  Bryophytes as Indicators
of air pollution In the Tyne Valley.  New Phytologlst 67:15-30.

Gilbert, 0. L.  1968.  Bryophytes as Indicators of air pollution 1n the
Tyne Valley.  New Phytologlst 67:15-30.

Gilbert, 0. L.  1970.  Further  studies on the effect of sulphur dioxide
on lichens and bryophytes.  New Phytologlst 69:605-627.

Gilbert, 0. L.  1974.  Lichens and air pollution, pp. 443-472.  jji The
Lichens.  V. Ahmadjlan and M. E.  Hale, eds.  Academic Press, New York.

Glass, N. R.  1979.  Environmental effects of Increased coal utilization:
Ecological effects of gaseous emissions from coal combustion.  Environ.
Health Perspec. 33:249-272.

Goettsche, D.  1972.  Vertellung  von Felnwurzeln und Mykorrhlzen 1m
Bodenprofll elnes Buchenund  Flchtenbestandes 1m Soiling.  D1ss. Univ.
Hamburg.

Gravenhorst, G., S. Belike,  M.  Betz, and H. Georgll.  1978.  Sulfur
dioxide absorbed 1n rain water, pp. 41-45.  In Effects of Acid
Precipitation on Terrestrial Ecosystems, T. TJ7 Hutchlnson,  and M. Havas,
eds.  Plenum Press, New York.

Grennfelt, P., C. Bengtsong, and  L.  Skarby.  1980.  An estimation of the
atmosphere Input of acidifying  substances to a forest ecosystem,  pp.
29-40.   In T. C. Hutchinson  and M. Havas, Eds.  Effects of  Acid
PrecipifaTion on Terrestrial Ecosystems.  Plenum Press, New York.
                                  3-70

-------
Grodzinska, K.  1979.  Tree bark  -  sensitive  biotest for environment
acidification.  Environ. Intern.  2:173-176.

Grzywacz, A. and J. Wazny.   1973.  The  impact of  industrial air
pollutants on the occurrence of several  important pathogenic fungi of
forest trees in Poland.  Eur. J.  For. Path. 3:129-141.

Guderian, R.  1977.  Air Pollution.  Springer-Verlag, New York.

Gustafson, F. G.  1956.  Absorption of  Co60 by leaves of young plants
and its trans!ocation through the plant.   Am. J.  Bot. 43:157-160.

Gustafson, F. G.  1957.  Comparative absorption of cobalt-60 by upper and
lower epidermis of leaves.   PI. Physiol.  32:141-142.

Hadfield, J. S.  1968.  Evaluation  of diseases of red spruce on the
Chamberlin Hill sale, Rochester Ranger  District,  Green Mt. Nat. Forest.
Unpub. Manuscript no. A-68-8 5230.   U.S.D.A.  Forest Service Amherst Field
Office, Amherst, MA.

Hagvar, S., G. Abrahamsen,  and A. Bokhe.   1976.   Attack by the pine bud
moth in southern Norway:  Possible  effect of  acid precipitation.  For.
Abstr. 37:394.

Haines, B., M. Stefani, and F. Hendricks.  1980.   Acid Rain:  Threshold
of leaf damage in eight plant species from a  southern Appalachian forest
succession.  Water, Air Soil Pollut.  14:403-407.

Hallgren, J. E. and K. Huss.  1975.  Effects  of SO? on photosynthesis
and nitrogen fixation.  Physiologia Plantarium 34:171-176.

Hamaker, J. W.  1972.  Decomposition, quantitative aspects, pp. 253-340.
Jhi Organic Chemicals in the Soil  Environment.  C. A. Goring and J. W
Hamaker, eds.  Marcel Dekker, Inc., New York.

Haughbotn, 0.  1973.  Neb0runders0kelsor i Sarpsborgdistriktet og
underjs"kelser over virkninger av forsurende nedfall pa jordas kjemiske
egenskaper. As-NLH, 1973.  151 pp.  (referenced in Horntvedt, 1975).

Hawksworth, D. L.  1971.  Lichens as litmus for air pollution:  a
historical review.  International Journal  of  Environmental Studies
1:281-296.

Hawksworth, D. L. and F. Rose.  1976.   Lichens as Pollution Monitors.
Studies in Biology No. 66.   Edward  Arnold Publishers, London.

Hay, C. J.  1975.  Arthropod stress,  pp.  33-34.   Jj^ Air Pollution and
Metropolitan Woody Vegetation. W.  H. Smith and L. S. Dochinger, eds.
U.S.D.A. Forest Service, Publ. No.  PLEFR-PA-1, Upper Darby, Penn.

Heagle, A. S.  1973.  Interactions  between air pollutants and plant
parasites.  Annu. Rev. Phytopathol  11:365-388.
                                  3-71

-------
Heagle, A. S., R.  B.  Philbeck, P. G. Feicht, and R. E. Ferrell.  1983.
Response of soybeans  to  simulated acid rain in the field.  J. Environ.
Qual. submitted.

Heggestad, H.  E. and  W.  W. Heck. 1971.  Nature, extent, and variation of
plant response to  air pollutants.  Adv. in Agron. 23:111-145.

Heinrichs, H., and R. Mayer.  1977. Distribution and cycling of major and
trace elements in  two central European forest ecosystems.  J. Envir.
Qual. 6:402-407.

Hemphill, D. D. and H. B. Tukey, Jr.  1973.  The effect of intermittant
mist on absisic acid  content of Euonymus alatus Sieb. 'Compactus'.  J.
Amer. Soc. Hort. Sci. 98:416-420.

Henderson, R.  E. et al.  1981.  An overview of atmospheric processes and
modeling of acid rain.   Final Report prepared for U.S. Dept. of Energy
under contract 80ET13800-TD14 by the MITRE Corp., McLean, VA.

Hill, A. C.  1971. Vegetation:  A sink for atmospheric pollutants.
Jour. Air Poll. Control  Assoc. 21:341-346.

Hill, D. J.  1971. Experimental study of the effect of sulphite on
lichens with reference to atmospheric pollution.  New Phytologist
70:831-836.

Hindawi, I. J., J. A. Rea, and W. L. Giffis.  1980.  Response of bush
bean exposed to acid  mist.  Am. J. Bot. 67:168-172.

Hoeft, R. G.,  D.  R. Keeney, and L. M. Walsh.  1972.  Nitrogen and sulfur
in precipitation  and  sulfur dioxide in the atmosphere in Wisconsin.  J.
Environ. Qual. 1:203-208.

Hoffman, W. A., Jr.,  S.  E. Lindberg, and R. R. Turner.  1980.  Some
observations of organic  constituents in rain above and below the forest
canopy.  Environ.  Sci. Technol. 14:95-100.

Houston, D. B. and L. S. Dochinger.  1977.  Effects of ambient air
pollution on cone, seed, and pollen characteristics in eastern white and
red pines.  Environmental Pollution 12:1-5.

Huffaker, C. B.  1974.  Some implications of plant-arthropod and
higher-level,  arthropod-arthropod  food links.  Environ. Entomol. 3:1-9.

Hull, H. M.  1974.  Leaf structure as related to penetration of organic
substances, pp. 45-93.  ^Absorption and Translocation of Organic
Substances in Plants. J. Hacskaylo, ed.  7th Ann. Symposium, Southern
Sec. Am. Soc.  PI.  Physiologists, Emory Univ., GA.  April 16.

Ibsen, H.  1866.   "Brand", Act V.
                                  3-72

-------
Inglis, F. and D.  J.  Hill.   1974.   The effect of sulphite and fluoride on
carbon dioxide uptake by mosses  in  the light.  New Phytologist
73:1207-1213.

Irving, P. M.  1979.   Response of field-grown soybeans to acid
precipitaion along and in combination with sulfur dioxide.  Ph.D.
Dissertation, University of Wisconsin, Milwaukee.

Irving, P. M. and D.  A. Sowinski.   1980.  Effects of H+, $04%
N03- and Nfy- concentrations and ratios in precipitation applied
to greenhouse-grown soybeans.  Argonne National Laboratory, Radiological
and Environmental  Research  Division Annual Report, ANL-80-115, Part III,
pp. 6-10.

Irving, P. M. and J.  E. Miller.  1980.  Response of field-grown soybeans
to acid precipitation alone and  in  combination with sulfur dioxide.
Proc. Int. Conf.  Ecol. Impact Acid  Precip., D. Drablos and A. Toll an
(eds.), SNDF Project, As-NHL,  Norway, pp. 170-171.

Irving, P. M. and J.  E. Miller.  1981.  Productivity of field-grown
soybeans exposed to acid rain and sulfur dioxide alone and in
combination.  J.  Environ. Qua!.  10(4):473-478.

Irving, P. M., C.  M.  Gradert,  and J. A. Maley.  1982.  Effect of
variations in rainfall chemistry on the productivity of bushbeans and
radishes.  Agronne National Laboratory, Radiological and Environmental
Research Division Annual Report, January-December 1981.

Jaakhola, S., H.  Katainen,  S.  Kellomaki, and P. Saukkola.  1980.  The
effcts of artificial  acid rain on the spectral reflectance and
photosynthesis of Scots pine seedlings, pp. 172-173.  In Ecological
Impact of Acid Precipitation.  D. Drabljls and A. Tollan, eds.
Proceedings of an International  Conference, Sandefjord, Norway.

Jacobson, J.  1980.  The influence  of rainfall composition on the yield
and quality of agricultural crops,  pp. 41-46.  In Ecological Impact of
Acid Precipitation.  D. Drablj6s  and A. Tollan, "eels.  Proceedings of an
International Conference, Sandefjord, Norway.

Jacobson, J. S.,  J. Troiano, L.  J.  Colavito, L. I. Heller, and D. C.
McCune.  1980.  Polluted rain and plant growth.  Jji Polluted Rain.  T. Y.
Toribara, M. W. Miller and  P.  E. Morrow, eds.  Plenum Pub. Corp., N. Y.

Jeffree, C. E., E. A. Baker, and P. J. Halloway.  1975.  Ultrastructure
and recrystallization of plant epicuticular waxes.  New Phytologist
75:539-541.

Johnson, A. H. and T. G. Siccama.   1983.  Spruce decline in the northern
Appalachians:  Evaluating acid deposition as a possible cause.  Ir[ Proc.
of Ann. Meeting of Tech. Assessment of Pulp and Paper Industry.  March,
Atlanta, GA.
                                  3-73

-------
Johnson, A. H., T. 6.  Siccama,  R.  S.  Turner, and D. G. Lord.  1983.
Assessing the possibility of a  link between acid precipitation and
decreased growth rates of trees in the  northeastern U.S.  J^n The Effects
of Acidic Deposition on Vegetation.   Am. Chem. Soc. Meeting, Los Vegas,
NV.  Ann Arbor Science, Ann Arbor, MI.  (In press).

Johnson, A. H., T. 6.  Siccama,  D.  Wang, R. S. Turner, and T. H.
Barringer.  1981.  Recent changes  in  patterns of tree growth rate in the
New Jersey pinelands:   A possible  effect of acid rain.  J. Environ. Qual.
10(4):427-430.

Johnson, D. W. 1975.  Processes of elemental transfer in some tropical,
temperate, alpine and northern  forest soils: factors influencing the
availability and mobility of major leaching agents.  Ph.D. thesis,
University of Washington, Seattle.  169 p.

Johnson, P. C.  1950.   Entomological  aspects of the ponderosa pine blight
study, Spokane, Washington.  Unpubl.  Report, U.S.D.A. Bur. Entomol. and
Plant Quar., Forest Insect Laboratory,  Coeur d" Alene, Idaho  15 pp.

Johnson, P. C.  1969.   Atmospheric pollution and com'ferophagous
invertebrates.  Proc.  20th Ann. Western For. Insect Work Conf., Coeur
d1Alene, ID.

Johnston, J. W., D. S. Shriner, C. I. Klarer, and D. M. Lodge.  1982.
Effect of rain pH on senescence, growth, and yield of bush bean.
Environ. Exp. Bot. 22:329-337.

Jones, F. G. W., R. A. Dunning, and K.  P. Humphries.  1955.  The effects
of defoliation and loss of stand upon yield of sugar beets.  Ann. Appl.
Biol. 43:63-70.

Jones, U. S., M. G. Hamilton, and  J.  B. Pitner.  1979.  Atmospheric
sulfur as related to fertility  of  ultisols and entisols in South
Carolina.  Soil Sci. Soc. Am. J. 43:1169-1171.

Jonsson, B.  1975.  Soil acidification  by atmospheric pollution and
forest growth, pp. 837-842.  Jji Proceed! ngs of the First International
Symposium on Acid Precipitation and the Forest Ecosystem.  L. S.
Dochinger and T. A. Seliga, eds.  Department of Agriculture, Washington,
D.C.

Jonsson, B. and R. Sundberg.  1972.   Has the acidification by atmospheric
pollution caused a growth reduction in  Swedish forests?  A comparison of
growth between regions with different soil properties.  Dept. of Forest
Yield Research. 'Research Notes -  NR20.  48 pp.

Jonsson, B. and L. Svensson.  1983.   Studi over luftfororeningarnas
inverkan pa skogeprodu.  Jni uppofoljning och revision av Jonsson och
Sundberg rapport 1972.  Secrages Lantbruks Universitat Nist fur Biometre
och Skogsindeling rapport.
                                  3-74

-------
Jordan, C.F., F. Golley,  J.  Hall  and J. Hall.  1980.  Nutrient scavenging
of rainfall  by the canopy of an Amazonian rain forest.  Biotropica.  (In
press).

Jordan, L. S., W. J.  Fanner, J. R. Goodin, and B. E. Day.  1972.
Volatilization and non-biological degradation of triazine herbicides in
vitro and in soils.  Res. Rev. 32:236-237.

Juniper, B.  E. and D. E.  Bradley.  1958.  The carbon replica technique In
the study of the ultra-structure  of leaf surfaces.  J. ultrastructure
Res. 2:16.

Karnosky, D. F. and G. R. Stairs.  1974.  The effects of SOo on In
vitro forest tree pollen  germination and tube elongation.  JournaT of
Environmental Quality 3:406-409.

Kelly, J.M.   1979.  Camp  Branch and Cross Creek experimental watershed
projects: objectives, facilities and ecological characteristics.
EPA-600/7-79-053, Washington, D.C.

Kettlewell,  H. B. D.   1973.  The  Evolution of Melanism.  Clarendon Press,
Oxford.

Kozel, P. C. and H. B. Tukey, Jr.  1968.  Loss of glbberelUns by
leaching from stems and foliage of Chrysanthemum morlfoliurn 'Princess
Ann.1  Amer. J. Bot.  55:1184-1189.

Kratky, B. A., E. T.  Fukunaga, J. W. Hylin, and R. T. Nakano.  1974.
Volcanic air pollution:  Deleterious effects on tomatoes.  J. Environ.
Qua!. 3:138-140.

Lacy, G. H., B. I. Chevone,  and N. P. Cannon.  1981.  Effects of
simulated acidic precipitation on Erwinia herbicola and Pseudomonas
syrlngae populations.  Phytopathology /i:»»a.

Ladlie, J. S., W. F.  Meggltt, and D. Penner.  1976.  Effect of soil pH on
mlcrobial degradation, adsorption, and mobility of metrlbuzln.  Weed Sci.
24:477-481.

Laurence, J. A.  1981. Effects of air pollutants on plant-pathogen
Interactions.  J. PI. D1s. Protect. 87:156-172.

Lawrey, J. D. and M.  E. Hale, Jr.  1981.  Retrospective study of lichen
lead accumulation in  the  northeastern United States.  The Bryologist
84:449.456.

Le Blanc, F. and D. N. Rao.  1975.  Effects of air pollutants on lichens
and bryophytes, pp. 237-272.  J!n_  Responses of Plants to Air Pollution.
J. B. Mudd and T. T.  Kozlowski, eds.  Academic Press, New York.
                                 3-75

-------
Lebrun, P.  1976.   Effects  e  cologiques de la pollution
atmospherique sur  les  populations et communautes de
microarthropodes corticoles (Acariens, Collemboles, et Pte rygotes).
Bull. Soc. Ecol.  7:417-430.

Lechowfcz, M. J.  1981.   The  effects of simulated acid precipitation  on
photosynthesis in  the  caribou lichen Cladina Stellaris (Opiz.) Brodo.
Water, Air, and Soil Pollution 14:133-lb/.

Lee, C.I. and H.  B. Tukey, Jr.  1972.  Effect of intermittent mist on
development of fall  color in  foliage of Euonymus alatus Sieb.
'Compactus'.  J. Amer.  Soc. Hort. Sci. 9/:9/-iui.

Lee, J. J. and 6.  E. Neely.  1980.  CERL-OSU acid rain crop study
progress report.  Air  Pollution Effects Branch, Corvallis Environmental
Research Laboratory.

Lee, J. J. and D.  E. Weber.   1979.  The effect of simulated acid rain on
seedling emergence and growth of eleven woody species.  Forest Science
25:393-398.

Lee, J. J., G. E.  Neely and S. C. Perrigan.  1980<  Sulfuric acid rain
effects on crop yield  and foliar injury.  U.S. Environ. Prot. Agency,
Publica. No. EPA-600/3-80-316, Corvallis, OR.  21 pp.

Lee, J. J., 6. E.  Neely,  S. C. Perrigan, and L. C. Grothaus.  1981.
Effect of simulated sulfuric  acid rain on yield, growth, and foliar
injury of several  crops.  Environ. Exp. Bot. 21:171-185.

Leonard, 0. A.  1958.   Studies on the absorption and translocation of
2,4-D in bean plants.   Hilgardia 28:115-160.

Leung, S., W. Reed,  S.  Cauchois and R. Howitt.  1978.  Methodologies  for
valuation of agricultural crop yield changes:  A review.  EPA-600/5-78-
018.  Corvallis Environmental  Research Laboratory, Corvallis, OR.

Likens, G. E.  1976.   Acid  rain.  Chemical and Engineering News
54:29-44.

Likens, G. E. and  T. J.  Butler.  1981.  Recent acidification of
precipitation in North America.  Atmos. Environ. 15(7):1103-1109.

Lindberg, S.E., R.C. Harriss,  R.R. Turner, D.S. Shriner, and D.D. Huff.
1979.  Mechanisms  and  rates of atmospheric deposition of selected trace
elements and to a  deciduous forest watershed.  Environmental Sciences
Division Pub!. no. 1299,  ORNL-TM/6674, Oak Ridge National Laboratory, Oak
Ridge, Tennessee.

Lindberg, S. E. and R.  C. Harriss.  1981.  The role of atmospheric
deposition in element  cycling in an eastern United States deciduous
forests.  Water, Air,  and Soil Pollut. 16:13-31.
                                 3-76

-------
Unskens, H. F.  1950.  Quantitative  bestlmmung der benetzbarkelt von
blattoberflachen.  Planta 38:591-600.

Llnskens, H. F.  1952.  Uber die  Anderung der benetzbakeH von
blattoberflachen und deren ursache.   Planta 41:40-51.

Llano, G. A.  1948.  Economic uses of lichens.  Econ. Bot. 2:15-45.

Lord, D. 6.  1982.  Root and Foliar Composition of Declining and Healthy
Red Spruce In Vermont and New Hampshire.  M. S. Thesis, U. Pennsylvania,
Philadelphia.  287 p.

Lovett, G. M., W. A. Relners, and R.  K. Olson.  1982.  Cloud droplet
deposition 1n subalplne balsam fir forests:  Hydrologlcal and chemical
Inputs.  Science 218:1303-1305.

Madgwlck, H. A. I. and J. D.  Ovlngton.  1959.  The chemical composition
of precipitation In adjacent forest and open plots.  Forestry 32:14-22.

Malhotra, S. S.  1976.  Effects of S02 on biochemical activity and
ultrastructural organization of pine  needle chloroplasts.  New
Phytologlst 76:239-245.

Manlon, P. D.  1981.  Tree Disease Concepts.  Prentice-Hall, Englewood
CUffs, New Jersey.  399p.

Manlon, P. D. and R. J. Bragg. 1982.  Effects of add precipitation on
Scleroderrfs canker disease of red pine, pp. 55-56.  Proc. N. Y. State
Symposium on Atmopsherlc Deposition.  Center for Environmental Research,
Cornell University, Ithaca, New York.

Manning, W. J.  1975.  Interactions between air pollutants and fungal,
bacterial and viral plant pathogens.  Environ. Pollut. 9:87-90.

Marsh, J. E. and T. H. Nash III.  1979.  Lichens In relation to Four
Corners power plant In New Mexico.  Bryologlst 82:20-28.

Martin, J. T. and B. E. Juniper.  1970.  The Cultlcles of Plants.  St.
Martins Press, New York.  247 pp.

Mattson, W. J. and N. D. Addy. 1975.  Phytophagous Insects as regulators
of forest primary production.  Science 190:515-522.

Matzlrls, D. E. and G. Nakos.  1977.  Effect of simulated acid rain on
juvenile characteristics of Aleppo pine (Plnus halepensls Mill.). Forest
Ecology and Management 1:267-272.

McFarlane, J. C. and W. L. Berry.  1974.  Cation penetration through
Isolated leaf cuticles (Apricot).  Plant Physlol. 53(5):723-727.
                                 3-77

-------
Mclaughlin, S. B.  1981.   SOg, vegetation effects and the air quality
standard:   Limits of interpretation and application, pp. 62-83.  In Proc.
of Proposed SOX and Particulate Standard, Atlanta, 6A, September TF-18.

Miller, P. R. and J. R. McBride.  1975.  Effects of air pollutants on
forests,  pp. 195-235.  lr± Response of Plants to Air Pollution, J. B.
Mudd and T. T. Kozlowski,  eds., Academic Press Inc., NY.

Miller, P. R., R. N. Kickert, 0. C. Taylor, R. J. Arkley, E. W. Cobb, D.
Z. Dahlsteln, P. J. Gersper, R. F. Huck, J. R. McBride, J. R. Parmeter,
J. R. Wentz, M. White and  W. W. Wilcox.  1977.  Photochemical oxldant air
pollution effects on a mixed conifer forest ecosystem.  EPA 600/3-77-104.
U.S. Environmental Protection Agency, Corvallis, OR.  339 p.

Miller, R. B.  1963.  Plant nutrients 1n hard beech III.  The cycle of
nutrients.  N.Z.J. Sci.  6:388-413.

Mollitor, A. V. and D. J.  Raynal.  1982.  Acid precipltaion and Ionic
movements in Adirondack forest soils.  Soil Sci. Soc. Am. J. 36:137-141.

Moser, T. J., T. H. Nash III, and W. D. Clark.  1980.  Effects of a
long-term field sulfur dioxide fumigation on Arctic caribou forage
lichens.  Can. Jour. Bot.  58:2235-2240.

Mueller, P. L., G. M. Hidy, K. Warren, T. F. Lawrey, and R. L. Baskett.
1980.  The occurrence of atmospheric aerosols In the northeastern United
States.  Annals of the New York Academy of Sciences 338:463-482.

Myttenaere, C., C. Daoust  and P. Roucoux.  1980.  Leaching of technetlum
from foliage by simulated  rain.  Environ. Exp. Bot. 20:415-419.

Nash, E. H.  1970.  The effects of effluents from a zinc smelter on
mosses.  Ph.D. thesis, Rutgers State University of New Jersey.

Nash, T. H. III.  1971.  Lichen sensitivity of hydrogen fluroide.  Bull.
Torrey Bot. Club 98:103-106.

Nash, T. J. III.  1973.  Sensitivity of lichens to sulfur dioxide.
Bryologist 75:333-339.

Nash, T. H. III.  1975.  Influence of effluents from a zinc factory on
lichens.  Ecol. Monogr. 45:183-198.

Nash, T. H. III.  1976.  Lichens as indicators of air pollution.
Naturwissenschaften 63:364-367.

Nash, T. H. Ill and E. H.  Nash.  1974.  Sensitivity of mosses to sulfur
dioxide.  Oecologia 17:257-263.

Nash, T. H., Ill and L. L. Sigal.  1979.-  Gross photosynthetic response
of lichens to short-term ozone fumigations.  The Bryologist 82:280-285.
                                  3-78

-------
Nash, T. T., Ill and L.  L.  Sigal.   1980.  Sensitivity of lichens to air
pollution with emphasis  on  oxidant  air pollutants.  j£ Proceedings of the
Symposium on Effects of  Air Pollutants on Mediterranean and Temperate
Forest Ecosystems.   Gen. Tech.  Rep. PSW-43.  Pacific Southwest Forest and
Range Exp. Stn, Forest Serv., U. S. Dept. Agric., Berkeley, CA.

Nichols, G. E.  1918. The  vegetation of northern Cape Breton Island,
Nova Scotia.  Trans. Conn.  Acad. Arts Sci. 22:249-467.

Nieboer, E., H. M.  Ahmed, K. J. Puckett, and D. H. S. Richardsnn.  1972.
Heavy metal content of lichens  in relation to distance from a nickel
smelter in Sudbury, Ontario.  Lichenologist 5:292-304.

Nieboer, E., D. H.  S. Richardson, K. J. Puckett, and F. D. Tomassini.
1976.  The phytotoxicity of sulphur dioxide in relation to measureable
responses in lichens, pp.  61-85.  ^n T. A. Mansfield (ed.), Effects of
Air Pollutants on Plants.  Cambridge University Press, Cambridge.

Nihlgard, B.  1970.  Precipitation, Its chemical composition and effect
on soil water in a  beech and a  spruce forest in South Sweden.  Oikos
21:208-217.

Novaskova, E.  1969.  Influence des pollutions industrielles sur les
communautes animals et 1'utilisation des animaux comme
bio-indicateurs.  Proc.  1st Eur. Congr.  Influence of Air Pollution on
Plants and Animals, Wageningen, 1968.  pp. 41-48.

Costing, H. J. 1956.  The Study of  Plant Communities.  W. H. Freeman Co.,
San Francisco, CA.   440  p.

Orgell, W. H. and R. L.  Weintraub.  1957.  Influence of some ions on
foliar absorption of 2,4-D. Bot. Gaz. 119:88-93.

Overrein, L. N., H. M. Seip, and A. Tollan.  1980.  Acid preciptation
effects on forest and fish. Final  Report, SNSF Project.  175 pp.

Parker, G. G., S. E. Lindberg,  andJ. M. Kelly.  1980.  Atmosphere-
Canopy Interactions of Sulfur in the Southeastern United States, pp.
477-493.  In Atmospheric Sulfur Deposition: Environmental Impact and
Health Effects.  D. S. Shriner, C.  R. Richmond, and S. E. Lindberg, eds.
Ann Arbor Science,  Ann Arbor, MI.

Patterson, D. T.  1975.   Nutrient return in the stemflow and throughfall
in individual trees in the  piedmont deciduous forest, pp. 800-812.  Jji
Mineral Cycling in  Southestern  Ecosystems.  F. G. Howell, J. B. Gentry,
and M. H. Smith, eds. USERDA Conf. 740513.

Penkett, S. A., B.  M. R. Jones  and  A. E. I. Eggleton.  1979.  A study of
S02 oxidation in stored  rain water  samples.  Atmos. Envir. 13:123-137.
                                  3-79

-------
Persson, H.  1980.  Death and replacement of  fine roots in a mature Scots
pine study.  In Structure and Function of Northern Coniferous Forests-An
Ecosystem Study.  T. Persson, ed.  Ecological  Bulletin (Stockholm)
32:251-260.

P1ke, L. H.  1978.  The Importance of epiphytic lichens 1n mineral
cycling.  The Bryologlst 81:247-257.

Plummer, 6. L.  1980.  Observations on lichens, granite rock outcrops and
acid rain.  Georgia J.  Science 38:201-202.

Przybylskl, Z.  1979.  The effects of automobile exhaust gases on the
arthropods of cultivated plants, meadows and  orchards.  Environ. Pollut.
19:157-161.

Puckett, K. J., E. Nleboer, W.  P.  Flora, and  D. H. S. Richardson.  1973.
Sulphur dioxide:  Its effect on photosynthetlc 14C fixation 1n lichens
and suggested mechanisms of phytotoxldty.  New Phytologist 72:141-154.

Puckett, K. J., E. Nieboer, W.  P.  Flora, and  D. H. S. Richardson.  1974.
Photosynthetic 1*C fixation by  the lichen Umbillcaria muhlenbergii
(Ach.) Tuck, following  short exposures to aqueous sulphur dioxide.  New
Phytologist 73:1183-1192.

Puckett, L. J.  1982.  Acid rain,  air pollution, and tree growth 1n
southeastern New York.   J. Environ. Qua!. 11(3) :376-381.

Pyatt, F. B.  1970.  Lichens as indicators of air pollution in a steel
producing town in South Wales.  Environ. Pollution 1:45-56.

Rao, D. N. and F. Le Blanc.  1966. Effects of sulphur dioxide on the
lichen alga, with special reference to chlorophyll.  Bryologlst 69:69-75.

Rao, D. N., 6. Robitaille, and F.  Le Blanc.   1977.  Influence of heavy
metal pollution on lichens and bryophytes.  J. Hattori Bot. Lab.
42:213-239.

Rapp, M.  1973.  Le cycle biogeochimique du soufre dans une foret de
Quercus ilex L. du sud  de la France.  Oecol.  Plant. 8:325-334.

Raynal, D.  1980.  Actual and potential effects of acid precipitation on
a forest ecosystem in the Adirondack Mountains.  ERDA 80-28, State Univ.
of New York.

Raynal, D. J., A. L. Leaf, P. D. Manion and C. J. K. Wang.  1980.  Actual
and potential effects of acid precipitation on a forest ecosystem in the
Adirondack Mountains.  New York State Energy  Research and Development
Authority, Research Report 80-28.

Raynal, D. J., J. R. Roman, and W. M. Eichenlaub.  1982a.  Response of
tree seedlings to acid  precipitation.  I. Effect of substrate acidity on
seed germination.  Environ, and Expt. Bot. 22(3):377-384.


                                  3-80

-------
Raynal, D. J., J. R. Roman,  and W.  M.  Eichenlaub.  1982b.  Response of
tree seedlings to acid precipitation.   II.  Effect of simulated acidified
canopy throughfall on sugar  maple seedling  growth.  Environ, and Expt.
Bot. 22(3) :385.

Rehfuess, K. E.  1981.  Uber die Wirkungen  der sauren Niederschlage in
Waldokosystemen.  Forstwissenschaftliches Centralblatt 100:363-381.

Rehfeuss, K. E., C. Bosch, and E. Pfannkuch.  1982.  Nutrient imbalances
in coniferous stands in southern Germany.   Int. Workshop on Growth
Disturbances of Forest Trees,  lUFRO/FFRJ-Jyvaskyla, Finland, Oct. 10-13.

Rentschler, I.  1973.  Significance of the  wax structure in leaves for
the sensitivity of plants to air pollutants, pp. A139-A142.  ^n
Proceedings of the Third International  Clean Air Congress.  Dusseldorf.
International  Union of Air Pollution Prevention Associations.
ORNL-TR-4120.   Oak Ridge National Laboratory Translation, Oak Ridge, TN.

Rhoades, F. M.  1981.  Biomass of epiphytic lichens and bryophytes on
Abies lasiocarpa on a Mt. Baker lava flow,  Washington.  The Bryologist
84:39-77:

Richardson, D. H. S.  1975.   The Vanishing  Lichens:  Their History and
Importance.  David and Charles, London.

Richardson, D. H. S. and E.  Nieboer.   1981.  Lichens and pollution
monitoring.  Endeavour 5:127-133.

Richardson, D. H. S. and C.  M. Young.   1977.  Lichens and vertebrates,
pp. 121-144.  In Lichen Ecology.  M. R. D.  Seaward, ed.  Academic Press,
New York.     ~

Roberts, B. A. and L. K. Thompson.   1980.   Lichens as indicators of
fluoride emission from a phosphate  plant, Long Harbour, Newfoundland
Canada.  Can.  J. Bot. 58:2218-2228.

Robitaille, G., F. LeBlanc,  and D.  N.  Rao.  1977.  Acid rain:  A factor
contributing to the paucity  of epiphytic cryptogams in the vicinity of a
copper smelter.  Rev. Bryol. Lichenol.  43:53-66.

Roman, J. R. and D. J. Raynal.  1980.   Effects of acid preciptation on
vegetation, pp. 4-1 - 4-63.   In Actual  and  Potential Effects of Acid
Precipitation in the Adironda"cR" Mountains.  New York State Energy
Research and Development Auth.  Report ERDA 80-28.

Rychert, R. C., and J. Skujins. 1974.  Nitrogen fixation by blue-green
algae-lichen crusts in the Great Basin Desert.  Soil Science Society of
America Proceedings 38:768-771.

Sargent, J. A. and G. E. Blackman.   1962.   Studies on foliar penetration.
I.  Factor controlling the entry of 2,4-dichlorophenoxy-acetic acid.  J.
Exp. Bot. 13:348-368.


                                 3-81

-------
Saunders,  P.  J.  W.   1970.  Air pollution in relation to lichens and
fungi.  Lichenologist 4:337-349.

Saunders,  P.  J.  W.   1971.  Modification of the leaf surface and its
environment by pollution, pp. 81-89.  Jto Ecology of Leaf Surface
Microorganisms.   T.  F. Preece and C. H. Dickinson, eds.  Academic Press,
New York.

Saunders,  P.  J.  W.   1973.  Effects of atmospheric pollution on leaf
surface micro-flora. Pestic. Sci. 4:589-595.

Saunders,  P.  J.  W.   1975.  Air pollutants, microorganisms and interaction
phenomena.  Environ. Pollut. 9:85.

Savage, K. E.  1973. Nitralin and trifluralin persistence in soil.  Weed
Sci. 21:285-288.

Schnaider, Z. and Z. Sierpinski.  1967.  Dangerous condition for some
forest tree species  from insects in the industrial region of Silesia.
Prace Instytut Badawczy  Tesnictwa (Warsaw) No. 316, pp. 113-150.

Schnepf, E.  1965.   Licht und elektronenmikroskopische beobachtungen an
des trichom-hydathodes  von Cicer arietinum.  Z. Pflanzenphysiol.
53:245-254.

Schutte, J. A.  1977.  Chromium in two corticolous lichens from Ohio and
West Virginia.  Bryologist 80:279-283.

Scott, B. C. and N.  S.  Laulainen.  1979.  On the concentration of sulfate
in precipitation. J. Applied Meteor. 18:138-147.

Sheridan, R. P. and  R.  Rosenstreter.  1973.  The effect of hydrogen ion
concentrations in simulated  rain on the moss Tortula rural is (Hedw.) Sm.
Bryologi st 76:168-173.

Shields, L. M., C. Mitchell, and F. Drouet.  1957.  Alga- and
lichen-stabilized surface crusts as soil nitrogen sources.  Am. J. Bot.
44:489-498.

Shields, W. J. and  S. D. Hobbs.  1979.  Soil nutrient levels and pH
associated with Arminariella mellea on conifers in northern Idaho.  Can.
J. For. Res. 9:45-48.

Showman, R. E.  1972.  Residual effects of sulfur dioxide on the net
photosynthesis and respiratory  rates of lichen thalli and cultured lichen
symbionts.  Bryologist  75:335-341.

Showman, R. E.  1975.  Lichens  as indicators of air quality around a
coal-fired power generating  plant.  Bryologist 78:1-6.

Shriner, D. S.   1974.   Effects of simulated rain acidified with sulfuric
acid on host-parasite interactions.  Ph.D. Thesis.  North Carolina State
University, Raleigh. 79 pp.


                                 3-82

-------
Shriner, D. S.  1975.   Effects  of  simulated rain acidified with sulfurlc
acid on host-parasite  interactions, pp. 919-925.  I_n The 1st Internat.
Symp. on Acid Precipitation and the Forest Ecosystem.  L. S. Dochinger
and T. A. Seliga, eds.  U.S.D.A. For.  Serv. Genl. Tech. Rep. No. NE-23,
Upper Darby, Pennsylvania.

Shriner, D. S.  1977.   Effects  of  simulated rain acidified with sulfuric
acid on host-parasite  interactions.  Water, Air, and Soil Pollut. 8:9-14.

Shriner, D. S.  1978a.  Effects of simulated acidic rain on host-parasite
interactions in plant diseases. Phytopathology 68:213-218.

Shriner, D. S.  1978b.  Interactions between acidic precipitation and
S02 or 03:  Effects on plant response.  Phytopathology News 12:152.
(Abstr.)

Shriner, D. S.   1981.  Terrestrial vegetation air pollutant
interactions: Nongaseous pollutants, wet  deposition.  Paper presented at
Int. Conf. on Air Pollutants and Their Effects on Terrestrial Ecosystems.
Banff, Alberta, Canada.  May 20-17, 1980.

Shriner, D. S. and J.  W. Johnston. 1981.  Effects of simulated,
acidified rain on nodulation of leguminous plants by Rhizobium spp.
Environ. Exp. Bot. 21:199-209.

Siccama, T. G.  1974.   Vegetation, soil and climate on the Green
Mountains of Vermont.   Ecol. Monographs 44:325-349.

Siccama, T. G., M. Bliss, and H. W. Vogelmann.   1982.  Decline of red
spruce in the Green Mountains of Vermont.  Bull. Torrey  Bot. Club:
109:162-168.

Sierpinski, Z.  1967.  Influences  of  industrial  air pollutants on the
population dynamics of some primary pine  pests.  Proc. 14th Congr.  Int.
Union For. Res. Organiz. 5(24):518-531.

Sierpinski, Z.  1970.  Economic significance  of noxious  insects  in  pine
stands under the chronic impact of the industrial air pollution.  Sylwan
114:59-71.

Sierpinski, Z.  1971.  Secondary noxious  insects of pine in stands
growing in areas with industrial air  pollution containing nitrogen
compounds.  Sylwan 115:11-18.

Sierpinski, Z.  1972a.  The economic  importance  of secondary noxious
insects of pine on territories with chronic influence of industrial air
pollution.  Mitt. Forstl. Bundesversuchsanst  Wien 97:609-615.

Sierpinski, Z.  1972b.  The occurrence of the spruce  spider
(Paratetranychus (Oligonychus)  ununquis Jacoby)  on Scotch pine in the
range of the influence of industrial  air  pollution, pp.  101-110.  Jji
Institute  Badawczego  Lesnictwa, Warsaw, Bull,  No. 433-434.
                                  3-83
  1+09-262 0-83-7

-------
Sierpinski,  Z. and J.  Chlodny.   1977.  Entomofauna of forest plantations
in the zone  of disastrous industrial pollution, pp. 81-150.  In
Relationship Between Increase in Air Pollution Toxicity and Elevation
Above Ground.  J.  Woldk, ed.  Institute Badawczego Lesnictwa, Warsaw.
Shields, W.  J. and S.  D. Hobbs.  1979.  Soil nutrient levels and pH
associated with Armillariella mellea on conifers in northern Idaho.  Can.
J. For. Res. 9:45-48.

Si gal, L. L. and 0. C. Taylor.   1979.  Preliminary studies on the gross
photosynthetic response of  lichens  to peroxyacetylnitrate fumigations.
The Bryologist 82:564-575.

Skye, E.  1968.  Lichens and air pollution.  Acta Phytogeogr. Suecica
52:1-123.

Skye, E.  1979.  Lichens as biological indicators of air pollution.  Ann.
Rev. Phytopath. 17:325-341.

SMA.  1982.   Acidification  Today and Tomorrow.  Swedish Ministry of
Agriculture.  Environment 1982 Committee, Stockholm, Sweden, 227 p.

Smith, R. A.  1872.  Air and Rain.  Longmans, Green and Company, London.

Smith, W. H.  1970.  Tree Pathology--A Short Introduction.  Academic
Press, New York. 309 pp.

Smith, W. H.  1976.  Air pollution—effects on the structure and function
of plant-surface microbial-ecosystems, pp. 75-105.  jji Microbiology of
Aerial Plant Surfaces. C.  H. Dickinson and T. F. Preece, eds.  Academic
Press, New York.

Smith, W. H.  1981.  Air Pollution  and Forest.  Springer-Verlag, New
York, 379 pp.

Sokolov, A.A.  1972.  (Chemical  composition of rainfall passed through
the birch and spruce canopies).  Lesovedenie 3:103-105 (Russian with
English summary; (referenced  in  Horntvedt, 1975).

Sollins, P., L.L.  Grier, F.M. McCorison,  K. Cromack, Jr., A.T. Brown, and
R. Fogel.  1979.  The internal nutrient cycles of an old growth Douglas
fir stand in western Oregon.  Ecological  Monographs.  (In press).

Steinnes, E. and H. Krog.   1977. Mercury, arsenic, and selenium fallout
from an industrial complex  studied  by means of lichen transplants.  Oikos
28:160-164.

Strand, L.  1980.  The effect of acid precipitation on tree growth, pp.
64-67.  In Ecological  impact of  acid precipitation.  D. Drabltfs and A.
Tollan, "eUs.  Proceedings of  an  International Conference.  Sandefjord,
Norway.
                                  3-84

-------
Stringer, P. W. and M. H.  L.  Stringer.   1974.  A quantitative study of
corticolous bryophytes In  the vicinity  of  Winnipeg, Manitoba.  Bryologlst
77:551-560.

Sundstrom, K. R. and J. E. Hallgren.  1973.  Using lichens as
physiological indicators of sulfurous pollutants.  Ambio 2:13-21.

Tamm, C. 0.  1951.  Removal of plant nutrients from tree crowns by rain.
Physiologia Plantarum 4:184-188.

Tamm, C. 0. and E. B. Cowling.  1976.   Acidic precipitation and forest
vegetation, pp. 854-855.  Jji  Proc.  First Internat. Symp. on Acid
Precipitation and the Forest Ecosystem.  U.S.D.A. Forest Service, Gen.
Tech. Report. NE-23, Upper Darby,  PA.

Tamm, C. 0. and 6. Wiklander.   1980.  Effects of artificial acidification
with sulfuric add on tree growth  in Scots pine forest.  In Ecological
Impact of Acid Precipitation.   D.  Drabljls  and A. Tollan, ed"s.
Proceedings of an International Conference, Sandejford, Norway.

Tempi1n, E.  1962.  On the population dynamics of several pine pests 1n
smoke-damaged forest stands.   Wissenschafthche Zeitschrlft der
Technischen Universltat, Dresden 113:631-637.

Torstensson, N. T. L.  1975.   Degradation  of 2,4-D and MCPA in soils of
low pH, pp. 262-265.  Jji Environmental  Quality and Safety.  F. Coulston
and F. Korte, eds.  Georg  Thieme,  Stuttgart, Germany.

Treshow, M.  1975.  Interaction of air  pollutants and plant diseases, pp.
307-334.  ]jn Responses of  Plants to Air Pollution.  J. B. Mudd and T. T.
Kozlowski, eds.  Academic  Press, New York.

Treub, M.  1888.  Notice sur la nouvelle flora de Krakatau.  Ann. Jour.
Bot. Buitenzorg 7:221-223.

Troiano, J., L. Colavito,  L.  Heller, and D. C. McCune.  1981.  Effect of
simulated acid rain and photochemical oxidant on seed development in
soybean.  Phytopathology 71:565.   (Abstr.).

Troiano, J., L. Colavito,  L.  Heller, and D. C. McCune.  1983.  Effects of
acidity of simulated acid  rain and its  joint action with ambient
photochemical oxidant on measures  of biomass and yield in soybean.
Agric-Environ.  In press.

Troiano, J., L. Heller, and J. S.  Jacobson.  1982.  Effect of added water
and acidity of simulated rain  on growth  of field-grown radish.  Environ.
Pollut. (Series A) 29:1-11.

Tukey, H. B., Jr.   1970.  The  leaching  of  substances form plants.  Ann.
Rev. Plant Physiol.  21:305-324.
                                 3-85

-------
Tukey, H. B., Jr.   1975.   Regulation of  plant growth by rain and mist.
Proc., Int. Plant  Prop. Soc.  25:403-406.

Tukey, H. B., Jr.   1980.   Some effects of rain and mist on plants with
implications for acid precipitation, pp. 141-150.  In Effects of Acid
Precipitation on Terrestrial  Ecosystems.  T. C. HutcKinson and M. Havas,
eds.  Plenum Press, New York.

Tukey, H. B., Jr.  and J.  V. Morgan.  1963.   Injury to foliage and its
effect upon the leaching  of nutrients from above-ground plant parts.
Physio!. Plant. 16:557-563.

Turk, R. and V. Wirth. 1975. The pH dependence of S02 damage to
lichens.  Oecol. 19:285-291.

Turk, R., V. Wirth, and 0. L. Lange.  1974.  Carbon dioxide exchange
measurements for determination of sulfur dioxide resistance of lichens.
Oecologia 15:33-64.

Tveite, B.  1980a.  Tree  growth  in field experiments, pp. 206-207.  I_n
Ecological Impact  of Acid Precipitation.  D. Orablos, and A. Tollan, eds.
Proc. of an International Conference, The Norwegian Interdisciplinary
Research Programme Acid Precipitation-Effects on   forest and Fish.
Sandfjord, Norway, March  11-14,  1980.

Tveite, B.  1980b.  Foliar nutrient levels in field experiments.  Ir±
Ecological Impact of Acid Precipitation.  D. Drablos, and A. Tollan, eds.
Proc. of an International Conference, The Norwegian Interdisciplinary
Research Programme Acid Precipitation-Effects on forest and Fish.
Sandfjord, Norway, March  11-14,  1980.

Tveite, B. and G.  Abrahamsen. 1980.  Effects of artificial acid rain on
the growth and nutrient status of trees,  pp. 305-318.  li± Effect of Acid
Precipitation on Terrestrial  Ecosystems.  T. C. Hutchinson and M. Havas,
eds. Plenum Press, New York.

Ulrich, B.  1981a.  Eine  okosystemare Hypothese uber die Ursachen des
Tannensterbens (Abies alba Mill.) Forstw. Cbl. 100, 228-236.

Ulrich, B.  1981b.  Bodenchemische und Umwelt-Aspekte der Stabilitat von
Waldokosystemen.  Library Environment Canada Translation no. OOENV
TR-2038.

Ulrich, B.  1982.   "Dangers  for  the forest ecosystem due to acid
precipitation.  Necessary counter-measures:   Soil liming and exhaust gas
purification" U.S. EPA translation TR-82-0111.  EPA/NCSU Acid
Precipitation Program.  North Carolina State University, Raleigh.

Ulrich,  B., R. Mayer, and P.  K.  Khanna.   1980.  Chemical changes due  to
acid precipitation in a loss-derived  soil in central Europe.  Soil Sci.
130:193-199.
                                  3-86

-------
U.S. Department of Agriculture.   1976.   Proc.  First  Int. Symp. Acid
Precipitation and the Forest Ecosystem.   L.  S. Dochinger and T. A. Seliga
(eds), USDA Forest Serv. Gen.  Tech.  Rep.  NE-23, Upper Darby, PA.

U.S. Environmental Protection Agency.   1978.   Air  quality criteria for
ozone and other photochemical  oxidants.   Env.  Criteria and Assess. Off.,
EPA/600-8-78-004, Research Triangle  Park, NC.  341 pp.

U.S. Environmental Protection Agency.   1982.   Air  quality criteria for
particular matter of sulfate oxides.  Env.  Criteria  and Assess Off.
Research Triangle Park, NC.  (In press).

Villemant, C.  1979.  Modifications  de  Tenclomocemose due pin sylvestre
en liaison avec ea pollution atmospherique  en  foret  de Roumare
(Seine-Maritime).  Doctural Disseration.   Pierre and Marie Curie
University, Paris, 161 pp.

Vogelmann, H. W., T. 6. Siccama, D.  Leedy,  and D.  C. Ovitt.  1968.
Precipitation from fog moisture  in the  Green Mountains of Vermont.
Ecology 49:1205-1208.

Way, T.  1855.  The atmosphere as a  source  of  nitrogen to plants.  Roy.
Agr. Soc. Jour. 16:249-267.

Weber, J. B.  1972.  Interaction of  organic pesticides with partial!ate
matter in aquatic and soil systems.  Adv. Chem. Ser. 111:55-120.

Weber, J. B. and S. B. Weed.  1974.  Effects of soil on the biological
activity of pesticides, pp. 223-255. Jji  Pesticides  in Soil and Water.
W. D. Guenzi, ed.  Soil Sci. Soc. Am.,  Inc., Madison, WI.

Wellburn, A. R., 0. Majernik,  and F. A. M.  Wellburn.  1972.  Effects of
S02 and N0£ polluted air upon the ultrastructure of  chloroplasts.
Environmental Pollution 3:37-49.

Wells, C.G., A.K. Nicholas, and  S.W. Buol.   1975.  Some Effects of
Fertilization on Mineral Cycling in  Loblolly Pine.   Pp. 754-764 in
Howell, F.G., J.B. Gentry, and M. H. Smith  (eds.)  Mineral Cycling in
Southeastern Ecosystems.  U.S. Research and Development Administrations.

Westburg, H. H., K. J. Allwine,  R. A. Rasmussen, and E. Robinson.
1976.  Int. Conf. National Sensing and  Assessment.   Inst. Electrical and
Electronic Engineers.  Washington, DC, Vol.  I, p.  2.

Wiackowski, S. K. and L. S. Dochinger.  1973.  Interactions between air
pollution and insect pests in  Poland.  2nd  Inter.  Congr. Plant Pathol.,
Univ. of Minneapolis, Minnesota.  Abstr.  No.  0736, p. 1.

Will-Wolf, S.  1980.  Effects  of a "clean"  coal-fired power generating
station on four common Wisconsin lichen species.   Bryologist 83:296-300.
                                  3-87

-------
Winner, W. E.  and J.  D.  Bewley.  1978a.  Terrestrial mosses as
bioindicators  of S02  pollution stress.  Oecologia 35:221-230.

Winner, W. E.  and J.  D.  Bewley.  1978b.  Contrasts between bryophyte and
vascular plant synecological  responses in an SOo-stressed white spruce
association in central Alberta.  Oecologia 33:311-375.

Winner, W. E., J. D.  Bewley,  H. R. Krouse, and H. M. Brown.  1978.
Stable sulfur  isotope analysis of S02 pollution impact on vegetation.
Oecologia 36:351-361.

Wood, T. and F. H. Bormann.   1974.  The effects of an artificial acid
mist upon the  growth  of  Betula alleghaniensis Britt.  Environ. Pollut.
7:259-268.

Wood, T. and F. G. Bormann.   1977.  Short-term effects of simulated acid
rain upon the  growth  and nutrient relations of Pinus strobus L.  Water,
Air, and Soil  Pollut. 7:479-488.
                                  3-e

-------
            THE ACIDIC DEPOSITION PHENOMENON AND ITS  EFFECTS
                   E-4.  EFFECTS ON AQUATIC CHEMISTRY

4.1  INTRODUCTION (J. N. Galloway)

     In the last decade, acidification of streams and lakes, with
subsequent biological damage,  has become a well  reviewed effect of
acidic deposition (NAS 1981,  U.S./Canada 1981, NRCC 1981).  However,
confusion, ignorance, and debate still cloud our knowledge of past,
current, and future trends in  the acidification  of aquatic systems, the
key processes that control the acidification, and the degree of
permanency of biological effects.

     This chapter critically reviews how aquatic chemistry responds to
acidic deposition.   Initially, basic definitions and  concepts regarding
acidic deposition, terrestrial and aquatic systems, measurements of
sensitivity using alkalinity,  and' the different  time  scales of
acidification are presented.   These definitions  are followed by a
detailed listing of the important characteristics of  terrestrial and
aquatic system subcomponents that ameliorate or  enhance the effect of
acidic deposition.  The theoretical and practical sensitivity of aquatic
systems to acidic deposition  is discussed with locations of sensitive
and affected sensitive systems documented.  Aquatic systems considered
include lakes, streams, and estuaries.  The degree of acidification of
these systems is examined in the light of existing models, and the
models are critically reviewed.  The role of S and N  in the
acidification process is addressed.  The status  of our knowledge on
acidification of aquatic systems is presented in the  context of asking
what will happen if depositions of S and N compounds  from the atmosphere
increase or decrease.  As a final section, the interaction of aquatic
acidification with the metal and organic biogeochemical cycles is
addressed and an assessment of knowledge is presented.

4.2  BASIC CONCEPTS REQUIRED TO UNDERSTAND THE EFFECTS  OF ACIDIC
     DEPOSITION ON AQUATIC SYSTEMS

     The following concepts concerning effects of acidic deposition on
aquatic systems will  serve as  a foundation for critically assessing our
current knowledge.

4.2.1  Receiving Systems (J. N. Galloway)

     Receiving systems are terrestrial,  wetland,  and  aquatic.  Their
component parts include:

    a.   Terrestrial  Components

        (1)   forest,  crop, or  grass canopy
                                  4-1

-------
        (2)   litter layer
        (3)   organic soil layer

        (4)   Inorganic  soil  layer
        (5)   bedrock

    b.  Wetland Components

        (1)   vegetation - mosses and other semi-submerged plants

        (2)   water - stream,  pond,  swamp

    c.  Aquatic Components

        (1)   stream

        (2)   lake

        (3)   sediment

     These systems and  their components are linked, so the effects of
atmospheric  deposition  on one component can cause secondary effects 1n
another component.  The hydro!ogle  pathway controls which components are
affected by  (or linked  to) other components.  For example, water (preci-
pitation) first hits the tree canopy, then travels through successive
layers of the terrestrial system before it enters wetlands adjacent to
the terrestrial system  and then finally the lake.  Therefore, the
effects of atmospheric  deposition on any one component of the
terrestrial-wetland-aquatic  system  depend not only on the composition of
the atmospheric deposition but also on the effect of the atmospheric
deposition on every system  'upstream1 from the component of Interest.
For example, the effect of atmospheric deposition on aquatic systems
depends on what is in the atmospheric deposition and its effect on all
components of the terrestrial and wetland systems that it contacts prior
to discharge into the aquatic system  (Figure 4-1).

     Therefore, when discussing effects on components of terrestrial,
wetland, or  aquatic systems  we are  incorrect to investigate only the
component of interest.   Rather, all components either directly or
indirectly linked to the specified  component should also be included in
the study.  This point will  become  especially pertinent in Section
4.3.2, when  it is shown that decreases in sulfur deposition from the
atmosphere may not result in decreases in lake acidification until the
terrestrial  system above the lake recovers.  Instances where the
terrestrial  system is less  important are  (1) lake systems with a large
lake/watershed area ratio and (2) lake and stream systems that receive
runoff or snowmelt that has  had little contact with the terrestrial
systems.

     The composition of aquatic systems is controlled by not only
physical and chemical processes but also  by biological processes.  In
discussing the concept of system sensitivity and determining the degree
of acidification, we cannot ignore  the biological component because,


                                 4-2

-------
                              TERRESTRIAL
                              ECOSYSTEM
                                AQUATIC
                              ECOSYSTEM
METEOROLOGIC
                                  V
                               GEOLOGIC
                                            BIOLOGIC
Figure 4-1.   Diagrammatic  model  of the  functional  linkages  between
             terrestrial and aquatic  ecosystems.   Vectors may  be
             meteorologic,  geologic,  or biologic  components moving
             nutrients  or  energy along  the  pathways  shown.  Adapted
             from Likens and Bormann  (1974).
                                  4-3

-------
depending on location,  type,  and  productivity, the biological component
can make waters more sensitive, less sensitive, more acidified, or less
acidified.  Specific details  on the importance of the biological systems
in mediating the chemical  response of an aquatic system to acidic
deposition can be found in Section 4.3.2.6.

     Additional details on the terrestrial systems are found in this
section (following)  and in Chapters E-2 and E-3 on soils and vegeta-
tion, respectively.   The next chapter, Chapter E-5, discuss the effects
of acidification of  aquatic systems on biota.

4.2.2  pH, Conductivity, and  Alkalinity (M. R. Church)

     Three important analytical measures for evaluating acidification of
ground or surface waters are  pH,  conductivity, and alkalinity.
Definitions of these three quantities are briefly given here.  A later
section (4.3.3.2.1)  examines  problems concerning the comparability of
historical and more  recent pH, conductivity, and alkalinity data.

4.2.2.1  j)H--In 1909 the Danish chemist S. P. L. Stfrensen introduced
the term pH" when he  used exponential arithmetic to express the
concentration of hydrogen Ions in aqueous solution.  He formulated his
definitive equation  as

     CH - icrp                                                    [4-i]

where CH was the hydrogen ion concentration and P was the hydrogen ion
exponent, which Sdrenson then wrote as PH and which we now write as
pH (Bates 1973).  For a number of reasons, too detailed to explore here,
pH as originally defined by Stfrensen is not a measure of either
hydrogen ion activity or concentration (Feldman 1956, Bates 1973).
Fortunately, this fact, 1n and of Itself, does not adversely affect
measurements of surface water acidification.  A practical (Feldman 1956)
or operational (Bates 1973) pH scale has been defined:

                     (Ex - EC) F
     P"(x) = PH(S) + —	                                    C4-2]
                     RT In 10

where pH(s) 1s the assigned pH of a standard solution, Es the emf
produced In a pH cell by the  solution, F the Faraday constant, R the
universal gas constant, T the temperature in °K, and Ex the potential
produced in the pH cell by an unknown solution X, which then by
definition has a pH  of  pH(x).

4.2.2.2  Conductivity—Conductivity (or specific conductance) measures a
solution's ability to conduct an  electric current.  This capacity is a
function of the Individual mobilities of the dissolved ions, the
concentrations of the ions, and the temperature of the solution.  As the
"ohm" Is the standard unit of resistance, the "mho" (ohm spelled
backwards) is the standard unit of conductance.  Conductivity is
conductance per unit length of a  substance of unit cross section and is
                                 4-4

-------
usually reported as ymho cnrl  or the equivalent ySiemens cm'1.
Distilled water may have a conductivity as low as 0.5 ymho cm~l, and
some naturally occurring surface waters in the United States may have
conductivities as high as 1500 ymhos cm'1 (Golterman 1969, American
Public Health Association 1976, Skougstad et al. 1979).

     The rationale for measuring conductivity in relation to
acidification of surface waters is  threefold.  First, low conductivity
values in surface waters generally  indicate a lack of buffering and thus
susceptibility to acidification (Ontario Ministry of the Environment
1979).  {In some cases, however,  organic compounds may contribute to
buffering but only very little to conductivity.)  Second, low
conductivity has been correlated with sparsity of fish populations in
low pH lakes (Leivestad et al. 1976, Wright and Snekvik 1978).  Third,
increases in conductivity over time in surface water under some
circumstances can be used to infer  acidification of that water body
(Nilssen 1980).  Hydrogen ions have extremely high mobilities in
solution and contribute greatly to  conductivity.  As a body of water
becomes acidified over time, increases in hydrogen ion concentrations
could lead to an appreciable increase in conductivity (e.g., from pH 5.0
to pH 4.5, an increase of approximately 7 ymho cnrl, us-jng a value
of 0.313 ymho cm'1 per yeq JT1 free H+; see Wright and Snekvik 1978).
In the absence of other data (e.g., pH, alkalinity, acidity) such a
change with time could possibly be  used to infer acidification,
provided, of course, no other  reasonable explanation was apparent.

4.2.2.3  Alkalinity—Al kalinity measures the ability of an aqueous
solution to neutralize acid.  For this reason it is also known as acid
neutralizing capacity, or ANC.  In  most natural freshwaters, buffering
is predominantly due to species of  the carbonate system (Stumm and
Morgan 1981).  In the very dilute surface waters often studied in
relation to acidification, total  inorganic carbon concentrations are
low; therefore, ANC due to the carbonate system is also low.  It is not
unusual to find in these systems  that other species, such as naturally
formed weak organic acids (when dissociated) and aluminum-hydroxy
compounds leached from soils and  sediments, contribute measurably to ANC
so that an appropriate expression for this quantity is
     ANC = (HC03~)  + 2 (C032') +  (A10H2+) + 2 (A1(OH)2+)


         + 4 (A1(OH)4~)  +  (RCOO~) +  (OH") - (H+)                    [4-3]


where (RCOO-)  represents dissociated organic acids (Bisogni and
Driscoll 1979).   In some waters organic acids may dominate both the pH
and buffering of natural waters.  North American areas with waters of
this type include parts  of the south and southeast, the upper midwest,
                                 4-5

-------
locations in the northeast (see General Introduction for map defining
regions) and the Atlantic  maritime provinces of Nova Scotia and
Newfoundland—all regions  noted for  their highly colored brownwater
lakes and streams.  Naturally acidic brownwater lakes and streams are
discussed further in Section 5.2.1.  For discussion of buffering due to
organic systems see Bisogni and Driscoll (1979), Wilson (1979), and
Section 4.6.3.2.

    The operational procedure for determining ANC is acidimetric
titration with strong acid to an  appropriate end point.  Methods for
performing such titrations and theoretical  treatment of-the pertinent
equilibria have been detailed in  many publications (e.g., Golterman
1969, American Public Health Association 1976, Loewenthai and Marais
1978, Skougstad et al. 1979, Stumm and Morgan 1981).

4.2.3  Acidification (J. N. Galloway)

     Acidification is defined as  the loss of alkalinity.  Those aquatic
systems for which acidic deposition-  may cause acidification or loss of
alkalinity, to levels that result in biological change are termed
sensitive.  Loss of alkalinity can be either chronic or acute,
identified as long-term acidification and short-term acidification,
respectively.  Short-term  acidification refers to the development of
strong acidity (i.e., alkalinity  < 0) during acid episodes (e.g., spring
snowmelt) lasting for periods of  days or weeks.  Because of the
relatively short exposure  periods, biological effects occur only at
those very low alkalinity  levels  (<  0).  Long-term acidification, on the
other hand, refers to the  gradual loss of alkalinity over periods of
years or decades.  As a result of chronic exposures, biological effects
may occur at alkalinity <  100 peg i~l (Chapter E-5, Section
5.10.4), and waters with alkalinity  < 200 ueq £-1 are generally
considered sensitive as defined in Section  4.3.2.6.1.

4.3  SENSITIVITY OF AQUATIC SYSTEMS  TO ACIDIC DEPOSITION  (J. N.
     Galloway and P. J. Dillon)

     The previous sections pertaining to aquatic systems have presented
concepts and definitions required to assess our knowledge of how aquatic
systems are affected by acidic deposition.  This section and the ones
following begin our assessment by identifying important components in
deposition processes and receiving  systems  that will control the
response of aquatic systems to acidic deposition.  Later sections then
examine what is known about this  response.

4.3.1  Atmospheric Inputs

     Five factors must be  considered when we  assess the role of
atmospheric deposition in  the acidification of aquatic and terrestrial
ecosystems.  These are the components  (total  vs wet vs dry) of the
deposition that are measured, the chemical  species in the deposition,
the concentration of the substances  in  the  deposition relative to their
loading (input rate), the  location of the deposition  (considering a


                                 4-6

-------
geographic scale as well  as considering  the  different components [e.g.,
leaf vs soil] of any system),  and  the temporal distribution of the
loadings.

4.3.1.1  Components of Deposition—To assess the  impact of acidic
deposition we must know the total  input  (wet and  dry).  A major part of
the current North American effort  regarding  deposition monitoring is
devoted to "wet-only" measurement.  These data are inadequate for
assessing impact on aquatic and terrestrial  ecosystems because they
underestimate total deposition, not  only near major point sources of
SOX, NOX (Dillon et al. 1982)  but  also in remote  areas (Galloway et
al . 1982a).  Relatively few attempts have been made to measure dry
deposition separately (Lindberg et al. 1982). In a few cases (e.g.,
Dillon et al . 1982) "calibrated" lakes and watersheds have been used to
infer dry or total deposition  of acidic  substances.  In other cases,
"bulk" deposition measurements (made with a  continuously open collector)
have been used.  Although these collect  an undefined portion of the dry
deposition, this information is more useful  for chemical budget
calculations than "wet only" measurements unaccompanied by dry
deposition measurements.   See  Chapter A-8 for further discussion of
deposition monitoring.

     In addition to H+ deposition, it is also important to measure the
contributions of SCty2- and N03~ and  NH4  (see Section 4.4.1 and
Chapter A-8).  In some systems, N03~ is  significantly utilized (chemically
or biologically), resulting in the internal  production of ANC (Harvey et al.
1981, Dillon et al. 1982).  In some  cases, $042-  is stored in terrestrial
watersheds by the process of sulfate adsorption (N. M. Johnson 1979), a process
that may also generate ANC if  the  S042~  is reduced or if strong acid is
simultaneously stored.  $042-  may  also be reduced in lakes, resulting in
production of ANC (Cook et al. 1981).  This  production of ANC is only
important on a long-term basis if  it is  net  production, i.e., a net
reduction of N03 and $64 on an annual basis.  In  other systems
S042- apparently acts as a conservative  substance within the limits
of error in the measurement of the dry deposition fluxes (Likens et al.
1977, Galloway et al. 1983c).
     Once wet or dry deposited,  $03 and SCty- have the same
pathways through the terrestrial  and  aquatic systems;  therefore, the
effect of S on aquatic systems  is not dependent on chemical speciation
or type of deposition.

     Virtually all  of the ammonium ion, NH4+, deposited from the
atmosphere on terrestrial  and aquatic systems is used chemically or
biologically in those systems (Likens et al. 1977, NAS 1981) in many
cases resulting in  a decrease in  ANC. NH4+ deposition is
"significant" (25 percent to 50 percent) relative to H+ deposition.
For example, at Harp Lake, Ontario, about 25 percent of the net input of
acid was from NH4+  deposition (Dillon et al. 1979).  Therefore,
measuring only free acid  (H+) is  inadequate for assessing the impact
of acidic deposition on systems.
                                 4-7

-------
    The input rate of basic cations is also important because without it
the net loss of base cations  from the watershed cannot be calculated.

4.3.1.2  Loading vs Concentration—Because the ANC of some components of
the systems receiving acidic  deposition is not renewed (other than over
geologic time), the total loading (or input rate) is the factor that
determines how long those components will be able to assimilate acidic
deposition.  The ability of some other components to assimilate acidic
deposition may depend on the  concentration as well as the total loadings
of the acids.  The assimilation capacities of components that have a
continually renewed ANC, (e.g., a lake's epilimnion that has ANC
produced through primary production) or those where reaction rates are
controlled by hydrologic factors (e.g., reaction between acidic
deposition and silicate bedrock) are sensitive to the amount of water
passing through components as well as the concentration of acid.

     In general, current measurements of acidic deposition include both
the concentration of important substances and the total loading rate of
those substances with the exception of dry deposition as discussed in
Section 4.3.1.1.

4.3.1.3  Location of the Deposition—Deposition of acidic substances is
well measured in most areas where the geological terrain has low
capability to neutralize acids and where the wet deposition is known to
be relatively high (> 20 meq  strong acid m~2 yr~l; see Chapter A-8).

     On a smaller scale, the  relative magnitude of deposition on
different components (leaf, soil, water surfaces, etc.) of specific
ecosystems is more poorly understood.  For example, the ability of the
vegetation in a terrestrial system, particularly the forest canopy, to
enhance deposition of acidic  substances relative to other components of
the system (e.g., bedrock or  soil surfaces, water surfaces) has been
demonstrated (Parker et al. 1980) but needs to be quantified in further
studies.

     Other factors, such as the relative deposition to the terrestrial
component of a watershed vs directly on the surface water, are also
important.  These factors determine the relative importance of the
pathways that the deposited  substances follow, which in turn controls
the overall assimilation capacity of the system.

4.3.1.4  Temporal Distribution of Deposition—To assess their impact on
receiving  systems, the input  rates of acids or acidifying substances
must be considered on a seasonal and a short-term (i.e., episodic) basis
as well as on a long-term  (annual) basis.

     Seasonal inputs are particularly important  in areas where snowpack
formation and subsequent release of a major portion of the annual
deposition during snowmelt to groundwaters and surface waters occur
(Jeffries et al. 1979, Galloway et al. 1980b).   In some cases (e.g.,
Central Ontario) during snowmelt  the ground is frozen.  As a result,
                                  4-8

-------
the release of Ions occurs at a time  when  the terrestrial system cannot
assimilate the ions as efficiently  as it can at other times.

     Short-term variations in deposition,  on even an episodic basis, may
be important in some instances.  Flow paths may be altered on a
short-term basis, resulting in shortened reaction times and less
assimilation of the acidic deposition.

     The seasonal variation in deposition  has been frequently
investigated; short-term variations are more poorly studied and need
further quantification.

4.3.1.5  Importance of Atmospheric  Inputs  to Aquatic Systems--

4.3.1.5.1  Nitrogen (N), phosphorus (P) and carbon (C).  Only recently
have researchers appreciated the importance of precipitation inputs of
various cations and anions, especially N and P to the nutrient balance
of inland freshwaters (e.g., Gorham 1958,  1961; Vollenweider 1968,
Schindler and Nighswander 1970, Likens 1974, Likens and Borman 1974).
Concentrations of inorganic and organic N  and Pin rain and snow may be
small, but the total input by storm,  by season, or by year may be a
significant source of these nutrients for  aquatic organisms,
particularly in nutrient-poor lakes (Likens et al. 1974).  Direct inputs
of nutrients in precipitation to lakes are particularly important in
areas with granitic geologic substrates, especially if the ratio of lake
surface area to terrestrial drainage  area  is large (Likens and Bormann
1974).  In addition, the gaseous exchanges of nitrogenous compounds in
many lakes may be important but are poorly understood (Likens 1974).

     Based on relatively few data,  some 50 percent of the P and 56
percent of the dissolved N for oligotrophic lakes may come from direct
precipitation (Likens et al. 1974).   With  human influences in the
watershed (urbanization, agriculture,  etc.) runoff inputs to aquatic
ecosystems Increase and direct precipitation inputs become much less
important to the total budget,  even though the absolute amount provided
by precipitation remains the same.  Where  terrestrial inputs of N and P
dominate, lakes are usually much more biologically productive, if not
eutrophic (Likens et al. 1974).

     Preliminary data suggest that  organic carbon inputs in
precipitation may be ecologically significant for some aquatic
ecosystems, particularly oligotrophic lakes.  Mean concentrations
averaged about 6 mg C £-1 in precipitation and accounted for 28
percent of the total allochthonous  inputs  of organic carbon for a small
oligotrophic lake in New Hampshire  (Jordan and Likens 1974).

     4.3.1.5.2  Sulfur.  Two sources  provide sulfur for surface waters:
rock weathering and atmospheric deposition.  In the absence of
significant sulfur sources in bedrock, atmospheric deposition is the
primary source (Cleaves et al.  1970).  This is especially true in areas
receiving acidic deposition, where  atmospheric sulfate becomes the
dominant anion in low alkalinity waters (Gjessing et al. 1976, Oden


                                 4-9

-------
1976a, Henriksen 1979,  Wright  et  al. 1980, Galloway et al. 1983c).  This
dependence is illustrated  by plotting the mean and range of excess
S042- (over and above that supplied by sea salt cycling) export from
watersheds across North America on a line that transects the region of
large atmospheric deposition of $042- (Figure 4-2).  The wet
deposition of excess S042~ at  each location is shown in the same
figure, with estimated total S042- deposition shown at four
locations.  There is a clear positive relationship between excess
S042~ deposition and SC)42~ in  the runoff.  The influence of the
increased S042- deposition on  aquatic chemistry is large, for on an
equivalent basis, the increase in the SO*2- has to be matched by an
increase in a cation, either protolytic  (proton donating) (e.g., H+,
Aln+) or non-protolytic (e.g., Ca2+, Mg2+, etc.)  (Galloway et al.
1983c).  An increase in the former will  result in loss of alkalinity
(acidification) of the waterbody. An increase in the latter will result
in a loss of cations from  the  terrestrial system.  Both effects can
potentially alter biological communities in the respective ecosystem
(see Sections 4.4 and 4.6).

4.3.2  Characteristics of  Receiving Systems Relative to Being Able to
       Assimilate Acidic Deposition

     The anthropogenic acids transported via  the  atmosphere may be
deposited directly into aquatic systems  (lakes, streams, wetlands) or
onto terrestrial systems that  drain  into the  aquatic systems.  Each of
the components or subsystems of these systems may be capable of
assimilating some or all of the acid  deposition received.  This section
discusses the factors that determine  the quantitative capability of the
subsystems to assimilate acidic deposition.

4.3.2.1  Canopy--Throughfal1  and  stemflow have elevated levels of most
elements relative to incident  rainfall  (Miller and Miller 1980) and
even,  in at least one report,  relative  to snowfall (Fahey 1979).  The
changes in chemical content result  from washdown  of particles filtered
from the atmosphere by the vegetation,  and from leaching  of the
vegetation (the crown in the  case of throughfall, the bark as well in
the case of stemflow).  The process of  particle washdown  is, of course,
completely independent of any  ability of the  canopy to assimilate acidic
deposition.  On the other hand,  leaching of cations from  the canopy may
represent a signficant assimilation capacity.  However, the relative
importance of each process is generally unknown.  Although there  are
conflicting reports, some generalizations may be  made.

     Stemflow has a lower pH than does  incident precipitation, either
because of leaching of organic acids or washdown  of acidic aerosols
(Miller and Miller 1980).

     Throughfall in deciduous  forests has usually been  found to have
elevated  pH and  increased cation  (Ca2+, Mg2+)  concentration  (Likens
et al. 1977, Cole and Johnson 1977).   The relative  importance of
washdown  of filtered particles and of cation exchange with the  leaf  is
unknown.  Direct uptake of S02 (Fowler  1980)  and  ammonium (Miller and


                                  4-10

-------
                                                 -3-
                                                        LABRADOR
                                                         ISLAND OF
                                                         NEWFOUNDLAND
HALIFAX


NEW BRUNSWICK


LAFLAMME


MAURICIE


ADIRONDACKS


N. OF OTTAWA


ALGONQUIN


HALIBURTON


SUDBURY


ALGOMA


THUNDER BAY


QUETICO


ELA
                               ,_  _
                                     ui baiu
Figure 4-2.   Mean and range  of basin specific yield of excess  sulfur
              (|—<•>—|) (U.S./Canada 1982) compared with atmospheric excess
              sulfur deposition (|—•—|) in  precipitation for 1980
              (Barrie and Sirois 1982).  Also  shown are the  ranges of wet
              plus dry deposition of sulfur (| — |) calculated  from the
              1980 measurements of SOX in the  atmosphere at  4 Canadian
              Acid Precipitation Network Stations (Barrie  1982).
                                     4-11

-------
Miller 1980)  also may contribute  to  the acidity of the throughfall.  The
throughfall  pH of in coniferous forests has been reported to be
decreased relative to pH  of  precipitation  (Horntvedt and Joranger 1976),
although cation content is increased.

     However, the amount  of  throughfall or stemflow is less than
incident precipitation (Ford and  Deans 1978, Miller and Miller 1980).
Therefore, an increase in concentration of substances in throughfall
relative to precipitation does not necessarily indicate that the canopy
has supplied  materials as a  result of either washdown or leaching.  The
loading of each substance beneath the canopy must be compared to that
above the canopy before the  occurrence of  either process can be
ascertained.

4.3.2.2  Soil--The surficial  material accumulated on the bedrock of
North America is extremely complex in both physical and chemical
properties.   This surficial  material assimilates acidic deposition
through dissolution, cation  exchange, sulfate adsorption, and biological
processes.

     In general, surficial materials containing carbonate minerals have
abundant exchangeable bases  and can  assimilate acidic deposition to an
almost unlimited extent.  Regions of North America with soils formed in
situ on limestone, dolomite,  or marble provide adequate neutralizing
capacity under all loading conditions.  Soils formed in situ on
carbonate-cemented, carbonate-interbedded, or carbonate clastic
sedimentary  rocks may have reduced assimilation capacity under very high
acidic deposition conditions, but effects  of acidic deposition on
streams and lakes are probably minimal.  As a result of the transport of
surficial material in the glaciated  areas, it is possible to find
carbonate-containing deposits on  non-carbonate bedrock.

     The ability of surficial materials that contain no carbonate
minerals to assimilate acidic deposition results from cation exchange
reactions, silicate-mineral  dissolution reactions and, in some cases, Fe
and Al oxide  dissolution. The result of these reactions is an increase
in the concentrations of  major cations (particularly Ca2+, Mg^+, and
possibly Na+  and K+), and Al  and  Fe  in the runoff water leaving the
watersheds.   This ability is affected by:

1)  the chemical nature of the surficial material, in particular the
    cation exchange capacity (CEC) and the base saturation (BS),

2)  the permeability of each layer of the  soil,

3)  the surface area (or  grain size) of the soil particles, and

4)  the amount (depth and/or mass) of soil in the watershed.

     The most important of these  factors are the CEC (the total amount
of cations that can be exchanged  for H+; Table 4-1) and the BS (the
proportion of the total exchangeable cations that consists of


                                  4-12

-------
TABLE 4-1.  TYPICAL CATION  EXCHANGE CAPACITIES OF SOIL COMPONENTS
                    (FROM MCFEE ET AL. 1976)
 SOIL COMPONENTS
                                                    (meq 100 g'1)
 Organic matter (humus)                                   200
 Silicate clays
     vermiculite                                         150
     montmorillorite                                     100
     kaolinite                                            10
     illite                                               30
 Hydrous oxide clays                                       4
 Silts and sands                                       negligible

 aVariation is commonly  40% of these  mean  values.
                              4-13

-------
Mg2+, Na+,  and K+).   The organic layer of the soil has a high CEC
(McFee et al.  1976).   Fresh organic litter has a substantial BS
component,  so  it has  the potential to assimilate acidic deposition if
the BS is high, but potential  is low if the BS is low.

     The permeability of the soil layers is also important because it
determines the contact time of the percolating water with the soil
particles.   Loosely-packed organic material in the upper layer is
usually highly permeable and so may provide little assimilation
capacity, especially  in cases  of high input of water.  As the surface
area of the soil particles in  the organic layer increases, the
permeability of the layer decreases, both factors increasing the H
assimilation capacity of the soil, whether it is a result of surface
cation exchange reactions or silicate or metal oxide dissolution
reactions.   However,  the proportion of the soil consisting of very small
particles (i.e., clays) may increase to the point where permeability of
a specific layer is decreased  very significantly.  In some cases,
impermeable layers may effectively eliminate the potential for
assimilation of acidic deposition by deeper soil layers.

     The depth of the surficlal material in a watershed is, of course,
also very important.   Areas with extremely shallow (1 m) till often have
only an organic layer and a well-weathered layer (horizon) that may have
little assimilation capacity left (i.e., have low BS).  Areas with deep
tills (e.g., till plains, kames, moraines, eskers, spillways, outwash,
and alluvial formations) will  almost always have high capacity for
assimilating acidic deposition because of their  moderate to high BS at
greater depth, combined with their large amounts of unweathered
material.  Further evaluation  of soils as they affect aquatic systems is
found in Chapter E-2.

     Another soil process  important in controlling the response of
aquatic systems to acidic deposition is sulfate adsorption.  Soils with
large sulfate  adsorption capacities will essentially act as sinks for
the atmospheric sulfur, preventing it from reaching the aquatic system.
Generally, soils in unglaciated regions have much greater sulfate
adsorption capacities (see Chapter E-2, Section 2.2.8) and hence will be
protected from extreme acidification of aquatic systems (relative to the
northeastern United States) until the sulfate adsorption capacity is
saturated.  Estimates of the length of that period are decades to
centuries (Galloway et al. 1983a).

4.3.2.3  Bedrock—The ability  of bedrock to neutralize acidic deposition
is controlled  by:

     1)  chemical composition  of the bedrock,

     2)  effective reaction surface area, and

     3)  retention time or contact time of water with the bedrock.
                                  4-14

-------
    Carbonate minerals in the bedrock result in rapid assimilation  of
the strong acids by dissolution and in production of bicarbonate  ion.
Bedrock types containing no carbonate minerals may neutralize acidic
deposition by the dissolution of silicate minerals, which is an
extremely slow process relative to carbonate dissolution.

     Massive, impermeable bedrock's effective surface area for chemical
reaction is minimal.  Acidic deposition contacts only the upper surface
layer, so the slow dissolution process will modify water chemistry  only
marginally, regardless of which silicate material is involved. Bedrock
exhibiting only jointing or fracturing will provide relatively greater
surface area for reaction, but complete assimilation will only occur  at
considerable depth, probably affecting the chemistry of the groundwater
pool but having little effect on stream and lake chemistry.  The  maximum
extent of surface reactions will be attained by silicate bedrock  having
a porous nature, e.g., weakly cemented sandstone.

     Slower movement of acidic waters through silicate bedrock will
result in greater assimilation.  Massive igneous beds will shed water
with only a short contact time, while more permeable sandstone beds will
increase contact time.

     Table 4-2 summarizes the assimilation capacity of various bedrock
types.  Surficial geology, including glacial deposits, soils, and
unconsolidated material, has a greater influence on a system's ability
to assimilate acidic depositon.  Bedrock influence on surface water
chemistry is mainly indirect through derived unconsolidated material.

4.3.2.4  Hydrology—

     4.3.2.4.1   Flow paths.   The extent of reaction of the strong acid
components of deposition  with each  component of the substrate  (i.e.,
bedrock, soil)  depends in most cases on the time of contact with that
substrate, thus the flow  path of water is important in determining the
total assimilating capacity  of the  terrestrial  system.   Time of contact
is important because only surface reactions (adsorption,  ion exchange)
occur rapidly for aluminosilicate minerals; slow diffusion processes
control  subsequent reaction  rates.   Reaction rates with carbonate
(bedrock, or in soil)  are rapid; therefore, these areas are not
sensitive to acidic deposition.  Because the groundwater pool  often has
a slow turnover rate (i.e.,  contact time is long), assimilation of  H+
is expected.

     The contact time of  runoff waters in the organic and inorganic
layers of the soil  profile depends  on many factors:   topography (e.g.,
basin slope, soil  thickness), meteorology (e.g., precipitation rate,
season), etc.  These factors affect the degree of soil  saturation, which
in turn determines contact time.

     In areas with snowpacks, contact time is reduced during snowmelt
because of the quick saturation of  the soils by the first stages of
melting.  In areas where  the soil freezes, contact time is even further
                                  4-15

-------
         TABLE 4-2.  BUFFERING CAPACITY OF VARIOUS BEDROCK  TYPES
                   (ADAPTED FROM HENDREY ET AL. 1980b)
    Buffering capacity
             Bedrock type
Low to none
Granite/Syenite or metamorphic
equivalent
Granitic gneisses
Quartz sandstones or metamorphic
equivalent
Medium to Low
Sandstones, shales, conglomerates or
their metamorphic equivalents (no
free carbonate phases)
High-grade metamorphic felsic to
intermediate volcanic rocks
Intermediate igneous rocks
Calc-silicate gneisses with no free
carbonate phases
Medium to high
Slightly calcareous rocks
Low-grade intermediate to mafic
volcanic rocks
Ultramafic rocks
Glassy volcanic rocks
'Infinite1
Highly fossiliferous sediments or
metamorphic equivalents
Limestones or dolostones
                                  4-16

-------
reduced.  In both cases,  the  impact of  snowmelt on runoff (and therefore
on stream and lake)  chemistry is great  (Jeffries et al. 1979,
Johannessen et al. 1980,  Overrein  et al. 1980).  In central Ontario, the
upper 1.0 to 1.5 m of the soil  is  usually frozen each winter (Jeffries
1981, pers. comm.) so spring  runoff may flow over the soil layer or
through only the top few  cm.   In other  areas (e.g., Adirondacks, White
Mountains in New Hampshire),  surface soil layers freeze only when little
snowpack develops during  winter.

     4.3.2.4.2  Residence times.   It is often assumed that headwater
lakes are more sensitive  to acidic deposition than are other lakes
(Gjessing et al. 1976, Minns  1981).  This assumption may arise, in part,
because headwater lakes

     a) often have longer hydrologic residence times than lakes
     downstream, simply because their total catchment area-lake area
     ratio is smaller (hydrologic  residence time is a function of lake
     volume rather than lake  area  so lake morphometry must also be
     considered);

     b) often are at higher elevations  (on a regional basis) and
     therefore have few or no soil deposits in their watersheds; and

     c) often have poorly developed soils in their watersheds.

     Lakes with smaller catchment  area-lake area ratios will usually
receive a greater proportion  of their total input of water via
deposition directly on the lake surface.  The acids in the deposition on
the lake surface have not been assimilated by any other system.  On the
other hand, even in systems with small  watersheds, assimilations of
hydrogen ion in the terrestrial systems can be >^ 50 percent on an annual
basis (Galloway et al. 1980a, Wright and Johannessen 1980, Jefferies et
al. 1981).  Therefore, as the catchment area-lake area ratio increases,
the ability of the overall  watershed (terrestrial catchment + lake) to
assimilate the acidic deposition falling on it increases.

     A long hydrologic residence time is favorable (i.e., makes a lake
less sensitive) 1f a major portion of the ANC that enters the lake
results from internal processes.   If water renewal rate is slow, the ANC
provided by processes such as primary production will build up from
year-to-year rather than  be lost from the lake via outflow.

     In summary, the relative importance of the ANC supplied by internal
processes in a lake vs the acid assimilation capability of the
terrestrial watershed will  determine, for a particular lake, whether a
long hydrologic residence time is  beneficial or detrimental.

4.3.2.5  Wetlands--Very little is  known about the role of wetlands in
assimilating acidic deposition. In addition to neutralization by
alkalinity present in the aqueous  component of the wetland, other
processes may contribute  to assimilation, including 1) reduction
reactions and 2) ion exchange reactions.


                                  4-17

-------
     Reduction reactions (e.g.,  N03-  reduction, S042- reduction,
Fe3+ reduction)  occur in the  aqueous  portion of the wetland under
anaerobic conditions, e.g., under  ice-cover during the winter.  They may
also occur in the sediments,  which are typically high in organic content
and are anaerobic at all times.  The  ANC  produced by these reduction
reactions may, however,  be temporary  (Section 4.3.2.6.2) if the
reactions are reversed when the  water is  oxic, or if the water is
removed (e.g., by evaporation) exposing the sediments to the atmosphere.
Some of the ANC produced is permanent if,  for example, sulfide produced
from SO^- reduction is  stored as  FeS.  In other cases, oxygen
demand in the wetland may be  high  enough  at all times because the high
organic content and relatively shallow water depth keep the aqueous
component anoxic.  The reduction processes may, in these cases, produce
permanent ANC.

     Cation exchange reactions with the sediments or detrital material
in the wetland may result in  significant  assimilation of strong acid if
the BS is appreciable.  However, this is  probably seldom the case.  In
fact, some wetlands, particularly  Sphagnum bogs, have been shown to
produce mineral  acidity  (Clymo 1963)  by means of cation exchange
reactions.

4.3.2.6  Aquatic—The ability of an aquatic system to assimilate acidic
deposition must be considered with respect to time frame, i.e., the
ability of the system to prevent long-term acidification vs short-term
acidification.  This, in turn, depends on  the system's ability to
assimilate acidic deposition  at  all times so that no fluctuations in pH
or alkalinity result in  mineral  acidity at any time, even during major
hydrologic events such as snowmelt or stormflow.

4.3.2.6.1  Alkalinity.  A threshold alkalinity, below which an aquatic
system would have the potential  for becoming acidic to a point where
biological effects might occur,  can be estimated.  The following
material provides support for a  quantitative estimate of such a
threshold.  It should be clearly recognized that the presentation
emphasizes the long-term (years/decades)  concept of lake and stream
acidification.  Also, the computed threshold ignores any acid
assimilation of acidic deposition  by  the  watershed, the importance of
which is noted in the preceding  sections.  As a result, the estimate to
follow represents an upper threshold  limit for systems receiving present
levels of acidic deposition.   Using this  threshold to estimate the
number of systems sensitive to the long-term inputs of acidic deposition
would likely result in an overestimate.   However, this same threshold
would underestimate the number of  aquatic systems that are susceptible
to short-term (days/weeks) acidification.

     The instantaneous ability,  i.e., excluding watershed influences,
of lake or stream water to assimilate acidic deposition is quantita-
tively measured as the ANC or alkalinity  of the water (Stumm and Morgan
1970).  HCOs" is an adequate  measure  of ANC in most lakes because
other contributing species  (e.g.,  ammonia--NH3, borate--B(OH)4~)
are of minor importance.


                                  4-18

-------
     Alkalinity frequently  has been used to assess the sensitivity of
lakes to acidic deposition, and  subjective criteria have been
established to "classify" lakes; e.g., lakes in Ontario were classified
as having extreme sensitivity if 0 to 40 yeq alkalinity £-! was
measured, moderate sensitivity if 40 to 200 yeq alkalinity £-1 was
measured, etc. (Anon  1981).  The boundary between "sensitive" and
"insensitive" that is commonly used is 200 yeq£-l of alkalinity
before the onset of acidification (Hendrey et al. 1980b).  Altshuller
and MacBean (1980) classified lakes as "susceptible" if alkalinity was
measured as < 200 yeq s,'1.  Calcite saturation index (CSI)~~a
measure of the degree of saturation of water with respect to CaC03
(calcite) that integrates alkalinity, pH, and Ca concentration—has
also been used (Kramer 1979, unpub. manuscript; Harvey et al. 1981).  In
another case (Minns 1981),  simple assessment of lake sensitivity has
been based on ionic strength (conductivity), with the unstated
assumption that ionic strength must be a good correlate of alkalinity.

     To understand why an alkalinity of < 200 yeq A'1 was selected
as indicative of sensitivity, it 1s first necessary to explore the
relationship between  pH and alkalinity in oligotrophic systems.  A
typical relationship  between alkalinity and pH for oligotrophic systems
1s obtained by plotting pH  versus alkalinity for 928 streams and lakes
in New York State (Figure 4-3, individual points not shown).  This
graphical relationship between pH and alkalinity will be used in two
ways.  First, it will illustrate dependence of pH changes on the
magnitude of the initial  alkalinity.  Secondly, it will be used to
relate alkalinity changes to pH  changes and subsequent biological
effects.  This linkage is needed because the literature on the chemical
effects of acidic deposition on  aquatic ecosystems uses alkalinity as
the critical variable and not pH.  However, those researchers
investigating biological  changes due to acidic deposition relate the
biological changes to changes in pH and not alkalinity.  Therefore, to
be able to relate changes in alkalinity of aquatic systems due to acidic
deposition to changes in biological systems due to changes in pH, we use
Figure 4-3.

     The logic behind alkalinities < 200 yeq £-1 as an upper
long-term sensitivity limit is as follows:

    o   On a regional basis, the maximum increase of $04, due to
        acidic deposition,  in aquatic systems is  ~ 100 yeq £-1
        (Harvey et al. 1981, NRCC 1981).  Therefore, the maximum
        alkalinity decrease that could have occurred over time is  -
        100 yeq £-1 (see explanation, Section 4.4.3).

    o   Biological effects  due to acidification become apparent as
        alkalinities  decline to  about 65 to 35 yeq £-1 and pH's
        between 6.5 and 6.0 (Figure 4-3; Chapter E-5, Table 5-16).

    o   Given the above two points, if systems with alkalinities
        < 200 yeq £-1 are acidified to the maximum amount
        (Aalkalinity = 100 yeq  £-!), then resulting


                                 4-19

-------
            -100
                            ALKALINITY (peq i'1
Figure 4-3.  The change in pH for a given change alkalinity at two
             alkalinity levels and an example of pH-alkalinity relation-
             ship for aquatic systems.  The alkalinity data were obtained
             by a single or multiple endpoint titrations using a pH meter.
             The solid S-shaped line represents the median values.  The
             dashed lines form a 68% band (analogous to one standard
             deviation).  Each line is a smoothed (cubic spline) moving
             average of five points of the appropriate percentiles  (2, 16,
             50, 84, 98) computed from the data at each 0.1 pH point.
             Adapted from Hendrey (1982).
                                 4-20

-------
        alkalinities will  be <  100  yeq A-lf which is near the
        threshold for biological effects on a long-term basis.  In
        addition, there is a short-term consideration.  If an aquatic
        system that was originally  at 200 yeq x,'1 is acidified to
        100 yeq £-1, there may  be no biological effects on a
        long-term basis, but there  could be some on the short-term
        basis.  During spring snowmelt, alkalinity reductions of > 100
        yeq £~1, lasting several weeks, have been reported
        (Galloway et al. 1980b, Galloway and Dillon 1982).

     This is graphically illustrated in Figure 4-3, where 200 yeq £-1
is represented by Xi.  If  acid  precipitation is added to the system,
the alkalinity could decrease to -  100 yeq fc-1 (X2).  This assumes
that acidification is immediate and that the only chemical effect of the
added acid is the loss of  HC03~.  Thus, it represents an upper limit
(see Section 4.4.3.1.1.1).   Note that this acidification has only
resulted in a small pH change (~ pH 7.4 to 7.1).  However, future
inputs of acidic deposition on  a long-term basis (years/decades) or any
short-term acidification (days/weeks) will probably result in large pH
decreases, with subsequent biological effects, because the system is now
poorly buffered.

     Systems of alkalinities ~  200  yeq £-1 that are exposed to
high current levels of acidic deposition (annual average pH _< 5.0) are
only minimally sensitive.   [Natural pH of precipitation in eastern North
America is most certainly  greater than pH 5.0 due to the absence of
large sources of sulfur and nitrogen oxides (Galloway and Whelpdale
1980).]  However, systems  with  < 100 yeq Jr1 of alkalinity that
are exposed to the same acidic  deposition are very sensitive.  For
example, point X2 represents a  system that started with 100 yeq
£  .  Point \3 represents  the same  system after it is exposed to
current levels of acidic deposition.  The subsequent large decrease in
pH and alkalinity will influence a  large variety of biological
responses.

     In summary, the boundary between sensitive and nonsensitive aquatic
systems is chosen to be 200 yeq £-1 after consideration of (1)
current levels of acidic deposition (annual average pH £ 5.0), (2) the
relationship between pH and alkalinity in oligotrophic systems, and (3)
the pH and alkalinity values at which changes in them will result in
biological effects.  The choice of  200 yeq £-1 of alkalinity
identifies all aquatic systems  possibly sensitive to long-term
acidification as a result  of current levels of acidic deposition but
underestimates those systems sensitive to short-term acidification.

     Although parameters such as alkalinity and calcium saturation index
(CSI) probably provide reasonable estimates of the instantaneous ability
of the lake water to neutralize acidic deposition, they have two
drawbacks:  the interpretation  of degree of sensitivity is subjective,
and the methodology relies on only  a static measure.  The renewal rate
of ANC is undoubtedly more important and ultimately governs a lake's
                                 4-21

-------
sensitivity.  Renewal  rate of ANC  in  a  lake  depends on the rate of ANC
supply from the lake's watershed which,  in turn, depends on many factors
(see preceding sections);  the importance of  direct atmospheric
deposition on the lake surface (since this will have alkalinity < 0)
relative to the total  water budget; and the  internal renewal rate of
alkalinity.  As long as the total  input of alkalinity, including
external and internal  sources, remains  positive, the lake will not
become acidic (i.e., will  not have mineral acidity) because alkalinity
is a conservative parameter.  However,  short-term acidification (e.g.,
at snowmelt) may occur in  lakes that  will never experience long-term
acidification.

4.3.2.6.2  Internal  production/con sumption of ANC.  The internal
production of alkalinity is usually overlooked in considerations of lake
sensitivity, but it may be very important, especially in lakes with low
alkalinity.  In the epilimnion, the major pathway for the production of
alkalinity is primary  production  (Brewer and Goldman 1976, Goldman and
Brewer 1980).  The generation of alkalinity  depends upon the use of
     as a nitrogen source  by algae:

     106 C02 + 16 N03~ + HP042' +  122 H20 +  18 H+  ->

                              + 138 02.                          [4-4]
Although it is well  known that NH4+  is  preferred over N03~
(Lui and Rolls 1972, McCarthy et al. 1977), mass balances of the two
species in most north-temperate lakes are  such that N03" "se
surpasses Nfy  use (Harvey et al.  1981;  Dillon, unpub. studies).
Any NH4  use results in a decrease in alkalinity in the lakewater
(D. W. Schindler, unpub. studies).  However,  a net gain of ANC can
result from NOs- uptake during primary  production since the
inorganic NOo" is converted to organic  nitrogen and stored
permanently in the lake's sediments.

     The uptake of N03~ (corrected for  uptake of NH4+) is
often in the range of 10 to 20 ueq £-1  over the summer in
oligotrophic north- temperate lakes (Dillon 1981).  The net uptake
calculated on a whole-year basis,  on the other hand, may be closer to 5
yeq £-1.  Even this lesser amount may be significant; e.g., in a
lake with mean depth of 10 m, this represents a production of 50 meq
alkalinity m~2 yr~l, an amount comparable  to  the deposition of
strong acids in many parts of eastern North America.

     Therefore, an increase in nutrient levels may increase the
alkalinity generation if N03" is used as the  N-source, on a net
basis, and the organic N is lost permanently  to the sediments.
Fertilization with NHA+, on the other hand, may result in lake
acidification (D. W. Schindler 1981, pers. comrn.).  Nutrient status is
therefore very important in determining  the sensitivity of a lake to
acidic deposition.
                                  4-22

-------
     In most lakes (I.e., those with vernal  circulation), the alkalinity
of the whole lake 1s uniform at the  Initiation of  stratification.   If no
Internal production or use of alkalinity occurred  In the hypollmnlon, It
would serve as a "reserve" pool  for  the  lake, because the external
Inputs of strong acids (atmospheric  deposition,  runoff) enter the
epil Imnion.  However, Internal  processes within  the hypollmnlon may also
deplete or produce alkalinity (Schlndler et  al.  1980, Cook 1981, Harvey
et al. 1981, Kelly et al. 1981,  unpub. manuscript).

     Acidification of lakes by acidic deposition results 1n Increased
transparency (Dillon et al. 1978,  Schlndler  et al. 1980, Harvey et al.
1981, Schindler and Turner 1982, Van 1983).  Therefore, hypolimnetlc
primary production (by phytopl ankton or  perlphyton), and associated
production of ANC, may be elevated relative  to non-acidic lakes of
equivalent nutrient and morphometrlc status.

     Under oxic conditions, respiration  of organic matter (produced
principally in the epilimnion and  metal imnion) results in a decrease in
alkalinity or depletion of ANC:


     C106H263°110N16P1 + 138 °2  *  106 C02  +  122  H2°
                                                                    [4-5]
                         + 16 HN03 + H3P04.

This reaction may occur in the hypol Imnetic  water or at the sediment
water Interface.  As mentioned earlier,  some of  the organic matter
produced in the lake is permanently  stored in the  sediments (i.e.,
respiration < production).

     Under anoxlc conditions, several microbial  processes that occur in
the hypol imnion (or in the surficial  sediments)  and that require organic
material produce alkalinity:
   o
$04   reduction

                         53 so42'  +  106  H+ +
                                                                    [4-6]
            106 C02 + 16 NH3 + H3P04 + 106 H20 + 53 H2S
NQ3~ reduction

     5 C6H12°6 + 24 N03~  + 24 H+ -»• 30 C02 + 12 N2 + 42 H20           [4-7]


Mn4+ reduction

     C106H263°110N16P1  +  236  Mn02 + 472 H+  ->
                  94.                                                [4-8]
            236 Mn2+ +  106 C02 + 8N2 + H3P04 + 366 H20
                                  4-23

-------
Fe3+ reduction
     c106H263°ll()Nl6pl + «4 FeOOH + 848 H+ ->
                                                                    [4-9]
            424 Fe2+ + 106  COg + 16 NH3 + ^04 + 742 H20.
     However, the alkalinity  produced by  some of these processes may
be temporary.  Fe2+ and  Mn2+  (and NH4+) production Is probably
largely temporary, with  the reverse  reaction occurring as soon as oxlc
conditions again prevail  at overturn.  NOa" reduction occurs In
hypollmnla or In lake sediments, but the  M2 evolution makes the
reaction Irreversible; therefore, this represents a source of
permanent alkalinity. SCty2-  reduction results 1n permanent
alkalinity 1f the $2- formed  Is Irreversibly lost to the sediments.
Any S2~ (HS-, H2$) left  1n  the water column at fall circulation is
re-oxidized to S042', with  concurrent loss of alkalinity.

     Therefore, the critical  factor  with  respect to the ability of a
lake's hypolimnlon to assimilate acidic deposition is its oxygen regime.
At the Experimental Lakes Area, Schlndler et al . (1980), Kelly et al .
(1981), and Cook (1981)  studied fertilized and unfertilized lakes that
had anoxic hypolimnia and consequent summer alkalinity production.
Increased S04Z" input resulted  in  increased alkalinity generation.
In Muskoka and Haliburton counties  (Dillon et al., unpub. results) and
in the Sudbury area (Van and  Miller 1982), most study lakes did not have
large anoxic zones in their hypolimnia and $04   loss was not
observed.  Fertilized lakes (Yan and Lafrance 1982) were an exception,
however.

     4.3.2.6.3  Aquatic  sediments.   The potential for lake sediments to
assimilate acidic deposition  is not quantitatively understood.  The same
mlcrobial processes that occur  In hypolimnia occur in lake sediments,
but the contribution of alkalinity  to  the overlying waters is controlled
by slow diffusion processes.

      That sediments also supply ANC by chemical pathways can be
inferred from neutralization  experiments  near Sudbury, Ontario
(Dillon and Smith 1981).  The acidified lakes studied had reduced pH
(of  ~ 4.0 to 4.5) in the upper  5 cm of the  sediments, with pH of 6.0
to 7.0 at greater depth.  Following neutralization of three study lakes
(with  CaC03 plus Ca(OH)2), the pH  of  the upper  sediments increased
to the same levels as the deep  sediments.  Sediment consumption of the
added ANC varied  from 33 to 60  percent of the  total added to the lake.
The  sediments were therefore able  to  supply 0.9 to 3.0 eq m~2 of BNC.
Over the subsequent five years, one of the  three neutralized lakes was
reacidlfied.  The pH of the upper  5 cm of sediment decreased to levels
comparable to those measured prior to  neutralization of the lake.

     The same processes that occur in  soils may occur  in lake  sediments.
Hongve  (1978) has suggested that cation  exchange  in lake sediments may
result  in acidification of lakewater by  Ca2+  exchange  for H+.  He
suggested, however, that the reverse process  will occur with increased


                                  4-24

-------
lake acidity.  These results were demonstrated  in laboratory experiments
only.

4.3.3  Location of Sensitive Systems (J.  N. Galloway)

     Identification of aquatic systems  sensitive to acidic deposition
ideally should take into account all factors  outlined  in Section 4.3.2.
Unfortunately, for most of these parameters,  regional  data are not
available nor do we have a clear understanding  of how  parameters
interact.  The alkalinity of a surface  water  does reflect a combination
of many relevant factors.  Aquatic systems with alkalinity < 200 yeq
£-1 have been defined in Section 4.3.2.6.1 as sensitive to
acidification by acidic deposition.   These systems can be located by
direct analyses of alkalinity over large  areas  or by use of geological
and soil maps to identify areas that will have  aquatic systems with low
alkalinities.  The advantage of the  first method is that the alkalinity
is determined by an actual measurement.   The  disadvantage is that
thousands of measurements have to be made of  lower order streams and
headwater lakes to determine the sensitivity  on a regional basis and
that in the absence of measurements  no  mechanism to estimate the
alkalinity exists.  The advantage of the  second method is that broad
regional determinations can be made, but  the  major disadvantage is that
no fine detail is available. Therefore, the proper way to address this
issue is to use the regional data on bedrock  and soil  characteristics to
determine general areas of sensitivity  and then to follow up with
alkalinity surveys in the regions designated  as sensitive.

     The state of our knowledge is illustrated  with four figures.  Using
bedrock geology as a criterion, Galloway  and  Cowling (1978) made a rough
approximation of sensitive areas in  North America (Figure 4-4).  Their
identification was improved by the addition of  soils and surficial
geology information to determine sensitive systems for eastern Canada
(NRCC 1981; Figure 4-5).  For the soils of the  eastern United States,
McFee (1980) used CEC and percent BS to indicate the areas where soils
would be expected to be sensitive to acidic deposition (Figure 4-6).
Significant portions of the soils within  the  areas designated
"sensitive" or "slightly sensitive"  would provide low  buffering capacity
and therefore would have little ability to neutralize  acidic inputs as
they passed through the soil.

     As a check on the use of soil  characteristics and bedrock geology
as predictors of low alkalinity waters, Hendrey et al. (1980b) using
Norton's (1980) methods, compared surface water alkalinities waters with
sensitivity predicted on the basis of geology on a county by county
(U.S.) basis; they found clear correlations.  Haines et al. (1983b)
surveyed New England lakes and compared alkalinities with predictions
(by Norton) of sensitivity on a drainage  basin  basis;  high correlations
existed.

     As mentioned earlier, instead of using maps of soil characteristics
and bedrock geology to predict areas of low alkalinity, actual values of
alkalinity may be measured and displayed  on a map.  Omernik and Powers
(1982) used such an approach, as is  shown in  Figure 4-7.


                                  4-25

-------
Figure 4-4.   Regions in North America containing lakes  that are potentially
             sensitive, based on bedrock geology, to acidification by
             acid precipitation.  Adapted from Galloway and Cowling (1978).
                                  4-26

-------
       HIGH  SENSITIVITY
            I Granite, granite gneiss,
             orthoquartzite, syenite

       INTERMEDIATE-HIGH SENSITIVITY
             Volcanic rocks, shales, greywacke
            jsandstones, ultramafic rocks,  gabbro,
             mudstone, and metamorphic equivalents
INTERMEDIATE-LOW SENSITIVITY
      Calcareous clastic rocks, carbonate rocks
      interbedded or interspersed with non-calcareous
      sedimentary, igneoussand metaporphic rocks
      Limestone, dolomite and metamorphic
      equivalents
Figure 4-5.   Map of areas containing  aquatic  systems  in  eastern  Canada that are potentially sensitive ,
             based on bedrock  geology and  surficial soils,  to acidic deposition.   Adapted from NRCC (1981).

-------
                                                   LEGEND

                                                NONSENSITIVE SOILS

                                                SLIGHTLY SENSITIVE SOILS

                                                SENSITIVE SOILS
Figure 4-6.   Regions with significant soil areas  that are potentially
              sensitive to acidic  deposition.  Adapted from McFee  (1980),
              See also discussion  of soil sensitivity in Chapter E-2,
              Section 2.3.5.
                                  4-28

-------
 LEGEND


-------
     The map is a useful  presentation of regions where waters of low
alkalinity might be found.   In essence, the map was created using a
predictive technique.   Specifically, existing data on surface water
alkalinity were compiled  and then correlated with geology, soils,
climatic, physiographic,  and human  factors.  These correlations were
used to predict mean annual  alkalinities for regions without data.
There are problems with this predictive technique, that while not
terminal, need to be realized.   First, if the compiled data are not
themselves representative of a region (e.g., if they are weighted
towards small  or large  watersheds instead of a representative mixture)
the resulting correlations  and predictions will also be biased.  Second,
it is difficult to estimate the  errors involved in the prediction.
Third, and as the authors note,  a certain degree of averaging was
required to create a map  on the  scale of the United States.  Therefore,
the ranges cited are for  the mean annual alkalinities of most surface
waters in a given region.  In areas where substantial heterogeneities in
soil, geology, elevation, etc. occur there may be large variations from
the mean.  Unfortunately, sensitive areas generally occur in regions
with large variations  in  elevation  and soil thicknesses.

     Some of these problems can  be  resolved by field testing the
predictions of the method.   However, these field tests will be site
specific.  A good test in one area  would not necessarily mean that all
regions would be as well  behaved.

     We know large regions  in North America contain aquatic systems with
low alkalinity that are presumably  sensitive to acidic deposition.
These regions are found throughout  much of eastern Canada; New England;
the Allegheny, Smoky,  and Rocky  Mountains; and the Northwestern and
North Central  United States (Galloway and Cowling 1978, NAS 1981, NRCC
1981) and parts of the south and east coasts of the United States
(Omernik and Powers 1982).   However, a large amount of more detailed
survey work is required to  determine the levels of alkalinity and the
degree of sensitivity  of  individual aquatic systems.

4.3.4  Summary—Sensi ti vi ty

     The sensitivity of aquatic  systems to acidic deposition depends on
the composition of the deposition,  the total rate of the loading (wet
plus dry deposition),  the temporal  distribution, and the characteristics
of the receiving system.

     Atmospheric deposition is a major source of chemicals to aquatic
systems.  The chemicals supplied by atmospheric deposition in important
quantities include P,  Ca, Mg, S, N, Pb, Zn, H,. Cl, Na, and Cd.  Of
these, the concentrations of Pb, Zn, Cu, Cd, S, NOX, and H, in
atmospheric deposition in eastern North America are apparently directly
controlled by anthropogenic activities (Chapters A-2 and A-3; see also
Sections 4.3.1.5 and 4.6.1.1).   The effects of S on aquatic systems are
independent of type of deposition (wet or dry) or chemical speciation
(e.g., S02, $04).  It  is  the total  deposition of the total element
that controls the effect  in aquatic systems.


                                 4-30

-------
     The effects of acidic deposition on aquatic systems depend upon the
characteristics of the receiving  systems.  Three such characteristics
determine the ability of the receiving  systems to assimilate acidic
deposition.  The characteristics  are size, composition, and hydrological
residence time.  The smaller the  receiving system, the less likely it
will be able to assimilate the  acidic deposition.  The greater the
watershed to surface water ratio, in general, the greater the ability to
assimilate acids.  The ability  of systems to assimilate the acidic
deposition also depends upon the  composition and characteristics of the
soil in exchange.  Small  systems  with calcareous rock, for example, are
much better able to assimilate  the acidic deposition than small systems
with granite bedrock and low CEC, percent BS, and sulfate adsorption
capacity.  The hydrologic residence time is also important.  For,
generally, the longer the acidic  deposition stays in contact with the
system the more it is assimilated. The longer the hydrologic time in
the terrestrial system, the less  the effect of the acidic deposition on
the aquatic system.  The aquatic  systems that tend to be the most
sensitive to acidic deposition  are located in the areas 'downstream of
terrestrial systems that are small, have slowly weathering soil and
bedrock, have short hydrologic  residence times, and are unable to
assimilate the acidic deposition  that falls into them.

     After consideration of the maximum loss of alkalinity that could be
caused by acidic deposition and the alkalinity range where biological
effects begin, sensitive aquatic  systems are defined as those alka-
linities < 200 yeq £-1 (see Section 4.3.2.6.1).  Such systems are
located through much of eastern Canada, New England, the Allegheny,
Smoky and Rocky Mountains, the  Northwestern and North Central United
States and parts of the south and east  coasts of the United States.

4.4  MAGNITUDE OF CHEMICAL EFFECTS OF ACIDIC DEPOSITION ON AQUATIC
     ECOSYTEMS

       The previous sections have laid  a foundation of important
definitions, concepts and characteristics of deposition and receiving
systems.  The following sections  discuss what is known about the degree
of acidification of sensitive systems,  and the methods used to determine
the degree and rate of acidification.

4.4.1   Relative Importance of  HNO^ vs  H?S04 (J. N. Galloway)
           is the more important in  acidification of aquatic systems
for two reasons.  First,  in most areas  impacted by acidic deposition,
atmospheric H2S04 loading exceeds the HN03 loading.  The second
reason that increases the importance of ^$04  relative to HN03 has
to do with the terrestrial  system associated with the aquatic system.
Specifically, for aquatic systems dependencies are controlled by the
ability of the terrestrial  system to retain $042- and N03~.  In
the case of $042-, because its concentration varies little
seasonally or year-to-year in  a given volume of surface water (e.g.,
stream or lake epilimnion), the spatial  variability is the most
important.  This variability is controlled by  the $042- adsorption
                                  4-31

-------
capacity (SAC)  of the soils  in  the  terrestrial  system.  In the
northeastern United States and  eastern  Canada,  the SAC of the soils is
low; thus the $042- concentrations  in the  surface waters are higher
than those in the mid-Atlantic  region of the United States, which has
soils with a higher SAC (Chapter E-2, Section 2.3.3.2; sulfate
adsorption in soils).

     In the case of N03~, regional  variability  is overwhelmed by
seasonal variability.  During most  of the  year,  the hydrologic residence
time in the soil is sufficient  to allow for rapid uptake of NOs"
(Likens et al.  1977).  Of the HQ-$~  released from the terrestrial
system to the aquatic system, most  comes during  periods of high flow
(spring snowmelt, large intense rainstorms).  During these types of
events the rate of nitrogen  transport through the system is faster than
the rate of uptake within the terrestrial  system.  In addition to
constraints of the hydrologic residence times on N03~ transport
through soil systems, a temperature dependency  also exists.  During warm
periods (e.g.,  summer), when biological acitivity is the highest,
N03~ is efficiently retained by the terrestrial  systems.  However,
during colder periods (e.g., winter) there is often a maximum of
     concentrations (Likens  et  al.  1977).
     Therefore, the two periods that allow a larger  flux  of  N from  the
soil system to the lakes with subsequent increased N03" concentrations
are winter base flow and spring snowmelt.  This is illustrated by a
37-month record of the N03- concentration in the outlet of Woods Lake,
a small oligotrophic lake in the Adirondack Mountains, NY, (Figure 4-8)
and by two of the inflows to Harp Lake in Southern Ontario (Figure 4-9).
The NOj values are highest in the winter and during  spring snowmelt
(usually in March and April).  Therefore, during most of  the year,
because H2$04 deposition is greater than HN03 deposition  in  most
areas and because there is greater retention of HN03 by the
terrestrial system, H2S04 is more important than HN03 in  causing
acidification of aquatic systems.

     Due to the increased importance of HN03 (relative to H2S04)
in acidification of the spring snowmelt and the effects of the large  pH
and alkalinity decreases (Table 4-3) on sensitive life forms and life
stages, it is necessary to explore further the relative roles of HN03
and H2S04 during spring acidification.  Galloway et  al .  (1980b)
studied the role of N03 as a source of acidity for Woods  and Panther
Lakes, acidified during the 1979 snowmelt. They found that in the two
lakes alkalinity decreased during snowmelt because of dilution of base
cations (Co) and an increase in HN03 in the lake epilimnion  (Figure
4-10).  Although $04 concentrations changed only slightly 1n Woods  and
Panther Lakes during snowmelt, S042~ still contributed to the
acidification in an Indirect manner, namely, by causing  long-term
alkalinity reductions (as opposed to episodic).  Thus, the episodic
reduction of alkalinity due to NOg is added to the~long-term
reduction in alkalinity due to $04 (see Sections 4.4.2 and 4.4.3) .
Galloway et al. U980b) concluded that the primary cause  of  the
increased N03 concentration was release from the snowpack.   An


                                  4-32

-------
u>
co
   Figure 4-8.  The concentration of N03 in the outlet of Woods Lake, Adirondack Mountains,  NY.   Adapted

                from Galloway and Dillon (1982).

-------
         o>
            2000
            1600
             1200
          •   800
          ro
         o
              400
                Ou
                      HARP INFLOW 3A
2000


1600


1200


 800


 400
                      HARP INFLOW 5
                       ^••••^••fc^™****"""*

                   1976     1977      1978      1979     1980
Figure 4-9.  Nitrate concentration in inflow 3A and inflow 5 to Harp
             Lake, Ontario, for a 4-year period (June 1976 - May  1980)
             Adapted from Galloway and Dillon (1982).
                                  4-34

-------
TABLE 4-3.   MAGNITUDE  OF  pH AND ALKALINITY DECREASES OF LAKES AND
     STREAMS DURING  SPRING SNOWMELT  (GALLOWAY ET AL. 1980a;
                  1980b;  JEFFRIES ET AL. 1979)

Adirondack Mountains, New York, USA
Panther Lake, 1979
Woods Lake, 1979
£ Sagamore Lake, 1979
Southern Ontario, Canada
Harp Lake #4, 1978
Dickie Lake #10, 1978
Paint Lake #1, 1978
Prior
PH
6.6
4.8
6.1
6.6
4.8
5.5
to Melt
Alk
yeq A-l
162
-38.5
28.8
108
-16
61
During Melt
pH Alk
yeq A'1
4.8
4.5
4.9
5.4
4.5
5.0
-18
-42.2
-16.7
8
-32
8
PH
1.8
0.3
1.2
1.2
0.3
0.5
A
Alk
yeq A-1
180
4.0
45.5
100
16
53

-------
.£»

CO
  I
  o?

  cr
  0)
       240
       200
       160
       -80
            m
                   RTOT
MIDWINTER
  THAW
SPRING
 "JAW
                       LEGEND

                	  PANTHER  LAKE

                	 WOODS LAKE
                         JL
                                                 J	L
                      M    A    M    J    J    A
                                     1978
                                                             0    N
        F    M     A   M
             1979
Figure 4-10.  Temporal trends in alkalinity at outlets of Woods and  Panther Lakes.   Adapted from
              Galloway et al. (1980a).

-------
analysis of two additional  snowmelt  periods  (1978, 1980) supports this
conclusion (Galloway et al.  1983b).

     The chemical  changes that accompany the decreases in pH and
alkalinity, however, are not consistent from study area to study area.
For example, Jeffries and Snyder  (1981) found that $042- levels
increase in several  streams  in the Muskoka-Haliburton area of Ontario at
peak flow during snowmelt.   On the other hand, Johannessen et al. (1980)
reported decreasing  $042- during  snowmelt in streams in Norway.
Three of six streams studied by Jeffries and Snyder (1981) exhibited
declining N03~ concentrations associated with peak H+ concentrations,
a finding opposite to that of Galloway et al. (1980b) in the
Adirondacks.

     In sunmary, during most of the  terrestrial biological productivity,
$04^" is the most important  anion causing acidification.  However, in
winter and in areas  with snowpacks,  in the spring N03 can become more
important both 1n an absolute sense  and relative to $042-.  The
effects of H2S04 and HN03, on acidification of aquatic ecosystems
are:

     0  H2S04 causes long-term (decades) alkalinity reductions on a
        regional basis.

     0  HN03 cause episodic  short term (weeks) alkalinity reductions
        that are in  addition to the  long-term reductions caused by
        H2S04.

4.4.2  Short-Term Acidification (J.  N. Galloway)

     Acidification of lakes  and streams during major hydrologic events
has been demonstrated in Norway (Gjessing et al. 1976, Henriksen and
Wright 1977, Johannessen et al. 1980), Sweden (Oden and Ahl 1970,
Hultberg 1977), Finland (Haapala  et  al. 1975), Ontario (Scheider et al.
1979, Jeffries et al. 1979,  Jeffries and Synder 1981) and the
northeastern United  States (Johannessen et al. 1980, Galloway et al.
1980b, 1983c).  The  hydrologic event leading to acidification has usually
been snowmelt; however, periods of heavy rain have resulted in decreases
in alkalinity and pH 1n at least  one case (Scheider et al. 1979).

     Episodic events have resulted in decreases in pH of greater than or
equal to one pH unit in several reported cases (Table 4-3).  For example,
the change in pH of  Harp Lake Inflow #4 during the snowmelt of 1978 was
1.2 pH units (Jeffries et al. 1979)  while the alkalinity decrease was 100
vieq A-l.  During the 1979 spring  snowmelt, the pH and alkalinity
decreases in Panther Lake epilimnion were 1.8 pH units and 180 yeq
A"1, respectively (Galloway  et al. 1980a,b).  Streams in Ontario,
Canada and New York, USA with lower  pre-melt pH's and alkalinitles had
correspondingly smaller decreases (Table 4-3).

     Based on the studies cited above and on other available data sets
(Leivestad and Muniz 1976,  Schofield 1980) it is reasonable to expect


                                 4-37

-------
that the pH levels during  spring  snowmelt may approach pH 4.3 to 4.9.
This is the same pH range  observed for chronic, long-term acidification
(Section 4.4.3;  Pfeiffer and  Festa 1980, Haines et al. 1983b).  The
difference is that in episodic acidification, aquatic systems with pH's
as high as 7.0 can be acidified to pH < 5.0.  While in long-term
acidification, aquatic systems with  pH's of > 6.5 are, on the average,
too well buffered to be acidified to pH < 5.0.

     To predict the importance of episodic events in aquatic ecosystems,
one must be able to evaluate  the  probability of chemical (pH, alkalinity,
aluminum, etc.)  change of  specific magnitude in a lake or stream for a
specified duration.  To construct a  model to make these predictions, one
must know the following:

     a)  the amount of the chemical  of interest and of water stored
         (mass/area in the
         watershed,
         the hydrological  pathway that the snowmelt follows (Section
         4.3.2.4), and the biogeochemical interactions that occur en
         route (e.g., ion  exchange,  biological uptake),

     c)  the volume of the lake that interacts with the runoff
         (determined partly by temperature patterns),

     d)  the chemical composition and amount of the snowmelt relative to
         the composition and  amount  of the baseflow.

     e)  the relative importance  of  natural vs anthropogenic induced
         changes in runoff composition during snowmelt.

     Factors a, d, and e have been measured in a  few studies (e.g.,
Johannessen et al. 1980, Galloway et al. 1982a).  Factors b and c have
rarely been investigated.   Before a  predictive model is available,
long-term (i.e., multi-year)  measurements of all  of these factors are
required in a variety of different geographical locations.

4.4.3  Long-Term Acidification  (J. N. Galloway)

     Affected sensitive aquatic  systems must have three characteristics.
First, they must have waters  with alkalinities <  200 yeq A""1
(Figures 4-4 to 4-7).  Second, they  must receive  acidic deposition  (pH £
4.5; Figure 4-11).  Third, they must be shown to  have been acidified by
acidic deposition.  Overlaying Figures 4-4 and 4-7  (sensitivity to  acidic
deposition) on Figure 4-11 (occurrence of acidic  deposition) suggests
regions that potentially may  have been  impacted.  Documentation that
aquatic systems have been  acidified  (lost alkalinity) by acidic
deposition has been provided  by  three  techniques: (1) analysis of
temporal trends in alkalinity and pH,  (2) paleolimnological analysis, and
(3) investigation of the importance  and  source of $04 in aquatic
systems.

     Studies that have used the  first  technique,  historical  pH/alkalinity
data,  to identify waters acidified by  acidic deposition are  reviewed  in

                                 4-38

-------
Figure 4-11.   pH from weighted average hydrogen concentration
              for 1980 for wet deposition samples (reproduced
              from Barrie et al.  1982)
                               4-39

-------
Section 4.4.3.1.   In  some of  the  studies, problems exist with the
comparison of old data  to recent  data.  The errors associated with the
comparison may preclude an  absolute statement that each of the aquatic
systems has been  acidified.   However, the fact that many studies point to
decreased alkalinity  is strong circumstantial evidence for acifidication
by acidic deposition.

     Supporting this  circumstantial evidence is the analysis of diatoms
in lake sediment  cores. While such analysis has been used successfully
in Scandinavia the technique  is still being developed for use in the
United States (Section  4.4.3.2).

     A technique  for  implicitly circumnavigating the problems of
incomplete or imprecise trend data has been proposed by Galloway et al.
(1983c).  The approach  is based on considerations of solution electrical
neutrality ( zc^  = z a.,-  where  c-j is the normality of the ith
cation and ai is  the  normality of the ith anion).  It is most
applicable to clear water lakes and streams (no organic ions) with no
source of sulfur  in the bedrock of the drainage basins.  Marine aerosol
content corrections should  also be performed.

     The basis for the  technique  is that the concentration of S042~
in clear water lakes  and streams  has increased due to atmospheric
deposition.  With the increase in S042~ has to come an increase in a
positive ion, H+, Ca2+, Mg2+, etc.   If H+ increases, the aquatic
system is acidified (i.e.,  alkalinity decreases).  If the concentration
of Ca or another  non-protolytic cation increases, only, then no loss of
alkalinity occurs.  For example,  Figure 4-12 shows the two extremes with
chemical changes  that can occur to an aquatic system when the $04
concentration increased by  a  factor of five.  In one extreme, the
increase in the $04 anion is  balanced by an increase in the
non-protolytic base cations (Alternative I, Figure 4-12).  In the second
extreme, the increase in $04  is balanced by an increase in H+, which
causes a reduction of alkalinity  (Alternative II, Figure 4-12).  Since
these are extremes, the real  world lies somewhere in between and depends
on the characteristics  of the soil and the hydrologic pathway.  In
sensitive systems (bedrock  and soil with low ANC and short hydrologic
path lengths), Alternative  2  appears to be a closer approximation to the
process that has occurred.  As support of this Henriksen (1982), in an
analysis of long-term time  series for the concentrations of Ca and Mg
over gradients of acidic deposition, concludes the increases in $04 in
lakes are balanced by increases in H+ (^60 percent) and increases in
base cations (£40 percent).  Therefore, in aquatic systems with a
predominant atmospheric source of S042~ and with alkalinities less
than 200  eq  -1, the increases in S042- will cause decreases
in alkalinity, i.e.,  acidification, although the magnitude/significance
of the decrease is dependent  on watershed characteristics.

     The beginning of this  section stated that affected sensitive systems
had three character!sties—they were sensitive, receiving acidic
deposition, and had been shown to be acidified.  Using the information on
temporal trends in Section  4.4.3.1.2 and studies on the role of


                                 4-40

-------
PRE ACIDIC
DEPOSITION PERIOD
                                 BASE CATIONS
                             S04
HC03-
             ALTERNATIVE 1
              ALTERNATIVE 2
ACIDIC
DEPOSITION PERIOD
BASE CATIONS
$04
HC03-
BASE CATIONS
H+
S04
     Figure 4-12.  Two extremes for the response of aquatic systems to a
                   5-fold increase in $04.  The length of the boxes relates
                   to yeq £~1.
                                        4-41

-------
Figure 4-13.   The  response of aquatic systems to atmospheric deposition
              of acidic and acidifying substances from local or regional
              (long-range) sources.  The numbers refer to the following
              references, which conclude that acidic deposition has
              caused acidification of aquatic systems.
              1.   Gordon and Gorham (1960), Beamish et al. (1975),  Dillon
                  et al. (1978)
              2.   Bobee et al. (1982)
              3.   Watt et al. (1979, 1983, Wiltshire and Machell  (1981)
              4.   Davis et al. (1978), Norton et al. (1981b)
              5.   Schofleld (1976c), Pfeiffer and Festa (1980),
                  Galloway and Dillon (1982), Galloway et al. (1983c)
              6.   Shaffer and Galloway (1983)
              7.   Gordon and Gorham (1963)
              8.   A. H. Johnson (1979)
              9.   Burns et al. (1981), Johnson et al. (1981)
              The  letters refer to the following references where
              possible acidification of aquatic systems has been studied
              but  not  found.
              A.   McCarley (1983)
              B.   Schindler and Ruszczynski (1983)
              C.   Logan et al. (1982)
              D.   Zeman  (1973), Feller and Kimmins (1979)
              E.   Melak et al. (1983)
                                  4-42

-------
                                                           IQ
                                                           CD
                                                            I
                                                            t— •
                                                            co
-P.
CO

-------
atmospheric sulfur in aquatic  systems,  Figure 4-13 indicates (by numbers)
the areas that have been  shown  to be acidified.  All numbers fall in
sensitive areas receiving acidic deposition.  In addition to the studies
showing where acidification  has occurred, the letters on Figure 4-13
represent studies where the  atmospheric sulfur has been shown not to have
acidified the aquatic system because of its low concentration.

     The maximum degree of acidification of freshwaters by acidic
deposition depends on the total increase in acid anions (primarily
S042-, Section 4.4.1). For  each peq r-1 increase in S042', the
maximum loss of alkalinity 1s  1 yeq jr1 (Section 4.4.1).  Studies
of $04 in aquatic systems across depositional gradients (Figure 4-2),
and sulfur budget studies for  watersheds and lakes (Galloway et al.
1983c, Dillon 1981) and determination of excess S042' in aquatic
systems    (Likens et al. 1977, Harvey  et al. 1981, NRCC 1981) indicate
that the maximum increase in $64, and therefore maximum loss of
alkalinity in  aquatic systems as a result of acidic deposition, is 100
peq JT • The actual loss  will  certainly be less and will depend on
the avail a- bility of base cations  in terrestrial systems receiving
acidic deposi- tion.  This maximum alkalinity decrease is merely a
boundary condition that can  be compared to measured or estimated degrees
of acidification.

4.4.3.1  Analysis of Trends based on Historic Measurements of Surface
Water QuaTity (M. R. Church) —

4.4.3.1.1  Methological problems with the evaluation of historical
trends.  In assessing the effects of acidic precipitation on the
chemistry of surface waters, investigators have  searched laboratory
records and the literature for historical data with which to compare
present day measurements.  The three water chemistry variables most
widely cited in this regard are pH, conductivity, and alkalinity
(Section 4.2.2).  A discussion of how methodology for their determination
has changed with time and the  comparability of historical and current
data are presented here.

     4.4.3.1.1.1  £H

     4.4.3.1.1.1.1  pH-early methodology--Many of the early measurements
of surface water pH In areas of North America and Scandinavia were made
colorimetrically with acid-base indicators.  Materials  for visual
colorimetry are inexpensive and readily portable and, thus, highly
amenable to use in rugged, remote  field locations, often the  site of
"acidification" problems and studies.   An excellent discussion of
acid-base colorimetric indicators  is  presented by Bates  (1973), who
recommends the works of Kolthoff (1937) and  Clark  (1928) for  even more
exhaustive accounts, descriptions,  and  discussions  of colorimetric
indicator use.

      Acid-base indicators are  weak  acids or bases that  change color with
the loss or gain of a proton (or protons).   Such behavior may be
represented by the simplified  equilibrium  formulation


                                  4-44

-------
     HIn (Color A)  £  In"  (color B) + H+ .            [4-10]

Indicators are used to measure  the pH of an unknown aqueous solution as
follows.  When the  optical characteristics or "color tone" of an unknown
(with indicator added)  match  the color tone of a standard reference
solution (to which  indicator  has also been added), then the two
solutions are assumed  to have the same pH.  Sometimes the color tone of
the unknown  solution  plus indicator is matched with calibrated colored
discs, each indicating a different pH.  The band of pH over which the
color change of an  indicator  is detectable (by a colorimeter or by the
human eye)  is called the transformation range.  For visual color
comparisons using two-color indicators, transformation range is generally
on the order of two pH units  (Golterman 1969, Bates 1973).  As Haines et
al. (1983a) noted the  best results are achieved near the mid-point of the
transformation range of each  indicator.

     The key assumption in indicator use is that identical color tone of
an unknown and a standard  solution of the same temperature to which
indicator has been  added implies identical pH, under some circumstances,
however, as Bates (1973)  explains, this is not true.

     One reason this assumption may be false can be explained with the
aid of the equation
      ,.„.,*„,,+ !.,£_+log JS-                            [4-11]
                                   HIn

where pa^ is defined by Equation 4-2, pKuin is the thermodynamic
dissociation constant of the acid form or the indicator,   is the
fraction of the indicator in the form In, and Yin andYHIp are
the activity coefficients of the dissociated and undissociated forms of
the indicator, respectively.   Color matching (by eye or instrument)
indicates only that the term log a/l-a is the same for the unknown
and the standard solutions.  However, if the activity coefficient ratio
(the last term in Equation 4-11) of the indicator is not the same in both
the standard and the unknown solutions, the pH of the solutions will not
be the same when the colors are identical.  This is called the "salt
error" and can be estimated by comparing the "true" or electrometric
(hydrogen electrode) pH of a series of solutions having different ionic
strengths with their respective values of pH as implied by use of the
indicator (Bates 1973).  Salt  effects can be minimized by adjusting the
ionic strengths of the buffer  solution or the unknown solution so they
are nearly equal.  Such adjustments, however, may cause changes in the
reference or unknown pH, introducing further uncertainties.

     Another potential source  of error is that the addition of an
indicator to a solution may actually change the pH of that solution.
This is most likely in poorly  buffered waters, such as those readily
susceptible to acidification.  To overcome this problem the pH of the
indicator solution can be adjusted  so it is close enough to the pH of the
unknown solution that no pH change  occurs when the two are mixed.  This


                                 4-45

-------
can be accomplished by  a  trial and error technique using portions of the
sample to be determined plus a variety of indicators.

     Bates (1973)  indicates that when the above cautions are observed and
correction or adjustment  is made for salt effects, an accuracy and a
precision of 0.05  to 0.1  pH unit can be expected "in properly
standardized routine measurements of buffered solutions."  It is likely
that colorimetric  determinations of pH made in the field, often under
adverse conditions and  often on poorly buffered solutions, probably
seldom  approach such accuracy or precision (Boyd 1977, 1980; Haines et
al. 1983a).  Indeed, many details (including exact methpdology) of
historical pH measurements made with indicators are often "lost in
antiquity," lending further uncertainty to their reliability.
Fortunately, many  of the  investigators of acidification trends in surface
water pH values appreciate these considerations (e.g., Wright 1977,
Overrein et al. 1980).

     4.4.3.1.1.1.2  pH-current methodology—Today, most pH measurements
are made electrometrically (potent?"ometrically) both in the laboratory
and, with the advent of more reliable portable pH meters, in remote field
locations as well.

     The "practical" or "operational" pH was defined in Equation 4-2 (see
Section 4.2.2.1).   To define standard potentials and set the pH scale,
cells of the following  type are used:

     Pt; H2(g), Soln. X | KCKsatd.)  I reference electrode.      [4-12]

The reference electrode is usually either a calomel or si Tver-siTver
chloride electrode (Bates 1973, Durst 1975), which is a primary cell.
For most day-to-day laboratory measurements and all field measurements
researchers use secondary cells in which the hydrogen gas electrode is
replaced by a glass electrode.  The proper use of commonly available
commercial pH assemblies  (cell plus meter circuitry) has been  discussed
in many books, journal  articles, and  laboratory manuals (e.g., Feldman
1956, Golterman 1969, Bates 1973, Durst 1975, American Public  Health
Association 1976,  Westcott 1978, Skougstad et al. 1979).

     An important potential  source of error in electrometric pH measure-
ments of surface waters is the residual liquid-junction potential.
Li quid-junction potentials arise at the point of contact of the reference
electrode and the  solution being tested.  Liquid-junction  potentials are
a  function of, among other things, the  ionic strength of the solution
being tested.  Therefore, the liquid-junction  potential formed in a high
ionic strength medium (e.g., buffer)  is different from  that formed  in a
low ionic strength medium (e.g., dilute acidification-prone surface
water).  The difference between  these liquid-junction potentials  is the
"residual liquid-junction potential"  (Bates 1973).   Such a potential can
introduce errors on the order  of 0.04 pH  unit  when  ignored in  measure-
ments of dilute precipitation  samples (Galloway et  al.  1979).  This type
of error can be minimized by  equalizing  the  ionic  strength of  the test
and reference solutions.   Three  ways  to do this are  to  1)  add  inert  salts


                                  4-46

-------
(e.g., KC1)  to the dilute test solution (this may Introduce Impurities,
thus altering the pH),  2}  dilute the standard solution (which alters Its
pH--a correction must be applied), or 3) use dilute strong add standards
(these are not normally reliable pH standards--they must be frequently
calibrated by tHratlon)  (Bates 1973, Galloway et al. 1979).

     Another potential  source of error 1n electrometrlc pH measurements
of dilute solutions Is  the streaming potential.  Errors arise when
measurements are made on dilute solutions while they are flowing or being
agitated.  Errors of this sort as large as 0.5 pH unit have been reported
for precipitation samples (Galloway et al. 1979).  To eliminate such
error, measurements should be made only on quiescent solutions.

     Under rigorous conditions 1n a properly equipped laboratory, routine
electrometrlc pH measurements can probably approach, at best, an accuracy
and a precision of +_ 0.02 pH  unit.  Most field measurements of the pH of
dilute surface waters probably have an accuracy and precision of no
better than +_ 0.05 unit.

     4.4.3.1.1.1.3  pH-comparability of early and current measurement
methods—Inasmuch as both colorlmetrlc and electrometrlc measurements
(using secondary cells) are based on operational or practical pH
(designated by the primary pH cell and scale), the methods are directly
comparable.  Attention  has been, and should continue to be, placed on the
limits of reliability of the  measurement methods as discussed above.

     4.4.3.1.1.1.4  pH-general  problems—Independent of the methodology
employed, several factors can influence pH measurements of surface waters
and the use of such measurements to estimate the degree of acidification
over time.  Principal among these factors 1s the variation 1n the pH of
surface waters over relatively short time Intervals.  The most dramatic
and Important "short-term" changes 1n surface water pH values are those
seasonal changes associated with spring snowmelt and ice-out periods,
during which pH may drop sharply due to release of add held 1n Ice and
snow (Wright 1977, Overreln et al. 1980, Galloway et al. 1980b, Hendrey
et al. 1980a).  Surface water pH values during the rest of the year may
be considerably higher  than those during snowmelt.  Obviously, time of
year must be taken Into account when we compare past and present pH
measurements in an effort to  assess acidification.

     Less Important but potentially meaningful effects are pH changes
associated with the uptake and release of 003 and/or HC03~ by
aquatic plants.  Most lakes studied In conjunction with acidification
problems are usually ol1gotroph1c, and these changes are probably small.
Yet another factor to consider (especially In streams) Is the occurrence
of local sources of groundwater high 1n C02«  One method sometimes used
to account for variable COg concentrations 1s to report the pH value
after a sample has been thoroughly agitated to equalize Its C02 partial
pressure with that in the laboratory.  It must be noted, however, that
the CO? concentration in a laboratory can vary considerably from day to
day and 1s nearly always well above that commonly considered to be the
global mean (Church 1980). A nunber of methods may be employed to


                                 4-47

-------
overcome this problem  and  to  insure comparability between laboratories
and within a laboratory on a  day to day basis.  These methods include
equilibrating solutions with  outside air or determining the partial
pressure of C02 in  solutions  or in the laboratory atmosphere.  Better
yet would be to equilbrate all samples by bubbling with bottled air of a
known and standardized C02 content.

     4.4.3.1.1.2 Conductivity

     4.4.3.1.1.2.1   Conductivity methodology--The apparatus for measuring
conductivity consists  of a cell of two electrodes (often platinum)  and a
Wheatstone bridge.   The latter is used to balance the resistance of
standard or unknown solutions in which the cell is immersed.  Solutions
of KC1 are used to  standardize the instrument by calculation of the cell
constant.  Important corrections due to temperature variation are also
required.  Conductivity is routinely reported as ymho cm~l at 25.0 C.
Detailed instructions  for  the measurement of the conductivity of surface
water samples can be found in standard laboratory manuals (e.g.,
Golterman 1969, American Public Health Association 1976, Skougstad et al.
1979).  The precision  of conductivity measurements of surface water
samples seems inversely related to the sample conductivity, with relative
standard deviations being  as  great as 10 percent at levels of conductiv-
ity as low as those often  reported in studies of acidification of surface
waters (American Public Health Association 1976, Skougstad et al. 1979).
Inasmuch as this figure pertains to measurements made under laboratory
conditions it is to be expected that measurements made with portable
battery-powered conductivity  meters in the field would be less precise.

     4.4.3.1.1.2.2   Comparability of early and current measurement
methods--Routine measurements of conductivity are always made with the
type of apparatus described above, so historical and recent data should
be roughly comparable, if  the instrumentation has been properly
calibrated and used.  Data published in the literature concerning
otherwise comparable lakes lying in acidic and unaffected areas show that
acidified lakes tend to  have  higher conductivities (Wright and Gjessing
1976, Dillon et al. 1979), most likely reflecting the higher hydrogen
(and  to a much lesser extent  sulfate and nitrate) ion concentrations
found in those lakes.   Continuous monitoring of some surface waters in
southern Norway has shown  increases in conductivity over a period of
decades coinciding  with decreases  in pH and increases in transparency of
lakes (Nilssen 1980),  all  changes associated with effects of acidic
deposition.

      It must be noted  here that many factors, not just inputs of acids,
may cause increases in the concentrations of dissolved salts, and thus
conductivity, in surface  waters.   In fact, increases 1n conductivity
certainly may be associated with both increases and decreases in pH and
alkalinity.  For this reason  observed Increases in conductivity should
not be  used by themselves to  infer that acidification has occurred.

      4.4.3.1.1.2.3  General  problems--Conductivity can be expected to
vary  seasonally  (e.g., it may be much higher  during snowmelt than at


                                  4-48

-------
other times).  Therefore,  comparison of historical and  recent
measurements to assess acidification should  take Into account time of
year when the measurements were made. Temporary changes 1n conductivity
of surface waters may also occur during rainfall events.  In short, any
factor that alters Ionic concentrations will  alter conductivity.

     4.4.3.1.1.3  Alkalinity.  Procedures  routinely used  to determine ANC
of surface waters have changed significantly over the years, so
estimating acidification as the decrease 1n  ANC with time may be
extremely difficult (Dillon et al.  1978, Ontario Ministry of the
Environment 1979, Zimmerman and Harvey 1979,  Jeffries and Zimmerman 1980,
National Research Council  of Canada 1981).

     4.4.3.1.1.3.1  Early methodology—Historically, acidimetrlc
titratlons have usually been performed to  an  endpoint of  pH 4.5
determined electrometrlcally or to  an end  point determined by a
colorlmetric Indicator (usually methyl  orange) or mixed indicators (e.g.,
bromcresol green-methyl orange). ANC measured in this  way has been
termed total fixed endpoint alkalinity or  TFE (Dillon et  al. 1978,
Ontario Ministry of the Environment 1979,  Jeffries and  Zimmerman 1980).
These procedures can lead to two types of  problems.

     First, the assumption that the endpoint of pH 4.5  Is close to the
equivalence point for tltratlon of  the predominating inorganic carbon
species Is true only for samples with relatively high total Inorganic
carbon content (I.e., ~ 2.5 mM total Inorganic carbon;  Golterman 1969)
and ANC.  For samples with relatively low  total inorganic carbon and ANC
(like those usually considered 1n studies  of surface water acidification)
the tltratlon endpoint should be at a higher pH (e.g.,  near pH 5.0 for
total Inorganic carbon of ~ 0.2 mM) to approximate the  equivalence
point (Golterman 1969, American Public Health Association 1976, Dillon et
al. 1978, Ontario Ministry of the Environment 1979, Zimmerman and Harvey
1979, Jeffries 1980, National Research Council of Canada  1981).  Methods
currently 1n routine use account for this  fact.

     Second, unless detailed notes  have been  kept of titratlons to some
endpoint determined with a colorlmetric Indicator, It may be impossible
to determine exactly what the pH was at the  finish of the titration.  For
example, the indicator methyl orange has a pKa of 3.5.  The transition
range for this indicator 1s usually given  as pH 4.5 to  3.1 (e.g., Bell
1967, Golterman 1969), over which range the  color changes from yellow to
orange to pink to red.  Careful analysts prepare standard solutions of
known pH to which Indicator Is added so they can tell by  comparison to
the sample being titrated precisely when the tltratlon  has reached the pH
that they have a priori selected as the endpoint.  Unfortunately, many
early tltratlon data are accompanied by notations only  to the effect that
such-and such an indicator was used.  In such cases It  may be Impossible
to determine the endpoint pH of the tltratlon.

     4.4.3.1.1.3.2  Current methodology--Determining ANC  of surface water
samples Is now commonly done by acldlmetric  tltratlon to  the (HC03~
- IF) equivalence point (inflection point) of the tltratlon curve.


                                 4-49

-------
This point can be readily determined by using differential electrometric
titration methods or Gran's  (1952)  procedure (see Stumm and Morgan 1981).
ANC determined in this fashion  is  termed  total  inflection point
alkalinity or TIP (Dillon et al. 1978, Ontario  Ministry of the
Environment 1979, Jeffries 1980).

     As discussed previously in Section 4.2.2.3  (and later in Section
4.6.3.2) organic compounds may  contribute  significantly to ANC in waters
low in total inorganic carbon and  of low  pH.  This contribution becomes
important in the pH range below that of the (HC03~ - H+) equiva-
lence point (see Bisogni  and Driscoll 1979, Wilson 1979).  Because of
this fact it is likely that  most TFE alkalinity  titrations fail  to
measure any possible contribution  of organics to the buffering of natural
waters.  Gran's procedure in which the solution  is titrated to quite low
pH values and total ANC determined by linear back extrapolation is able
to account for such buffering,  should it exist.

     4.4.3.1.1.3.3  Comparability  of early and current measurement
methods--As stated by the Ontario  Ministry of the Environment (1979)

     "Almost all past water  quality surveys conducted on Precambrian
     Shield waters have employed a TFE method [note added:  i.e., to
     ~ pH 4.5] and hence, unrealistically  high estimates of acid
     buffering capabilities  will be drawn  from the data.  This fact
     is true no matter whether  an  extremely sensitive potentiometrie
     TFE procedure was used  or  an  insensitive field titration
     (indicator, eye dropper titrant addition, etc.) method was
     employed.  The most  appropriate use  of the  TFE alkalinity data
     obtained from past surveys is (a) to  define water systems which
     are not acid susceptible,  (b)  to.suggest where further sampling
     is warranted, and (c) in the  case of high  quality potentiome-
     tric data, to infer  relative  levels  of susceptibility for lakes
     in the 0-20 mg a~l (CaC03) (0-400 yeq rl)
     alkalinity range. The data base on "absolute" or correct water
     alkalinity values is very  small; there is a great need for
     improving this situation as quickly  as possible."

     Discussions of the differences between TFE  and TIP have been pre-
sented by Zimmerman and Harvey  (1979), Jeffries  (1980), and the National
Research Council of Canada (1981). As pointed out in some detail by the
National Research Council of Canada (1981), a rigorous correction may be
made from TFE to TIP in the  situation where (1)  TFE has been precisely
determined to a known endpoint  pH  and (2)  the total inorganic carbon
concentration of the sample  is  known (or can be  closely estimated).
Unfortunately, in most cases TFE has historically been determined using
colorimetric titration procedures, and total inorganic carbon concentra-
tions are not known.  Conversion of such  values  to accurate TIP alka-
linities has proven to be nearly impossible (Jeffries 1980).  Obviously,
such difficulties exist in comparing past and present data regardless of
whether the data come from Canada, Scandinavia,  or the United States.

     As always, when one  compares  samples  taken  years apart care must be
taken so that short-term  variability in ANC (e.g., due to snowmelt,

                                  4-50

-------
rainstorms, uptake of \\CQ^~ by aquatic plants) will not distort
evaluations of long-term trends.

     4.4.3.1.1.4  Summary of measurement techniques.  Each of the three
types of measurements (i.e., pH, conductivity, alkalinity) discussed here
has something to recommend its continued use  in the study of surface
water 'acidification1.  Conductivity  seems to be the least informative of
the measurements, but it is likely that historical measures of this
variable are the most accurate and consistent (with current data) of the
measures discussed.  Although in comparison to current pH data historical
measures of pH are somewhat unreliable, a relative wealth of pH
measurements exists in comparison to  early data for conductivity and
alkalinity.  As discussed above, early measurements of alkalinity are
often of little use due to procedural problems.  In addition, they are
relatively scarce.  Knowledge of the  alkalinity of surface waters and
changes in alkalinity with time, however, are important considerations in
the study of 'acidification1.

     It is clear that in relation to  historical data no overall best
analytical measurement of surface water chemistry exists for evaluating
acidification of lakes and streams.   The authors recommend that pH,
alkalinity, and conductivity should continue  to be routinely measured in
surface water acidification studies,  taking into careful consideration
the detailed sampling and analytical  techniques outlined in the articles
and manuals referenced above.

4.4.3.1.2  Analysis of trends.

     4.4.3.1.2.1  Introduction.  Numerous studies of temporal trends in
the pH, alkalinity, or conductivity of selected North American surface
waters have appeared in the peer reviewed scientific literature or in
readily available technical reports.  The following is a brief review of
the material presented in these reports and articles.

     In considering each of these studies the critical reader should bear
in mind all of the potential problems of bias (in both sampling and
chemical analysis) that may or may not have been taken into account,
reported, and discussed by the principal investigators.  As an example of
the kinds of problems that may exist  with regard to unbiased sampling,
Figures 4-10 and 4-14 serve to illustrate the kinds of seasonal
variations that may occur in alkalinity and pH at the outlets of
Adirondack lakes.  Not shown in these figures are the kinds of shorter
term variations that may occur over a day due to biological activity or
the longer term variations that may result from extended periods of
either drought or higher than  usual precipitation.  Given the kinds and
ranges of variation that occur it is  clear that significant potential
often exists for sampling bias and resulting  misinterpretation of
observed temporal "differences" in pH or alkalinity.  This potential is,
of course, greatest when data  from two discrete points in time are
compared rather than a more complete  time series of data.
                                  4-51

-------
4-52

-------
     Each of the following reviews presents the  pertinent  information
given by the authors in their original  manuscripts.  The authors may
possess considerably more information concerning their research than they
were able to present in their original  publications.  The location and
evaluation of such unreported information is clearly outside of the scope
of this review.  Only that information  presented in the original
technical report or journal  article is reviewed  here.  In  some cases the
information presented by the original authors does not demonstrate
"beyond a shadow of a doubt" that their sampling was completely unbiased.
But this does not mean then  that their  sampling  was necessarily biased,
and it is not the duty or intent of this reviewer to focus unduly on such
omissions or to speculate irresponsibly on their importance.  Major
critical discussions are presented here only on  important  points of
reasonable debate for which  sufficient  information was presented by the
authors.

     4.4.3.1.2.2  Canadian studies

South-Central Ontario (Beamish and Harvey 1972).

     Beamish and Harvey (1972) were the first investigators to present
evidence of decreases 1n lake pH in North America attributable to acidic
precipitation.  They studied chemistry  changes and loss of fish
populations 1n lakes of the  La Cloche Mountains, an area that has
quartzlte geology and that receives acidic precipitation.  The acidity of
the precipitation is directly attributable to smelters at  Sudbury,
Ontario, 65 km to the northeast.  During the period of their study
(1969-1971) Beamish and Harvey found the pH of rainwater ranged from 3.6
to 5.5 and the pH of melted  snow ranged from 2.9 to 3.8.

     The authors began their study with Lumsden  Lake, a small oligotro-
phic lake 1n a watershed devoid of either human  habitation or industry.
The study was then expanded  to include  a total of 150 lakes in the
region.  For some of these other lakes  earlier (pre 1968) data were
available from studies performed by the Ontario  Department of Lands and
Forests.

     In all of the studies,  samples were taken between April and November
(most often in August and September).   Beamish and Harvey  (1972) measured
pH 1n the field with a Sargent-Welch Model  PBL portable pH meter
standardized at pH 7.0 and 4.0 before and after  each series of readings.
Prior to 1970 they repeated  their pH measurements on shore with a Fisher
Model 310 expanded scale pH  meter.   All  measurements were made promptly
in the field to avoid the kind of pH changes they observed with time
(probably due to C02 degassing).  In studies prior to 1968 the Ontario
Department of Lands and Forests measured pH with a HelHge comparator
(Beamish and Harvey 1972).  No other details of  sampling or analytical
procedures were given.

     Beamish and Harvey (1972) found "little vertical stratification" in
pH in Lumsden Lake and nearby George Lake and only "some seasonal
variation."  Their principal  finding with regard to lake chemistry was


                                 4-53

-------
that for lakes In and to  the east of the La Cloche Mountains pH had
decreased with time (Table 4-4).  For 11 lakes sampled prior to 1961  H+
concentration had increased 10- to 100-fold by 1971.  The average change
in the mean annual  pH for all 22 lakes was minus 0.16 unit.  The authors
found that 26 lakes in a  region just north of the La Cloche Mountains
were less acidic and had  apparently experienced lesser decreases in pH
(Table 4-5).  They attributed these facts, at least partially, to the
presence of outcrops of carbonate-bearing rocks in that area.  The
authors concluded that "the increases in acidity appear to result from
acid fallout in rain and  snow.  The largest single source of this acid
was considered to be the  sulfur dioxide emitted by the metal smelters of
Sudbury, Ont." (Beamish and Harvey 1972).

South-Central Ontario (Beamish et al. 1975).

     Beamish et al. (1975) reported on the relationship between various
fish populations and water chemistry in George Lake, Ontario, for the
period 1967-1973.  In that report they cited evidence for a trend of pH
decrease in the lake.

     Over the period 1968 to 1973 they measured pH electronic trie ally in
the field or in the laboratory within 12 hours of sampling.  From a
regression of 28 such measurements plus one measurement "using a dye
indicator method" in 1961 they arrived at a linear decline in lake pH of
0.13 unit per year, on the average.  The correlation coefficient for this
regression was 0.85.  Discarding the 1961 data point, they arrived at a
linear mean annual  decline of 0.13 with a correlation coefficient of 0.65
(Beamish et al. 1975). In their report they provided no other details of
their sampling methods or analytical procedures.

South-Central Ontario (Dillon et al. 1978).

     As part of a study on the effects of acidic precipitation on lakes
in south-central Ontario, Dillon et al. (1978) collected alkalinity data
for four lakes for which  some historical data existed.  These lakes were
Walker Lake, Clear Lake,  Harp Lake, and Jerry Lake.  Precipitation in the
region has a mean pH between 3.95 and 4.38.

     The authors sampled  Clear Lake three times in the period June-August
1977 and found TIP alkalinities ranging from 2 to 25 (yeq x,"1).
This was a decrease from  a TIP alkalinity of 33 (yeq &"1) reported
for the year 1967 by Schindler and Nighswander (1970).

     Dillon et al. (1978) reported TFE alkalinities (measured
potentiometrically to pH  4.5) of 153 (yeq r*1) for the epilimnion
and 130 (yeq &-1) during  a non-stratified period for Walker Lake in
1976.  These were decreases from TFE values of approximately 180 (yeq
A'1) during 1974 (from unpublished data of the Ontario Ministry of
the Environment) and approximately 400 (yeq £-1) during 1971
(several samples on a single date; Michalski 1971).

    The authors did not find any noticeable differences between the TFE
alkalinities of Harp Lake (137 to 152 yeq r-1) or Jerry Lake (137

                                 4-54

-------
TABLE 4-4.  EARLIEST AND 1971  pH MEASUREMENTS ON LAKES IN AND TO THE
     EAST OF THE LA CLOCHE  MOUNTAINS (BEAMISH AND HARVEY 1972)
Lake
Broker3

Carlyle

David

Free! and

George
Grey3

Johnnie

Kakakise

Killarney

L * F 24

Lumsden

Lumsden II

Lumsden III

Mahzenazing3

Nellie

Norway

O.S.A
Spoon3
Sun fish3

Threenarrows

Township
Attlee

Carlyle

Stalin and Goschen

Klllarney

Killarney
Sale

Goschen and Carlyle

Killarney

Killarney

Carlyle

Klllarney

Klllarney

Klllarney

Carlyle and Hunboldt

Roosevelt

Klllarney

Klllarney
Kilpatrlck and
Humboldt
Hunboldt

Klllarney, Roosevelt,
and Stalin
Date
Sept/61b
Aug/71
May/68b
Aug/71
^ * •
Aug/61b
Aug/71
June/69
SepV71
•»••— r ~ L.
Sept/61b
SepV71
Sept/59b
SepV71
Aug/61b
Aug/71
June/68b
Aug/71
Aug/69
Sept/71
w
Sept/67b
Aug/71
Sept/61b
Aug/71
June/69
OcV71
June 69
Oct/71
Sept/61b
Aug/71
Sept/69
Aug/71
Sept/69b
Aug/71
/61b
SepV71.
Sept/61b
Aug/71
Sept/61b
Apr/71
T * •
Nov/69b
Aug/71
PH
6.8
4.7
5.5
5.1
5.2
4.3
5.2
4.8
6.5
4.7
5.6
4.1
6.8
4.8
6.0
5.7
4.5
4.4
6.0
5.0
6.8
4.4
4.6
4.0
4.6
4.0
6.8
5.3
4.5
4.4
4.5
4.5
5.6
4.3
6.8
5.5
6.8
4.4
5.2
5.2
Avg
annual
change In
pH units
-0.21

-0.13

-0.09

-0.20

-0.18
-0.13

-0.20

-0.10

-0.05

-0.25

-0.24

-0.30

-0.30

-0.15

-0.05

0.00

-0.12
-0.12
-0.24

0.00

                               4-55

-------
                         TABLE 4-4.  CONTINUED


Lake
Tyson9
Unnamed Lake3
(46001I30"NS1°24'W)

Mean of 22 lakes


Township
Sale and Humboldt

Killarney




Date
Aug/55b

June/69
Oct/71



pH
7.4

5.7
5.2

Avg
annual
change in
pH units
-0.16

-0.25

-0.16
aLocated east of the  La  Cloche Mountains.
bpH determined by the Ontario Department of Lands and Forests.
                                  4-56

-------
TABLE 4-5.  EARLIEST AND 1971  pH  MEASUREMENTS ON LAKES NORTH OF THE
           LA CLOCHE MOUNTAINS (BEAMISH AND HARVEY 1972)
Lake
Anderson

Annie

Bear

Brazil

Deerhound

El i zabeth

Frank

Fox

Griffin

Hannah

Han wood

Lang

Leech

Little Bear

Little Hannah

Little Panache

Long

Loon

Plunge

St. Leonard

Township
Merritt

Bevin and Sale

Roosevelt and Dieppe

Foster

Curtin

Foster

Goschen

Gosch

Merrit

Foster, Truman,
Curtin and Roosevelt
Roosevelt

Curtin

Roosevelt

Roosevelt

Truman

Louise and Dieppe

Eden, Waters, and
Broder
Merritt and Foster

Roosevel t

Foster

Date
Aug/60a
Oct/71
/61a
Aug/71
Aug/68a
Aug/71
Aug/67a
Aug/71
Sept/68a
Aug/71
Sept/68a
SepV71
/60a
Oct/71
July/60a
SepV71
Aug/60a
Oct/71
Aug/68a
Aug/71
Aug/67a
OcV71
Aug/68a
Oct/71
Aug/67a *
OcV71
Aug/68a
OcV71
Aug/68a
Aug/71
July/68
May/70
Nov/69a
SepV71
Sept/68a
OcV71
Aug/68a
OcV71
Sept/68a
Aug/71
pH
7.4
6.4
5.6
4.7
6.5
6.3
7.5
6.7
7.0
6.7
6.5
7.5
6.9
5.6
6.1
5.3
7.8
6.7
7.0
6.7
7.0
6.0
6.5
6.8
6.5
6.0
6,5
5.7
7.5
6.5
8.5
7.8
6.5
6.8
6.5
6.5
6.6
6.0
6.8
6.7
Avg
annual
change in
pH units
-0.09

-0.09

-0.07

-0.20

-0.10

+0.33

-0.03

-0.07

-0.10

-0.10

-0.25

+0.10

-0.13

-0.27

-0.33

-0.35

+0.15

0.00

-0.20

-0.03

                               4-57

-------
                         TABLE 4-5.  CONTINUED


Lake
Simon

Spring

Stratton

Walker

WMtefish

White Oak

Mean of 26 lakes


Township
Graham

Merritt

Fo ster

Truman and Roosevelt

Whitefish Indian
Reserve
Tilton and Halifax




Date
Aug/60a
Sept/71
Aug/66a
Oct/71
Sept/68a
Aug/71
Aug/68a
Aug/71
Aug/60a
Oct/71
Nov/69a
Oct/71



pH
6.1
6.4
7.0
6.2
7.0
6.7
6.5
6.3
6.3
6.4
4.2
4.1

Avg
annual
change in
pH units
+0.03

-0.16

-0.10

-0.07

+0.01

-0.05

-0.08
apH determined by the  Ontario Department of Lands and Forests.
                                  4-58

-------
 to 168 yeq £-1)  in 1978 and earlier values reported by  Nicholls
 (1976).

      Dillon et al. (1978) discussed in detail  their analytical
 methodology but  did not give any details of their sampling  procedures or
 any information  on possible short-term variations in alkalinity.

 Halifax, Nova Scotia (Watt et al. 1979).

      Gorham (1957) reported on the chemistry of 23 lakes near Halifax,
 Nova Scotia, sampled in December 1955.  Twenty-one years and two weeks
 later, Watt et al. (1979) attempted to sample  these same lakes to look
 for water chemistry changes that may be associated with sulfur emissions
 from Industrial   sources near Halifax.   They found one lake  to be filled,
 one to be inaccessible, and five to have significant local  disturbances-
 leaving 16 lakes to be compared to the 23 studied by Gorham.

      Watt et al.  (1979) took considerable care to sample in the manner
 Gorham (1957)  used.  They measured pH  with a Fisher Accumet Model 230 pH
 meter before and after sample C02 equilibration with the laboratory
 atmosphere and stated that "since both studies used glass-electrode pH
 meters, the combined error for the pH  differences should be less than +
 0.07"  (Wattetal. 1979).  They also measured  specific  conductivity,  ~~
 alkalinity and acidity, even though the last two variables  were not
 determined by Gorham (1957).

      Watt et al. (1979) performed variance analysis on  the  samples from
 the  16 lakes and found that pH differences associated with geology had
 not changed since the study by Gorham  (1957) but that pH values of the
 lakes did differ significantly from those found in 1955.  They found
 current pH values from 3.89 to 6.17 (before air equilibration).  In 1955,
 pH values in these lakes ranged from 3.95 to 6.70  (before air equilibra-
 tion)  (Gorham  1957).   Watt et al.  (1979)  plotted 1977 pH values vs 1955
 pH values (Figure 4-15) and found that all  points  were below the 1:1
 line, that the pH drop was significant to the   P < 0.001 level, and that
 the  slope was  significantly less than  one (P < 0.001).  They also found
 that conductivity in  the lakes increased significantly  (P < 0.001) over
 the  21-year period.  The authors reported that recent pH data from other
 Nova  Scotia lakes and from lakes in New Brunswick  and on Prince Edward
 Island, when compared with data reported  by Hayes  and Anthony (1958),
 tend to confirm  a trend towards lake acidification in these areas.

     Watt et al. (1979)  did not measure precipitation pH but did note
that mean sulfur emissions from the Halifax metropolitan area were
approximately  double  1n 1977 the amount they were  in 1955.  The authors
concluded that it was "clearly unnecessary  to  look beyond local  sources
 (i.e., to long-range  atmospheric transport)  for an explanation of the
acidic condition of lakes  in the Halifax  area"  (Watt et al. 1979).

Nova Scotia  and  Newfoundland (Thompson  et al.  1980).

     Thompson  et al.  (1980)  reported temporal  trends in the pH of Nova
Scotia and Newfoundland rivers.   In  their report they discussed data

                                 4-59
   409-262 0-83-9

-------
         7.0
         6.0
     £  5.0

      Q.




         4.0
         3.0
            3.0
           4.0
   5.0

pH 1955
6.0
7.0
Figure 4-15.
Relationship between pH values  for 16  lakes  (near Halifax,
Nova Scotia) in 1977 and 1955.   Dashed line  is  line  of no
change; all  values are below this  line and drop in pH  is
significant  to p < 0.001 level.   Slope of least-squares
equation (solid line) is significantly less  than that  of
dashed line  (p < 0.001) indicating greater pH  declines in
in higher pH lakes.  Adapted from Watt et al.  (1979).
                                  4-60

-------
 given  by  Thomas  (1960) for the years 1954-56 and more recent data
 reported  by the  Water Quality Branch of Environment Canada.   The more
 recent data are  stored in the data archive NAQUADAT.

     Three Nova  Scotia rivers were studied—the Tusket River,  the Medway
 River,  and the St. Mary's River.  Samples were taken approximately
 monthly in 1955  (Thomas 1960) and in the years 1965-74.  Samples were
 kept tightly stoppered in the dark, and "the pH's used for comparison
 were measured in the laboratory, at room temperature" (Thompson et al.
 1980).  Thompson et al. (1980) compared the discharges on days of sam-
 pling  to  mean annual discharges and concluded that "although sampling in
 various years was commonly biased toward either high or low  flow, there
 was no consistent relationship between mean pH and such bias ... the
 calculated pH's are reasonable, representative and comparable."  No other
 information was provided on sampling or analysis.  The value of discharge
 weighted mean pH of the rivers decreased from roughly 5.2 to 4.4 (Tusket
 River), 5.7 to 4.9 (Medway River), and 6.2 to 5.5 (St. Mary's  River).

     The  three Newfoundland rivers studied were the Isle Aux Morts River,
 the  Garnish River, and the Rocky River.  Sampling and analysis were as
 for the Nova Scotia rivers.  Although plots of discharge weighted mean
 annual  pH of these rivers over the period 1971-78 appear quite variable,
 the  authors believe that these data together with the data for the Nova
 Scotia  rivers indicate a general steady decrease in pH until 1973 and a
 steady  increase afterwards*  The increase is apparently attributed to
 decreased acid loading to the Atlantic Provinces since 1973  "presumably
 because of changed weather patterns"  (Thompson et al. 1980).  The authors
 presented no appropriate statistical  evidence in support of  any of the
 "apparent" trends.

     4.4.3.1.2.3  United States studies.

 New  England (Maine)  (Davis et al. 1978).

     Davis et al. (1978)  studied 1936 pH readings taken from 1368 Maine
 lakes during the period 1937-74 in an effort to see if they  could find pH
 decreases associated with the acidic precipitation of that area (4.4 < pH
 < 5.0 since at least 1956; Cogbill 1976; Likens 1976).   Samples and data
 were from a variety of sources (Davis et al.  1978)  but apparently most
 samples were taken over the deepest portion of each lake,  near mid-day,
 during  the simmer.  Wallace-Pieman colorimetry was used to  measure pH
 "until  the 1960's";  then  pH was measured with portable meters.  "The two
 methods were found to agree within 0.1  pH units (sic)"  (Davis et al.
 1978).

     The authors noted initially that the mean  pH of  196  samples from
 1937-42 was 6.81 and that the mean for 289 samples from 1969-74 was
 6.09—a 5.2-fold acidity  increase. -They also noted that most of the
 change  seemed to occur in  the early  1950's and that overall  the change
might have been  greater if it had not been for  some cultural  euthrophi-
 cation beginning in  the 1950's.   The  authors  realized that these pre-
liminary results might have been affected  by  regional edaphic differences


                                  4-61

-------
in lake types and also by  differences  1n precipitation  acidity across
the state.  Amounts and seasonal  patterns of precipitation also may have
played a part (Davis et al.  1978).   In an attempt to minimize such
potential  regional  distortions, they analyzed the data by using three
procedures based on H+ concentration changes in individual lakes.

     They found 258 lakes  had pH  readings separated by at least a year.
There was a mean of 2.9 readings  per lake and a mean of 12.7 years
between successive readings  (pairs) for a total of 376 "pairs during the
period 1937-74.

     Procedure I of Davis  et al.  (1978) was as follows.  They used data
pairs to calculate slopes  (H+ concentration vs time) for individual
lakes and then mean slopes from 1937-74.  The mean slopes were added to
obtain a total H+ concentration change for the entire period.  Given a
starting pH of 6.89 (mean  of 123  v.alues 1937-42), the final (1974) pH
would be 5.79, an increase on acidity of 12.6 times.  Using a t-test, the
authors also found that the  mean  annual increase in H+ concentration
based on the mean slopes for each year was significantly different from
zero change with p < 0.0001.  The authors noted, however, that this
procedure more strongly weights data pairs with long time separations,
thus possibly invalidating the use of a t-test.

     The second procedure  Davis et al. (1978) used was to average the 376
single slope values.  This gave a mean of 1.15 x 10-7 M yr-l ^+
concentration change.  By  t-test, this mean is significantly different
from zero at p < 0.1, but  not at  p < 0.05.  If a disproportionately
greater decrease in pH occurred in the 1950's (as the authors
hypothesized), this procedure would give greater weighting to the more
frequent data pairs beginning about that time and would thus overestimate
total change (Davis et al. 1978).

     Procedure III the authors used was to weight each data pair (H+
concentration) slope linearly in  inverse proportion to the time interval
between each reading.  Tliese weighted  slopes were then averaged for each
year that they applied. Using an initial pH of 6.89 in 1937, the authors
noted that pH decreased by 1950 to only 6.83.  By 1961, however, the pH
had decreased to 5.91, so  73 percent of the increase in acidity occurred
in this latter time period.   The  authors believed that this 73 percent
increase 1n acidity was actually  an underestimate for this time period.

     Davis et al. (1978) also discussed some alkalinity data they had for
44 of the 258 lakes cited  above.  These data were from the period
1939-71, a total of 96 values and 52 pairs.  No information was given on
the analytical method(s) used to  determine alkalinity.  Applying their
Procedure I to those data, they obtained a decrease of about 6.34 ppm (as
CaC03; from 11.82 to 5.48  ppm, typically; corresponding to a decrease
of 127 yeq &'1 from 236 to 109 yeq A"1) over the period.  This was much
less than.expected from pH changes from the same period and from observed
relationships between pH and alkalinity.  The authors noted that  "the
discrepancy may be due in  large part to the inadequate sampling and great
variance of the alkalinity data,  including the fact that 67 percent of
the pairs had their initial  member in  1960 or later" (Davis et al. 1978).

                                 4-62

-------
     The authors concluded from their study that between the years
1937-74 H+ concentration in Maine lakes increased about 10"6 M and pH
decreased from about 6.85 to 5.95.  Further,  nearly three-quarters of
this change occurred in the 1950's.  "This is the first demonstration of
a  pH decrease due to acidic precipitation on a large  region of lowland
lakes in the United States" (Davis et al. 1978).

New England (Maine, New Hampshire, Vermont) (Norton et al. 1981a).

     Norton et al. (1981a) measured pH in 94 New England lakes (82 in
Maine, 8 in New Hampshire, 4 in Vermont)  for which historical pH existed
from the period 1939-46.  The lakes sampled were small, oligotrophic-
mesotrophic, and located in forested areas on non-calcareous bedrock.
The recent sampling (1978-80) was done during July-October but not on the
same monthly dates as the historic sampling.   These samples were
collected at 1 m depths, and the lakes were stratified at the time of
sampling.

     The pH values of the recent samples  were measured in the field with
(1) a portable pH meter with combination  electrode, and (2) a Hell ige
color comparitor.  Except for three spurious  cases of low pH lakes, the
authors found that "reasonable agreement  exists  for these two methods,
especially at higher pH's" (Norton et al.  1981a).

     The authors presented their results  in plots of  (1) old colorimetric
pH vs recent colorimetric pH, and (2)  recent colorimetric pH vs recent
electrometrlc pH (Figures 4-16 and 4-17).   They  concluded that their
study "confirms the results of Davis et al. (1978) regarding an overall
decrease in the pH of Maine lakes"  (Norton et al. 1981a).

New England (New Hampshire) (Hendrey et al. 1980b, Burns et al. 1981).

     During 1936-39 the New Hampshire Department of Fish and Game
conducted a biological  survey of waters in the White Mountains of that
state.  Their survey Included measurement of  pH  of headwater streams and
measurement of alkalinity and pH for small  lakes.  In 1979 Burns et al.
(1981) resampled 38 of these waters and made  determinations of alkalinity
and pH (note:   the data for this study were also presented and discussed
by Hendrey et al. 1980b).  Since at least 1955-56 this area has been
receiving precipitation with a weighted annual pH less than 4.5 (Cogbill
and Likens 1974).

     The sampling rationale and analytical  methodology used by Burns et
al. (1981)  were exactly the same as used  In their study of North Carolina
streams.  A detailed discussion of these methods Is presented in that
section of this review.

     Burns et al. (1981)  found that 90 percent of the 38 samples showed a
decrease in pH between  the late 1930's and 1979  (mean pH 6.66 in 1936-39
and mean pH 6.06  1n  1979).   Mean H+ concentration was 0.22 (yeq
r1) in 1936-39 and  0.87  (yeq JT1)  in  the  1979 samples.   A
t-test showed  this  increase to be  significant at the p < 0.02.  "However,


                                  4-63

-------
          8.5





          8.0





      1  7.5





      S  7.0
      i—i
      Q£
      O


      S  6.5
      o

      O


          6.0
         5.5
                                   I	  I
            4.5    5.0    5.5     6.0    6.5    7.0     7.5


                          NEW COLORIMETRIC  (pH)
Figure 4-16.   Old lake water  pH  (colonmetric) vs recent lake water pH
              (colorimetric).  Adapted from Norton et al. (1981a).
                                  4-64

-------
          7.5r-
          7.0
      5  6.5

      o
          6.0
      _
      O
      o
          5.5
          5.0 -
          4.5
                              I	I	I	I
I
             4.5     5.0     5.5     6.0     6.5     7.0     7.5


                            NEW ELECTRODE  (pH)
Figure 4-17.   Recent  lake surface water electrode pH vs recent colorimetric
              pH.
                                 4-65

-------
when the errors associated with comparing  the colorimetric data to the
electrometric data are considered, the difference 1n pH between the
1960's (sic—the authors meant 1930's, Burns pers,,, comm.) and 1979 may
not be significant" (Burns et al. 1981).   The authors had historical
alkalinity values for only five lakes In New Hampshire.  Alkalinity
decreased at all five sites (mean decrease 103 percent of original), but
the authors noted that there were not enough samples to make a valid
statistical comparison.  (See also the review of the North Carolina study
by the same authors for a critical discussion of comparison of their
alkalinity values with historical measurements.)

New York (Schofield 1976a).

     Schofield (1976a)  reported on a 1975  survey of water chemistry and
fish status of 217 Adirondack lakes located at elevations greater than
610 m.  For 40 of these lakest pH data exist from the period 1929-37.
Frequency distribution plots (Figure 4-18) of lake pH for the two data
sets illustrate the apparent pH decrease with time (Schofield 1976a).
During the period September 5, 1974-April  9, 1975 the weighted mean pH of
precipitation on this area on a storm-by-storm basis was 4.23 (range 3.94
to 4.83)  (Schofield 1976b).  Schofield (1976a) did not present any
information on sampling or analytical methodology for pH for the data
sets, stating only that they were "comparable data."

New York (Pfeiffer and Festa 1980).

     In the simmer of 1979 the New York Bureau of Fisheries Lake
Acidification Studies Unit sampled 396 ponded Adirondack waters.  For 138
of these waters historical pH data from the period 1930-34 existed.  As
part of their report on the acidity status of Adirondack lakes, Pfeiffer
and Festa (1980) compared the pH values of these Takes in 1979 to the
values of the period 1930-34.

     The 1979 sampling  was done via helicopter and samples were taken at
a depth of 1 m.  No information was given  on the sampling during the
period 1930-34.  For the samples taken In  1979, pH was determined in the
laboratory, using both a pH meter and a Hellige colorimetric comparitor.
These determinations were made on the samples after each sample had been
equilibrated with the laboratory atmosphere.  The only information given
on the pH determinations of the 1930-34 samples was that the measurements
were made using a Hellige comparitor.

     Pfeiffer and Festa (1980) reported that their colorimetric and
electrometric measurements on the samples  taken in 1979 disagreed
markedly and that the Hellige comparitor consistently overestimated pH
throughout the range of sample values and  especially drastically at the
lower values.  Schofield (1981)  compared Hellige comparitor measurements
to pH meter measurements for similar samples, concluding that agreement
between the two methods was much better than found by Pfeiffer and Festa
(1980) and that the discrepancies found by these authors were due "to
errors in pH meter measurements."
                                 4-66

-------
                20
                10 •-
                             1930's
                        r-n
                                                8
                20
                10
                                  1975
                                                8
                                 NO FISH PRESENT

                                 FISH PRESENT
Figure  4-18.
Frequency distribution of pH fish population status in 40
Adirondack lakes greater than 610 m elevation, surveyed
during the period  1929-37 and again in  1975.  Adapted from
Schofield (1976a).
                    4-67

-------
     To minimize any potential bias In the comparison of pH measurements
over time, Pfelffer and  Festa  (1980) used only colorlmetrlc measurements
In their data analysis.   They  presented their results graphically (Figure
4-19).  They concluded that  "historic readings obtained in the 1930's
were generally higher than comparable current determinations for the same
group of waters.  This reflects a general deterioration of water quality
during the 40-year time  frame  between samplings" (Pfelffer and Festa
1980).  The authors did  not  explain the apparent differences between the
pH distribution of the lakes on which they reported and the pH
distribution of the lakes described by Schofield (1976a).  Like Schofield
(1976a), however, they attributed the observed deterioration of water
quality to the acidic precipitation in the region.

New Jersey (A. H. Johnson 1979).

     Searching for evidence  of temporal trends, A. H. Johnson (1979)
examined 17 years of pH  data for two small headwater streams (McDonalds
Branch and Oyster Creek)  1n  the New Jersey P1ne Barrens.  Precipitation
In the area had a mean pH of 4.4 1n 1970, 4.25 for seven months in 1971,
and 3.9 from May 1978 to April 1979.  Nearly all of the data for the study
came from two sources: U.S.  Geological Survey sampling and analyses from
1963-78 and a University of  Pennsylvania trace metal study 1n 1978-79. The
US6S samples were collected  randomly with a frequency of 2 to 12 per year.
This sampling was not biased seasonally for McDonalds Branch but was
slightly biased consistently throughout the study towards a greater
representation of Spring samples for Oyster Creek.  The  University of
Pennsylvania samples were collected weekly in McDonalds Branch only from
1978 through 1979.  Johnson  presented little information on sample pH
analyses except that "all pH values were measured with a glass
electrode."

     Johnson (1979) had  varying levels of confidence in the pH data.
Those data he considered most  reliable were from samples on which cations
balanced anions within 15 percent and calculated conductance balanced
measured conductance within  15 percent.  He performed regressions of
stream pH vs time for different groups of data  (Table 4-6 and Figure
4-20) and found for most groups that a significant decrease existed.
Johnson noted no evidence that oxidation of geological sulfides, changes
in land use, or changes  1n the amount of precipitation were responsible
for the long-term trends. He  concluded "it appears that the decrease 1n
stream pH is a real phenomenon and not attributable to differences or
bias in sampling or measurement.  The data collected to date are
consistent with the post illation of an atmospheric source for the
increased H+."

Pennsylvania (Arnold et  al.  1980).

     In an effort to assess  temporal changes 1n pH and alkalinity of
Pennsylvania surface waters, Arnold etal. (1980) examined five existing
water quality data bases. Nearly all of the data examined were from
streams.  Arnold et al.  found  314 Instances where data were taken at
least one year apart at  the  same location or "sufficiently close


                                 4-68

-------
5.5
6.0         6.5         7.0          7.5

        DETERMINED COLORIMETRICALLY  (pH)
8.0
 Figure 4-19.   Cumulative  comparison of historic and recent pH values for a
               set of 138  Adirondack lakes.  Adapted from Pfeiffer and
               Festa (1980).
                                   4-69

-------
  TABLE 4-6.   REGRESSIONS OF STREAM pH ON TIME: N IS THE NUMBER OF SAMPLES,
    r IS THE  CORRELATION COEFFICIENT, AND P IS THE LEVEL OF SIGNIFICANCE;
        an  AND ai ARE COEFFICIENTS IN THE REGRESSION pH = an + aix,
     WHERE  x  IS THE NUMBER OF MONTHS AFTER JUNE 1963 (A. H. JOHNSON 1979)
     Data source
                                      A yeq H+
                                      per liter
                                       (1963-
                                        1978)
USGS data, 1963-78

USGS data + UP data3

USGS data, anion equiva-
  lents balance cation
  equivalents; measured
  and calculated specific
  conductances are equal
All USGS data

USGS data, anion equiva-
  lents balance cation
  equivalents; measured
  and calculated specific
  conductances are equal
 McDonalds  Branch,
 New Jersey Pine
 Barrens

 90   4.42    -0.0022

100   4.49    -0.0030

 36   4.35    -0.0012
 Oyster Creek,
 New Jersey Pine
 Barrens

 78   5.10  -0.0047

 26   4.89  -0.0027
-0.22    0.05

-0.32    0.01

-0.29    nsb
-0.56   0.01

-0.53   0.01
+57

+80

+29
+48

+26
  ^Includes all data collected by the U.S.  Geological  Survey  (USGS)  from
   1958 to 1978 and the monthly average pH  of University of Pennsylvania
   (UP) samples.

  bNot significant.
                                    4-70

-------
  6


  5


  4


^6


  5


  4
                I

            OYSTER CREEK
            •
                      A
                              •   .
                I            I
            MCDONALDS BRANCH
              1960
                                1970
1980
Figure 4-20.
       Stream pH 1979.  Closed circles represent samples in which
       anion and cation equivalents balanced and calculated and
       measured specific conductances were equal.  Open circles
       are samples for which the chemical analyses were incomplete
       or for which discrepancies in anion and cation and con-
       ductivity balances could not be attributed to errors in pH.
       The closed triangle represents the average pH determined in
       a branch of Oyster Creek in a 1963 study.  Open triangles
       are monthly means of pH data collected weekly from May 1978
       to January 1979 during a University of Pennsylvania trace
       metal study.
                                  4-71

-------
(generally within one mile with no major tributaries or Influences
between)."  Of these  314 cases, 107  (34 percent) showed decreases 1n  pH,
alkalinity, or both.  The mean pH of the "earliest" of these 107 cases  of
decrease was 7.31 (range 5.8 to 8.8), whereas the mean pH of the "most
recent1 was 6.94 (range 4.9 to 8.3).  The mean change 1n pH was a de-
crease of 0.37 unit,  and the range of change was -1.3 to +0.2 units.  For
alkalinity, the mean  of the "earliest" samples was 834 (yeq a"1)
(range 100 to 4000 yeq  A'1), and the mean of the "most recent" was
532 (yeq jr1) (range  40 to 3720 yeq  A"1).  The mean net
change was a decrease of 302 (yeq &'1) and the range was (-2100 to
+360 yeq A"1).  The average time span between the "earliest" and
"most recent" samples was 8 1/2 years; the range was 1 to 27
years.  Arnold et al. (1980) concluded that "although the data upon which
this report Is based  are not sufficiently strong to define statistically
valid relationships,  1t seems clear  that there Is a definite overall
trend toward Increasing acidity 1n many Pennsylvania streams ...."

     Although the authors presented and discussed the means and ranges  of
pH and alkalinity decreases for those cases where decreases were found
(34 percent of the total), they did not present or dfscuss the overall
changes for the 314 total cases examined.  If 34 percent of the total
cases decreased, then 66 percent must have remained the same or
Increased.  This, plus  the fact that five separate data bases were used,
that very little Information was presented concerning sampling, and that
no Information was presented about analytical procedures gives rise to
some serious questions  concerning this study.  Also of concern Is the
fact that decreases over a period as short as one year are considered
part of a "definite overall trend" (Arnold et al. 1980).  Yet another
consideration 1n studies such as this has been noted by Schofleld (1981);
"It 1s obvious that detection of significant, long-term pH changes 1n
acidifying systems, still In a bicarbonate buffered state, cannot be  made
reliably because normal metabolism Induced changes 1n C02 levels would
likely obscure any pH change resulting from decreased alkalinity.  Thus
Interpretations of long-term pH changes 1n the range of 6-7 must be
viewed with caution."  Given the possible variability (not to mention
potential bias) In data taken from a variety of sources (perhaps arrived
at by a variety of procedures), the mean decreases In pH and alkalinity
of only those cases that did decrease In the study seem not so profound.
They may, Indeed, only  represent Inherent scatter In such a data set.  To
cite these as evidence  of a "definite overall trend" (Arnold et al.
1980)  seems premature.

North Carolina (Hendrey et al. 1980b, Burns et al. 1981).

     In the period 1961-64 the North Carolina Division of Inland
Fisheries measured the  pH and alkalinity of a number of North Carolina
headwater mountain streams.  Burns et al. (1981) resampled 38 of these
streams In 1979, attempting to discern any changes 1n stream chemistry
that might have occurred In association with the acidic precipitation
that falls 1n the area  (weighted annual pH 4.7 to 5.2 1n 1955-56 and
< 4.5 In 1979).  The  data discussed  by Burns et al. (1981) were also
presented and discussed by Hendrey et al. (19805).


                                 4-72

-------
     Burns et al. (1981) used detailed maps to resample at exactly  the
locations of the original samples.  The authors considered the  possible
sampling bias inherent in representing by a single sample  the chemistry
of a stream "where pH could fluctuate daily as well  as  seasonally.   It
was assumed that daily and seasonal  fluctuations were random and  normally
distributed if the new samples were  taken during the day and at the same
time of year as the previous ones."

     Significant differences existed in the analytical  methods  used for
the early and recent data sets.  For the 1961-64 samples,  pH was  measured
with a Hellige colorimetric kit and  alkalinity was determined by
titration to a colorimetric (methyl-orange)  endpoint.   For the  1979
samples, pH was determined electrometrically and alkalinity by  Gran's
plots.  The authors compared pH measurements by Hellige kit to  those with
their pH meter and found that they "agreed to within +  0.15 of  a  pH  unit"
(Burns et al. 1981).  To correct for the possible ove7titration of
alkalinity (past the true equivalence point to some  arbitrary endpoint)
and thus the overestimation of alkalinity for the 1961-64  samples,  the
authors subtracted 32 (yeq jrl) from each of the historical values.
Unfortunately, this may not have been a valid procedure.   Crude
calculations using the mean pH and alkalinity of the 38 early samples
indicate that the equivalence point  for the alkalinity  titrations is near
pH 5.0.  A correction of 32 (peq jr1) assumes that the  actual
titration endpoint was at pH 4.5. No records exist  to  indicate that this
was the case (Burns, personal communication).  All  that is known  is  that
a methyl-orange technique was used.   The pKa of methyl-orange is  3.5 and
its transition range (yellow-orange-pink-red)  is roughly pH 4.5 to 3.1
(Bell 1967, Golterman 1969).  Not only is it impossible to know what
color (and thus pH) the original analyst used as an  endpoint, it may be
that a color transition is not even  observable at the pH 4.5 point
assumed by the authors.  If, for example, the titration endpoint  for the
early samples was actually pH 4.0, a correction of 100  (peq A'1)
would be required.  This level of uncertainty in the correction between
techniques is quite large compared to the mean differences in alkalinity
found by the authors--!46 (yeq £-1)  in 1961-64 compared to 80
(yeq £"i) by Gran's titration in 1979.  This would seem to cast
doubt upon the authors' findings that "the decrease  in  alkalinity between
the 1960's and 1979 was statistically significant at the 0.02 probability
level using a t-test"  (Burns et al.  1981).  The authors did not find a
significant temporal  trend in pH (mean 6.77  in 1961-64  and mean 6.51 in
1979).

Florida (Crisman et al. 1980).

     Crisman et al.  (1980)  reported  pH changes in 13 poorly buffered
oligotrophic lakes (known as the Trail  Ridge lakes)  in  northern Florida.
They monitored the lakes quarterly (1978-79)  and found  a mean annual pH
of 4.98.  The mean annual  precipitation pH at the time  of  the study was
4.58.  "Comparison of the present data with  that collected over the past
20 years indicates that the mean pH  of the Trail  Ridge  lakes has declined
an average of 0.5 pH  units (sic)  since 1960"  (Crisman et al. 1980).  The
                                 4-73

-------
authors neither presented further information on their sampling or
analytical  methods for pH,  nor  did they present any historical data or
their sources for such data.

California (McColl 1981).

     The San Francisco Bay  area of California receives part of its water
supply from two Sierra Nevada reservoirs—Pardee and Hetch Hetchy.  These
reservoirs are located in an area underlain principally by Mesozolc
granite and are subject to  acidic deposition resulting from NOX and
SOg pollution generated in  the  Bay area (McColl 1930, 1981).  Measure-
ments of pH have been made  weekly in untreated reservoir outlet waters
for the two reservoirs since 1954. Alkalinity has been measured weekly
(by titration to a pH 4.5 endpoint) in Pardee outlet water since 1944.
McColl (1981) reported on results of analyses of these data up to the
year 1979.

     McColl (1981) performed linear regressions of both the pH data (as
annual average H+ concentration)  and the alkalinity data vs. time.  The
results of the regression analyses are shown in Figures 4-21 and 4-22.
The Increases in (H+) and decreases in alkalinity are clear.  Further
analyses by McColl showed that  (1) mean annual (H+) of the two
reservoirs was correlated (r =  0.51, p < 0.02), (2) that rates of
Increase of (H+) did not vary significantly on a seasonal basis, and
(3) yearly precipitation did explain a small percentage of the variance
in mean annual (H+) of the  release water but that time was by far the
most important factor.

     McColl (1981) considered the possible Influence of logging and
mining within the reservoir watersheds on the observed trends in (H+)
and alkalinity, concluding  that these activities could not account for
the trends.  He similarly considered and dismissed as unimportant the
possible effects of atmospheric increases 1n C02.

     McColl (1981) concluded from his analyses "It Is clear that the
(H+) of waters in both reservoirs has Increased since at least 1954, 1f
not 1944.  On the basis of Indirect evidence and correlative data
discussed ... I conclude that the most likely cause is the Increased
acidity of atmospheric depositions, especially those resulting from
emissions of nitrous oxides by  automobiles."

4.4.3.1.3  Summary—trends  in historic data.  Numerous studies have
examined temporal changes in surface water chemistry 1n areas that have
"sensitive" terrain and that receive precipitation more acidic than pH
4.7.  A consistent (and sometimes major) drawback of these studies 1s a
lack of clear documentation of  the "historic" data used.  Often It is
unproven that these crucial data are unbiased, either by sampling or by
the analytical procedures used.  Many authors recognize this problem;
Davis et al. (1978) stated of their work, "the unconventional and
imperfect means which we used to reconstruct the pH history of Maine
lakes were made necessary by the deficiencies of the only data set
                                  4-74

-------
               200
               160
               120
               80
               40
                       LEGEND    •
                  - o HETCH  HETCHY
                    • PARDEE
6.7


6.8

6.9

7.0

7.1
7.2
7.3
7.4
7.6
7.8
                   1955   1960  1965   1970   1975  1980

                                 YEAR
Figure 4-21.   Increasing acidity at Pardee and Hetch  Hetchy,  shown  by
              hydrogen ion activity vs  year,  for the  period  1954-79.
                                 4-75

-------
           en
           ro
          o
          o
           (O
          o
20


18


16


14


12


10
                                               i    i
                1945 1950 1955 1960 1965 1970 1975 1980


                                 YEAR
Figure 4-22.  Decreasing alkalinity at Pardee, shown by alkalinity as
                    vs year, for the period 1944-79.
                                  4-76

-------
available."  However,  In every case reviewed here, the scientists who
performed these studies concluded that pH and/or alkalinity decreased 1n
at least some of the waters studied.

     Surely In regions where acidic substances are deposited there exist
lakes and rivers (1n otherwise undisturbed watersheds) that have not
experienced recent decreases 1n  pH or alkalinity.  Such an occurrence,
however, Is not a valid argument against the phenomenon of surface water
acidification.  The body of evidence In toto from the studies reviewed
above Indicates that in some regions of acidic deposition, otherwise
undisturbed lakes and  streams are being acidified.  Particularly
noteworthy by Its absence Is any body of data Indicating consistent
decreases In alkalinity or pH of surface waters In otherwise unaffected
regions (see Section 4.4.3.3) jiot receiving acidic deposition.

     This reviewer Is  unaware of any natural process that would cause
decreases In pH and/or alkalinity at the rates Indicated by the studies
(of apparently otherwise unaffected regions) reviewed here.  Until
appropriate evidence 1s presented In support of some such natural process
or until some other reasonable explanation of the data presented above Is
put forth, the only logical  conclusion Is that acidic deposition 1n these
regions Is causing acidification of some surface waters.  Furthermore, 1t
Is only reasonable to  assune that In regions of similar sensitivity that
receive similar levels of acidic deposition other surface waters are
being acidified.

4.4.3.2  Assessment of Trends Based on Paleollmnologlcal Technique (R. B.
         Davis and u.  5. Anderson)—

     To assess the Impact of acidic precipitation and associated
pollutants on lake ecosystems, scientists have begun to analyze the
record contained In the lake sediment (Norton and Hess 1980, Davis et al.
1980). The sediment contains a diversity of physical, chemical, and
biological evidence which starts thousands of years ago deep In the
sediments and proceeds upward toward the sediment surface to cover the
period of the Industrial revolution and recent technological activities.
By applying paleollmnologlcal techniques Including the dating of the
sediment (B1rks and Blrks 1980), researchers can reconstruct
chronological sequences of pollution Inputs to lakes (e.g., lead) and
responses of the lake  biota (e.g., plankton).  Among the specific studies
being carried out Is the Identification and enumeration of the many kinds
of diatom remains (their siliceous shells) preserved In the sediments.
Diatoms are sensitive  Indicators of water pH; the various species differ
1n that each 1s more or less restricted to a different pH range.  By
careful study of these pH relationships for present-day diatom
assemblages, 1t Is possible to calibrate the sedimentary diatom record so
that the past pH of lake waters  can be Inferred.  Thus, a dated record of
lake acidification can be constructed by studying sediments cores.

     The paleollmnologlcal  approach 1s useful for assessing the Impact of
acidic precipitation,  because for the vast majority of acidification-
susceptible lakes no record of past, direct pH measurement exists.  Where
such direct data exist, they (1) postdate 1920, (2) are usually only for

                                 4-77

-------
one year or a short series  of consecutive years, (3) are ordinarily only
for mid-summer when pH's are highest, and (4) are usually made by
colorimetric pH indicators  (pre    1965) which themselves may alter the
pH of poorly buffered  waters.  Paleolimnological pH reconstruction
provides a nearly continuous record based on a  single technique, from the
present to the past (including pre-1920).   It solves the difficult
problem of the direct  sampling of  short-term variation in pH by
integrating daily, seasonal, and annual variation in single sediment
samples encompassing an entire year or  small nunber of years' deposits.

4.4.3.2.1  Calibration and  accuracy of  paleolimnplogical reconstruction
of pH history^Davis  et al.  (1983) have been calibrating the sedimentary
en atom record of pH by deriving "transfer functions" (Webb and Clark
1977) from the study of subfossil  diatoms in surface-sediments (uppermost
0.5 or 1.0 cm) from the deepest part of 31  lakes in northern New England
and 36 lakes in Norway.  Davis et  al. have  developed regression equations
relating these subfossil diatom assemblages to  pH of the surface waters
in the lakes.  The regression coefficients  are  used as transfer functions
to infer down-core pH. These regressions have  standard errors (se)
ranging from + 0.23 to + 0.54 pH units.  The errors for the New England
data are greater than  th"ose for Norway, partly  because the pH readings in
New England are limited to  mid-simmer.  Regressions on Nygaard's (1956)
alpha index, based on  Hustedt  (1937-39) pH  preference categories, provide
less precise pH inferences, especially  for  lakes pH < 6.2.  This probably
is a result of the semi-quantitative nature of  HustecTt's categories, the
uncertainty in assigning individual taxa to categories and possibly
additional uncertainties.   Several factors  responsible for variance in
the surf ace-sediment data sets would have remained more or less constant
at any given lake during the  past  two or three  centuries.  For example,
elevation and lake morphometry would have been  constant, and
concentrations of certain elements in the water (e.g., K and Cl) are
likely to have changed little.  Thus, any relative changes in pH inferred
down-core at individual lakes are  probably  more accurate and precise than
the regression statistics for the  surf ace-sediment data would suggest.

4.4.3.2.2  Lake acidification determined by paleolimnological
recons true 11 on".  Quantifying  this  pal eol imnological approach and applying
it to lakes affected by acidic precipitation are quite recent techniques.
The methods are time-consun ing, and  pH  reconstructions have been
completed for only a small  number  of lakes. In southern Norway, recon-
structions for seven acidic (pH <  5.7)  lakes indicate that acidification
started between 1850 and 1930  (different dates  at different lakes) and
that the total decrease  in  pH by 1980 was 0.06  to 0.83 units (depending
on lake; average decrease 0.40 units)  (Davis et al. 1983). Before this
acidification, these lakes  were "naturally" all quite acidic (pH 5.0 to
6.0) and were highly susceptible to  further acidification.   In southern
Sweden, Aimer et al. (1974) estimated a pH  decrease from "about 6.0 to
4.5" for Stora Skarsjon occurring  between 1943  and 1973.  Also in
southern Sweden, Renberg and Hellberg  (1982) report for Gardsjon a pH
decrease from 6.1 to 4.5 starting  in the 1950's; in Harvatten a decrease
from 5.9 to 4.1 (no dates); and  in Lysevatten from 6.2 to 5.3 (no
starting date) until liming occurred in 1974.


                                  4-78

-------
     The results for the northeastern United States are, so far, less
clear.  Reconstructions  for 6 acidic lakes in northern New England (Davis
et al. 1983)  indicate that acidification started between 1900 and 1970
(different dates at different lakes) and that the total decrease in pH by
1980 was 0.20 to 0.35 units (depending on lake; average decrease 0.26
units).  However, in at  least three cases the pH decrease may have
resulted in part from a  recovery  from an earlier, mild eutrophication
(and elevated pH) associated with limbering or other disturbance.  Del
Prete and Schofield (1981)  report a pH decrease of 0.6 units for another
Adirondack lake, but this is based on only one sample (0 to 1 cm vs all
deeper samples), and the deeper sediment is not dated.  In two other
Adirondack lakes, the diatom record indicated no pH change.

     To clarify these relationships, Davis et al. are continuing
paleolimnological studies of pH change in acidification-susceptible lakes
in the northeastern United States.  An impediment to this research in the
United States is the scarcity of  recently well-monitored (for pH)
acidification-susceptible lakes.  This hampers efforts to develop more
precise transfer functions for inferring past pH.

4.4.3.3  Alternate Explanations for Acidification-Land Use Changes (S. A.
         Norton)—

     Land use changes may directly affect the pH (and related chemistry)
of surface waters in a number of  ways, including variations in the
groundveter table; accelerated mechanical weathering or land scarifica-
tion; decomposition of organic matter; long-term changes in vegetation;
and chemical  amendments.  Details of each are presented below.

4.4.3.3.1  Variations in the groundwater table.  The water table in
mineral or organic soils generally marks a transition from aerobic to
anaerobic conditions. This transition is particularly sharp in
saturated, organic-rich  soils.  With a lowering of the groundwater table
due to drought, lowered  lake levels, or drained terrestrial systems
(e.g., bogs), previously anaerobic and reduced material is exposed to
oxygen.  The following types of reactions may occur:

     FeS2 + 02 + H20 * Fe(OH)a or FeO(OH) or Fe20a + 2H+ + S042"

             02 + H20 ->  Mn02 + H2X

     Organic matter + Decay •*• NOs" + H+ + C02

The associated H* production is commonly accompanied by accelerated
loss of cations from the ecosystem (Likens et al. 1966, Damman 1978).

4.4.3.3.2  Accelerated mechanical weathering or land scarification.
These processes may result from logging, fires, slope failure, and other
disturbances of the land surface. The exposure of relatively unweathered
material to chemical weathering results in accelerated leaching of
cations from watersheds.  If  uptake of nitrogen from decaying organic
material occurs, the pH  of surface waters may rise along witn cation
concentrations.  This results  in  eutrophi cation trends in downstream

                                 4-79

-------
waters (Pierce et al.  1972).   Readjustment of the system may take
decades, with concurrent long-term changes in surface water chemistry,
including pH.

4.4.3.3.3  Decomposition of organic matter.  Long-term trends in the net
production of biomass  result  in  near steady-state chemistry for aquatic
ecosystems but, depending on  the direction of the trend (aggrading or
degrading), the water  quality parameters have different values (Nilsson
et al. 1982).  The net loss of organic matter generally  results in
accelerated production of nitric acid, C02, and increases in
cations—all other conditions being kept the same.  However, a change in
stored biomass is generally accompanied by other changes such as changes
in canopy interception of aerosols, changes in evapotranspiration, or
changes in surface water temperatures, so the individual effects are
difficult to sort out.

4.4.3.3.4  Long-term changes  in  vegetation.  Long-term changes in
vegetation bring about various physical and chemical changes in the soils
and watershed which result in long-term changes in  surface water
chemistry.  For example, Harriman and Morrison (1980) have demonstrated
that spruce reforestation in  Scotland resulted in acidification of
streams and increased  export  of  cations.  It is not clear whether this is
due to indigenous tree-related processes as compared to the pre-existing
peaty soil vegetation  or is due  to changes in aerosol capture of acidic
components or to changes in hydrology.

     Certain vegetation types (e.g., conifers) produce abundant humic
material which can produce acidity.  Thus, the appearance of these
vegetation types in a  succession could yield long-term declines in pH as
well as dissolved organic material concentrations.  The appearance of
Sphagnum sp. because of changes  in the moisture regime could also result
in acidification of surface waters due to the highly effective cation
exchange capacity of Sphagnum with associated release of H'1".  Malmer
(1974) reviewed the Swedish literature relating to  reversion of farmland
to forests and finds that the chemical changes (increased organic
content, lower pH, lower exchangeable metals) are the same as those that
have also been attributed to  acidic precipitation.

4.4.3.3.5  Chemical amendments.   Adding some fertilizers (such as
ammonium phosphate) to agricultural soils has an acidifying effect on
soils, and this could  be transmitted to surface waters (along with
elevated levels of phosphate).  This potential acidification is generally
recognized and the affected soils are amended with  a base, CaCOa, with
subsequent elevation of pH.   In  regions where agriculture is on the wane
and reforestation is underway, the implicit cessation of CaCOs
application might result in a decline in the pH of  surface waters,
erroneously suggesting natural acidification.

4.4.3.3.6  Summary—effects of land use changes or  acidification.
Drabltfs et al. (1980)  examined historical land-use  changes in southern
Norway and their relationship to regional lake acidification and
decreasing fish populations.   They found no relationship.  Thus, although


                                 4-80

-------
changing land use may locally  alter  the  pH regime of lakes and streams,
It appears clear that regional lake  acidification and episodic pH
depression occur 1n response to  Increased atmospheric deposition of
strong acid, primarily ^$04.

4.4.4  Summary—Magnitude of Chemical Effects of Acidic Deposition
       (J. N. Galloway)

     The aquatic systems that  are most likely to be influenced by
atmospheric deposition are those with alkalinity of less than 200 peq
&"1.  Large areas of Canada and  the  United States contain such
systems.  For example, approximately 80  percent of New England, by virtue
of the geology, has surface waters with  less than 200 ueq £-1.
Eastern Canada provinces range from  90 percent (Quebec) to 20 percent
(New Brunswick).

     Of the aquatic systems that are potentially susceptible to
acidification (Figures 4-4 to  4-7),  only ones located in eastern North
America and small regions of western North America are receiving acidic
deposition (pH^B.O;  Figure 4-11; see also Chapter A-8, Section 8.4).

    Acidification of aquatic systems receiving acidic deposition has been
noted in several instances (Figure 4-13).

    Acidification of aquatic systems and acidic deposition is supported
by the following lines of evidence:

    o   Due to acidic deposition,  $04 concentrations have increased in
        aquatic systems in most  of eastern North America.  The increase
        in $04 has to have been  matched  by an increase in Cp or
        H+.  Since aquatic systems with  original low alkalinities are
        characterized by watersheds  with low CB/H+ ratios in the
        soil, a large portion  of the increase in $04 will have to be
        matched by an increase in H+, i.e., decreased alkalinity.

    o   Although there are problems  with comparing old and" new data,
        overall, the analysis  of temporal records shows decreases in
        alkalinity and pH in aquatic systems of eastern North America
        receiving acidic deposition.

    0   The limited application  of paleolimnologic indicators shows
        acidification of aquatic systems.

    o   Acidified aquatic systems are only found in areas receiving
        acidic deposition (pH  £  5.0).  In areas not receiving acidic
        deposition (pH >_ 5.0), acidification of sensitive aquatic
        systems is not found.

    0   No other possibilities exist to  explain the regional scale of
        acidification that has occurred.  For example, changing land use
        is at times advanced as  one  explanation.  However, in areas with
        identical changes in land use, it is only those areas receiving
        acidic deposition that are acidified.

                                  4-81

-------
           is the primary  cause  of  the long-term acidification of
aquatic systems on a regional basis.  The maximum decrease in alkalinity
that can occur due to acidic  deposition depends on the maximum long term
increase in S042-.  In the northeastern United States and southeatern
Canada this is about 100 yeq  «,-!.   The actual decrease in alkalinity
depends on how much of the increased S042" is balanced by increases in
base cations.  One estimate (Henriksen 1982) is that for a 100 yeq fc-1
increase in S042- and N03~ there will be an approximately 60 yeq £-1
decrease in alkalinity.  The  pH  change associated with an alkalinity
decrease of 60 yeq Jr* can range a  few tenths of a pH unit to 2 pH
units.  Those systems with the lowest initial alkalinities will show the
greatest loss of alkalinity due  to  acidic deposition because of the
scarcity of exchangeable cations in the terrestrial system.

     The time scales of long-term acidification are on the order of years
to decades in areas with low  $04 adsorption capacities (Northeast and
North Central United States)  and decades to centuries in areas with high
$04 adsorption capacities (Southeast United States).  If acidity of
deposition decreases, it is reasonable to believe the time scales of
recovery will be of similar magnitude (Galloway et al. 1983a) .

     In addition to long-term acidification (years and decades) by
H2S04» short-term acidification  (days to weeks) occurs as a result of
the combined action of ^$04  and HNOa in areas that develop acidic
snowpacks or receive a large  amount of rain over a short period of time.
Losses of alkalinity of 200 yeq  jr1 and reduction of pH from 7.0 to
4.9 have been reported due to the action of both $042- and
4.5  PREDICTIVE MODELING OF THE EFFECTS OF ACIDIC  DEPOSITION ON SURFACE
     WATERS (M. R. Church)

     The predictive modeling of the effects of acidic  precipitation on
the chemistry of natural waters is an extremely complicated task requir-
ing a great amount of data, knowledge, insight, and  skill.  Two avenues
exist for approaching the problem — empirical  modeling  and mechanistic
modeling.  Each approach has its advantages and disadvantages.

     Empirical models, in general, have two principal  advantages.  First,
they integrate the processes between inputs and outputs, thus eliminating
the need for precise knowledge of the behavior of  controlling mechanisms.
Second, they are usually very simple computationally.   Empirical models
do have certain drawbacks,  however.  One drawback  is that long periods of
data may be required to verify that an observed relationship between
inputs and outputs represents a steady state.  Other drawbacks include
the problems of verifying the validity of applying a relationship
observed in one geographic area to another area and  extrapolating from
one observed loading rate (or regime) to another.  Finally, because they
are almost always based on assumptions of steady state, empirical models
possess no time component;  they cannot predict the time required to reach
a new output level .
                                  4-82

-------
     Mechanistic models,  of course, have a different set of pros and
cons.  The principal  attraction of mechanistic modeling 1s that 1f
accurate mathematical representations of all (or the most Important) of
the phys1cal/chem1cal/b1olog1cal  processes Involved can be devised and
properly related to one another,  then a variety of extrapolations may be
made with confidence.  Such extrapolations Include the application of the
model (with appropriate calibration) to a variety of geographic areas;
the use of the model  to estimate  rates of change (e.g., of the alkalinity
or pH of a lake);  and the prediction of responses to almost any loading
scenario.

     Along with this potential for widespread application, however, go
certain problems.   The first, and perhaps the most obvious, 1s that the
knowledge may not exist to allow  formulation of accurate representations
of all (or even the most  Important) phys1cal/chem1cal/b1olog1cal
processes of Interest. Second, mechanistic models (especially of
lake-watershed eco-systems)  require extensive calibration for the region
to which they will  be applied.  Such calibration can be very time
consuming and expensive.   Third,  to be used predlctlvely, mechanistic
models that operate with  a relatively short time-step (say, less than one
week) require a correspondingly fine-scale source of predicted Input.
This requires a separate  method (or model) to generate Inputs of
precipitation form, amount,  and quality as stochastic variations around
annual (or even seasonal)  means.  This task, by Itself, 1s somewhat
Involved and time consuming. The last drawback to the mechanistic
approach 1s that as the representations of controlling processes become
more detailed and Intertwined, the time and effort required to perform
the calculations Increases substantially, even to the point where
significant amounts of computer time may be needed to perform long-term
simulations.

     A variety of  models  exist or are currently being developed to deal
with the problem of predicting the effects of various levels of acidic
deposition on the  chemistry of surface waters (e.g., Aimer et al. 1978;
Henriksen  1980, 1982; Chrlstophersen and Wright 1981; Thompson 1982;
Chen et al. 1982;  Chrlstophersen  et al. 1982; Schnoor et al. 1982).  The
models range from simple  empirical approaches to very computationally
complex formulations. A comprehensive review of all of these efforts 1s
beyond the scope of this  chapter. Instead, a fairly complete yet brief
review 1s presented of those three empirical models that are, so far, the
best known and most referenced of existing approaches.

4.5.1  Almer/Dlckson Relationship

     Aimer et al.  (1978)  plotted  lake pH vs lake sulfur loading (gm S
m"z yr"1 "concentration of 'excess1 sulfur multiplied by yearly
runoff) for Swedish lakes.  They found "tltratlon curve"-type patterns
for data from sets of lakes occurring In areas of similar bedrock.  They
plotted two curves (Figure 4-23):  One for waters "with extremely
sensitive surroundings" and one for waters with "slightly less sensitive
surroundings" (Aimer et al.  1978).  The authors did not define any
                                 4-83

-------
              7.Or
Figure 4-23.
              6.0
              5.0
              4.0
                 0               1               2

                   EXCESS  S  IN  LAKE  WATER  (g nT2 yr"1)
The pH values and sulfur loads in lake waters with
extremely sensitive surroundings (curve 1) and  with
slightly less sensitive surroundings (curve 2).   Load =
concentration of "excess" sulfur multiplied by the yearly
runoff.  Adapted from Aimer et al.  (1978).
                                   4-84

-------
objective method for classifying lakes with regard to their  surroundings
and responses to sulfur loadings (e.g., "extremely sensitive" or
"slightly less sensitive").  This limits their approach  as a general tool
for predicting the pH of lakes as a function of sulfur loadings.

     At first glance, using such a treatment of data  might seem to be a
way to help determine the levels of sulfate deposition (to watersheds)
that may have adversely affected lake water quality (pH).  Closer
examination of the approach, however, demonstrates that  care must be
taken in making such an application.  For example, the quantity "excess S
in lake water" must be carefully distinguished from the  quantity "total
excess S deposited".   Unfortunately, confusion about this question and
the original designation of the abscissa of Figure 4-23  has  led to
several mislabelings of reproductions of the original  figure (e.g., Glass
1980, Loucks et al. 1981, U.S./Canada 1981).  If Figures 4-24 and 4-25
(adapted from Aimer et al. 1978) can be compared" (note that  they
represent data roughly four years apart), they show that the relationship
is quite variable for the regions of Sweden for which the "Almer/Dickson
Relationship" was derived.  Not only is more excess sulfur deposited than
shows up in lake water (indicating some sulfate retention),  but also the
isopleths of the two plots are not parallel, indicating  that this
retention is different in different regions.

     As this example illustrates, the crux of the problem in applying the
"Almer/Dickson Relationship" is the translation of the abscissa of Figure
4-23 from a representation of "excess S in lake water" to some more
primary or causative factor (e.g., area! rate of total excess sulfur
deposition, areal rate of wet excess sulfur deposition,  concentration of
sulfate in precipitation, pH of precipitation, etc.). Such  a translation
requires quantitative knowledge of the relationships  among such things as
concentrations in lake waters, concentrations in precipitation, ratios of
wet to dry deposition, amounts of precipitation, amounts of  runoff, etc.
In turn, the statistical estimation of these types of relationships for
any region requires large amounts of data for that specific  region.

     Beyond the problems described above, other pertinent factors
involved in the use of the "Almer/Dickson Relationship"  must be
considered.  It is important to note that several  assumptions are
inherent in the approach.

     First, Aimer et al. (1978) assumed that within each of  the two sets
of lakes represented by the curves of Figure 4-23, initial (e.g., prior
to deposition of strong acids) steady-state values of alkalinity were all
the same.  Second, they assumed that the current pH values and the
current excess sulfur concentrations they observed in lake water were
both at steady state.  No evidence was offered in support of either of
these assumptions.  Finally, there is the problem of  hysteresis.  No data
exist to indicate that as a result of decreases in S042~ loading
rates, previously acidified lakes wouTd "return" along the curves of
Figure 4-23 to higher steady-state pH values.  Conditions extant have not
permitted such observations to be made, and there is  perhaps no clear
scientific consensus on this problem.


                                  4-85

-------
                             1.5
Figure 4-24.
Atmospheric load of "excess" sulfur from precipitation  and
dry deposition, 1971-72 (g S nr2 yr~l).   Dry deposition
calculated from a deposition velocity of 0.8 cm s"l.
Adapted from Aimer et al.  (1978).
                                  4-86

-------
Figure 4-25.
"Excess" sulfur in lake  water per year  (g  S  m~2 yr"1).
(Concentration of "excess  sulfur multiplied  by the yearly
runoff.)  Adapted from Aimer et al.  (1978).
                                   4-87

-------
     As a minimum condition, before the Aimer/Dick son Relationship can be
applied to the problem of  predicting the effects of changes in acidic
deposition on the chemistry of  surface waters in any geographic region,
reliable quantitative  relationships between primary factors (e.g., wet
sulfate deposition)  and sulfate concentrations in surface waters must be
developed.  Further, all assumptions inherent in the approach require
testing and validation.

4.5.2  Henriksen's Predictor Nomograph

     The contributions of  Henriksen (1979, 1980, 1982) to the empirical
study of the effects of atmospheric and edaphic factors on the chemistry
of oligotrophic lakes  in Scandinavia are well known.  Among his
contributions is the "predictor nomograph"--an empirical relationship
intended to be used as a tool  in  predicting effects of varying levels of
acidic deposition on the pH of  lakes.

     Using  data from  719  lakes in southern Norway (Wright and Snekvik
1978), Henriksen (1980) compared  the concentration of excess (above sea
salt contributions)  calcium plus  excess magnesium with excess sulfate
concentrations in the pH ranges 4.6 to 4.8 and 5.2 to 5.4 (see Figure
4-26) and found "highly significant" linear correlations.  Axes of excess
calcium concentration  (parallel to the axis of excess calcium plus
magnesium) and excess  sulfate  in  precipitation and pH of precipitation
(both parallel to the axis of  excess sulfate in lake water) complete the
predictor nomograph.  These final axes were developed from local
empirical relationships.  Henriksen (1980) used an independent data set
from a survey of 155 Norwegian  lakes to test his nomograph and found that
it correctly predicted pH  groupings approximately 85 percent of the time.
Henriksen (1982) concluded that the relationships depicted by the
predictor nomograph corroborated  his hypothesis that for the lakes he
studied (clear headwater oligotrophic lakes on granitic or siliceous
bedrock) "acidified waters are the result of a large scale acid base
titration."  He further concluded that the nomograph was capable of
predicting the effects that a  change in precipitation pH might have on
the pH status of lakes of  the  type he studied in the region he studied.

     As with all predictive constructs, or models, a number of key
assumptions (all clearly recognized and noted by Henriksen 1980, 1982)
are involved in the use of the predictor nomograph.

     One assumption or condition  for using the model is that it not be
used for lake waters with  significant concentrations of organic acids.
This is because (1)  these  acids may affect lake pH independent of
precipitation acidity and  (2)  analyses for calcium and magnesium include
these ions bound to organics;  thus ionic concentrations of excess Ca
plus excess Mg+2 may be overestimated.

     A second factor in the use of the nomograph  involves the possible
increased leaching of base cations from soils by acidic precipitation.
In his original work, Henriksen (1980) assumed no increased leaching of
base cations but noted the possible  importance such an event would hold


                                  4-88

-------
            300
        o-  200
        
-------
for use of the nomograph.   He has subsequently  studied  this question in
more detail, using data from lakes in  North  American and Scandinavia
(Henriksen 1982).

     He examined data from lakes in areas  of similar geology over a
gradient of deposition acidity,  and he also  compared time trend data of
calcium and magnesium in certain waters.   Unfortunately, he found no
clear cut answer to the question.  In  some cases, there was evidence of
increase in base cation concentrations (up to 0.63 for  Lake
Rishagerodvatten, Sweden).  In other cases}  there was none. In an effort
to overcome these difficulties and conflicting  data, Henriksen (1982)
used his best judgment to  designate a  maximum value of  "base cation
increase factor" of 0.4 yeq (Ca* + Mg*)/yeq  SO**.  That is, for
every yeq JT1 increase in  excess sulfate ($04*) concentration in
a lake, a maximal increase in excess calcium plus magnesium (Ca* + Mg*)
concentration would be 0.4 yeq £~1. It must be noted that in at
least one case, Henriksen  (1982) found a greater increase factor than
this—0.63 for Lake Rishagerodvatten,  Sweden.  Care should be exercised
in the application of this "base cation increase factor" for predictive
purposes.  It may vary significantly from  region to region (or watershed
to watershed within a region) as a function  of  soil chemical properties
(e.g. sulfate adsorption capacity, cation  exchange capacity, base
saturation), soil depth, and the path  of precipitation  through the soil.
In fact, it seems reasonable to  assume that  for some regions initially
experiencing acidic deposition,  the "increase factor" may be as high as
1.0.  Certainly more quantitative research is needed on this question.

     Another condition noteworthy in the use of the predictor nomograph
is the premise that all data used in its construction and verification
represent steady state conditions.  Due to the  large number of lakes and
deposition events and periods sampled, the data requirements to verify
this condition for the nomograph are astronomical and virtually
impossible to satisfy.  As an article  of faith  it must  be assumed that
the data employed do represent steady  state  conditions.  For many of the
lake data (especially at the "edges" or extremes of conditions) this
probably is not a bad assumption.  Lake data representing transitory
conditions are, perhaps, more suspect.

     A final question to consider in regard  to  the predictor nomograph is
its application to geographic regions  other  than (but similar to) the one
for/from which it was developed.  This is  always a key  question with such
empirical models.  Even if the general approach is accepted as sound,
common sense dictates that the empirical relationships  found in southern
Norway and Sweden may not pertain to even  seemingly analogous conditions
elsewhere.  (Certainly this is true of the axes relating precipitation
chemistry to excess sulfate concentrations in lakes.  Most acidic
precipitation in North America contains relatively more nitric acid than
does acidic precipitation  in Scandinavia.)  The inconsistencies
encountered by Bobee et al. (1982) and Haines et al. (1983b) in
attempting to apply the nomograph to lakes in Quebec and New England,
respectively, should be noted in this  regard.  It may very well be that
                                  4-90

-------
the predictor nomograph will  have to be modified  to accommodate local
relationships for whatever region for which application is attempted.

4.5.3  Thompson Cation Denudation Rate Model  (CDR)

     As seen in the previous  discussions  of the.Almer/Dickson
relationship and the Henriksen predictor  nomograph, the quantification of
the interrelationships of sulfate loading, base cation concentrations,
and surface water pH seem to  hold promise for understanding and
predicting surface water chemistry in some situations. These
interrelationships have been  explored also by Thompson (1982), who has
related surface water pH to excess sulfate loading and the rate of cation
loss from watersheds (the Cation Denudation Rate  or CDR).  As with the
prior models, her approach is restricted  to relatively unbuffered surface
waters with low concentrations of organic acids in areas with
acid-resistant bedrock, till, and soils.

     Thompson's model derives from charge balance and holds that a plot
of excess sulfate concentration vs the sum of base cation concentrations
yields a series of lines representing constant bicarbonate concentration.
If C02 partial pressure is constant,  then each line also represents
constant pH.  If CDR (concentration x discharge * watershed area) is
plotted against atmospheric excess sulfate loading rate (equivalent to
acid loading) and if runoff is specified  at 1 m yr"1, then an
equivalent representation applicable to lakes or  streams is generated
(Thompson and Hutton 1981, Thompson 1982) (see Figure 4-27).

     A number of important assumptions apply  to this approach.  First,
all non-sea salt sulfate must come from atmospheric loading alone.
Second, all sulfate deposited in a watershed  must flow through the
watershed without being retained (on a net basis).  Third, all sulfate
must be accompanied by protons as it enters and leaves the watershed.
The difficulties with each of these assumptions and the everyday
application of such a model have been thoroughly  described in the
preceding discussions of the  Almer/Dickson Relationship and the Henriksen
predictor nomograph.  Another difficulty  or necessary assumption relates
to both the constancy and quantification  of PQQ2  in any set °f waters
to which the model  may be applied.  Significant variations in C02
partial pressures in surface  waters are well  known.

     Yet another point worth  considering  is the fact that Thompson (1982)
tested this approach in some  highly colored lakes and rivers of Nova
Scotia (Figure 4-28).  Although she noted that the pH values of these
rivers "have been thought to  be dominated by  naturally-occurring organic
acids", Thompson (1982) feels that "their low pHs can be explained quite
well on the basis of simple inorganic chemistry."  A way to resolve this
question is through Gran titrations for weak  and  strong acids.
Apparently, such a study has  not been conducted.  The CDR model has yet to
be thoroughly verified with any other data sets.
                                  4-91

    409-262 0-83-10

-------
                                                                     10
                                                                       -2.5
                                        PC02
                                       RUNOFF =  1 m yr
                                                                            -1
Intercepts  are at
       £.5  and 10
        Figure  4-27.
                     ACID LOAD (meq nr2 yr'1) or EXCESS S042' (yeq jf
A plot of the model  that relates pH and sum of cations to
excess $04 - in concentration units, or pH and CDR to rate
of excess $04 - loading in rate units.
                                          4-92

-------
          WALLACE.
             200
  CM
  I
   ^    METEGHA
   g      LA HAV
      PIPERS HOLE

       ST.


           TUSKEI.
N.E. POND, MEDWAYt
           MERSEY- •
          ROSEWAY. -
                       RUNOFF  =  1 m

                       PpC02 = 2'5
                0
                                EXCESS S042" (meq m'2
    Figure 4-28.  CDR plot for rivers with mean runoff near J m yr-1, 1973
                  excess $04 - loads, and mean or median river pH.
                                         4-93

-------
4.5.4  Summary of Predictive Modeling

     As is evident in the preceding  discussions, there is still much to
learn about a number of key factors  that influence the ways in which
lakes/watersheds respond to acidic deposition, and thus the ways in which
these responses may be modeled and predicted, even on the most basic
levels.  Factors that appear to  be of  primary importance but about which
our knowledge is still  inadequate include:  1) the ability of soils to
retain sulfur inputs from atmospheric  deposition; 2) the effects of
acidic inputs on cation exchange and leaching from soils; 3) the
mobilization of aluminum compounds from soils due to acidic deposition;
4) the effects of acidic inputs  on mineral  weathering; 5) the presence or
absence of hysteresis in those processes and their effects as a function
of increasing or decreasing inputs of  acids (Galloway et al. 1983a).

     In short, predictive modeling of  the acidification of surface waters
is still in an infant stage.  Some interesting ideas have been put forth
and some progress is being made  but  there is still a very long way to go
before any model will be able to be  used with quantitative confidence.
Certainly none of the three models discussed briefly here have been
verified adequately for "off-the-shelf" application in North American
waters.  Such an application without a clear recognition and statement of
all the assumptions and limitations  contained in these approaches would
violate virtually every rule concerning the prudent use of predictive
models (Reckhow and Chapra 1981, Bloch 1982).

4.6  INDIRECT CHEMICAL CHANGES ASSOCIATED WITH ACIDIFICATION OF SURFACE
     WATERS

     Acidic deposition is composed of  NH4+, S042-, N03", H+, and basic
cations.  The previous sections  have discussed the chemical effects
acidic precipitation has in aquatic  systems by direct altering the
concentrations of these same chemicals.  There are additional indirect
effects on other chemicals.  Specifically,  the addition of acidic
deposition to terrestrial and aquatic  systems can disrupt the natural
biogeochemical cycles of some metal  and organic compounds to such a
degree that biological effects occur.  The  following three sections
discuss these chemical effects and assess the state of our knowledge.
The first section (4.6.1) focuses on metals in general; the second
(Section 4.6.2) specifically on  aluminum.   Elevated levels of aluminum 1n
acidified surface waters have been demonstrated to be toxic to aquatic
biota (Chapter E-5, Section 5.6) and thus are of particular concern.
Potential interactions between acidic  deposition and organic carbon
cycles are discussed in Section  4.6.3.

4.6.1  Metals (S. A. Norton)

     The impact of acidic precipitation or, more broadly, atmospheric
deposition on metal mobility  in  aquatic ecosystems may be divided  into
four areas:
                                  4-94

-------
     1)   Increased loading of metals from atmospheric deposition  to
         terrestrial and aquatic ecosystems.

     2)   Direct effects of atmospheric deposition on metal  release
         rates from or to aquatic ecosystems.

     3)   Secondary effects of atmospheric deposition on metal  release
         rates from or to aquatic ecosystems—both positive and
         negative.

     4)   Changes  1n aqueous spedatlon of metals and consequent
         biological  effects.

4.6.1.1   Increased Loading of Metals From Atmospheric Deposition—In many
Instances enhanced loadings of metals are associated with elevated levels
of NH4+, S042-, N03% and H+ In acidic deposition.  Although
this excessive of metals Is apparently related to Industrial activities,
historic measurements of metals 1n atmospheric deposition are not
sufficient for establishment of temporal trends. Indirect evidence for
Increasing atmospheric deposition of metals Is as follows:

     a)   Contemporary variations In atmospheric deposition of metals
         (e.g., Pb and Zn) are closely related to the geographic
         distribution of fossil fuel consumption, smelting, and
         transportation  (which uses the  Internal combustion engine)
         (Lazrus et al. 1970).  Where these sources are absent, metal
         deposition rates are lower (Galloway et al. 1982b).  Thus, as
         fossil fuel consumption and other processes expand, Injection
         of metals Into  the atmosphere Increases and atmospheric
         deposition Increases.

     b)   Ombrotrophlc peat bogs, those having no source of water  other
         than precipitation, receive all their nutrients and
         non-essential metals from atmospheric deposition.   Some
         elements are relatively immobile (e.g., Pb) and, after
         deposition, do not chemically migrate as the peat accumulates.
         Increased concentrations of lead in recent peat in eastern
         Massachusetts  (up to 1.2 x over background) suggest increases
         in atmospheric  deposition of at least 3.5 x over the past few
         decades  (Hemond 1980).  Absolute chronology in accumulating
         peat generally  can only be estimated; thus absolute increases
         cannot be rigorously established.  Other elements (e.g., Zn  and
         Cu) are  Increased  in concentration 1n modern peat as compared
         to "old" peat, but chemical mobility at the low pH of peat
         interstitial  waters, variable redox conditions, and biological
         recycling do not permit the precise calculation of absolute
         increases of atmospheric  deposition of these metals.

     c)  'Continuously'  accumulating  snow is believed to record or
         reflect changes in the chemistry of atmospheric deposition of
         metals.  However,  fractional melting, ablation, erosion and
         deposition of  snow, and other factors obscure absolute


                                  4-95

-------
    deposition  rates.  Nonetheless, it 1s clear that the deposition
    of Pb  and Zn  (fossil fuel-related elements) has greatly
    accelerated over the last 100 to 150 years 1n areas as remote
    as Greenland  (Herron et al. 1976, 1977).  The relative
    Increases depend on background (pre-pollution) values and the
    emission (and subsequent deposition) rates for specific
    metals.

d)  Galloway and  Likens (1979) showed higher concentrations of Pb,
    Au, Ag, Zn, Cd, Cr, Cu, Sb, and V 1n modern-sediments relative
    to older sediments of relatively undisturbed lakes.  Norton et
    al. (1981a) and Johnston et al. (1981) demonstrated that
    concentrations of Pb, Zn, Cu, Cd, and V are higher In modern
    sediments than 1n older sediments and established that the
    ubiquitous  (1n northern New England) and essentially
    synchronous (ca. 1860-80) Increases correlate with the initial
    rapid  Increase 1n the consuuptlon of fossil fuel in this
    country.  Because these lakes are relatively undisturbed, these
    changes are interpreted to be caused by increases 1n the rate
    of atmospheric deposition of these metals, starting prior to
    1860.

e)  Hanson et al. (1982) have shown that Pb concentrations in the
    organic soil  horizons of high elevation spruce/fir forests of
    New England,  New Brunswick, and Quebec are related to the pH of
    precipitation.  Low pH Is associated with high Pb (Lazrus et.
    al. 1970).  Groet (1976) demonstrated spatial variation in the
    northeastern  United States of concentrations of heavy metals 1n
    bryophytes, mosses, and liverworts  (known concentrators of
    atmospherically-deposited metals).  Highest concentrations are
    related  to  regional industrialization.

f)  The litter, fermentation, and  humic layers of organic soils of
    fir forests represent successively  longer time period and
    progressively more decayed material.  The concentration of
    lead,  which is chemically immobile  (probably because of
    adsorption),  Is highest in the fermentation layer (but nearly
    the same as 1n the Utter layer), suggesting increased
    deposition  of Pb  (Reiners et al. 1975, Hanson et al. 1982).
    Although Pb can be removed mechanically by erosion andvertlcal
    displacement, rates of deposition can be derived if the age of
    litter Is  known and mechanical erosion is nil.  Siccama et al.
    (1980) studied white pine forest soils in central Massachusetts
    collected at  two times (separated by 16 years) and found a
    higher rate of Pb accumulation 1n recent litter.  Many workers
    have demonstrated spatial and  temporal trends for other
    elements (e.g., Zn) which parallel  those for Pb, but
    quantitatively assessing increased  deposition rates cannot be
    done because  of the nonconservative nature of the other
    elements.
                             4-96

-------
4.6.1.2  Mobilization of Metals by Acidic  Deposition--The  stoichiometry
of chemical  weathering reactions and cation  exchange and experimental
evidence (e.g., Cronan 1980),  suggests that  increasingly acidic
precipitation should increase  the release  of cations (any  positively
charged aqueous species) from  soils and aquatic  sediments.  Empirical
evidence from the United States for accelerated  release of cations due to
acidic deposition over a long  time period, however, is rare.  Oden
(1976a) cited evidence for long-term increasing  Ca concentrations in
Swedish rivers, but long-term  land use changes on the scale of 10 to 100
years (including vegetation succession)  (Nilsson et al. 1982) may cause
similar results (Section 4.4.3.3).

     Paleolimnological evidence from sediment cores (Hanson et al. 1982)
indicates that detritus deposited in lakes has been, in undisturbed
watersheds, progressively more depleted in recent time with respect to
easily mobilized elements, e.g., Zn, Mn, Ca, and Mg.  These decreases in
concentration start as early as about 1880 and are interpreted to result
from increased leaching of these elements  from the terrestrial ecosystem.
Similar changes are not seen in areas that have  only recently received
acidic deposition (e.g., Swedish Lappland, Norton, unpub.  data).
Deposition rate and concentration data for sediments from  undisturbed
lakes in New England indicate  continuously increasing values for Pb for
all lakes for about 100 years.  The values for Zn increase continuously
to the present for lakes with  a pH > 6.0 and decrease in younger
sediments for lakes with pH <  5.5, suggesting recent acidification of
those lakes with decreasing Zn (as well  as Ca, Mn, and possibly Mg).

     Field and laboratory soil lysimeter studies by Cronan and Schofield
(1979) and Cronan (1980) indicate that modern soil solutions have
chemistry (e.g., Al  concentrations) that is  inconsistent with the
historical soil horizon development.  This is interpreted  to be due to
more acidic influx to the soil from acidic precipitation,  causing Al
leaching where before Al was accumulating.

     Episodic decreases in the pH of surface waters (linked
quantitatively to meteorological events) are commonly accompanied by
increases in dissolved Al (Schofield and Trojnar 1980) and other
elements, suggesting the direction of changes to be expected in the
mobilization of metals from soils, bedrock,  and  sediments  as
precipitation becomes more acidic (Norton 1981).

     Data sets for metal concentrations of lake  waters versus pH suggest
that, because of solubility relationships, mobility of certain metals
(Al, Zn, Mn, Fe, Cd, Cu) should be relatively greatly increased with
increasingly acidic precipitation (Norton et al. 1981b, Schofield, 1976b,
Wright and Henriksen 1978).  Other metals (K, Na, Ca, Mg), the
concentration of which is in the ^ 0.1 ppm range, will also be affected
but to a lesser degree relative to initial concentrations.

     Accelerated cation release (from aquatic sediments) has also been
demonstrated during experimental acidification of surface  waters.  In the
field, Hall and Likens (1980)  observed increased release of Al, Ca, Mg,


                                -  4-97

-------
K, Mn, Fe, and Cd due to artificial  acidification of streams.  In
isolated columns in lakes and  in whole lake acidification experiments,
Schindler et al. (1980)  observed increased leaching of Fe, Mn,and Zn from
the sediments.  Andersson et al. (1978), Hongve (1978), Davis et al .
(1982), and Norton (1981)  demonstrated in laboratory sediment/water core
microcosms that accelerated  leaching of metals from sediment occurs
during acidification.

4.6.1.3  Secondary Effects of  Metal  Mobil izati on—Secondary effects of
acidic deposition may lead to  increased or decreased metal mobility.  For
example, the release of Hg from sediments and soils and production of
methyl mercury is promoted by  more acidic waters (Wood 1980).

     Secondary effects may be  operative but have not been demonstrated.
For example, increases of Pb (as Pb2+) and S042" may result in immobil-
ization of both Pb2+ and S042~ as the insoluble salt, PbSO/j.  Similarly
Nriagu (1973) has suggested  that excess Pb2+ may immobilize P042~.  This
could cause a reduction in available phosphate for aquatic ecosystems.
Al sulfate minerals (Nordstrom 1982) are now suggested as being a control
on Al and/or S042-.  Increased A13+  in acidified soil waters could also
immobilze phosphate. Alternatively,  desorption from or solution of FeOOH
from "B" soil horizons in well drained soils could liberate adsorbed
phosphate.  These potentially  important mechanisms have not been thoroughly
investigated in the context  of acidic precipitation.  Very probably P04
availability will be strongly  affected by increased concentrations of
     and Al3+ in soil  and and  surface waters.
4.6.1.4  Effects of Acidification  on  Aqueous Metal Speciation--The
chemical form of dissolved metals  is  important  in determining the total
mobility of a metal  and the biological effects  related to acidification
of aquatic ecosystems.   In general, most metals are complexed less at
lower pH values because few HC03~, C032~, OH" and other weak acid
ligands are present.  Limits for concentrations of metals for toxicity to
organisms (Gough et al. 1979)  are  generally based on experiments where
the water chemistry is not well  characterized so such limits are probably
excessively high.  Some toxicity limits  have been defined for "soft" and
"hard" water (e.g.,  Howarth and  Sprague  1978).   The upper limits for
toxicity for hard water are generally much higher than for soft water,
reflecting the probable importance of speciation.

4.6.1.5  Indirect Effects on Metals in Surface  Waters—The rate of
deposition of several  metals from  the atmosphere is increased due to
anthropogenic activities.  The metals include Pb, Au, Ag, Zn, Cd, Cr, Cu,
Sb, and V.  Primary and secondary  effects of acidic deposition on metal
mobility include increased solubility of Al , Zn, Mn, Fe, Cd, Cu, K, Na,
Ca, and Mg.  These metals are mobilized  by acidic deposition both from
the terrestrial system and from  lake  sediments.

     As aquatic systems acidify, speciation of  metals changes.  The
direction of changes is generally  to  a more biologically active species.
                                  4-98

-------
4.6.2  Aluminum Chemistry in  Dilute  Acidic  Waters  (C. T. Driscoll)

4.6.2.1  Occurrence, Distribution, and Sources  of  Aluminum--Aluminum is
the third most abundant element within the  earth's crust (Garrels et al.
1975).  It occurs primarily  in aluminosilicate  minerals, most commonly as
feldspars in metamorphic and  igneous rocks  and  as  clay minerals in
well-weathered soils.  In high elevation,  northern temperate regions, the
soils encountered are generally podzols (Buckman and Brady 1961).  The
podzolization process involves mobilizing  aluminum from upper to lower
soil horizons by organic acids leached from foliage as well as from
decomposition in the forest  floor (Blcornfield 1957; Coulson et al.
1960a,b; Johnson and Siccama  1979).   Aluminum largely precipitates in
lower soil horizons (Ugolini  et al.  1977).   Ugolini et al. (1977) have
observed that during podzolization little  aluminum mobilizes from the
adjacent watershed to surface waters,   Stumm and Morgan (1970) report a
median aluminum value of 10  yg Al jT1  for  terrestrial waters, while
Bowen (1966) gives an average concentration of  240 g Al  -1 for
freshwaters including bogs.   It is noteworthy that values of aluminum
reported for circumneutral waters are generally greater than levels
predicted by mineral equilibria (Jones et  al. 1974).  Because of the
tendency for aluminum-hydroxy cations to polymerize through double OH
bridging when values of .solution pH  exceed about 4.5  (Smith and Hem
1972), a considerable fraction of the "dissolved"  aluminum reported in
many analyses of natural water having near-neutral or slightly acidic pH
may consist of suspended microcrystals of  aluminum hydroxide.  Hem and
Robertson (1967) have shown  that crystals  having a diameter near 0.1  m
were relatively stable chemically.   Filtration  of  samples through 0.4
 m porosity membranes, a common practice in clarifying natural water
prior to analysis, may fail  to remove such material (Kennedy et al.
1974).  However, the concentrations  of dissolved aluminum are generally
low in most circumneutral natural  waters due to the relatively low
solubility of natural aluminum minerals.

     Superimposed on the natural podzolization  process is the
introduction of mineral acids from acidic  deposition  to the soil
environment.  It has been hypothesized that these  acids remobilize
aluminum soluble previously  precipitated within the soil during
podzolization or held on soil exchange sites (Cronan and Schofield 1979).
Elevated levels of aluminum  have been reported  in  acidic waters within
regions susceptible to acidic deposition (Table 4-7).

     Many investigators have  observed an exponential increase in aluminum
concentration with decreasing solution pH  (Hutchinson et al. 1978,
Dickson 1978a, Wright and Snekvik 1978, Schofield  and Trojnar 1980,
Vangenechten and Vanderborght 1980,  Hultberg and Johansson 1981, Driscoll
et al. 1983).  This phenomenon is characteristic of the theoretical and
experimental solubility of aluminum  minerals.  Researchers have
hypothesized several mechanisms for  the solid phase controlling aluminum
concentrations in dilute water systems, including  poorly crystallized 1:1
clays (Hem et al. 1973) kaolinite (Norton  1976), aluminum trihydroxide
(May et al. 1979, Johnson et al. 1981, Driscoll et al. 1983), basic
aluminum sulfate (Eriksson 1981) and exchange on soil organic matter


                                  4-99

-------
                               TABLE 4-7.  ALUMINUM CONCENTRATION IN DILUTE ACIDIC WATERS
o
o
Location
Lakes
Sweden
Norway
Scotland
Bel gi urn
USA
Canada
Canada
Streams
USA
USA
Description

Swedish West Coast, 1976
Regional Survey, 1974-77
Southwestern Scotland, 1979
Moorland pools Northern
Belgium, 1975 - 1979
Adirondacks, 1977-1978
Ontario various locations, 1980
Sudbury, Ontario
Adlrondacks, 1977-1978
Adirondacks, 1977
pH

4.0
4.2
4.4
3.5
3.9
4.1
4.3
4.0
4.4
Range

- 7.4
- 7.8
- 6.4
- 8.5
- 7.2
- 6.5
- 7.0
- 7.6
- 6.5
Al
ng

10
0
25
300
4
6
150
92
Range
Al a-1

- 670
- 740
- 310
- 8000
- 850
- 856
- 1150
- 1170
100 - 1000
Reference

Dick son 1978a
Wright et al. 1977
Wright et al. 1977
Vangenechten and
Vanderborght 1980
Driscoll 1980
Kramer 1981
Scheider et al .
1975
Driscoll 1980
Schofield and
                                                                                         Trojnar 1980

-------
                                                   TABLE 4-7.   CONTINUED
I
I—1
o
Location Description
Streams (cont.)
USA Hubbard Brook stream order








average
pH

1
2
3
2
3
3
4
4
5

Range

4.73
4.94
5.09
5.19
5.54
5.46
5.51
5.58
5.68
4.90
Al Range Reference
yg Al A-l

710 Johnson et al.
320 1981
210
200
150
190
180
160
150
230
      Groundwaters

      Sweden

      USA
West Coast,  1977-1978

Hubbard Brook  seepwater,  1979
3.8 - 5.7

4.6 - 6.5
100 - 2600

  0 - 700
Hultberg and
Johansson 1981
Mulder 1980

-------
(Bloom et al. 1979).  Johnson et al.  (1981)  and Driscoll et al. (1983)
compare and discuss solution  characteristics of New Hampshire and
Adirondack waters, respectively, with the  theoretical  solubility of a
variety of aluminum minerals.  Eriksson  (1981) observed that calculated
values of aquo aluminum in soil  solutions  from Sweden  were similar to
values predicted from mineral solubility reported by van Breemen (1973)
for Al (OH) $04, at a given pH.   This lead Eriksson (1981) to suggest
that atmospheric deposition of sulfate has acidified and transformed
aluminum oxides to basic aluminum sulfate  in Swedish soils.  Unfortu-
nately, Eriksson (1981) failed to consider fluoride, sulfate, and organic
complexation reactions when computing aquo aluminum levels.  Therefore,
as suggested by Nordstrom (1982), it  is  doubtful that  aluminum sulfate
minerals (e.g., jurbanite, alunite, basaluminite) control aquo aluminum
levels in waters acidified by acidic  deposition.  In actuality it is
extremely difficult to identify a specific solution controlling phase.
Analysis of soils and sediments  by x-ray diffraction has failed to
confirm the presence of hypothesized  solution controlling minerals of
aluminum (Driscoll et al. 1983).

4.6.2.2  Aluminum Speciation--Dissolved  monomeric aluminum occurs as aquo
aluminum, as well as hydroxide,  fluoride,  sulfate, and organic complexes
(Robertson and Hem 1969, Lind and Hem 1975). Past investigations of
aluminum in dilute natural waters have often ignored non-hydroxide
complexes of aluminum (Cronan and Schofield  1979, N. M. Johnson 1979,
Eriksson 1981).  Driscoll and coworkers  (Driscoll et al. 1980, Driscoll
1980, Driscoll et al. 1983) have fractionated Adirondack waters into
inorganic monomeric aluminum, organic monomeric aluminum, and acid
soluble aluminum.  They observed that inorganic monomeric aluminum levels
increased exponentially with  decreasing  solution pH.   Organic monomeric
aluminum levels were strongly correlated with total organic carbon (TOO
concentration but not pH.  Acid soluble  aluminum levels were relatively
constant and not sensitive to changes in either pH or  TOC.  Driscoll et
al. (1983) reported that organic complexes were the predominant form of
monomeric aluminum in Adirondack waters, on  the average accounting for 44
percent of monomeric aluminum.  Aluminum fluoride complexes were the
second major form of aluminum and the predominant form of inorganic
monomeric aluminum, accounting for an average of 29 percent of the
monomeric aluminum.  Aquo aluminum and soluble aluminum hydroxide
complexes were less significant than  aluminum fluoride complexes.
Aluminum sulfate complexes were  small  in magnitude.

4.6.2.3  Aluminum as a pH Buffer--Pilute water systems are character-
istically low in dissolved inorganic  carbon  (DIG) due  to limited contact
with soil.  Because dilute waters are inherently low in DIG, they are
limited with respect to inorganic carbon buffering capacity.
Consequently, non-inorganic carbon acid/base reactions, such as
hydrolysis of aluminum and protonation/deprotonation of natural organic
carbon, may be important in the  pH buffering of dilute waters.

     Several researchers have investigated organic carbon, weak acid/base
systems in dilute waters.  Dickson (1978a) observed that elevated levels
of aluminum increased the BNC of Swedish lakes.  Waters were strongly
                                  4-102

-------
buffered by aluminum in the pH range 4.5 to 5.5   The  BNC  of  aluminum was
particularly evident when acidified lakes were treated  with  base  (limed).
Aluminum BNC was comparable in magnitude to hydrogen  ion  and inorganic
carbon BNC; therefore, the presence of aluminum  substantially increased
base dose requirements and the cost associated with the restoration of
acidified lakes.

     Johannessen (1980) investigated non-hydrogen/inorganic  carbon
buffering in Norwegian waters.  While reiterating the importance  of
aluminum as a buffer in dilute acidified waters,  she  also evaluated the
role of natural organic acids.  Natural  organic  matter  reduced the degree
to which aluminum hydrolyzed in the pH range 5.0 to 5.5,  presumably due
to complexation reactions, and therefore decreased the  buffering  of
aluminum.  Natural  organic matter also participated in  proton
donor/acceptor reactions; the extent to which total organic  carbon (TOO
would dissociate/associate protons was 7.5 peq mg organic carbon-1.
Johannessen (1980)  concluded that organic carbon was  the  most important
weak acid/base system in acidic Norwegian waters because  of  the high
organic carbon concentration relative to aluminum.

     Glover and Webb (1979)  evaluated the acid/base chemistry of  surface
waters in the Tovdal region of southern Norway.   The  BNC  of  hydrogen ion
was small compared to the BNC of weak acid systems.   These investigators
suggested that of the total  weak acid BNC, 40 to 60 yeq £~*  could
be attributed to dissolved aluminum and silicon,  while  20 to 50 yeq
JT1  could be attributed to natural  organic acids.  Solution
titrations were characterized as having a major  proton  dissociation
constant (Ka) of 10~6 to 5 x 10'7, in addition to some  less  well
defined iom'zation at higher pH values.

     In a comparable study, Henriksen and Seip (1980) evaluated the
strong and weak acid content of surface waters in southern Norway and
southwestern Scotland.  In addition to a titrametric  analysis, the
aluminum, dissolved silica,  and TOC content of water  samples were
determined.  Weak acid concentrations, determined by  a  Gran  (1952)
calculation, were evaluated by multiple regression analysis.   Most of the
variance in the weak acid concentration could be explained by the
aluminum and TOC content of the waters.   Thus, it was concluded that the
weak acid content of acidified lakes in southern  Norway and  Scotland was
largely a mixture of aluminum and natural  organic acids.

     Driscoll  and Bisogni (1983)  quantatively evaluated weak  acid/base
systems buffering dilute acidic waters in the Adirondack  region of New
York State.  Natural organic acids were  fit to a  monoprotic  proton
dissociation constant model  (pKa = 4.41), and the total,  organic  carbon
proton dissociation/association sites were observed to  be empirically
correlated to TOC concentration.   Aquo-aluminum  levels, calculated from
field observations, appeared to fit an aluminum  trihydroxide  solubility
model.

     Calculated buffering capacity (B) is plotted against pH  in Figure
4-29 for a hypothetical system that has  some properties in common with


                                  4-103

-------
Adirondack waters (Driscoll  and  Bisogni 1983).  Buffering capacity is
defined as the quantity of  strong acid or base  (mols £~M which would
be required to change the pH of  a liter of  solution by one unit.
Conditions specified for the construction of  Figure 4-29 are indicated in
the figure title.  Aluminum species  may dominate  the buffer system at low
pH if these conditions are  fulfilled, suggesting  that the lower limit of
pH observed in acidic waters with elevated  aluminum levels may be
controlled by the dissolution of aluminum.  At  higher pH values the
buffer system is dominated  by inorganic carbon  and would be even more
strongly dominated if carbonate  solids were present.

     It is noteworthy that  aluminum  polymeric cations and particulate
species, that may occur in  acidic solutions,  provide some solution
buffering (both ANC and BNC). However, these large units may be slow to
equilibrate with the added  titrant.   Therefore, ANC and BNC
determinations have limitations  in acidic waters  due to heterogeneity
phase problems.

4.6.2.4  Temporal and Spatial Variations  in Aqueous Levels of Aluminum--
Pronounced temporal and spatial  variations  in levels of aqueous aluminum
have been reported for acidic waters.  Schofield  and Trojnar (1980)
observed that high aluminum levels occurred during low pH events in
streams, particularly during snowmelt.  Driscoll  et al. (1980) also
observed this phenomenon but attributed aluminum  increases to inorganic
forms of aluminum.  During  low flow  conditions, neutral pH values were
approached in streams (pH 5.5 to 7.0) and inorganic monomeric aluminum
levels were low.  During summer  months, levels  of TOC in streams
increased and organically complexed  aluminum  levels increased.  As
mentioned previously, levels of  organic monomeric aluminum were strongly
correlated with surface water TOC  (Driscoll et  al. 1983).

     Johnson et al. (1981)  studied temporal and spatial variations in
aluminum chemistry of a first-through-third order stream system in New
Hampshire.  Observations of temporal variations in aluminum were similar
to those reported for the Adirondacks (Driscoll et al. 1980, Schofield
and Trojnar 1980).  Johnson et al.  (1981) reported decreases in hydrogen
ion and aluminum levels with increasing  stream  order.  They suggested a
two-step process for the neutralization of  acidic deposition.  Mineral
acidity entering the ecosystem from  atmospheric deposition was converted
to a mixture of hydrogen ion and aluminum BNC (acidity) in headwater
streams and was subsequently neutralized  through  the dissolution of basic
cation (Ca2+, Mg2+, Na+, K+) containing minerals  within the soil
environment.

     Johnson et al. (1981)  observed  a shift in  aluminum  speciation with
increasing stream order.  Aquo aluminum and aluminum hydroxide complexes
decreased substantially with increasing  stream  order.  Alumino-fluoride
complexes remained constant throughout the  experimental reach.
Alumino-organic  forms increased in  concentration with decreasing
elevation.
                                  4-104

-------
       CQ
       Q.
                                    ALUMINUM
                                    CARBONATE
                                    ORGANIC  SOLUTES
                                    WATER
                                    PH
Figure 4-29.
Buffer capacity diagram for dilute Adirondack water systems
(Driscoll and Bisogni  1983).   Equilibrium with aluminum
trihydroxide (pKso = 8.49), organic solutes (CTorg =
2 x 10~5, pKorg =4.4)  and atmospheric carbon dioxide
(Pco2 = 10~3-5 atm) were assumed.
                                  4-105

-------
     Driscoll (1980)  has evaluated  temporal and spatial variations In
aluminum levels In acidic lakes.  During summer stratification, monomeric
aluminum levels were  low in  the upper waters and increased in
concentration with depth. Low aluminum levels reported in the upper
waters during the summer coincided  with elevated pH and ANC values.  The
increased pH and ANC  values  were attributed to algal- assimilation of
nitrate (Brewer and Goldman  1976).   During ice cover, pH (and ANC) values
were low and aluminum levels high directly under the ice.  The pH values
increased and aluminum values decreased with depth.  The clinograde
distribution of pH and aluminum observed during ice cover periods has
been attributed to reduction processes in sediments (e.g.,
denitrification).  These processes  generate ANC, which diffuses into the
lower waters.  During fall and spring turnover, aluminum is evenly
distributed throughout the water column of acidic lakes.  Aluminum levels
were particularly high during the spring season because of inputs of low
pH, high aluminum stream water associated with spring snowmelt.

     Few studies have considered temporal and spatial variations in
aluminum chemistry of groundwaters.  Hultberg and Johansson (1981) have
observed acidification events in groundwater chemistry in Sweden.  They
hypothesized that much of the atmospheric input of sulfur was retained
within the terrestrial ecosystem as reduced sulfur forms.  During
extremely dry conditions, the water table was lowered and pools of
reduced sulfur within the soil become oxidized by molecular oxygen
entering the zone. Very low values (< 4.0) and very high aluminum levels
(> 40 mg Al £-1) have been reported in groundwater by Hultberg and
coworkers (Hultberg and Wenblad 1980, Hultberg and Johansson 1981) when a
prolonged dry period  was followed by a rainfall event.  It is difficult
to conclusively attribute groundwater to atmospheric deposition of
sulfate.  A possible  source  of the  acidity in the groundwater studied by
Hultberg and Johansson (1981) was the oxidation of reduced iron minerals,
likely to have been present  naturally in the upper part of the zone of
saturation.  This oxidation  would have occurred when the water table
declined due to dry weather  and molecular oxygen entered the zone.  The
hydrogen ion produced by iron oxidation with molecular oxygen would not
be significantly mobilized in the groundwater until the water table
increased again to a  more normal level.

4.6.2.5  The Role of  Aluminum in Altering Element Cycling Within Acidic
Waters—In acidic water systems conditions or supersaturation with
respect to aluminum trihydroxide have been reported (Driscoll et al.
1983).  During conditions of supersaturation, aluminum will hydrolyze,
forming partial!ate aluminum oxyhydroxide.  The acid-soluble aluminum
fraction mentioned earlier would include the mi croc rystal line hydroxide
particles and their polymeric hydroxycation precursors.  Smith and Hem
(1972) observed that  during  the polymerization process, aluminum
hydroxide units displayed metastable ionic solute behavior until they
contained from 100 to 400 aluminum  atoms.  When particles developed to
that size their behavior was characteristic of a suspended colloid.
Mi croc rystalline particles have a very large specific surface area and
may adsorb or co-precipitate organic and inorganic solutes.  The cycling
                                  4-106

-------
of orthophosphate (Huang 1975,  Dickson  1978a), trace metals (Hohl and
Stumm 1976)  and dissolved organic carbon  (Dickson 1978a, Davis and Gloor
1981, Driscoll  et al.  1983,  Hall et al. 1982) within acidic surface
waters may be altered  by adsorption on  aluminum oxyhydroxides.  However,
few studies have addressed this specific  hypothesis.

     Huang (1975) studied the adsorption  of orthophosphate on
AlpOs-  He observed an adsorption maximum at pH 4.5.  While Huang
(1975) studied the adsorption of high  levels of orthophosphate (10-4 to
10-3 M), much higher than would be  observed in natural dilute water
systems, his observations of phosphate  aluminum interactions may be
generally applicable.

     Dickson (1978a) observed that  when acidic lake water, elevated in
aluminum, was supplemented with orthophosphate (50 and 100 yg P
£'*), dissolved phosphorus was  removed  from solution.  The removal of
phosphorus was most pronounced  at pH 5.5.  Dickson (1978a, 1980)
suggested that aqueous aluminum may substantially alter phosphorous
cycling within acidic  surface waters through adsorption or precipitation
reactions.  This hypothesis is  noteworthy because phosphorus is often the
nutrient limiting algal growth  in dilute  surface waters (Schindler 1977).
Any decrease 1n aqueous phosphorus  induced  by adsorption on aluminum
oxyhydroxides may result in a decrease in algal growth and an accompanied
decrease in algal generated ANC (see Section 4.7.2).  Any decrease in ANC
inputs would result in an aquatic ecosystem more susceptible to further
acidification.

     Aluminum forms strong complexes with natural organic matter (Lind
and Hem 1975).  Complexation substantially  alters the character of
natural organic acids.  Driscoll et al. (1983) observed that DOC was
removed from the water column of an acidic  lake after CaC03 addition.
They hypothesized that DOC sorbed to the  particulate aluminum that had
formed within the water column  shortly  after base addition.  Driscoll
(1980) observed decreases in water  column TOC during conditions of super
saturation with respect to A1(OH)3  in an  acidic lake.  He hypothesized
that natural organic carbon was scavenged from solution by particulate
aluminum formed in the water column.  Davis (1982) has studied the
adsorption of natural  dissolved organic matter at the  Y- A1203/
water interface.  He observed that  natural  organic matter adsorbs by
complex formation between the surface hydroxyls of alumina and acidic
functional groups of organic matter.  Davis (1982) indicated that DOC
adsorption was maximum at pH 5. Davis and  Gloor (1981) reported that DOC
associated with molecular weight fractions  greater than 1000 formed
strong complexes with  the alumina surface,  but low molecular weight
fractions were weakly  adsorbed. Davis (1982) suggests that under
conditions typical for natural  waters almost complete surface coverage by
adsorbed organic matter can be  anticiapted  for alumina.  Organic coatings
may be important with  respect to subsequent adsorption of trace metals
and anions.

     Hall et al. (1982) observed a  decrease in DOC levels of a third
order stream in New Hampshire after aluminum chloride (A1C13) addition.


                                 4-107

-------
In addition,  a reduction in  surface  tension  occurred at the air-stream
interface and was attributed to  a  decrease in the solubility of DOC due
to interactions with aluminum.

     DOC loss to acidic waters  is  significant in several respects.  DOC
represents a  weak base that  serves as  a component of solution ANC
(Johannessen  1980, Driscoll  and Bisogni 1983).  DOC also serves as an
aluminum complex!ng ligand.   Complexation of aluminum by organic ligands
mitigates aluminum toxicity  to  fish  (Baker and  Schofield 1980).
Therefore, any loss of DOC may  translate to  an  environment less
hospitable to fish.

4.6.3  Orgam'cs (C. S. Cronan)

4.6.3.1  Atmospheric Loading of Strong Acids and Associated Organic
Microponutants--This first  subsection deals with the association (but
not necessarily interaction) between anthropogenic strong acids and
organic micropollutants introduced to  aquatic systems via long-range
transport and wet/dry deposition processes.  Methods for isolating and
characterizing organic micropollutants in natural samples have been
described by  Gether et al.  (1976)  and  Heit et al. (1980).  These methods
were used by  Lunde et al. (1976) to  identify a  wide range of organic
pollutants in rain and snow  samples  from Norway, including alkanes,
polycyclic aromatic hydrocarbons,  phthalic acid esters, fatty acid ethyl
esters, and many other chemicals of  industrial  origin.  Concentrations
ranged from one to several  hundred ng  £~1, with PCB concentrations
registering five times higher than freshwater or seawater.

     In a related study, Alfheim et  al. (1978)  examined the access of
certain non-polar organic pollutants to lakes and rivers in Norway.
Results indicated that PCB  concentrations  in water samples from a lake  in
southern Norway were considerably  lower than in melted snow from the same
area.  Two explanations were offered to account for these observations:
(1) the PCB's in the water  column  may  have been associated with
particulate matter, preventing  them  from being  detected in the dissolved
phase, and (2) terrestrial  humic substances  may have complexed the PCB's
and related pollutants, thereby reducing their  leaching into lakes and
rivers.

     The studies by Heit et al. (1981) focused  on the historical patterns
of organic pollutant deposition to remote  Adirondack lakes.  Using lake
sediment cores and advanced analytical techniques, they found the
following results.  First,  all  of  the  nonalkylated 3- to 7-ring parental
PAH's, with the exception of perylene, decreased in concentration with
sediment depth.  Surface concentrations of many of these compounds
approached or exceeded levels reported for sediments from urban and
industrialized areas, while baseline levels  lower in the core were
similar to those reported for pristine areas such as in northern Ontario.
Overall, the data indicated that all of the  parental PAH compounds except
perylene entered these Adirondack  lakes primarily through anthropogenic
rather than natural processes.
                                  4-108

-------
     These investigations by  Alfhelm et al. (1978, 1980) and Heit et al.
(1981) have shown that a  broad  range of organic micro-pollutants may
originate in industrial centers and be carried downwind to remote
ecosystems by long-range  atmospheric transport.  Thus, similar patterns
and processes may contribute  to the atmospheric transport and deposition
of both anthropogenic strong  acids and organic micro-pollutants.

4.6.3.2  Organic Buffering Systems—Organic and/or aluminum weak add
buffer systems may dominate the acid-base  chemistry of surface waters in
watersheds characterized by the following  kinds of features:  granitic
bedrock, thin or impermeable  surf Ida! deposits, steeper slopes, high
water tables, or extremely permeable siliceous surficial deposits.  In
such  soft water ecosystems, organic and aluminum weak acids may provide
the only buffering protection against further acidification by
anthropogenic strong adds.  Likewise, natural himic materials may
themselves have sufficiently  low pka constants that they contribute to
the free acidity of surface waters.  Organic weak anlons may be
particularly significant in providing ANC  below pH 5.0, with the greatest
buffer intensity for the organlcs exhibited  in the range of pH 4.5
(Figure 4-29) (Driscoll 1980).

  The organic species responsible for contributing to the buffer capacity
of these soft water lakes Include a range  of hydrophlUc and hydrophoblc,
low and high molecular we.ight compounds.   These organic solutes may range
from  simple carboxyllc acids  like malic acid to complex poly phenolic
compounds like the model  fulvic acid described by Schnitzer (1980).  On
the average, these organic adds 1n natural  waters might be expected to
contribute 5 to 10 yeq of anionlc charge per mg carbon  (Driscoll 1980;
Cronan, unpub. data), and perhaps 5 to 20  yeq per mg organic carbon In
total acidity (Schnitzer 1978, Henriksen and Seip 1980).  Historically,
organic acid buffer systems were probably  relatively common in soft water
aquatic systems.  However, the relative importance of aluminum buffering
Section 4.6.2.3) may have increased  recently in those soft water lakes
that  have experienced modern  acidification from atmospheric deposition
(Henriksen and Seip 1980).

4.6.3.3  Organo-Metalic Interactions—Acidification of  surface waters may
affect metal-organic associations and trace  metal speciation. Stability
constants for metal-fulvic acid (FA) complexes have been shown to
decrease with decreasing pH.   For example, the conditional stability
constant for Pb2+-FA at pH 5.0 is 4.1, whereas it 1s 2.6 at pH 3.0;
likewise, the Zn2+-FA stability constant at pH 5.0 1s 3.7, but is 2.4
at pH 3.0 (Schnitzer 1980).  Because  of  this effect of  pH on
metal-organic complexation, one might expect lake acidification to result
in decreased concentrations of organlcally-complexed metals and
correspondingly higher concentrations of  free  inorganic trace metals.
Simultaneously, the decreases in pH  could  lead  to  increased protonation
of organic acid functional groups, thereby increasing the hydrophobic
character of the organic adds.  This process  could affect  the adsorption
of humlc materials on mineral surfaces and could also affect  Interactions
between hum1c/fulv1c monomers.  The net result of  this  could  be to
                                  4-109

-------
increase clay interlayer adsorption  of fulvic  acids  (with associated clay
degradation) and to increase the  polymerization and  settling of aquatic
humic materials (Schnitzer 1980).

     Along similar lines, there may  be very  important biological
consequences resulting from acidification of natural waters containing
metal-organic complexes.  Driscoll et al. (1980) and others have already
shown that free inorganic species concentrations of  trace metals like
aluminum are significantly more toxic than are the organically-complexed
forms.  Thus, where atmospheric deposition leads to  a shift from
organically complexed to free inorganic species of trace metals, there
may be attendant impacts on aquatic  biota.

4.6.3.4  Photochemistry--Another  interaction that has been described is
the effect of decreasing pH on the coloration  or light absorption of
aquatic humic materials.  For instance, Schindler (1980) and Schindler
and Turner (1982) found that lake coloration and extinction coefficients
decreased with decreasing pH, even though no measurable change in the DOC
occurred.  This change in lake transparency  resulted in an increase in
primary productivity in the experimental lake.  In addition, the
acid-induced increases in transparency accelerated the rates of
hypolimnion heating and thermocline  deepening; at the same time, there
was no significant effect on the  lake's total  heat budget.  In terms of
processes, the data were interpreted to indicate that acidification
caused a qualitative change in the structure of aquatic humus and its
ability to absorb light.  Aimer et al. (1978)  also found evidence of
changes in lake transparency associated with lake acidification in
Sweden; however, they observed lower concentrations  of DOC in transparent
acidified lakes.  According to their data, this scavenging of organic
carbon from the lake water column may have been largly due to the
formation of insoluble organic-aluminum coloids and  the subsequent
sedimentation of these particulates  to the lake bottom.

4.6.3.5  Carbon-Phosphorus-AIuminum  Interactions--The potential impact of
acidic deposition upon aluminum leaching and phosphorus availability has
been discussed in Section 4.6.2.5 and described by Dickson (1980) and
Cronan and Schofield (1979).  As  Dickson (1980) has  shown experimentally,
increased concentrations of inorganic aluminum in freshwaters may cause
increased precipitation of aluminum  phosphates from  the water column,
resulting in decreased biological availability of phosphorus.  However,
where humic materials are present, the organic ligands will tend to bind
the aluminum preferentially, leaving the phosphorus  uncomplexed.  There-
fore, one would assume that where one finds  increased  concentrations of
aquatic humic materials these will tend to decrease  the toxic potential
of aluminum leached from soils and will tefid to preserve the availability
of phosphorus in aluminum-rich waters.

4.6.3.6  Effects of Acidification on Organic Decomposition in Aquatic
Systems--Lake and stream acidification associated with atmospheric
deposition may also cause reductions in the  rate of  organic matter
turnover and may ultimately lead  to  decreased  nutrient cycling and
                                  4-110

-------
availability (Chapter E-5, Section 5-8).  Traaen (1980)  found that
organic matter decomposition was retarded at pH 4.0 to 4.5 compared  to
control streams and suggested that this effect could be important for
lakes dependent upon allochthonous inputs of carbon.  Friberg et al.
(1980) observed that leaf litter decay was much slower in an acid stream
(pH 4.3 to 5.9) than in a paired stream at pH 6.5 to 7.3.  This  was
interpreted to indicate that stream acidification caused biotic
disturbances among the aquatic decomposer populations.  Finally,  Francis
and Hendrey (1980) compared the decomposition rates for leaf litter  in
three nearby lakes at pH 5.0, 6.0, and 7.0.  Results indicated that
decomposition of beech leaves was inhibited considerably and bacterial
populations were approximately an order of magnitude lower in the most
acidic lake.  These studies suggest a need to investigate what holistic
import reduced organic matter turnover in acidified aquatic systems  will
have.

4.7  MITIGATIVE STRATEGIES FOR IMPROVEMENT OF SURFACE WATER QUALITY
     (C. T. Driscoll and G. C. Schafran)

4.7.1  Base Additions

     The most effective means of regulating acidification would  be to
control hydrogen ion inputs.  For atmospheric inputs this involves many
political, social, economic, and energy related considerations.   An
alternative strategy is to symptomatically treat acidified waters by
chemical  addition.  Various substances have been proposed for use as
neutralizing agents (Grahn and Hultberg 1975);  however,  only lime [CaO,
Ca(OH)2 and limestone (CaC03)] have been used to any extent.  Two
base addition strategies have been practiced:  direct lake addition  and
watershed/stream addition.  While direct lake addition is the less
expensive approach, the relative effectiveness of the two strategies has
not been evaluated.  In addition, the positive and negative consequences
of these strategies have not been fully evaluated.

     A variety of methods for the treatment of acidic waters associated
with mine drainage have been researched and developed (Hodge 1953,
Pearson and McConnell 1975a,b).  Because mine drainage is often
extremely acidic and contains elevated levels of hydrolyzing metals,  it
is extremely difficult to extrapolate base addition concepts and
technology developed for mine drainage to dilute acidic  waters.
Therefore, this critical assessment will address only base addition  to
dilute water systems.  Fraser et al. (1982) and Fraser and Britt (1982)
compiled a detailed review of base addition technology and effects which
should be referred to for information beyond the scope of this document.

4.7.1.1  Types of Basic Materials--Several  types of basic materials  have
been used or proposed for neutralizing acidified surface waters.   These
materials include calcium oxide,  calcium hydroxide, calcium carbonate,
sodium carbonate, olivine, fly ash,  and industrial  slags (Grahn  and
Hultberg 1975).  There are many considerations  in selecting a base
material  to be used in neutralization.  Scheider et al.  (1975) have
summarized these considerations.
                                  4-111

-------
     1)   It must be readily  available  in large quantities.

     2)   It should be  relatively  inexpensive.

     3)   It must be safe  to  handle and store using conventional safety
         precautions.

     4)   It should have a high neutralization potential; i.e., a small
         quantity of chemical should be capable of neutralizing a large
         quantity of water.

     5)   Adding a known quantity  of chemical must produce a
         predictable change  in pH.  This is critical if pH sensitive
         organisms are already living  in the lake.

     6)   It must be amenable to a relatively simple application
         technique such that a large quantity of chemical could be
         applied in a  short  period of  time with a minimum of labor and
         equipment.

     7)   It must provide  for a natural deficiency in the aqueous acid
         neutralizing  capacity; i.e.,  it should be a normal component of
         the pH buffer system.

     8)   It should not initiate any significant ion exchange process in
         the lake sediment which  could impair the quality of the lake
         water.

     9)   It must not add  any extraneous contaminants to the lake water.

     Calcium oxide (quicklime, CaO) and calcium hydroxide (hydrated
lime, Ca(OH)2)  have been  used to  neutralize acidified surface waters.
These materials are relatively inexpensive and effective.  Lime is
generally used  in a powdered form and  is very soluble when added to
water.  Because lime is a soluble strong base, it readily increases the
pH of dilute solutions.   If  the solution is in contact with atmospheric
carbon dioxide  after strong  base  addition, the pH will slowly decrease.
This is because atmospheric  carbon dioxide will dissolve into solution,
neutralize the  hydroxide, and eventually form a bicarbonate solution.

                      ($2 introduction

          Ca2+  + 20H"  = Ca2+ + 20H" +  2C02 = Ca2+ + 2HC03-
Acidified waters generally have a low aqueous  buffering capacity.  As a
result, large increases in pH will  occur  upon  addition of typical
quantities of strong base (200 to 400 yeq £'1  yr"1).  Lake water
pH values which were below 5.0 prior to neutralization may increase to
above 10.0 immediately after strong base  addition.  This may result in
pH shock to organisms.  These problems are accentuated within certain
                                  4-112

-------
microenvironments,  particularly  if mixing is incomplete.  As a result,
dosage control  must be carefully monitored.

     Calcium oxide  is an  extremely corrosive material that generates
considerable heat when contacting water, which makes handling and
storage very difficult.  Calcium hydroxide is less hazardous and does
not generate heat upon contact with water.

     Calcium carbonate, commonly referred to as limestone, is a slightly
soluble base.  Dissolution  of limestone is slow, and a maximum pH of 8.3
is realized when an aqueous system is  in equilibrium with CaCOa and
atmospheric C02 (Stumm and  Morgan 1970).  The dissolution kinetics of
limestone are a function  of solution characteristics, impurities in the
stone, and the surface area of the stone (Pearson and McDonnell 1975a).
Limestone commonly  contains a significant amount of magnesium (often
called dolomitic limestone).   The greater the magnesium component in the
limestone the slower the  dissolution rate.  For acidified surface water
applications, enhanced dissolution rates of slightly soluble bases are
generally desirable.  Therefore, it is best to use a high purity stone
(e.g., low magnesium content).  Limestone can be obtained in a variety
of sizes.  Powdered limestone (agricultural limestone, 0 to 1 mm) is
often used in water neutralization efforts.  Dissolution is enhanced
because of the large surface area associated with the small particles.
Larger stone (0.5 to 2 cm)  may be used for limestone barriers in streams
(Section 4.7.1.3.2) or limestone contactors in springs.

     An important consideration  with regard to limestone dissolution  is
solution characteristics.  Dissolution rates are greatest in solutions
of low pH, low dissolved  inorganic carbon, and low calcium.  This
condition is characteristic of dilute  acidified waters.  Another
important consideration is the presence  of hydrolyzing metals  (Al, Fe,
Mn) and dissolved organic carbon.  Upon  increases in pH, these
components may deposit on the surface  of the stone,  inhibiting
dissolution and therefore decreasing the effectiveness of the base.
Pearson and McDonnell (1975a) observed that the dissolution rate of
CaCOa decreased by  up to  80 percent when CaC03 was coated with iron
and aluminum.

     Calcium carbonate is generally  favored for use  as a base  because
inorganic carbon is directly supplied  upon dissolution and dissolution
rates are relatively slow.  Aquatic  organisms are less prone to pH shock
with CaCOa treatment than with strong  base addition.
    Sodium carbonate (NagCOa, soda ash)  is a soluble  base which has
been used as a neutralizing agent (Lindmark 1981).  Sodium  carbonate  is
readily soluble and directly applies dissolved inorganic carbon to
solution.  Therefore, it is an effective base because there are minimal
losses due to incomplete dissolution while fluctuations in  pH are less
extreme.  Sodium carbonate is generally  an expensive  base and therefore
might not be used in lieu of calcium base sources (Ca(OH)2,
(see Section 4.7.1.2.2).
                                  4-113

-------
     Olivine (Mg,  Fe)2  Si04,  is  a  natural  silicate mineral that has
been used in neutralization efforts  (Hultberg and Andersson 1981).
Olivine is a continous  reaction  series  in  which magnesium and ferrous
iron can freely substitute for each  other.  Upon dissolution of Peg
$1*04, iron will oxidize and precipitate as Fe(OH)3 and thereby
contribute to the  acidification  of water.   Therefore, the effectiveness
of olivine as a neutralizing  agent increases with increasing magnesium
content.  Olivine  is a  slightly  soluble mineral; therefore, dissolution
characteristics and application  difficulties associated with biological,
chemical fouling will be in some ways similar to those associated with
     Fly ash is a material  of diverse chemical composition.  Western
coals have been found to produce  fly  ash  that is characteristically
basic (enriched in calcium)  while combustion of eastern coals generally
results in an acidic fly ash (enriched in iron) (Edzwald and OePinto
1978).  Basic fly ash has been shown  to be effective  in the neutraliza-
tion of acidified waters.  Neutralization by fly ash  is accomplished by
the release of hydroxyl  ion rather than inorganic  carbon to solution.

     Fly ash is a waste byproduct so  finding a way to use  it is
desirable.  Waste deposits  of basic fly ash are primarily  located in the
mi dwe stern United States while most of the acidic  waters are located in
the northeast.  Costs of transporting fly ash would probably be
prohibitive and certainly less economical than using  alternative
neutralizing agents located in the northeast.  Another problem
associated with fly ash is  trace  metal contamination.  Edzwald and
Depinto (1978) have indicated that release of trace metals to solution
from fly ash is comparable  to that released from sediments upon
acidification to pH 4.0.

     It has been proposed that industrial slags could be used in the
neutralization of acidic waters (Grahn and Hultberg 1975). One type of
slag formation is the use of calcium  carbonate to  produce  metals from
ores.  Basic slags formed in this and other processes vary considerably
with respect to physical and chemical  properties (Grahn and Hultberg
1975).  Basic slags are largely composed  of calcium (CaO)  and silicon
(SiOo) oxides.  While basic slags may contain similar calcium (CaO)
levels, dissolution rates and therefore neutralization characteristics
can vary considerably.  The dissolution rate of CaO within a slag is a
function of the manner in which CaO is bound to Si02  (Grahn and
Hultberg 1975).  Slags that increase  solution pH to the 6.0 to 8.0 range
and have long term neutralizing properties are the most desirable for
lake and stream management applications.   The determination of slag
dissolution characteristics may be accomplished through laboratory
testing.  The trace metal content of  slags may be  high; therefore,
potential for metal leaching exists.

     Costs associated with fly ash or basic slags, should  they be found
acceptable for use, would be largely  attributed to transportation and
                                  4-114

-------
handling, as these materials are waste  products.  Resistance to the use
of these materials may be encountered if  a substance the public
perceives as "waste" is recommended for application to pristine waters.

4.7.1.2  Direct Water Addition of Base—Direct water addition of base is
the most common management strategy for acidified lakes.  It has been
practiced in Sweden, Norway, Canada, and  in  the United States.  The
sources and sinks of hydroxide within an  acidified lake environment are
not quantitatively known; therefore there is no rational means of
computing base dose requirements.  Likewise,  there is no accepted method
for applying base to acidified lakes.

4.7.1.2.1  Computing base dose requirements.  Addition of base to
acidified waters should not be done arbitrarily.  For cost effective
use, a rational method for base dose determinations should be used;
however, to date none have been developed.   Hydroxyl ion sinks are
gaseous, aqueous, and solid in nature.  These sinks include atmospheric
carbon  dioxide, aqueous hydrogen ion,  aluminum, inorganic carbon, and
organic carbon, as well as exchange with  lake sediments.

     It is desirable to impart sufficient inorganic carbon ANC to a
water so that future inputs of acid may be neutralized without a drastic
decrease in pH.  Consumption of base by base neutralizing components
must be realized before residual  ANC can  be  imparted to water.  A
description of the aqueous base neutralizing capacity (BNC aq) can be
described by thermodynamic expressions:

     BNC aq = 2[H2C03] + [HC03']  + 3[Al+3] +  2[Al(OH)+2]

            + [A1(OH)2+] + 3[A1-F] + 3[A1-S04] + [RCOOH] + [H+]

            - [AKOHU-] - [OH-]

where   Al-F is the sum of all  aqueous  aluminum-fluoride complexes
          (mols £-1),

        Al-SO^ is the sum of all  aluminum -  sulfate complexes
          (mols r1), and

        RCOOH is the dissolved organic  carbon base neutralizing capacity
          (mols £-1).

Driscoll et al. (1983) found that aquo-aluminum levels in Adirondack
waters appear to follow an aluminum trihydroxide solubility model.  The
speciation of aluminum can be calculated  with aluminum, fluoride,
sulfate, and pH determinations as well  as pertinent thermodynamic
equilibrium constants.  Dissolved organic carbon can exert some base
neutralizing capacity in dilute waters.   From observations of Adirondack
waters, Driscoll and Bisogni (1983)  developed an empirical formulation
relating aquatic humus (dissolved "organic carbon, DOC) to the mols of
proton-dissociable organic acid/base:
                                 4-115

-------
          CT = 2.62 x 10-6  (DOC) + 7.63 x 10-6

where   DOC is the dissolved organic carbon concentration (mg C £-1)
        and

        Cj is the total,  organic carbon proton dissociation/
        association sites (mols £-1).

     A monoprotic proton  dissociation constant (pKa=4.4) was also
developed for Adirondack  surface waters.  From these relationships the
contribution of BNC from  aquatic humus may be quantified:
                    CT x
          [RCOOH] = 	
                       KJ.  PUT I .
                     d   L n  J


     Other metals (Cu, Mn, Zn,  Ni,  Fe)  are  not  included in the BNC
equation due to the low concentrations  usually  found in natural waters.
Collectively, BNC realized from these metals  is not substantial compared
to other aqueous components.  However,  these  metals may exert
substantial BNC when concentrations are high.   High concentrations would
most likely be found in acidified waters located near large industrial
areas, where atmospheric  deposition of  metals is high.  This condition
has been observed in the  Sudbury region of  Ontario, Canada, where levels
of copper and nickel have been  observed at  concentrations greater than
1.0 mg jr1 (Scheider et al.  1975, Van and Dillon 1981).

     If equilibrium with  atmospheric carbon dioxide is assumed, an upper
limit of the aqueous BNC  may  be estimated.  Driscoll and Bisogni (1983)
have made such an analysis to neutralize a  "typical" Adirondack lake
(Table 4-8) to pH 6.5 (Table 4-9).   It  is apparent that a substantial
portion of the aqueous BNC is associated with the hydrolysis of
aluminum, and this should not be overlooked when one computes base dose
requirements for acidified waters.

     In determining BNC of an aquatic system, one must consider the lake
sediment as well as the overlying water.  One of the consequences of
lake acidification is the accumulation  of organic sediments.  These
sediments have considerable  exchange capacity and contribute
significantly to the overall BNC of the aquatic system.  During the
acidification process, BNC associated with  sediment exchange sites
buffers the overlying water.  Upon  neutralization, the sediment
exchanges back into the water column, consuming added base.
Neutralization of the water  occurs  quickly  after base addition, whereas
the exchange with the sediment  may  be slow.

     Base additions (CaC03 and/or Ca(OH)2)  of 477, 196 and 477 yeq
£-1 were applied to Middle,  Lohi, and Hannah  Lakes, respectively, in
the Sudbury region of Ontario,  Canada (Dillon and Scheider 1983).  Of
                                  4-116

-------
         TABLE 4-8.  COMPONENTS OF BASE NEUTRALIZING CAPACITY IN
                      TYPICAL ADIRONDACK LAKE WATER
                (ADAPTED FROM DRISCOLL AND BISOGNI 1983)
Parameter                                         Value

pH                                                4.95
Inorganic Monomerlc Aluminum                      0.2 mg Al  &
Aluminum Fluoride forms                           0.105 mg Al
Aluminum Sulfate forms                            0.005 mg Al
Free Aluminum                                     0.04 mg Al J
A1(OH)2+                                          0.03 mg Al J
AKOH)2+                                          0.02 mg Al
TOC                                               5.0 mg C H'
           TABLE 4-9.  AMOUNT OF BASE REQUIRED TO NEUTRALIZE
                      BASE NEUTRALIZING CAPACITY OF
                 TYPICAL ADIRONDACK LAKE WATER TO pH  6.5
                (ADAPTED FROM DRISCOLL AND  BISCOGNI 1983)
Acid component                              Base  required  (eq £~
Hydrogen Ion                                      1.1  x  (lO"5)
Carbonate                                         1.3  x  (10~5)
Aluminum                                          2.0  x  (10~5)
Organic Carbon                                    0.4  x  (10~5)
                                  4-117

-------
these applications 161, 86 and 148 yeq £-1 (or 34, 44 and 31
percent, respectively, of the base applied) were consumed by reactions
with lake sediments.  Therefore, sediment reaction would appear to be a
major component of overall-lake base demand.

     Determining sediment base demand of a lake is difficult; no
accepted methods are  available.  Scheider et al. (1975) determined the
base demand of sediments from Sudbury lakes by titrating sediments with
Ca(OH)2 to a pH of 8.0 and arbitrarily assuming a reactive layer of 5
cm in the lakes.  Intralake variations in sediment base demand up to a
factor of 10 were noted.

     Through studies  of base application to improve fish production in
southeastern U.S. lakes, Boyd (1982) has developed a table in which
sediment pH and texture are used to calculate base dose requirements.

     Menz and Driscoll (1983) used experimental data obtained from
Sudbury, Ontario and  Adirondack, NY, liming experiments to develop a
sediment base-demand  model.  The base-demand of sediments (meq m~2) as
a function of the increase in ANC (due to base addition) of the water
column was fit to a Langmuir-type model:
                    SDmax  + ANCa
                       K  +  ANCa

where:  SD is the sediment  demand of base  (meq m~2),

        SDmax is the maximum demand of base  (meq m~2) »

        ANCa is the increase in water column ANC after base addition
                 (yeq r1), and

        K is the sediment demand constant  (yeq jr1).

This sediment demand model  was  coupled to  aqueous thermodynamic
calculations (see above)  to determine the  overall base demand of a lake.
Base dose calculations using the simle model proposed by Menz and
Driscoll (1983) depend on the volume of water to be treated, the
sediment surface area, the  solution water  quality, the length of time
over which the lake is to be treated, and  the ANC the lake is to be
increased to after treatment.

     Another element of uncertainty in base  dose calculations is base
dissolution efficiency.  For soluble bases (e.g., Ca(OH)?, Na2COa)
a dissolution efficiency  of 100 percent may  be a reasonable assumption.
However, the dissolution  efficiency of sparingly soluble bases (e.g.,
CaCOs, olivine) will depend on  the method  of application, the size and
the impurity content of the base, and the  extent of base-particle
coating (e.g., Al , Fe, organic  matter) that  impeded dissolution.
Driscoll et al . (1982) observed an accumulation of CaCOa coated with
                                  4-118

-------
organic detritus and metals  within the sediments of a limed lake.
Conversely, Dillon  and Schelder  (1983) observed complete dissolution of
CaC03 after application to Sudbury lakes.

     To develop a rational means for determining base dose requirements,
further research is needed to enhance our quantitative understanding of
components that exert a base demand in acidic lakes and of base
application efficiency.

      Base dose application  rates have been reported in the literature.
In southern Sweden, direct water addition doses needed for neutraliza-
tion have been noted: 200 to 400 yeq &'1 annually (Bengtsson
et al. 1980), which corresponds  to 1000 to 1500 yeq ha'1 of
watershed yr-1.  Blake (1981) has reported dose requirements of
7340 yeq CaCOa ha of lake surface area"1 for initial treatment of
Adirondack lakes.  The period of time over which these levels are
effective was not reported.

4.7.1.2.2  Methods  of base application.  Managing acidified waters by
adding chemicals Is a relatively new concept that has been practiced to
only a very limited degree.  Chemical addition strategies have generally
evolved through trial and error, and there is no single, accepted method
for applying chemicals to surface waters.  The following are some of the
reported methods of chemical application.

     It has been suggested that  waters amenable to neutralization should
be ranked so lakes used for  fishing and recreation are treated first
(Blake 1981).  These waters  are  generally accessible lakes, which are
less costly to treat than remote waters.  To determine the benefit
derived from neutralization, a cost-benefit ratio can be used.  This
cost-benefit ratio (Blake 1981)  might compare the cost of neutralization
to the value derived by anglers. Lakes with a low cost-benefit ratio
might be considered for lake neutralization programs.  Lakes having long
retention times should be favored over those with shorter retention
times (< 1 yr).  Because lakes with short retention times experience a
relatively fast "washout" of base-induced ANC, these systems are
susceptible to reacidification  and  the effective period of
neutralization is short.

     Once lakes to be neutralized are selected, application procedures
must be planned.  The method of  application and the location of base
addition should be optimized for the maximum dissolution of base, worker
safety, and minimum cost.

     Several ideas for the  optimum  placement of CaCOs have been
presented in the literature.  Sverdrup and coworkers (Bjerle et al.
1982, Sverdrup 1982, Sverdrup  and Bjerle 1982) have developed a model to
describe CaCOs dissolution  after application in acidic lakes.  The
major parameters influencing CaC03  dissolution are particle settling
depth and solution characteristics.  Sverdrup  (1982) indicates that
particles larger than 60 ym in  diameter dissolve to only a limited
extent in dilute acidic systems  and therefore are of little use in lake
                                  4-119

-------
liming.  Calcium carbonate  resting on  (or within) lake sediments has
very slow dissolution  rates.  This may be attributed to burial, limited
turbulence,  or coating of CaC03 particles by hydrous metal oxides or
organic matter.  Therefore, dissolution during water column
sedimentation should be maximized for  the most efficient application of
base.  Sverdrup's (1982) calculations  suggest that CaCOs should be
applied in the deepest portion of a lake.

     Driscoll et al. (1982),  however,  indicate that turbulence will
enhance dissolution.  They  suggest CaCOs should be placed in the
littoral zone where turbulence will enhance the dissolution rate.
Within the littoral  zone, areas that are sandy and not laden with
organic detritus provide the  best location for CaC03 placement.  If
CaC03 is placed in organic  sediments,  particles may become buried or
coated with metal and/or organic matter.  If applied to the littoral
region, CaCOa should be dispersed so only a thin layer accumulates on
sediments.  This will  ensure  that a large surface area of base directly
contacts the water and increases dissolution efficiency.  Driscoll et
al. (1982) observed that when CaCOa was applied in a thick {> 0.5 cm)
layer coating by organic detritus and  metals curtailed dissolution; when
deposits were spread thin  (<  0.5 cm) the CaC03 dissolved before
becoming coated.

     The application of base  materials has been accomplished in several
ways.  Transport and application vehicles include trucks, boats,
helicopters, and airplanes.   The accessibility of the water to be
neutralized largely determines the method selected.  Two prevalent
methods of application are  by boat or  helicopter.

     Application by boat is usually limited to readily accessible lakes
and ponds.  For an efficient  operation, base transported by truck must
by easily transferred  to a  boat.  This necessitates unloading the truck
close to the water.  Lime  (Ca(OH)2) transported in bags is a commonly
used base in boat application.  These  bags are loaded onto the boat and
then emptied as the boat moves slowly  through the water.  Calcium
carbonate may also be  applied in this  manner.  Scheider et al. (1975)
mixed lake water and base  on  board a boat, water was pumped into a
hopper where base was  poured  from a bag and mixed, with the resulting
slurry discharged into the  backwash of the boat.  Using one 5 m boat and
a five-man crew, approximately 7.3 x 103 kg was applied in an average
working day.

     Large amounts of  powdered CaCQ$ have been applied to an
Adirondack lake by using a  pontoon barge ( ~ 3 x 103 kg capacity).
(Driscoll et al. 1982). The  base was  transported to the application
site and washed off the barge with water supplied by a gasoline-powered
fire pump.  In this manner a  three-man crew can apply 30 x 103 kg of
CaC03 in an average working day.

     Helicopters have  been  used to. transport large quantities of base to
remote areas.  Blake (1981) has discussed different methods of
helicopter application. Transporting  bagged lime by helicoper into
                                  4-120

-------
lakes in the winter was not a  viable  application method due to the
considerable labor required, extremely low temperatures, and swirling
snow that made flying difficult.   Another attempted procedure involved
mixing water and lime in a  fire-fighting water bucket and spreading the
slurry on the lake surface.  This  technique proved inadequate because
mixing equipment and a large crew  were needed.  In addition,
transporting large quantities  of water was required.  The most practical
method was direct lime application with a "bucket" ( - 1 x 103 kg
capacity) suspended from a  helicopter.  Upon flyi.ng over a lake the
pilot opened a trap door, thereby  dropping the lime to the lake.  The
most efficient variation of this operation involved two buckets, with
one in use while the other  was being  filled.
    In Norway, agricultural  limestone  (CaCOs) has been applied on a
frozen lake (Hinckley and Wisniewski 1981).  After limestone was applied
by a manure spreader in a 2  meter wide strip along the shoreline, a snow
blower blew the limestone and snow mixture into a 10-meter strip.  Upon
ice melt the base mixed with the lake  water, resulting in
neutralization.

     Sodium carbnoate (soda  ash) is not generally used as a neutralizing
material.  However, Lindmark (1981) has hypothesized that the sodium
from soda ash will exchange  with cations on  sediment exchange sites.
Treated sediments containing sodium may exchange with inputs of base
neutralizing capacity (e.g.  H+,  Al) and serve to buffer the lake
against reacidification.  Lindmark (1981) suggests that calcium binds
irreversibly with sediment exchange sites; therefore, calcium treatments
will not introduce the sediment  buffering that sodium treatments may
provide.  Lindmark (1981) argues that  the effectiveness of soda ash
offsets its higher chemical  cost (Table 4-10) and is therefore
economically competitive with more conventional basic materials (e.g.,
Ca(OH)2, CaCOs).  Lindmark' s hypothesis is controversial because
monovalent cations do not compete effectively with polyvalent cations
for sites on an exchanger in dilute solutions.

     To neutralize with soda ash a 10  percent solution of sodium
carbonate has been applied to sediments of an acidified lake (Lindmark
1981).  The sodium carbonate was mixed on land and pumped to a moveable
raft, and a land-based compressor that supplied air to the raft.  From
the raft, air and the sodium carbonate solution were piped to a chemical
rake (10 m wide) that moved  along the  lake bottom.  Bubbles of
compressed air were released 15  cm below the sediment surface, helping
to break up the sediment. Sodium carbonate  was injected directly within
the sediments.  In this manner,  good contact with the base was assured.
Unfortunately, data are not  currently  available to evaluate the cost-
effectiveness of sodium carbonate treatment.  Since soda ash is a
relatively expensive base, more  research is  needed before this
technology can be evaluated  as a potential lake management tool.

       Neutralizing acidified waters through base addition is a
relatively new strategy that has not yet been extensively practiced.
Application methods must be  chosen that will be compatible with the


                                  4-121

-------
      TABLE 4-10.  CHEMICAL COST COMPARISON OF NEUTRALIZATION AGENTSa
Chemical
CaC03
Ca(OH)2
Ca(OH)2
CaO
Na2C03
(Mg,Fe)2Si04b
H3P04
Form Equivalent
supplied weight
(g eq-1)
bags (325 50
mesh)
bulk 37
bags 37
bulk 28
bulk 53
bulk (100 86
mesh)
agricultural 5.75C
grade (70%
solution)
Costa
Mass Equivalence
basis basis
(dollars x (dollars
10-3 j
-------
constraints inherent with each site.   If base addition  becomes  a more
widespread procedure to mitigate acidification,  new  techniques  for
application will be developed, along  with the refinement of existing
methods.

4.7.1.3  Watershed Addition of Base—Watershed addition of base,
including stream treatment, is a relatively  new  strategy that has been
evaluated to only a limited degree.   Research addressing watershed
addition of base has been conducted  largely  by Swedish  scientists
(Bengtsson et al. 1980, Hultberg and  Andersson 1981).   This discussion
will essentially reflect upon the Swedish experience, in addition to
addressing pertinent biogeochemical concepts.

4.7.1.3.1  The concept of watershed application  of base.  The concept of
base addition to watersheds was developed to overcome the potential
introduction of BMC (H+, Al) to a neutralized lake by ground-water and
streams (Section 4.4.1.4).  Watershed/stream base treatment
theoretically should enhance the neutralization  of ground and stream
waters and result in a more complete  and compatible  neutralization.

     There is considerable experience to draw upon with respect to
neutralization of soils, since applying lime (agricultural  grade
CaC03) is a common agricultural practice. However,  forest ecosystems
are considerably different than agricultural  ecosystems, and it is
difficult to extrapolate from one to  the other.

     The acid/base chemistry of soil  systems is  extremely complex, with
reactions such as cation exchange, mineral dissolution, and biological
uptake all influencing soil solution  acidity. In forest soils  derived
from silicate bedrock, the bulk of the cation exchange  capacity may be
attributed to natural organic matter  and to  a lesser extent clay
minerals (e.g., kaolinite, vermiculite, illite).  The exchangeable
cations are largely basic cations (Ca, Mg, Na, K) and/or acidic cations
(Al, H).  At near neutral pH values,  the exchangeable cation pool is
largely comprised of basic cations.   As soil  pH  decreases,  the
exchangeable acidity (Al, H) is thought to increase. Another reaction
of interest is biological uptake of  cations.   Forest biomass requires
cationic nutrients (e.g., Ca, Mg, K)  for growth.  An aggrading  forest
will assimilate basic cations and tend to deplete soil  pools.

     Cation exchange and biological  uptake reactions are significant
considerations with regard to watershed liming.   Forest soils are
generally nutrient poor and elevated  in exchangeable acidity.  Upon
addition of base [Ca(OH)2, CaCOs] to  a forest soil,  a considerable
shift in ionic equilibria would ensue.  The  introduction of elevated
levels of calcium would result in a Ca2+ - H+ exchange  on soil
exchange sites.  The release of protons would neutralize the associated
hydroxide or carbonate introduced in  the liming  process.  Biological
uptake of calcium may result from calcium addition;  this would  generate
protons as well and neutralize the associated basic  anion.
                                  4-123

   409-262 0-83-11

-------
     Terrestrial  acid/base  reactions are much more complicated and more
poorly understood than  aquatic acid/base reactions.  It is difficult to
evaluate,  much less quantify, perturbations in acid/base chemistry that
result from watershed liming.  As a result, assessing the efficiency of
a watershed liming program  is difficult.

     Stream neutralization  techniques have also been attempted.  Stream
neutralization is of interest because streams are valuable aquatic
resources and maintaining stream water  quality is of concern.  An
important consideration is  the fact that acidic streams may flow into
acidic lakes and influence  lake biogeochemistry.  When an acidified lake
is limed,  it will still  experience the  introdution of BNC (Al, H) from
stream inputs.  Aquatic organisms (particularly fish) that use the
stream for feeding or reproduction may  be adversely affected by the
extensive aluminum hydrolysis resulting from the introduction of acidic
stream water to a neutralized lake.  Stream (and watershed) liming could
help minimize this water quality problem.

4.7.1.3.2  Experience in watershed liming.  Experiments with watershed
liming are limited to those conducted in Sweden (Bengtsson et al. 1980).
Agricultural lime (powdered CaC03, 0 to 0.5 mm) has been transported
to the watershed in large trucks and applied as a slurry with a sprayer
truck.  In this manner  one  man  is able  to apply 40 x 103 kg 9f CaC03
per day (Hinckley and Wisniewski 1981).  The CaCOa dose required to
achieve adequate neutralization of water systems is generally two orders
of magnitude greater than that of direct water addition (Bengtsson et
al. 1980).  This is undoubtedly due to  the many base consuming processes
that occur within forest soil systems.  Application rates are generally
in the range of 5000 to 7000 kg CaC03 ha~* yr-1.  Powdered olivine
(0 to 1 mm), a magnesium iron silicate, has also been used as a base in
watershed application experiments (Hultberg and Andersson 1981).

     Water quality information  resulting from watershed application
experiments has not been published; however, authors indicate that
watershed liming efforts have been  successful (Bengtsson et al. 1980,
Hultberg and Andersson  1981).   Hultberg and Andersson reported that some
damage to the terrestrial environment may be associated with liming.
Sphagnum moss was severely  damaged as a result of CaCOa addition.
Damage to lichens, spruce needles,  and  other types of moss was also
observed.  Similar damage occurred with olivine application experiments;
however, the extent of  damage was considerably less than that from
      addition.
     There are problems associated with stream neutralization practices.
 It is reasonable to say that no cost effective method  of  achieving
 stream neutralization has been developed.   Problems  center around the
 drastic temporal changes in water flow and water quality  that occur  in
 headwater streams.  During spring and autumn,  water  flow  and solution
 BNC are high.  During summer,  water flow and BNC are low.  A successful
 neutralization scheme must adequately account for the  tremendous
 temporal fluctuation in base dose requirements of acidic  streams.
                                  4-124

-------
     Four approaches have been attempted to achieve  stream
neutralization.  The simplest approach is CaC03  addition to the
streambed (Hultberg and Andersson 1981).  This has been attempted with
both coarse (5 to 15 mm) and fine (0  to 0.5 mm)  CaCOa-  Coarse CaCOs
will tend to stay in the stream bed,  but neutralization is generally
inadequate because of the rather low  surface area of the stone.  Fine
CaC03 will more readily dissolve (due to a greater surface area) but
is more influenced by stream turbulence. Powdered CaC03 tends to be
transported to pools, where it settles within organic detritus, or it
can be washed into a lake.  In these  sites CaCOa is  ineffective in
supplying BNC to streams.

     Another approach to achieve stream neutralization is the limestone
barrier.  Driscoll et al. (1982) constructed a limestone barrier of
perforated 55-gallon drums filled with CaCOa (5  to 15 mm), in an
attempt to neutralize an acidic stream.   The barrels were placed across
the width of the stream, 2-barrels high  with loose limestone filling
spaces between the barrels.  Screens  were placed upstream to filter out
debris that might clog the pores of the barrier.  Stream neutralization
was accomplished for approximately one week, largely due to fine
material associated with the larger stone.  The coating of the stone by
hydrolyzed aluminum, iron, and organic detritus  quickly curtailed
further neutralization.  The coating  diminished  calcium carbonate
dissolution and rendered the barrier  ineffective as  a means of achieving
neutralization.

     Diversion wells have been used to treat acidic  streams in Sweden
(Swedish Ministry of Agriculture Environment Committee 1982).  Diversion
wells consist of a cylinder embedded  within a stream bank or channel and
filled with CaC03-  A pipe diverts stream water, by  gravity, to the
cylinder.  Water is introduced through the bottom of the cylinder and
flows upward through the CaCOa bed.   Water neutralized by passage
through the cylinder bed overflows back  into the stream, increasing the
ANC.  The upflow velocity results in  particle abrasion, which aids to
restrict particle coating.  A series  of diversion wells may be placed in
a stream such that the inflow pipes will  be located  at various levels of
stream stage.  Thus during high-flow  conditoins  several diversion wells
would be operating and treating a large  volume of flow.  As stream flow
decreases, the stream depth would decrease;  therefore the number of
operating wells and volume of water treated would decrease.

     A fourth type of stream neutralization, automated base addition
systems, is the most effective means  of  supplying ANC to acidic streams.
However, they are extremely expensive in terms of both capital and
operating costs.  Swedish scientists  have used river silos (cylindrical
storage bins) to accomplish stream neutralization (Hinckley and
Wisniewski 1981).  These silos hold 30 x 103 kg  Of base and can meter
up to 1300 kg day"1 of base into a stream.   Each silo costs
approximately $20,000 (1981 dollars).  The rate  at which base is metered
from the silo to the stream is activated by pH or flow sensing devices.
An automated system, like the river silos,  would seem to be the best
                                 4-125

-------
means of applying an adequate  base dose  to varying water flow and
quality conditions.

     In addition to  cost,  however, there are several problems associated
with automated stream treatment systems.  The  silos may be used only in
easily accessible streams, and the automated operation is not entirely
reliable and will malfunction  occasionally.  Also, stream base addition
requirements are considerable  during  high flow conditions; silos must be
constantly refilled  during spring and autumn (Hinckley and Wisniewski
1981).  These problems are not severe in themselves, but they imply that
stream neutralization efforts  may be  interrupted periodically.
Interruption of base addition  will most  likely occur during high flow,
low pH conditions (spring, autumn, and winter) when water quality
conditions are most  critical.   Periodic  discontinuities in base addition
have severe implications for aquatic  organisms.  The response of water
quality and aquatic  organisms  to acute fluctuations in ANC, from
equipment malfunctions, needs  to be evaluated  before automated base
addition systems are implemented as part of a  stream management
program.

4.7.1.4  Water Quality Response to Base  Treatment—Lake water
neutralization "by base addition may be accomplished by direct base
addition or by watershed/stream input neutralization.  Few studies of
the water quality response in  groundwater or streams as a result of
neutralization are reported in the literature.  Likewise, an evaluation
of lake neutralization by watershed/stream input neutralization has not
been made.  As a result, this  discussion of the water quality response
to base treatment reflects only the results reported from direct base
addition studies.

     0   Transparency increases immediately following base addition
         especially  in colored waters (Van and Dillon 1981, Hultberg and
         Andersson 1981).  This appears  to be  due to the removal of dis-
         solved organic matter by co-precipitation with metals (Yan and
         Dillon 1981).  The long term consequence, however, has been the
         reduced transparency  in neutralized lakes.  Decreases in
         epilimnion  thickness  and decreased hypolimnetic temperatures
         have been associated  with these transparency changes (Yan and
         Dillon 1981).  Upon reacidification,  transparency has
         increased.

     o   A natural consequence of base addition is the resulting
         increase in pH.  Response of pH is dependent on the
         neutralizing agent used.  When  soluble base such as Ca(OH)2
         is applied, pH increases  sharply and  a maximum pH is realized
         shortly after addition. This increase in pH is followed by a
         decline in  pH due to atmospheric carbon dioxide influx.  When
         equilibrium with C02  is approached, stabilization of pH
         results.  If acidic inputs  are  significant through either
         streamflow, groundwater infiltration, or sediment cation
         exchange, a gradual but steady  decrease in pH will occur.
                                  4-126

-------
     When a slightly soluble base (e.g.,  CaC03)  is  added  to an aquatic
system, the pH increase is less dramatic.  With  calcium carbonate
addition the rate of pH increase depends  on particle  size and degree of
water contact.  Increases in stone surface area  exposed to solution
enhance dissolution rates, resulting in a more rapid  pH increase.  Acid
neutralizing capacity also increases after base  addition  (Bengtsson et
al. 1980).  ANC increases are initially considerable  but may decrease
significantly with slight decreases in pH.

     o   Increases in dissolved inorganic carbon result from
         neutralization.  Increases in pH from less than  5.0 to greater
         than 6.5 cause dissolved inorganic carbon  equilibria to shift
         from a ^003 (C02[aq] + ^003) dominated system  to a
         bicarbonate dominated system. If the environment is open to
         atmospheric carbon dioxide, increases in dissolved inorganic
         carbon will result.  When a noncarbonate base (e.g., CafOH^)
         is added, the increase in inorganic carbon is due entirely to
         atmospheric CO?, whereas when a  carbonate  base (e.g.,
         03003) 1S added, inorganic carbon increases  are  due to the
         base itself as well as atmospheric 003.

     °   Trace metals concentrations generally decrease after base
         additions to acidified waters.  Metals  found in elevated
         concentrations in acidified waters include Al, Mn, and Zn.  Of
         these trace metals aluminum is probably of the most concern,
         with concentrations of 0.2 to 1.0 mg Al  &"1  commonly
         observed (Driscoll 1980).  As solution  pH  increases, due to
         base addition, aluminum hydrolyzes and  precipitates.  It has
         been observed that aluminum in hydrolyzed  forms  is toxic to
         fish (Driscoll et al. 1980).   In Swedish liming experiments,
         fish kills were experienced shortly after  base application
         (Dickson 1978b).  Fish stocking  should  be  attempted only after
         hydrolyzed aluminum has settled  from the water column.

     Addition of base generally results in decreased  concentrations of
other trace metals in addition to aluminum (Scheider  et al. 1975, Yan et
al. 1977, Driscoll et al. 1982).  Sediment trap  analyses support water
column data, showing a rapid accumulation of metals in traps following
an increase in pH.  Decreases in trace metal  levels from the water
column may be explained by hydrolysis  and precipitation, or adsorption
on hydrous aluminum oxides formed by base addition.   Adsorption on metal
precipitates is also considered to be  a mechanism by  which dissolved
organic carbon and phytoplankton are removed from the lentic
environment.

     Sulfate response to neutralization appears  to  be minimal.
Comparing lakes that had been neutralized to control  lakes showed no
significant variation in temporal  changes in sulfate  (Scheider et al.
1975).

    Basic cation chemistry, excluding  the cation associated with base
addition, appears to be unaltered by neutralization.  Levels of calcium


                                  4-127

-------
are observed to Increase,  as expected, due to dissolution of
calcium-based chemical  neutralizing agents.  The temporal increase in
calcium concentration will  depend on the dissolution rate of base.
Calcium levels increase quickly with soluble bases (Ca(OH)2) and more
slowly with slightly  soluble bases (CaCC^).  Once the initial
dissolution has occurred,  calcium levels peak in concentration and then
decline due to export from the lake or exchange with sediments.

     A major problem  associated with lake neutralization is the
potential for reacidification.  Reacidification results in the
resolubilization of trace  metals  (Al, Mn, and Zn) which are presumably
introduced from the lake sediments (Driscoll et al. 1982).
Reacidification does  not result in an immediate reintroduction of
dissolved organic carbon (DOC).   It appears that DOC must be
reintroduced to the water  column  from terrestrial inputs (e.g., stream
and groundwater inflows) and therefore takes considerable time to
appear.  The loss of  DOC implies  that there are few available organic
ligands to complex trace metals,  particularly aluminum, that enter the
water column during reacidification.  This translates to a decrease in
water quality, particularly with  respect to potential for aluminum
toxicity to fish.

     Another consideration is  input of stream water (and groundwater) to
neutralized lakes.  The introduction of acidic water to a neutralized
lake results in a localized metal hydrolysis region at the stream (and
groundwater)--!ake interface.  This may have implications with respect
to aluminum toxicity  to fish,  particularly those fish that associate
with stream systems for reproduction and feeding.  If aluminum is
hydrolyzing in this environment it may be unsuitable for habitation by
fish.  Any program to stock fish  in a neutralized lake must consider
problems associated with acidic stream/groundwater quality entering the
lake environment.

4.7.1.5  Cost Analysis, Conclusions and Assessment of Base Addition--

4.7.1.5.1  Cost analysis.   It  is  extremely difficult to make a cost
comparison of different acidified lake management strategies.  It is
relatively easy to tabulate capital, chemical, labor, and operating
costs, but any economic evaluation must be based on the effectiveness of
the treatment.  Little is  known of the effectiveness and efficiency of
various treatment strategies.  As a result, any economic evaluation of
management strategies for  acidified waters should be viewed with
caution.

     Costs of chemicals that  have been proposed for use  in neutraliza-
tion efforts are listed in Table  4-10, which shows the considerable
range  in chemical costs.  This tabulation  is somewhat misleading  because
it does not incorporate application efficiency into the analysis.
Soluble bases  (Ca(OH)2, CaO,  Na2COa)  are undoubtedly the most
efficient means to add base, while  slightly soluble bases (CaCOs,
MgFeSi04) and phosphorus (Section 4.7.2.2) are potentially less
                                  4-128

-------
efficient.   Very little is  known  about the relative efficiency of
neutralization strategies,  and without such an understanding chemical
cost comparisons are difficult.

     Costs involved in neutralization efforts will vary greatly with
lake location and accessibility.  Blake  (1981) determined costs for six
accessible ponds treated (by  boat)  in 1977-78 and four remote ponds
treated (by helicopter)  in  1978-79, totaling 79 ha and 39 ha,
respectively.  Neutralization cost  for accessible ponds was $131 ha'1
while cost for the remote ponds was $341 ha"1.  These were
experimental efforts,  so costs may  be substantially reduced if base
addition is implemented on  a  routine basis.  Costs for liming remote
ponds by helicopter on a routine  basis were estimated to average $247
ha"1.  This was based on the  following costs: helicopter - $250
hr-1, lime - $44 x 10-3 kg-1  delivered onsite, travel expenses -
$100 day-1, the ability to  apply  4.5 x 103 kg of lime hr"1, and the
use of an eight-man ground  crew at  $35 day1 person"1 (Blake 1981).
Neutralization of a series  of lakes has  been shown to be the most
efficient operation.  Four  ponds  treated in 1977 by a three-man crew
cost approximately $74 ha"1 (Blake  1981).

     Costs associated with  application by boat are not detailed in the
method described by Scheider et al. (1975).  However, a comparative cost
analysis may be determined.  A five-man  crew using a 5-meter boat was
able to apply 7.3 x 103 kg  day"1  of hydrated lime.  Since the major
costs of base addition are  associated with labor and the cost of base, a
reasonable comparative estimate can be formulated.

     Using chemical, labor, and transportation cost data obtained by the
above and other investigators, Menz and  Driscoll (1983) estimated the
costs of neutralizing acidic Adirondack  lakes through a program of base
addition.  In this analysis lakes were subdivided as accessible (those
lakes with access by road so they can be treated by boat) and
inaccessible (those lakes with no road access and requiring helicopter
treatment).  Costs to treat accessible lakes for a 5-year treatment
period were approximately $50.75  per surface hectare.  Chemical
transportation cost to the  site represented the major component of cost
for the treatment of accessible lakes.   The cost to treat inaccessible
lakes for a 5-year treatment period was  approximately $500 per surface
hectare.  Most of this cost was associated with the cost of applying the
chemical.  It is noteworthy that  costs vary, from lake to lake, with the
desired target pH (and ANC),  and  with the treatment period.  Overall
results were derived from water quality  data from 777 of the 2,877
Adirondack lakes sampled to date  (Pfeiffer and Festa 1980). The
estimated annual cost for a 5-year  base  addition program for the lakes
in this sample would be in  the range of  2 to 4 million dollars,
depending on the specific target  pH of the water.

     Presently, the support for most lake neutralization programs comes
directly from government agencies.  Sweden has been the most active,
with over 300 individual projects involving 1000 lakes and waterways
                                  4-129

-------
(Bengtsson et al.  1980).   As  concern for the problem increases, private
groups (i.e., sportsman clubs, lake associations) may become actively
involved in neutralization programs.  However, limited resources will
probably prevent the neutralization and management of all acidified
lakes.

4.7.1.5.2  Summary - base  additions.  Base addition is currently the
most viable strategy available for managing acidified lakes.  Methods
used to compute base application  requirements are crude due to our lack
of understanding of the efficiency of treatment techniques and sediment
interactions.  A benefit associated with base addition is the alteration
of the chemical  environment (e.g., increases in pH and calcium,
decreases in trace metal levels).  Such a chemical alteration might
result in an environment more hospitable to desirable aquatic biota
(e.g., decreases in Sphagnum, increases in fish populations).  However,
in addition to the benefits associated with base addition, there are
costs.  These costs include financial as well as environmental costs.
Environmental costs include pH shock associated with dramatic increases
in pH, the problems associated with aluminum hydrolysis at the
stream-neutralized lake interface, and the potential for lake
reacidification.  These and other environmental costs have not been
fully evaluated prior to base addition of acidified lakes.

     Base addition has become a popular strategy to mitigate water
quality problems associated with  acidification.  However, before base
addition is implemented as a regional, acidified lake management
alternative it should be more thoroughly evaluated.

4.7.2  Surface Vlater Fertilization

     Soft water lakes are  generally thought to be phosphorus growth
limited (N/P > 12).  As a  result, fertilization by phosphorus addition
might serve as a means of  restoring acidified lakes.  However, this
hypothesis has been researched and evaluated only to a limited degree.
This analysis is a summary of the limited studies on nutrient addition
to acidified waters, as well  as an extrapolation of some concepts
pertinent to natural waters.  Further research is needed to effectively
evaluate lake response and consequences associated with nutrient
addition.

4.7.2.1  The Fertilization Concept—The concept of phosphorus addition
as a strategy for the management  of acidified lakes is twofold:

     1.  To supply ANC through biological uptake of nitrate; and

     2.  To increase aquatic biomass and species diversity.

     The idea of supplying ANC through biological uptake of nutrients,
is summarized by the following  stoichiometric expression (Stumm and
Morgan 1970).
                                  4-130

-------
     106 C02 + 16 N0a~ + H2P04~  + 122 H20  +  17  H+  (+ Trace Elements, Energy)

     photosynthesis +  * respiration

     C106 H262 °110 N16 p (alQal protoplasm) +  138 02
     This is a generalized relationship and may  vary  significantly from
ecosystem to ecosystem, as well  as temporally  within  a  given aquatic
system.  Regardless of the inadequacies of the stoichiometric
expression, it provides a framework through which  microbially mediated
changes in solution acid/base chemistry might  be understood.

     The stoichiometric expression suggests that uptake of nutrients by
algae will  result in the consumption of protons  or the  generation of ANC
within the aquatic environment.   This essentially  results from the
assimilation of nitrate as a nitrogen source.  For the  organism to
maintain an electroneutrality balance,  the uptake  of  nitrate must be
countered by an equivalent cation uptake (or anion release).  In the
above expression this is realized through  hydrogen ion  uptake.

     This expression is somewhat simplistic, for in actuality a number
of additional  factors should be  considered.

      1)  Although nitrate nitrogen is  generally the  predominant
          nitrogen source in aerobic waters, uptake of  ammonium or
          organic nitrogen could occur. Under these  circumstances the
          stoichiometry would significantly change.   In fact, assimila-
          tion of ammonium as a  nitrogen source  would result in
          consumption of ANC (Brewer and Goldman 1976).

      2)  Plants require certain cations as nutrients (e.g., Ca£+ t
          Mg2+, Fe).  The uptake of cations by algal  protoplasm would
          diminish the quantity  of ANC  generated through photo-
          synthesis.

      3)  Although carbon fixation through photosynthesis results in
          generation of ANC, respiration will  result  in consumption of
          ANC.  This process may partially account for  why acidic lakes
          have a higher ANC in summer months than  in  winter months.
          Therefore, only net removal of reduced nitrogen associated
          with algal material through lake outflow or permanent burial
          in sediments will  result in a net production  of microbially
          mediated ANC.

     The concept of acid neutralizing changes  generated by phytoplankton
growth has  been studied by Brewer and Goldman  (1976).   Such processes
may be important in dilute water acid/base chemistry.  According to the
above stoichiometric expression,  4.8 x  10~3 yeq  of ANC  would be
generated per microgram of net algal  biomass produced,  or 5.5 x 10-1
yeq of ANC. would be generated per vg of net phosphorus  fixed by
algal uptake.
                                  4-131

-------
     The second reason for nutrient addition is to replenish  the biomass
of acidified lakes.  Hendrey et al. (1976)  have suggested  that
phytopiankton biomass is reduced by lake acidification.  Dillon et  al.
(1979) suggest that phytoplankton biomass is better correlated with
total  phosphorus levels than with pH.   However, acidification may alter
phosphorus cycling (Section 4.7.2.2).   Nutrient addition may  help
replenish phosphorus lost (possibly)  by acidification  and  increase  the
productivity and species composition  of these lakes.

4.7.2.2  Phosphorus Cycling in Acidified Water—Phosphorus cycling  is
reasonably well understood in circumneutral  lakes  (Hutchinson 1975).
Generally phosphorus will  enter a lake through direct  atmospheric
precipitation, groundwater, or stream  flow.   It may be exported from the
lake by groundwater or stream flow.  Within  the lake,  phosphorus may be
assimilated by phytoplankton or macrophytes.   Once in  the  form of
particulate phosphorus, it may be consumed  by organisms, released to the
water  by oxidation reactions, or lost to the sediments.  Within the
sediments, phosphorus may be released  by decomposition processes.  This
released phosphorus may bind with aluminum,  calcium, or iron  or diffuse
vertically back into the water column.

     In acidified waters aluminum might alter phosphorus cycling through
precipitation or adsorption reactions.  Aluminum can directly
precipitate with orthophosphate to form A1P04 (varacite).   A  more
plausible mechanism by which aqueous  phosphorus levels might  be
regulated is adsorption on hydrous aluminum  oxides (Huang  1975).  The
adsorption is pH dependent with a maximum near pH  4.5.  It is likely
that increases in pH of acidic water result  in the formation  of hydrous
aluminum oxides.  These oxides would  serve  as an adsorbent that could
effectively scavenge phosphorus from the water column.

     Upon nutrient addition to an acidic lake, competition between algae
and aluminum for a given phosphorus molecule will  ensue.   It  is
difficult to state how phytoplankton uptake  of phosphorus  is  altered by
the presence of aqueous aluminum.   This competition is undoubtedly
complicated and altered by environmental  conditions such as pH, general
water chemistry, light,  and temperature.

     Although changes in water quality may result  on a short  term basis,
most of the added phosphorus will  be lost to  the sediments  (Schindler et
al.  1973, Scheider et al.  1976).   The  degree  to which  sedimented
phosphorus diffuses back to the water  column  is virtually  unknown for
acidic lakes.  However,  because these  systems  are  generally aerobic,
have reduced decomposition rates,  and  undoubtedly  contain  significant
levels of amorphous iron and aluminum  oxides  that  potentially bind
phosphorus,  it is doubtful  that significant vertical diffusion of
phosphorus occurs.   If fixed nitrate associated with algal  uptake of
phosphorus is lost from the system, applying  phosphorus has been
efficient from the standpoint that ANC was produced in the water column.
However,  if fixed nitrate  reaches  the  sediment,  is  oxidized,  and
diffuses  to the water column while the associated  phosphorus  remains in
the  sediment, phosphorus application would be  inefficient  (no net
                                 4-132

-------
generation of ANC to the water column resulting).  Schindler et al.
(1973) have indicated that  nitrogen sedimentation and removal are less
efficient than phosphorous  sedimentation and removal.

4.7.2.3  Fertilization Experience  and Water Quality Response to
Fertilization—As mentioned previously there has been limited experience
with fertilization of acidic lakes.  Most of the work has been accom-
plished by Canadian scientists (Scheider et al. 1975, 1976; Scheider and
Dillon 1976; Dillon et al.  1977, 1979).  Generally a desirable water
column phosphorus level  is  chosen  for a particular lake, and a model
such as that of Dillon and  Rigler  (1974) is used to calculate the
required phosphorus dose.   Usually ^04 is applied because of its
low cost, ease of handling, and  solubility (Table 4-10).  Application is
usually made in the late spring  or early summer; periodic additions may
be made throughout the summer to enhance assimilation efficiency.

     Nutrient addition has  generally been used to increase the standing
crop of food chain components within a lake.   To accomplish this,
phosphorus addition has generally  been practiced after liming.
Phosphorus consuming reactions are minimized by precipitating aluminum
with base and allowing aluminum  to settle out  of the water column prior
to any phosphorus addition.

     Few data have been reported on ANC changes as a result of
phosphorus addition.  However, Dillon and Scheider (1983) observed
decreases in inorganic nitrogen  (largely nitrate) and increases in total
organic nitrogen folowing nominal  orthophosphate additions of 10 to 15
yg p £-1 to neutralized lakes (Hannah and Middle) in the Sudbury
region of Ontario, Canada.   They calculated the theoretical  increase in
ANC resulting from observed changes in nitrogen chemistry for fertilized
lakes (Hannah and Middle) in comparison to a neutralized lake that
received no phorphorus addition  (Lohi Lake).   The ANC generated from
nitrogen transformations for the fertilized lakes was 2 to 8 yeq
£~1 greater than the control lake. In addition, the ANC generated
from nitrogen transformations declined dramatically after phosphorus
additions were terminated.

     Observed changes in aquatic biota have been more significant.
Small additions of total phosphorus resulted in significant  increases in
phytoplankton biomass of neutralized Canadian  lakes (Dillon  et al.
1979).  No single observation in phytoplankton species composition  has
been reported.  Shifts to communities dominated by Chrysophytes
(Langford 1948), by blue greens  (Smith 1969),  and by different groups
in different years of fertilization (Schindler et al. 1973)  have been
reported.  Shifts in green  or bluegreen algae  dominance can  generally be
attributed to the nitrogen  to phosphorus ratio within the lake
(Schindler 1977).

     Dillon et al. (1979) observed changes in  phytoplankton  resulting
from  small levels of phosphorus  added to a limed lake (Middle Lake,
Ontario).  In the first year after addition blue-green algae biomass
increased significantly. The second year after fertilization, green


                                  4-133

-------
algae were generally dominant.  Fertilization of a second lake (Hannah
Lake, Ontario)  resulted  in  an  increase in biomass but no change in the
structure of the phytoplankton community.  Although increases in
phytoplankton biomass were  evident, no conclusions with regard to
changes in zooplankton population could be made from this study.

     In enclosure experiments  within a limed lake, Scheider et al.
(1975) observed that fertilization with phosphorus and wastewater
effluent resulted in an  increase in the standing stock of bacteria,
phytoplankton,  and zooplankton.  Hultberg and Andersson (1981)
investigated nutrient addition as a means of supplementing liming
efforts in Sweden.  They reported few results except for a shift in lake
phytoplankton from Peridineans to primarily chlorophyceans, which they
attributed in part to fertilization.

     Little work has been done with water chemistry response to
phosphorus addition.  Dickson  (1978b) has observed the precipitation of
phosphorus added to acidic  lake water; precipitation was most dramatic
at pH 5.5.  The presence of DOC inhibited the precipitation of
phosphorus by aluminum.   Scheider et al. (1975) observed decreases in
phosphorus added to enclosure  experiments.  They attributed this to
precipitation of the phosphorus by metals.

4.7.2.4  Summary-Surface Water Fertilization—It is difficult to
critically assess phosphorus addition as a management strategy to
improve the water quality of acidic lakes because the general process
has not been effectively evaluated.  While the chemical costs associated
with  phosphorus addition are low  (Table 4-10) applications may not be
efficient, particularly  in  view of potential interactions with aluminum
(Schindler et al. 1973,  Scheider  et al. 1976).   In the few studies
conducted, the benefits  accrued to the ecosystem have not been
evaluated.

4.8   CONCLUSIONS

      Acidification of lakes and  streams, with resultant biological
damage, has been widely  acknowledged in the last decade  (NAS 1981, NRCC
1981, U.S./Canada 1982).  Assessing  causal  relationships remains
difficult, however, because effects of acidic deposition on any  one
component of the terrestrial-wetland-aquatic  systems depend on not only
the composition of the atmospheric deposition but also on the effect of
the atmospheric deposition on  every  system  upstream from the component
of interest.  Composition of aquatic systems  results, moreover,  from
biological  processes  in addition  to  chemical  and physical  processes;
thus, assessing results of acidification on all  three processes  is
required.   Our  knowledge of past,  current,  and  future acidification
trends, of critical processes  that control  acidification, and of the
degree  of  permanency  of biological  effects  remains  incomplete and
subject to  debate.
                                  4-134

-------
of aquatic chemistry and acidic deposition,  the chapter listed those
characteristics of terrestrial  and aquatic systems that ameliorate or
enhance the effect of acidic deposition.  It then discussed aquatic
systems'  theoretical and practical  sensitivity to acidic deposition and
identified locations of sensitive and affected systems.  The chapter
also considered the interaction of aquatic acidification with the metal
and organic biochemical cycles  and then concluded by discussing
alternative methods for improving water quality where acidification has
occurred.

     The following statements summarize the  content of this chapter.

    0   Each of several components of aquatic or terrestrial systems may
        assimilate some or all  acidic deposition falling in a watershed.
        These components are vegetative canopy, soils, bedrock,
        hydrology, wetlands, or an aquatic system itself (Section
        4.2.1).

    0   Soils assimilate acidic deposition through dissolution, cation
        exchange, and biologic  processes.  Generally, soils containing
        carbonate materials have abundant exchangeable bases and can
        assimilate acidic deposition  to an almost unlimited extent.
        Soils that contain no carbonate materials can assimilate acidic
        deposition because of cation  exchange reactions, silicate-
        mineral dissolution reactions, and in some cases Fe and Al oxide
        dissolution.  Assimilation ability is affected by soil chemical
        nature (especially CEC  and BS), the  permeability at each layer,
        the surface area of the soil  particles, and the amount of soil
        in the watershed (Section 4.3.2).

    0   Hydrology, specifically flow  paths and residence times, can
        determine the extent of reactions between strong acid components
        of deposition and each  component the water contacts.  Flow paths
        and residence times are controlled by many factors, including
        topography and meteorology (Section  4.3.2.4).

    0   Alkalinity or acid neutralizing capacity (ANC) determines a
        lake's instantaneous ability  to assimilate acidic deposition,
        but the ANC renewal  rate depends upon the ANC supply rate from
        the watershed.  In addition,  internal production of alkalinity
        is important, especially in lakes with low alkalinity.  Because
        biological processes can alter the relative amounts of acidity
        and alkalinity within the body of water, nutrient status is very
        important in determining the  sensitivity of a lake to
        acidification (Section  4.3.2.6).

    0   Aquatic systems sensitive to  acidification by acidic deposition
        are commonly waters with pH and alkalinities towards the lower
        end of the spectrum. The .boundary between sensitive and
        insensitive that is used is 200 yeq  £-1 of alkalinity
                                  4-135

-------
(Section 4.3.2.6.1).   This concentration  is chosen because in
North America acidic  precipitation  has  resulted in about 100
peq r-1 of potential  acidification  of surface water
(Harvey et al. 1981,  Church and  Galloway  1983), and because
biological effects due to acidification begin when aquatic
systems reach alkalinities of  ~  100 yeq £-1.

Regions in North America  contain aquatic  systems sensitive to
acidification.  These regions  are found throughout much of
eastern Canada;  New England; the Allegheny, Smokey, and Rocky
Mountains; and the Northwest and North  Central United States
(Galloway and Cowling 1978, NAS  1981, NRCC 1981).  However, a
large amount of more  detailed  survey work is required to
determine the levels  of alkalinity  and  degree of sensitivity
(Section 4.4.3).

Studies uniformly point to acidification  of some surface waters
in eastern Canada and the northeastern  United States (Section
4.4.3.1.2).

Although changing land use may locally  alter the pH regime of
lakes and streams, it appears  that  regional lake acidification
and episodic pH depression occur in response to increased
atmospheric deposition of strong acid,  primarily H2S04
(Section 4.4.3.3).

Addition of acidic deposition  to terrestrial and aquatic systems
can disrupt the natural biogeochemical  cycles of some metal and
organic compounds to  such a degree  that they can cause
biological effects (Section 4.6).  The  chemical form of
dissolved metals is important  in determining the total mobility
of a metal and the biological  effects related to acidification
of aquatic ecosystems.  Acidification increases the
concentration of many metals in  surface waters and changes
speciation toward more biologically active  forms.

Waters may be treated with base substances  to neutralize the
effects of acidic deposition.   Only lime  and limestone have been
used to any extent in either direct lake  additions or
watershed/stream additions.  Several other  materials have been
proposed, but tests for effectiveness and operability must be
conducted.  Organic carbon addition and surface water
fertilization have also been proposed but also must be tested
(Section 4.7).
                          4-136

-------
4.6  REFERENCES

Alfheim, I., J. Gether, and G. Lunde.  1978.  Access of some  atmospheric
non-polar organic micropollutants to lakes and rivers.   FR  14/78,  SNSF,
Oslo-As, Norway.

Alfheim, I., M. B. Sti6bet, N. Gjrfs, A. Bjjrtrseth,  and S.
Wilhelmsen.  1980.  Analysis of organic micropollutants in  aerosols,  pp.
100-101.  j£ Ecological Impact of Acid Precipitation.  Proceedings of an
International Conference,  Sandefjord, Norway,  March 11-14,  1980.
Drablj6s, D. and A. Tollan, eds.  SNSF-project, Oslo-As, Norway.

Aimer, B., W. Dickson, C.  Ekstrom, E. Hornstrom,  and U. Miller.  1974.
Effects of acidification on Swedish lakes.  Ambio 3(l):30-36.

Aimer, B., W. Dickson, C.  Eckstrom, and E. Hornstrom.  1978.   Sulfur
pollution and the aquatic  ecosystems, pp.  273-311.   Ir±  Sulfur  in the
Environment, Part II. Ecological  Impacts.   J.  0.  Nriagu,  ed.   J. Wiley
and Sons, Inc., New York,  NY.

Altshuller, P. A. and G. A. MacBean.  1980.  Second Report  of  the
U.S.-Canada Research Consultation Group on Long Range Transport of Air
Pollutants, November 1980.  40 pp. (Available  from Environ. Can.
Ottawa).

American Public Health Association.  1976.  Standard Methods  for the
Examination of Water and Wastewater.  Fourteenth  Edition.

Andersson, G., S. Fleischer, and W. Graneli.  1978.  Influence of
acidification on decomposition processes in lake  sediment:  Verh.
Internat. Verein. Limnol.  20:802-807.

Anon.  1981.  Acid sensitivity survey of lakes in Ontario.  Ontario
Ministry of the Environment.  APIOS 002/81. 144  pp.

Arnold, D. E., R. W. Light, and V. J. Dymond.   1980. Probable effects
of acid precipitation on Pennsylvania waters.   EPA-600/3-80-012.   21  pp.

Baker, J. P. and C. L. Schofield.  1980.  Aluminum  toxicity to fish as
related to acid precipitation and Adirondack surface water  quality, pp.
292-293.  In Ecological Impact of Acid Precipitation.  Proceedings of an
International Conference,  Sandefjord, Norway,  March 11-14,  1980.
Drabljds, D. and A. Tollan, eds.  SNSF-project, Oslo-As, Norway.

Barnes, I.  1964.  Field measurement of alkalinity and  pH.  U.S. Geol.
Survey Water Supply Paper  1535-H.  17 p.

Barrie, L. A.  1982.  Environment Canada's long-range transport of
atmospheric pollutants programme:  Atmospheric studies, pp. 141-161.   lr^
Acid Precipitation:  Effect on Ecological  Systems.   F.  M. D'ltri,  Ed.
Ann Arbor Science, Ann Arbor MI.
                                  4-137

-------
Barrie, L. A. and A.  Sirois.   1982.  An analysis and assessment of
precipitation chemistry  measurements made by CANSAP (The Canadian
Network for Sampling  Precipitation):   1977-1980 Report AQRB-82-003-T.
Atmospheric Environment  Service, Downsview, Canada, 163 p.

Bates, R. G.  1973.   Determination of  pH, theory and practice.  John
Wiley and Sons,  New York,  NY.

Beamish, R. J. and H. H. Harvey.  1972.  Acidification of the La Cloche
Mountain lakes,  Ontario  and resulting  fish mortalities.  J. Fish. Res.
Board Can. 29:1131-1143.

Beamish, R. J.,  W. L. Lockhart, J. C.  Van Loon, and H. H. Harvey.  1975.
Long-term acidification  of a lake and  resulting effects of fishes. Ambio
4:98-102.

Bell, J. A. (ed.).  1967.  Chemical Principles in Practice.  Addison-
Wesley, Reading, MA.   273  pp.

Bengtsson, B., W. Dickson, and P. Nyberg.  1980.  Liming Acid Lakes in
Sweden.  Ambio 9:34-36.

Birks, H. J. B.  and H. H.  Birks.  1980.  Quaternary palaeocology.
Edward Arnold (Publishers) Limited, London.

Bisogni, Jr., J. J. and  C. T.  Driscoll, Jr.  1979.  Characterization of
the pH buffering systems in dilute Adirondack surface waters.  Research
Project Technical Completion Report, OWRT Project No. A-073-NY.

Bjerle, I., G. Rochelle  and H. Sverdrap.  1982.  Limestone dissolution
in acid lakes, Vatten 38:156-163.

Blake, L. M.  1981.   Liming Acid Ponds in New York.  ln_ Proceedings for
the Effects of Acid Precipitation on Ecological Systems:  Great Lakes
Region, Institute of  Water Research, Michigan State University, East
Lansing, MI.  April  1-3, 1981.

Bloch, A.  1982.  Murphy's Law, Book Three.  Price/Stern/Sloan, Inc.,
Los Angeles, 93 pp.

Bloom, P. R., M. C.  McBride and R. M.  Weaver.  1979.  Aluminum organic
matter in acid soils:  Buffering and solution aluminum activity.  Soil.
Sci. Am. J. 43:488-493.

Bloomfield, C.  1957.  The possible  significance of polyphenols in soil
formation.  J. Sci.  Food Agric. 8:389-392.

Bobee, B., Y. Grimard, M.  Lachance and A. Tessier.  1982.  Nature et
Etendue De L'acidification Des Lacs  Du Quebec, Rapport Scientifique No
140 pour le Ministere de 1'Environnement  du Quebec, Service de la
qualite des eaux, 234 pp.
                                  4-138

-------
Bohn,  H.,  B. McNeal  and G.  O'Connor.  1979.  Soil Chemistry, Wiley
Interscience, New York,  N.  Y., p. 329.

Bowen, H.  J. M.   1966.   Trace Elements in Biochemistry.  Academic Press,
New York,  New York.

Boyd,  C.  E.  1977.  Evaluation of a water analysis kit.  J. Environ.
Qual.  6:381-384.

Boyd,  C.  E.  1980.  Reliability of water analysis kits.  Trans. Amer.
Fish.  Soc. 109:239-243.

Boyd,  C.  E.  1982.  Liming  fish ponds.  J.  Soil and Water Conserv.,
37:86-88.

Brewer, P. G. and J. C.  Goldman.  1976.  Alkalinity changes generated by
phytoplankton growth.  Limnol. Oceanogr. 21:108-117.

Buckman,  H. 0. and N. C. Brady.  1961.  The Nature and Properties of
Soils.  MacMillian,  New York, NY.

Burns, D.  A., J.  N.  Galloway, and G.  R. Hendrey.  1981.  Acidification
of surface waters in two areas of the eastern United States.  J. Water,
Air, and Soil Pollution 16:277-285.

Chen,  C.  W., J. D. Dean, S. A. Gherini, and R. A. Goldstein.  1982.
Acid rain model:   Hydrologic module.  J. Env. Eng. Div. ASCE
108:455-472.

Christophersen, N. and R. F. Wright.  1981. Sulfate budget and a model
for sulfate concentrations  in stream  water  at Birkenes, a small forested
catchment in southernmost Norway.  Wat. Resour. Res.  17:377-389.

Christophersen, N.,  H. M. Seip, and R. F. Wright.  1982.  A model for
streamwater chemistry at Birkenes, Norway.  Wat. Resour. Res.
18:977-996.

Church, M. R.  1980.  Relative rates  of uptake and release of oxygen and
inorganic carbon in laboratory phytopl ankton cultures and in a Piedmont
Virginia stream.   Ph.D. Thesis, University  of Virginia, Charlottesville.
286 pp.

Church, M. R. and J. N. Galloway.  1983.  Application of the
'acidification indicator' and  'predictor nomograph' of Henricksen to
data from two Adirondack lakes.   In press.

Clark, W. M.  1928.  The Determination of Hydrogen Ions.  The Williams
and Wilkins Co.,  Baltimore, MD.

Cleaves, E. T., A. E. Godfrey, and 0. P. Bricker.  1970.  Geochemical
balance of a small watershed and  its  geomorphic  implications.  Geo. Soc.
of Amer. Bull. 81:3015-3032.
                                  4-139

-------
Clymo, R. S.  1963.   Ion exchange  in Spaghnum and its relation to bog
ecology.  Ann.  Bot.  27(106):309-324.

Cogbill, C. V.   1976.   The  history and character of acid precipitation
in eastern North America,  pp.  363-370.  J_n Proc. of the First Internat.
Symp. on Acid Precipitation and  the Forest Ecosystem.  L. S. Dochinger
and T. A. Seliga, eds.   USDA For.  Ser. Gen.  Tech. Rep. NE-23.

Cogbill, C. V.  and G.  E. Likens.  1974.   Acid precipitation in the
northeastern United States.  Water Resources Res. 10:1133-1137.

Cole, D. W. and D. W.  Johnson.  1977.  Atmospheric sulfate additions and
cation leaching in a Douglas Fir ecosystem.  Wat. Res. Res. 13:313-317.

Cook, R. B.  1981.  The Biogeochemistry  of Sulphur in Two Small Lakes.
Ph.D. Thesis, Columbia University.  234  pp.

Coulson, C. B., R. I.  Davies and D. A. Lewis.   1960a.  Polyphenols in
leaves, litter and superficial humus  from mull  sites.  J. Soil Sci.
11:20-29.

Coulson, C. B., R. I.  Davies and D. A. Lewis.   1960b.  Polyphenols in
plant, humus and soil  II.  Reduction  and transport by polyphenols of
iron in model soil.  J. Soil Sci.  11:30-44.

Crisman, T. L., R. L.  Schulze, P.  L.  Brezonik,  and S. A. Bloom.  1980.
Acid precipitation:  The biotic response in  Florida lakes, pp. 296-297.
In Ecological Impact of Acid Precipitation.  Proceedings of an
International Conference, Sandefjord,  Norway, March 11-14, 1980.
Drabltfs, D. and A. Tollan,  eds.   SNSF-project,  Oslo-JU,
Norway.

Cronan, C.  S.  1980.  Consequences of sulfuric  acid inputs to a forest
soil, pp. 335-344.  In Atmospheric Sulfur Deposition.  D. S. Shriner, C.
R. Richmond, and S. TT7 Lindberg, eds.   Ann Arbor Science, Ann Arbor,
MI.

Cronan, C.  S. and C. L. Schofield.  1979. Aluminum leaching response to
acid  precipitation: Effects on high-elevation  watershed  in the
northeast.  Science 204:304-306.

Damman, A.  W. H.  1978.  Distribution and movement of elements in
ombrotrophic peat bogs.  Oikos 30:480-495.

Davis,  J.  A.  1982.  Adsorption of natural  dissolved  organic matter  at
the oxide/water interface.  Geochim.  et Cosmochim.  Acta  46:681-692.

Davis,  J.  A. and R. Gloor.  1981.  Adsorption  of dissolved organics  in
lake  water by aluminum  oxide.  Effect of molecular  weight.   Environ.
Sci.  Tech.  15:1223-1229.
                                  4-140

-------
Davis, A. 0., J.  N.  Galloway,  and  D.  K.  Nordstrom.  1982.  Lake
acidification:  Its  effect on  lead in the  sediment of two Adirondack
lakes.  Limnol. and  Oceanog. 27:163-167.

Davis, R. B. and F.  Berge.  1980.   Atmospheric  deposition in Norway
during the last 300  years  as recorded in SNSF lake sediments.  II.
Diatom stratigraphy  and inferred pH,  pp. 270-271.  J^n Ecological Impact
of Acid Precipitation.   Proceedings of an  International Conference,
Sandefjord, Norway,  Marcholl-14, 1980.  Drabljbs, D. and A. Tollan,
eds.  SNSF-project,  Oslo-As, Norway.

Davis, R. B., M.  0.  Smith, J.  H. Bailey, and S. A. Norton. 1978.
Acidification of Maine  (U.S.A.)  lakes by acidic precipitation.  Verh.
Internat. Verein. Limnol.  10:532-537.

Davis, R. B., S.  A.  Norton, C. T.  Hess,  and D. F. Brakke.  1983.
Paleolimnological reconstruction of the  effects of atmospheric
deposition of acids  and heavy  metals  on  the chemistry and biology of
lakes in New England and Norway.  lr^  Proc.  III. Intern. Symp.
Paleolimnol., Devel. Hydrobiol., J. Merilainen, R. W. Battarbee, and P.
Huttunen, eds.  W. Junk, The Hague (in press).

Davis, R. B., S.  A.  Norton, D. F.  Brakke,  F. Berge, and C. T. Hess.
1980.  Atmospheric decomposition in Norway during the last 30 years as
recorded in SNSF lake sediments, pp.  274-275.  J^n IV. Synthesis, and
comparison with New  England.   In D. Drabltfs and A. Tollan, eds.
Ecological Impact of Acid  Precipitation, Proc.  Int. Conf. Ecol. Impact
Acid Precip., SNSF Project, Oslo,  Norway.

Daye, P. G.  1980.  Attempts to  acclimate  embroys and alevins of
Atlantic salmon (Salmo  salar)  and  rainbow  trout (Salmo gairdneri) to low
pH.  Can. J. Fish Aquat. Sci.  37:1035-1038.

Del Prete, A. and C. Schofield.   1981.  The utility and diatom analyses
of lake sediments for evaluating acid precipitation effects on dilute
lakes.  Arch. Hydrobiol. 91(3):332-340.

Dickson, W.  1978a.   Some  effects  of  the acidification of Swedish lakes.
Verh. internat. Verein. Limnol.  20:851-856.

Dickson, W.  1978b.   Liming, pp. 37-41.  In Limnological Aspects of Acid
Precipitation. G. R. Hendrey,  ed.   BrookKaven  National Laboratory.
Sagamore Lake Conference Center, Sept. 25-28.

Dickson, W.  1980.  Properties of  acidified water, pp. 75-83.  J^n
Ecological Impact of Acid  Precipitation.   Proceedings of an
International Conference,  Sandefjord, Norway, March 11-14, 1980.
Drabljls, D. and A. Tollan, eds.  SNSF-project,  Oslo-As, Norway.
                                  4-141

-------
Dillon, P. J.  1981.  Mass balance models:   An  explanation of  the
observed chemistry of lakes near Sudbury,  Ontario.  Jtj^ Studies of lakes
and Watersheds Near Sudbury, Ontario:   Final  Limnology  Report  of The
Sudbury Environmental Study.  Ont. Min.  of Environ.

Dillon, P. J. and F. H. Rigler.   1974.   A  test  of a simple nutrient
budget model predicting the phosphorus  concentration  in lake water. J.
Fish Res. Bd. Canada 31:1771-1778.

Dillon, P. J. and P. J. Smith.   1981.   Sediment chemistry and  rates of
sedimentation in lakes near Sudbury, Ontario.   Ir± Studies of Lakes and
Watersheds Near Sudbury, Ontario:   Final Limnology Report of the Sudbury
Environmental Study.  Ont.  Min.  of Environ.

Dillon, P. J. and W. A. Scheider.   1983.   Modeling the  reacidification
rates of neutralized lakes  near  Sudbury  Ontario.  J£  Modeling  of Total
Acid Precipitation Impacts, J. L.  Schnoor,  Ed.  Ann Arbor Science, Ann
Arbor, MI.  (in press).

Dillon, P. J., D. S. Jeffries, and W. A. Scheider.  1982.  The use of
calibrated lakes and watersheds  for estimated anthropogenic deposition
near a large point source.   Water, Air,  and Soil Pollut. 18:241-258.

Dillon, P. J., D. S. Jeffries, W.  Snyder,  R.  Reid, N. D. Van,  D. Evans,
J. Moss, and W. A. Scheider.  1978. Acid  precipitation  in south-central
Ontario: Recent observations.  J.  Fish.  Res.  Board Can.  35:809-815.

Dillon, P. J., N. D. Van, W. A.  Schieder,  and N. Conroy.  1977.  Acidic
lakes in Ontario:  Characterization, Extent and Responses to Base and
Nutrient Additions.  Ontario Ministry on the  Environment, Rexdale,
Ontario.

Dillon, P. J., N. D. Van, W. A.  Schieder,  and N. Conroy.  1979.  Acidic
lakes in Ontario, Canada; characterization  extent and responses to base
and nutrient additions.  Arch. Hydrobiol.  Beih. 13:317-336.

Drabljfc, D., I. Sevaldrud,  and J.  A. Timberlid.  1980.   Historical
land-use changes related to fish status  development in  different areas
in southern Norway, pp. 367-369.  Ir± Ecological Impact  of Acid
Precipitation.  Proceedings of an  International Conference, Sandefjord,
Norway, March 11-14. 1980.   Drablj6s, D.  and A.  Tollan,  eds.
SNSF-project, Oslo-As, Norway.

Driscoll, Jr., C. T.  1980.  Chemical Characterization  of Some Dilute
Acidified Lakes and Streams in the Adirondack Region  of New York State.
Ph.D. Thesis, Cornell University,  Ithaca,  NY.

Driscoll, C. T. and J. J.  Bisogni. 1983.   Weak acid/base systems in
dilute acidified lakes and  streams in the  Adirondack  region of New York
State.  In Modeling of Total Acid  Precipitation Impacts.  J. L. Schnoor,
ed.  Ann~~A>bor Science, Ann Arbor, MI (in  press).
                                  4-142

-------
Driscoll, Jr.,  C. T.,  J.  P.  Baker, J. J. Bisogni, Jr., and C. L.
Schofield.  1980.  Effect of aluminum speciation on fish in dilute
acidified waters.  Nature 284:161-164.

Driscoll, C. T., J. P. Baker,  J. J.  Bisogni  and C. L. Schofield.  1983.
Aluminum speciation and equilibria in dilute acidic surface waters of
the Adirondack  region  of  New York State.   In Geologic Aspects of Acid
Rain.  0. P. Bricher,  ed. Ann Arbor Science, Ann Arbor, MI (in press).

Driscoll, C. T., J. R. White,  G. C.  Schafran, and J. D. Rendall.  1982.
Calcium carbonate neutralization of  acidified surface waters.  J. Lnvir.
Eng. Div. ASCE  108:1128-1145.

Durst, N. R.  1975. Standarization  of  pH  measurements.  National Bureau
of Standards Special Publication 260-53.   U.S. Government Printing
Office, Washington, D.C.

Edzwald, J. K.  and J.  V.  DePinto.  1978.   Recovery of Adirondack Acid
Lakes with fly  ash treatment.   Final Report, Engineering Foundation
Grant No. RC-A-76-4.'

Eriksson, E.  1981. Aluminum  in groundwater possible solution
equilibria.  Nordic Hydrology  12:43-50.

Fahey, T. J.  1979. Changes in nutrient content of snow water during
outflow from Rocky Mountain  coniferous  forest.  Oikos 32:422-428.

Feldman, I.  1956.  Use and  abuse of pH measurements.  Anal. Chen.
28:1859-1866.

Feller, M. C. and J. P. Kimmins.  1979.  Chemical characteristics of
small streams near Honey  in  southwestern British Columbia Water.  Res.
15:2:247-258.

Ford, E. D. and J. D.  Deans.  1978.  The effects of canopy structure on
stemflow, throughfall  and interception  loss  in a young Sitka Spruce
plantation.  J. App. Ecology 15:905-917.

Fowler, D.  1980.  Removal of  sulphur and  nitrogen compounds from the
atmosphere in rain and by dry  deposition,  pp. 22-32.  In Ecological
Impact of Acid Precipitation.   Proceedings of an International
Conference, Sandefjord, Norway, March 11-14, 1980.  Drabltfs, D. and A.
To!Ian, eds. SNSF-project, Oslo-As,  Norway.

Francis, A. J.  and G.  R.  Hendrey.  1980.   Integrated Lake-Watershed
Acidification Study (ILWAS).  Quarterly Progress Report No. 5, Electric
Power Research Institute, Palo Alto, CA.

Fraser, J. E. and D. L. Britt.  1982.   Liming of acidified waters:  a
review of methods and  effects  on aquatic ecosystems.  Fish and Wildlife
Service.  FWS/OBS-80/40.13,  p. 189.
                                  4-143

-------
Fraser, J. E., D.  Hinckley,  R.  Burt,  R. Rodensky Severn and J.
Wisniewski.  1982.  Feasibility  study to utilize liming as a technique
to mitigate surface water acidification.  Palo Alto, CA.  Electric Power
Research Institute; EPRI  EA-2362; p.  101.

Friberg, F.,  C. Otto,  and B.  S.  Svensson.   1980.  Effects of
acidification on the dynamics of allochthonous leaf material and benthic
invertebrate  communities  in  running waters, pp. 304-305.  _l£ Ecological
Impact of Acid Precipitation.  Proceedings  of an International
Conference, Sandefjord, Norway,  March 11-14, 1980.  Drabltfs, D. and A.
Toll an, eds.  SNSF-project, Oslo-As, Norway.

Galloway, J.  N. and E. B. Cowling.  1978.   The effects of precipitaion
on aquatic and terrestrial ecosystems:  A proposed precipitation
chemistr network.   J.  of  the Air Pollution  Control Assoc. 28:229-235.

Galloway, J.  N. and P. J. Dillon.   1982.  Effects of acidic deposition:
the importance of  nitrogen.   Proceedings 1982 Stockholm Conference on
the Acidification  of the  Environment.

Galloway, J.  N. and G. E. Likens.   1979.  Atmospheric enhancement of
metal  deposition in Adirondack  Lake sediment.  Limnol. Oceanog.
24:427-433.

Galloway, J.  N. and G. E. Likens.   1981.  Acid precipitation:  The
importance of nitric acid.  Atmos.  Environ. 15:1081-1085.

Galloway, J.  N. and D. M. Whelpdale.  1980.  An atmospheric sulfur
budget for eastern N.  America.   Atmos.  Environ. 14:409-417.

Galloway, J.  N., B. J. Cosby, Jr.,  and  G. E. Likens.  1979.  Acid
precipitation:  Measurement  of  pH and acidity.  Limnol. Oceanogr.
24:1161-1165.

Galloway, J.  N., G. E. Likens,  W. C.  Keene, and J. M. Miller.  1982a.
The composition of precipitation in remote  areas of the world.  Journal
Geophysical Research 87:11:8771-8786.

Galloway, J.  N., S. A. Norton and M.  R. Church.  1983a.  Freshwater
acidification from atmospheric  deposition of  (^04:  A conceptual
meodel.  Environmental Sciences and Technology  (in  press).

Galloway, J.  N., C. L. Schofield, G.  R. Hendrey, E. R. Altwicker and D.
E. Troutman.   1980a.  An  analysis of  lake acidification using  annual
budgets.  J[n Proceedings  of  International Conference on the Ecological
Impact of Acid Precipitation.  Sandefjord,  Norway,  March 11-14.  pp.
254-55.

Galloway, J.  N., C. L. Schofield, G.  R. Hendrey, N. E. Peters  and A. H.
Johannes.  1983b.  Lake acidification during  spring snowmelt:  Processes
and Causes,   (in press).
                                  4-144

-------
Galloway, J. N., C. L.  Schofield,  G.  R.  Hendrey, N. E. Peters, and A. H.
Johannes. 1980b.  Sources  of acidity  in  three lakes acidified during
snowmelt, pp. 264-265.   Ir± Ecological  Impact of Acid Precipitation.
Proceedings of an International  Conference, Sandefjord, Norway, March
11-14. 1980.  Drablj6s,  D.  and A. Tollan,  eds.  SNSF-project,
Oslo-As, Norway.

Galloway, J. N., C. L.  Schofield,  N.  E.  Peters, G. R. Hendrey and E. R.
Altwicker.  1983c.  The effect of atmospheric  sulfur on the composition
of three Adirondack lakes.  Canadian  Journal of Fisheries and Aquatic
Sciences 40 (in press).

Galloway, J. N., J. D.  Thornton, S. A. Norton, H.  L. Volchok and R. A.
N. McLean.  1982b.  Trace  metals in atmospheric deposition:  A review
and assessment.  Atm. Env:  16:7:1677-1700

Garrells, R. M., F. T.  Mackenzie and  C.  Junt.  1975.  Chemical Cycles
and the Global Environment.  Wm. Kaufmann Inc., Los Altos, CA.

Gether, J., N. Gjtfs, and G. Lunde. 1976.  Isolation and
characterization in precipitation of  organic micropollutants associated
with long-range transport in air.  FR 8/76, SNSF,  Oslo-As, Norway.

Gjessing, E. T., A. Henriksen, M.  Johannessen, and R. F. Wright.  1976.
Effects of acid precipitation on freshwater chemistry, pp. 64-85.  In
Impact of Acid Precipitation on Forest and Freshwater Ecosystems in
Norway.  F. Braekke, ed.  Sur Nedbors Virkning Pa  Skog og Fisk Res. Rep.
6/76.

Glass, G. E.  1980.  Susceptibility of aquatic and terrestrial resources
of Minnesota, Wisconsin and Michigan  to impacts  from acid precipitation:
Information requirements,  pp. 112-113.  J_n Ecological  Impact of Acid
Precipitation.  Proceedings of an International  Conference, Sandefjord,
Norway, March 11-14. 1980.  Drabltfs,  D.  and A. Tollan, eds.
SNSF-project, Oslo-As,  Norway.

Glass, G. E. and T. G.  Brydges.  1981.  Problem  complexity  in  predicting
impacts from altered precipitation chemistry.  Presented at the
Symposium on Acid Precipitation and Fishery Impacts  in North America.
Aug. 1-5, 1981.  Cornell University,  Ithaca,  NY.

Glover, G. M., and A. H. Webb.  1979.  Weak and  strong acids in the
surface waters of the Tovdal region in southern  Norway.  Water Res.
13:781-783.

Goldman, J. C. and P. G. Brewer.  1980.   Effect  of nitrogen source and
growth rate on phytoplankton-mediated changes in alkalinity.   Limnol.
Oceanogr. 25:352-357.

Golterman,  H. L.  1969.  Methods for  Chemical  Analysis of Fresh Waters.
IBP Handbook No. 8. Blackwell Scientific Publications, Oxford.
                                  4-145

-------
Gordon, A. J. and E. Gorham.   1960.   The  influence of  smelten fumes on
the chemical  composition  of lake  waters near  Sudbury,  Ont. and upon the
surrounding vegetation.   Com.  J.  of  Botany  38:477-487.

Gordon, A. J. and F. Gorham.   1963.   Ecological  aspects of air
pollution, from an iron-scintering plant  at Wawa, Ontario.  Canadian
Journal of Botany 41:1063-1078.

Gorham, E.  1957.  The chemical  composition of lake waters in Halifax
County, Nova Scotia.  Limnol.  Oceanogr. 2:12-21.

Gorham, E.  1958.  The influence  and importance  of daily weather
conditions in the supply  of  chloride, sulphate and other ions to fresh
waters from atmospheric precipitation.  Philo. Trans.  Roy. Soc. London
2418:147-178.

Gorham, E.  1961.  Factors influencing  supply of major ions to inland
waters, with special reference to the atmosphere.  Geol. Soc. Arner.
Bull. 72:795-840.

Gough, L. P., H. T. Shacklette,  and  A.  A. Case.  1979. Element
concentrations toxic to  plants,  animals,  and  man.  U.S. Geol. Surv.
Bull. 1466.  80 p.

Grahn, 0.  1980.  Fishkills  in two moderately acid lakes due to high
aluminum concentrations,  pp.  310-311.  Jji Proceedings  of the
International Conference  on  Ecological  Impact of Acid  Precipitation.  D.
Drabltfs and A. Tollan, eds.   SNSF Project,  Sandefjord, Norway, March
11-14.

Grahn, 0. and H. Hultberg.  1974. 01 igotrophication,  a
self-accelerating process in  lakes subjected  to excessive  supply of acid
substances.  Ambio 3:93-94.

Grahn, 0. and H. Hultberg.  1975. The neutralizing  capacity of 12
different lime products used for pH  adjustment of  acid water.  Vatten
2:120-132.

Gran, G.  1952.  Determination of the equivalence  point  in
potentiometric titrations.  Part II.  Analyst 77:661-671.

Groet, S. S.  1976.  Regional  and local variations in  heavy metal
concentrations of bryophytes in the  northeastern United  States.  Oikos
27:445-456.

Haapala,  H., P.  Sepponen, and E. Meskus.   1975.  Effect  of spring  floods
on  water  acidity in the Kiiminkijoki area, Finland.   Oikos 26:26-31.

Haines,, T. A.  1982.  Vulnerability  of lakes and streams  in  the
northeastern United States to acidification from long  range  transport of
air pollution.   Unpub. Proc. of U.S. EPA Acidic Deposition Effects
Program Review,  Raleigh, NC.  7 p.


                                  4-146

-------
Haines, T. A., J. J. Akielagzek, S. A. Norton,  and R. B.  Davis.   1983a.
Errors in pH measurement with colorimetric indicators in  low alkalinity
waters.  Hydrobiologia (in press).

Haines, T. A., J. J. Akielaszek and P. J. Rago.  1983b.   A regional
survey of chemistry of headwater lakes and streams in New England:
Vunerability to Acidification Air Pollution and Acid Rain Series,  Fish
and Wildlife Service.

Hall, R. J. and G. E. Likens.  1980.  Ecological  effects  of experimental
acidification on a stream ecosystem, pp. 375-376.  lr± Ecological  Impact
of Acid Precipitation.  Proceedings of an International Conference,
Sandefjord, Norway, March 11-14, 1980.  Drablj6s,  D. and A. Tollan,
eds. SNSF-project, Oslo-As, Norway.

Hall, R. J., C. T. Driscoll and G. E. Likens.   1982.   Physical, chemical
and biological consequences of episodic aluminum  additions to a stream
ecosystem.  Limno. Oceanogr. (submitted for publication).

Hanson, D. W., S. A. Norton, and J. S. Williams.   1982.   Modern and
paleolimnological evidence for accelerated leaching and metal
accumulation in soils in New England, caused by atmospheric deposition.
Water, Air, and Soil Poll. 18:227-239.

Harriman, R. and B. Morrison.  1980.  Ecology  of  acid streams draining
forested and non-forested catchments in Scotland, pp. 312-313.  Ir±
Ecological Impact of Acid Precipitation.  Proceedings of  an
International  Conference, Sandefjord, Norway,  March 11-14, 1980.
Drablj6s, D. and A. Tollan, eds.  SNSF-project,  Oslo-As, Norway.

Harvey, H. H., R. C. Pierce, P. J. Dillon, J.  P.  Kramer,  and D. M.
Whelpdale.  1981.  Acidification in the Canadian  Aquatic  Environment:
Scientific Criterion for an Assessment of the  Effects of  Acidic
Deposition on  Aquatic Ecosystems.  Nat. Res. Coun.  Canada Report No.
18475.  369 pp.

Hayes, F. R. and E. H. Anthony.  1958.  Lake water and sediment.   I.
Characteristics and water chemistry of some Canadian  east coast lakes.
Limnol. Oceanogr. 3:299-307.

Heit, M., Y. L. Tan, C. K. Klusek, H. L. Volchok, and J.  C. Burke.
1980.  The Origin and Depositional History of Trace Elements and
Polynuclear Aromatic Hydrocarbons in Two Remote Lakes in  the Adirondack
Acid Rain Region.  U.S. Dept. of Energy Report  EML-381, New York.

Heit, M., Y. Tan, C. Klusek, and J. C. Burke.   1981.   Anthropogenic
trace element and polycyclic aromatic hydrocarbon levels  in sediment
cores from two lakes in the Adirondack acid lake  region.   Water, Air
Soil Pollut. 15:441-464.
                                  4-147

-------
Hem, J. D. and C. E. Roberson.  1967.  Form and stability of aluminum
hydroxide complexes in dilute solution.  I). S.  Geol.  Survey  Water  Supply
Pap. 1827-A, 55 p.

Hem, J. D., C. E. Roberson, C. J. Lind and W. L. Polzer.   1973.
Chemical interactions of aluminum with aqueous  silica at  25°C.   U.  S.
Geol. Survey Water Supply Pap. 1827-E, 57 p.

Hemond, H. F.  1980.  Biogeochemistry of Thoreau's Bog, Concord,
Massachusetts.  Ecol. Mon. 50:507-526.

Hendrey, G.  1982.  Based on data from the New  York  State Department of
Conservation and various other sources contained in  the Acidification
Chemistry Information Data Base (ACID).  Available from Brookhaven
National Laboratory.

Hendrey, G. R. and F. A. Ventucci.  1980.  Benthic plant  communities in
acidic Lake Golden, New York:  Sphagnum and the algal  mat, pp. 314-315.
jji Proceeding of the International Conference on Ecological  Impact  of
Acid Precipitation.  D. Drabltfs and A. Toll an,  ed.  SNSF  Project.

Hendrey, G. R., J. N. Galloway, and C. L. Schofield.   1980a. Temporal
and spatial trends in the chemistry of acidified lakes under ice cover,
pp. 266-267.  In Ecological Impact of Acid Precipitation. Proceedings
of an International Conference, Sandefjord, Norway,  March 11-14, 1980.
Drabltfs, D. and A. Tollan, eds.  SNSF-project,  Oslo-JU, Norway.

Hendrey, G. R., J. N. Galloway, S. A. Norton, C. L.  Schofield, P. W.
Schaffer, and D. A. Burns.  1980b.  Geological  and hydrochemical
sensitivity of the eastern United States to acid precipitation.
EPA-600/3-80-024.  100 pp.

Hendrey, G. R., K. Baalsrud, T. S. Traaen, M. Laake,  and  G.  Raddum.
1976.  Acid precipitation.  Some hydrobiological changes. Ambio
5:224-227.

Henriksen, A.  1979.  A simple approach for identifying and  measuring
acidification in freshwater.  Nature 278:542-544.

Henriksen, A.  1980.  Acidification of freshwaters--a large  scale
titration, pp. 68-74.  lr± Ecological Impact of Acid Precipitation.
Proceedings of an International Conference, Sandefjord, Norway,  March
11-14. 1980.  Drablj6s, D. and A. Tollan, eds.  SNSF-project,
Oslo-As, Norway.

Henriksen, A.  1982.  Changes on base cation concentrations  due  to
freshwater acidification.  NIVA Report OF-81623, Norwegian Institute  for
water research.

Henriksen, A., and H. M. Seip.  1980.  Strong and weak acids in  surface
waters of Southern Scotland.  Water Res. 14:809-813.
                                  4-148

-------
Henriksen, A. and R. F. Wright.  1977.  Effect of acid precipitation  on
a  small acid lake in southern Norway.  Nord. Hydro 1 .  8:1-10.

Herron, M. M., C. C. Langway, H. V. Weiss,  0. P.  Hurley,  R. 0. Kerr,  and
J. H. Cragin.  1976.  Vanadium and other elements in  Greenland ice
cores.  CERL Rep. 76-24.

Herron, M. M., C. C. Langway, Jr., H. V. Weiss, and J. Cragin.   1977.
Atmospheric trace metals and sulfate in the Greenland ice sheet.
Geochim. Cosmochim. Acta 41:915-920.

Hinckley, D. and J. Wisniewski.  1981.  Scandinavian  Trip Report:
Liming Activities in Norway and Sweden.  International Research  and
Technology Corporation, McLean, VA.

Hodge, W. W.  1953.  Waste disposal problems in the coal  mining
industry.  Ir^ Industrial Wastes:  Their Disposal  and  Treatment.
Rudolfs, W., ed.  Reinhold New York, NY.

Hohl, H. and W. Stumm.  1976.  Interactions of Pb2+ with  hydrous
T-A1203.  J. Coll. Int. Sci. 55:281-288.

Hongve, D.  1978.  Buffering of acidified lakes by sediments.  Verh.
Internal. Verein. Limnol .. 20:743-748.

Horntvedt, R. and E. Joranger.  1976.  Kjemisk innhold i  nedbo/r under
traer og annen vegetasjon: juni-oktober 1974.  SNSF-project TN 20/76.
21 pp.

Howarth, R. S. and J. B. Sprague.  1978.  Copper  lethality to rainbow
trout in waters of various hardness and pH.  Water Res. 12:455-462.

Huang, C. P.  1975.  Adsorption of phosphate at the hydrous
        electrolyte interface.  J. Coll. Int. Sci. 53:178-186.
Hultberg, H.  1977.  Thermally stratified  acid  water  in late winter— a
key factor inducing self-accelerating processes which increase
acidification.  Water Air Soil Pollut. 7:279-294.

Hultberg, H. and I. B. Andersson.   1981.   Liming of acidified lakes and
streams.  Aspects on induced physical-chemical  and biological changes.
Swedish Water and Air Pollution Research  Institute, Gothenberg, Sweden.

Hultberg, H. and S. Johansson.  1981.  Acid  groundwater.  Nordic
Hydrology, 12:51-64.

Hultberg, H. and A. Wenblad.  1980.   Acid  groundwater in  southwestern
Sweden, pp. 220-221.  Jji Ecological  Impact of Acid Precipitation.
Proceedings of an International  Conference,  Sandefjord, Norway, March
11-14. 1980.  Drabltfs, D. and A. Tollan, eds.   SNSF-project,
Oslo-As, Norway.
                                  4-149

-------
Hustedt, F.  1937.  Systematise he imd okologische untersuchungen  uber
die diatomeenflora von Java,  Bali und Sumatra.   Arch.  Hydrobiol.  Suppl.
15 and 16.

Hutchinson, G. E.  1975.  A Treatise on Limnology.  Vol.  1,  Part  2,
Chemistry of Lakes.  Wiley-Interscience, New York.

Hutchinson, T., W. Gizyn, M.  Havas,  and V.  Zobens.  1978.   Effect of
long-term lignite burns on arctic ecosystems of the Smoking  Hills,
N.W.T. lr± Trace Substances in Environmental  Health XII.   D.  D.  Hemphill,
ed., University of Missouri,  Columbia,  MO,  pp.  317-332.

Jeffries, D. S. and W. R. Snyder.  1981.  Atmospheric  Deposition  of
Heavy Metals in Central Ontario.   Wat.  Air  Soil  Pollut. 15:127.

Jeffries, D. S. and A. P. Zimmerman.  1980.   Comments  on  the analysis
and sampling of low conductivity  natural waters for alkalinity.   Can.  J.
Fish. Aquat. Sci. 37:901-902.

Jeffries, D. S., C. M. Cox, and P. J. Dillon.   1979.   Depression  of pH
in lakes and streams in central Ontario during  snowmelt.  J. Fish. Res.
Board Can. 36:640-646.

Jeffries, S., W. A. Scheider, and W. R. Snyder.   1981. Stream
chemistry: the interaction of precipitation and geochemical  processes  in
watersheds near Sudbury, Ontario.  Ir± Studies of Lakes and  Watersheds
near Sudbury, Ontario: Final  Limnology  Report of the Sudbury
Environmental Study.  Ont. Min. of Environ.

Johannessen, M.  1980.  Aluminum, a  buffer  in acidic waters. _^n_
Proceedings of the International  Conference on  Ecological Impact  of Acid
Precipitation, D. Drablos and A.  Tollan, eds.,  SNSF Project, Oslo,
Norway pp. 222-223.

Johannessen, M., A. Skartveit, and R. F. Wright.  1980.   Streamwater
chemistry before, during, and after  snowmelt, pp. 224-225.   Jhi
Ecological Impact of Acid Precipitation.  Proceedings  of  an
International Conference, Sandefjord, Norway, March 11-14,  1980.
Drablj6s, D. and A. Tollan, eds.  SNSF-project,  Oslo-Xs, Norway.

Johnson, A. H.  1979.  Evidence of acidification of headwater streams  in
the New Jersey pinelands.  Science 206:834-836.

Johnson, A. H. and T. G. Siccama.  1979.  Effect of vegetation  on the
morphology of Windson soils,  Litchfield, Connecticut.  Soil  Sci.  Am. J.
43:1199-1200.

Johnson, D. W. and D. W. Cole. 1980.  Anion mobility  in  soil.
Relevance to nutrient transport from .forest ecosystems.   Environ.
Internat. 3:79-90.
                                  4-150

-------
Johnson, N. M.  1979.  Acid rain:   Neutralization  within  the Hubbard
Brook Ecosystem and regional  implications.   Science 204:497-499.

Johnson, N. M., C. T. Driscoll,  J.  S.  Eaton,  G.  E. Likens and W. H.
McDowell.  1981.  Acid rain,  dissolved aluminum  and chemical weathering
at the Hubbard Brook Experimental  Forest,  New Hampshire.  Geochim.
Cosmochim. Acta 45:1421-1437.

Johnston, S. E., S. A. Norton,  C.  T.  Hess,  R. B. Davis, and R. S.
Anderson.  1981.  Chronology  of  atmospheric deposition of acids and
metals in New England, based  on  the record in lake sediments.  Ann Arbor
Science, Ann Arbor, MI.  14 pp.

Jones, B. F., V. C. Kennedy and  G.  W.  Zellweger.   1974.   Comparison of
observed and calculated concentrations of  dissolved Al and Fe in stream
water.  Water Resour. Res.  10:791-793.

Jordan, M. and G. E. Likens.  1974.  An organic  carbon budget for an
oligotrophic lake in New Hampshire, U.S.A.   Verh.  Internet. Verein.
Limnol. 19: In press.

Kahl, J. S., S. A. Norton,  and J.  S.  Williams.   1982.  Chronology,
magnitude, and paleolimnological record of changing metal fluxes related
to atmospheric deposition of  acids  and metals in New England {ext.
abs.).  Am. Chem. Soc., Ann.  Mtg.,  Env. Chem. Div.  pp. 486-489.

Kelly, C. A., J. W. M. Rudd,  R.  B.  Cook, and  D.  W. Schindler.  1981.
The potential  importance of bacterial  processes  in regulating rate of
lake acidification, (submitted  to  Limnol.  Oceanogr).

Kennedy, V. C., D. F. Jones,  and G. W. Zellweger.  1974.  Filter
pore-size effects on the analysis  of Al, Fe,  Mn  and Ti in water.  Water
Resour. Res. 10:785-790.

Kolthoff, I. M.  1937.  Acid-Base  Indicators. The Macmi11 an Co., New
York, NY.

Kramer, J. R.  1981.  Aluminum:  chemistry, analysis and biology.
Environmental  Geochemistry  Report,  1981/2,  Department of Geology,
McMaster University.

Langford, R. R.  1948.  Fertilization  of lakes in Algonquin Park,
Ontario.  Trans. Amer. Fish.  Soc.  78:133-144.

Lantzy, R. F. and F. T. MacKenzie.   1979.   Atmospheric trace metals:
global cycles and assessment  of  man's  impact. Geochim. Cosmochim. Acta
43:511-525.

Lazrus, A. L., E. Lorange,  and J.  P. Lodge, Jr.  1970.  Lead and other
metal  ions in United States precipitation.  Envir. Sci. Tech.  4:55-58.
                                  4-151

-------
Leivestad, H. and I.  Muniz.   1976.   Fish  kills at low pH in a Norwegian
river.  Nature 259:391-392.

Leivestad, H., 6. Hendrey,  I.  P. Muniz, and E. Snekvik.  1976.  Effects
of acid precipitation on freshwater  organisms, pp. 86-111.  jji Impact of
Acid Precipitation on Forest and Freshwater Ecosystems in Norway.  F. H.
Braekke, ed.  SNSF-Project  1432 Aas,  Norway.

Likens, G. E.  1974.   Water and nutrient  budgets for Cayuga Lake, New
York.  Cornell Univ.  Water  Resources  and  Marine Sciences Center.
Ithaca, NY.  Tech. Report 82,  91 pp.

Likens, G. E.  1976.   Acid  precipitation.  Chem. Eng. News 54:29-44.

Likens, G. E. and F.  H.  Bormann.   1974.   Linkages between terrestrial
and aquatic ecosystems.   Bioscience  24(8):447.

Likens, G. E., F. H.  Bormann,  and  N.  M. Johnson.  1966.  Nitrification:
Importance to nutrient losses from a cutover  forested ecosystem:
Science 163:2105-2106.

Likens, G. E., F. H.  Bormann,  R. S.  Pierce, J. S. Eaton, and N. M.
Johnson.  1977.  Biogeochemistry of  a Forested Ecosystem.  Springer-
Verlag, New York.  164 pp.

Likens, G. E., J. S.  Eaton,  and J. N. Galloway.  1974.  Precipitation as
a source of nutrients for terrestrial and aquatic ecosystems.  Proc.,
Atomic Energy Comm. Symposium on Precipitation Scavenging.  Champaign,
Illinois (ERDA Symposium Series 41).

Lind, C. J. and J. D. Hem.   1975.  Effects of organic solutes on
chemical reactions of aluminum.  U.S. Geol. Surv. Wat. Sup. Pap. 1827-G.

Lindberg, S. F., R. C. Harriss, and  R. R.  Turner.  1982.  Atmospheric
deposition of metals for forest vegetation.   Science 215:1609-1611.

Lindmark, G.  1981.  Restoration of  acidified lakes according to the
Contracid--A presentation.   Scandinavian  Fertilizers AB.  Institute of
Limnology, University of Lund, Sweden.

Loewenthal, R. E. and G. V.  R. Marais. 1978. Carbonate Chemistry of
Aquatic Systems:  Theory and Application.  Vol.  1.  Ann Arbor Science,
Ann Arbor, MI.

Logan, R. M., J. C. Derby,  L. C. Duncan.   1982.  Acid precipitation and
lake susceptability in the  Central Washington Cascades.  Env. Sci. Tech.
16 (11) 771-775.

Logan, J. A. 1982.  Sources and  sinks for nitrogen oxides in the
troposphere.  Submitted to  Advances  in Chemistry (Amer. Chem. Soc.).
                                  4-152

-------
Loucks, 0. L., D. Rapport,  W. Swanson,  and R.  W. Miller.   1981.
Assessment of Sensitivity Measures  for  Evaluating  Resources at Risk from
Atmospheric Pollutant Deposition, APP-0080-1980a,  ERL-Duluth, U.S. EPA.

Lui, N. S. T. and 0. A. Rolls.  1972.   Nitrogen metabolism of aquatic
organisms.  II. The assimilation of nitrate, nitrite and  ammonia by
Biddulphia aurita.  J. Phycol. 8:259-264.

Lunde, G., J. Gether, N. Gjj6s, and  M-B  Stj6bet  Lande.   1976.  Organic
micro-pollutants in precipitation in Norway.   FR 9/76, SNSF,
Oslo-JU, Norway.

Malmer, Nils.  1974.  On the effects of water,  soil and vegetation of an
increasing atmospheric supply of sulphur.   National Swedish Environ.
Prot. Bd., SNV PM 402E.  98 p.

May, H. M., P. A. Helmke and M. L.  Jackson.  1979.  Gibbsite solubility
and thermodynamic properties of hydroxyaluminum ions in aqueous
solutions at 25°C.  Geochim. et Cosmochim. Acta. 43:861-868.

McCarley, S.  1983.  Sensitivity of lakes  and  streams  in  National Park
to Acidification by Acidic  Deposition.   M.S. Thesis, Department
Environment Sciences, University of Virginia.

McCarthy, J. J., W. R. Taylor, and  J. L. Taft.  1977.  Nitrogenous
nutrition of the phytoplankton in the Chesapeake Bay.  1.  Nutrient
availability and phytoplankton preferences.  Limnol. Oceanogr.
22:996-1011.

McColl, J. G.  1980.  A survey of acid  precipitation in northern
California.  Final report to California Air Resources  Board, Univ. of
California, 93 pp.

McColl, J. G.  1981.  Increasing hydrogen  ion  activity of  water in two
reservoirs supplying the San Francisco  Bay Area, California.  Water
Resour. Res.  17:1510-1516.

McFee, W. W.  1980.  Sesitivity of  soil  regions to long-term acid
precipitation, pp. 495-506.  In Atmospheric Sulfur Deposition:
Environmental Impact and HealtFf Effects.  D. S. Shriner,  C. R. Richmond,
and S. E. Lindberg, eds.  Ann Arbor Science, Ann Arbor, MI.

McFee, W. W., J. M. Kelly,  and R. H. Beck.  1976.  Acid precipitation
effects on soil pH and base saturation  of  exchange sites.  Wat. Air Soil
Pollut. 7:401-408.

Melak, J. M., J. L. Stoddard, and D. L. Dawson.  1983.  Acid
precipitation and buffer capacity of lakes in  the  Sierra  Nevada
Mountains, California.  Proceeding  of Symposium on Acid Precipitation
from International Symposium on Hydrometeorology,  Denver,  CO 1982.
                                  4-153

-------
Menz, F. C. and C. T.  Driscoll.   1983.  An  economic evaluation of liming
to neutralize acidified Adirondack  surface  waters.  Water Resources
Research (in review).

Michalski,  M. F. P.  1971.   Status  of enrichment  of Walker's Lake,
Township of Lake of Bays, 1971.   Ont. Win.  Environ.  23 pp.

Miller, H.  G. and J. D. Miller.   1980.  Collection and retention of
atmospheric pollutants by vegetation, pp. 33-40.  In Ecological Impact
of Acid Precipitation.  Proceedings of  an International Conference,
Sandefjord, Norway, March 11-14,  1980.  Drablj6s,  D. and A. Tollan,
eds. SNSF-project, Oslo-JU,  Norway.

Minns, C. K.  1981.  Acid rain:   A  preliminary  estimate of the risk to
Ontario's inland fisheries.   Can. M. S. Rep.  Fish. Aquat. Sci.
1622:VI+36p.

Mulder, J.   1980.  Neutralization of Acid Rain  in the Hubbard Brook
Experimental Forest.  Prakijkverslag, Soil  Chemistry, unpublished
report, Wageningen, The Netherlands.

National Academy of Science.  1981.  Atmospheric-biosphere interaction:
Toward a better understanding of the ecological consequences of fossil
fuel combustion.  A report  prepared by  the  National Research Council,
National Academy of Sciences National Academy Press.  263 p.

National Atmospheric Deposition Program.  1982.  Distribution of  surface
waters sensitive to acidic  precipitation:   a  state level atlas.   NADP
Tech. Rept. No. IV, S.A. Norton editor, Department of Geological
Sciences, University of Maine, Orono, ME, 72  unnumbered pages.

National Research Council  Canada.  1981.  Acidification in the Canadian
Aquatic Environment.  NRCC  Publication  No.  18475. 369 p.

Nicholls, K. H.  1976.  Comparative limnology of  Harp and Jerry Lakes,
adjacent cottaged and uncottaged lakes  on southern Ontario's Precambrian
Shield.  Ont. Min. Environ.  87 pp.

Nilssen, J. P.  1980.  Acidification of a  small watershed  in southern
Norway and some characteristics of  acidic aquatic environments.   Int.
Revue ges. Hydrobiol. 65:177-207.

Nilsson, S. I., H. G. Miller, and J. D. Miller.  1982.  Forest growth  as
a possible cause of soil and water  acidification: an examination of the
concepts.  Oikos 39:40-49.

Nordstrom, D. K.  1982.  The effect of sulfate on aluminum
concentrations in natural  waters:  some stability relations  in the
system Al 203-503-^0 at 298°K.  Geochimica  et Cosmochimica Acta
46:681-692.
                                  4-154

-------
Norton, S. A.  1976.  Changes  in  chemical  processes in soils caused by
acid precipitation.   J_n Proceedings of the  First International Symposium
on Acid Precipitation and the  Forest  Ecosystem, L. S. Dochinger and T.
A. Seliga, eds. USDA Forest Service General Technical Report NE-2, pp.
711-724.

Norton, S. A.  1980.  Geologic factors controlling the sensitivity of
aquatic ecosystems to acidic precipitation, pp. 521-531.   In D. S.
Shriner et al., eds.  Atmospheric Sulfur Deposition.  Ann "Rrbor Science,
Ann Arbor, MI.

Norton, S. A.  1981.  The effects of  acidification on the  chemistry of
ground and surface waters.  Proc. of  Intern.  Symp. on Acidic
Precipitation and Fishery Impacts in  Northeastern North America, Am.
Fisheries Soc. Symp. 20 pp.

Norton, S. A. and C. T. Hess.   1980.  Atmospheric deposition in Norway
during the last 300 years as recorded in SNSF lake sediments.  I.
Sediment dating and chemical stratigraphy,  pp. 268-269.  In Ecological
Impact of Acid Precipitation.   D. Drabluls and A. Tollan, ed"s.  Proc.,
Int. Conf. Ecol. Impact Acid Precip., SNSF Project, Oslo,  Norway.

Norton, S. A. and J. S. Kahl.   1982.  Reduction of terrestrial
alkalinity production—a major mechanism of surface water  acidification
(abs.).  Symp. on Acidic Precip.  and  Atmos. Dep.—A Western Perspective.
Bellingham, WA.

Norton, S. A., R. B. Davis, and D. F. Brakke. 1981a.  Responses of
Northern New England lakes to  atmopsheric inputs of acid and heavy
metals.  Completion Report Project A-048-ME,  Land and Water Resources
Center, University of Maine, Orono.   90  p.

Norton, S. A., C. T. Hess, and R. B.  Davis.  1981b.   Rates of
accumulation of heavy metals in pre-  and post-European sediments in New
England lakes, pp. 409-421. j£ Input of Atmospheric  Pollutants to
Natural Waters.  Eisenreich, S. J., ed.   Ann Arbor Science, Ann Arbor,
MI.

Nriagu, J. 0.  1973.  Lead orthophosphates—II.   Stablility of
chloropyromorphite at 25°C.  Geoch. et  Cosmoch.  Acta  37:367-377.

Nygaard,  G.  1956.  Ancient and recent  flora of  diatoms  and
Chrysophyceae  in Lake Gribso.   Folia Limnol. Scand. 8:32-94.

Oden,  S.   1976a.  The acidity problem—an outline of concepts, pp.  1-36.
In  Proc.  of the First Intern.  Symp. on  Acid Precip.  and  the Forest
"Ecosystem.  Dochinger, L.  S. and T. A.  Seliga, eds.   U.S.  Dept. Agric.
For. Serv. Gen. Tech. Rep. NE-23.

Oden,  S.   1976b.  The acidity problem - an outline  of concepts.   Wat.
Air Soil  Poll. 6:137-166.
                                  4-155

    409-262 0-83-12

-------
Oden, S. and T. Ahl.   1970.   The acidification of Scandinavian waters.
Ymer. Arsboh. 103-122.

Oner-nick, J. M. and C. F.  Powers.  1982.  Total alkalinity of surface
waters - a national map.   EPA-600/D-82-333, Sept. 1982, Corvallis, OR.

Ontario Ministry of the Environment.   1979.  Determination of the
susceptibility to acidification of poorly buffered surface waters.
Water Resources Branch, Box  213, Rexdale, Ontario.  21 pp.

Overrein, L. N., H. M. Seip, and A. Toll an.  1980.  Acid
precipitation—Effects on  forest and  fish.  Final Report of the
SNSF-Project 1972-1980.

Parker, G. G., S. E.  Lindberg,  and J.  M. Kelly.  1980.  Atmosphere-
canopy interactions of sulfur in the  southeastern United States, pp.
477-493.  J_n Atmospheric Sulfur Deposition:  Environmental Impact and
Health Effects. D. S. Shriner, C. R.  Richard, and S. E. Lindberg, eds.
Ann Arbor Science, Ann Arbor, MI.

Pearson, F. H. and A. J.  McDonnell.   1975a.  Use of crushed limestone to
neutralize acid wastes.  J.  of the Environ. Engr. Div. ASCE 101:139-158.

Pearson, F. H. and A. J.  McDonnell.   1975b.  Limestone barriers to
neutralize acid streams.  J. Envir. Eng. Div. ASCE 101:425-440.

Pfeiffer, M. H. and P. J.  Festa.   1980.  Acidity status of lakes in the
Adirondack region of New York in  relation  to fish resources.  N. Y.
State Dept. Envir. Cons.,  FW-P168(10/80),  36 p. + appendices.

Pierce,  R. S., C. W. Martin, C.  C.  Reeves,  G.  E. Likens, and H. Bormann.
1972.  Nutrient loss from clearcuttings  in  New  Hampshire.  Symp. on
Watershed in Trans., Ft. Collins,  CO.  Am.  Water Res.  Assoc. and Col.
St. Univ.  pp. 285-295.

Reckhow, K. H. and S. L. Chapra.   1981.   Selecting a model, lake line.
Newsletter of  the North American Lake Management Society, 1; March 1981.

Reiners, W. A., R. H. Marks, and P.  M. Vitousek.   1975.  Heavy metals  in
subalpine and  alpine soils of New Hampshire.   Oikos 26:264-275.

Renberg, I. and T. Hellberg.  1982.   The pH history of lakes in
southwestern Sweden, as calculated from the subfossil  diatom  flora on
the sediments.  Ambio 11:30-33.

Robertson, C.  E. and J. D. Hem.  1969.  Solubility of  aluminum  in the
presence of  hydroxide, fluoride, and sulfate.   U.S. Geol.  Surv. Water
Supply Pap. 1827-C, Washington, DC.

Scheider, W. and P. J. Dillon.  1976.  Neutralization  and  fertilization
of acidified lakes near Sudbury, Ontario.   Wat. Pollut.  Res. Can.
11:93-100.
                                  4-156

-------
Scheider, W., J.  Adamski,  and  M.  Pay!or.   1975.  Reclamation of
acidified lakes near Sudbury,  Ontario.  Ontario Ministry of the
Environment, Rexdale, Ontario,  Canada.

Scheider, W. A.,  B. Cave,  and  J.  Jones.   1976.  Reclamation of acidified
lakes near Sudbury, Ontario, by neutralization and  fertilization.
Ontario Ministry  of the Environment,  Rexdale, Ontario, Canada.

Scheider W. A., D. S. Jeffries, and P. J.  Dillon.   1979.   Effects of
acidic ppt on Precambrian  freshwater in southern Ontario.  J. Great
Lakes Res. 5:45-51.

Schindler, D. W.  1977.  Evolution of phosphorus limitation in lakes.
Science 195:260-262.

Schindler, D. W.   1980.  Ecological  effects  of experimental whole-lake
acidification. J_n Atmosphere  Sulfur Deposition, Environmental Impact
and Health Effects.  Shriner,  D.  S.,  C. R. Richmond,  and S. E. Lindberg,
eds.  Ann Arbor Science, Ann Arbor,  MI.

Schindler, D. W. and J. E. Nighswander.   1970.  Nutrient supply and
primary production in Clear Lake, Eastern  Ontario.  J. Fish. Res. Board
Can. 27(11):2009-2036.

Schindler, D. W. and T. Ruszczynski.   1983.   A test of limnological data
from the Experimental Lakes Area, Northwestern Ontario for evidence of
acidification.  Can. Tech. Rep. Fish. Aquat. Sci.  1147.  Western Region,
Department of Fisheries and Oceans,  Winnepeg, Canada.

Schindler, D. W.  and M. A. Turner.  1982.  Biological, chemical, and
physical responses of lakes to experimental  acidification. Water, Air,
Soil Pollut. 18:259-271.

Schindler, D. W., F. A. J. Armstrong, S.  K.  Holmgren, and  G. J.
Brunskill.  1971.  Eutrophication of lake 227, experimental lakes area,
northwestern Ontario, by addition of phosphate and nitrate.  J. Fish.
Res. Bd. Canada 28(11):1763-1782.

Schindler, D. W., H. Kling, R. W. Schmidt, J. Prokowpowich, V. E. Frost,
R. A. Reid, and M. Copel.   1973.   Eutrophication of Lake 227 by addition
of phosphate and nitrate:   The second, third, and  fourth years of
enrichment, 1970, 1971, and 1972.  J. Fish.  Re.s Bd.  Canada
30:1415:1440.

Schindler, D. W., R. Wagemann, R. B. Cook, T. Ruszcczynski,  and J.
Prokopowich.  1980.  Experimental acidification  on Lake 223,
Experimental Lakes Area:  Background data and the  first  three years  of
acidification.  Can. J. Fish.  Aquat. Sci. 37:342-354.

Schnitzer, M.  1978.   Humic substances:  chemistry  and reactions, pp.
1-64.  J_n  Soil Organic Matter.  Schnitzer, M. and  S.  U. Khan,  eds.
Elsevier,  Amsterdam.


                                  4-157

-------
Schnitzer, M.  1980.   Effect  of  low  pH on  the chemical structure and
reactions of humic substances, pp. 203-222.  In Effects of Acid
Precipitation on Terrestrial  Ecosystems.   Hutclvinson, T. C. and M.
Havas, eds.  Plenum Press,  New York.

Schnitzer, M. and S.  U.  Kahn.  1972.  Humic  Substances in the
Environment.  Marcel  Dekker,  Inc., New York.

Schnoor, J. L., G. R. Carmichael,  F.  A. Van  Schepen.  1982.  An
integrated approach to acid rainfall  assessments.   Chapter 13 _I_n Vol II.
Energy and Environmental Chemistry:   Acid  Rain.  L. H. Keith, ed. Ann
Arbor Science, MI.

Schofield, C. L.  1976a.  Dynamics and management of Adirondack fish
populations.  Final Rept.  for Project F-28-R, Dept. of Environ.
Conservation, Albany, NY.   27 p.

Schofield, C. L.  1976b.  Acid precipitation:   Effects on fish.  Ambio
5:228-230.

Schofield, C. L.  1976c.  Lake acidification in the Adirondack Mountains
of New York:  Causes and consequences, p.  477.  ^Proceedings of the
First International Symposium on Acid Precipitation and the Forest
Ecosystem.  L. S. Dochinger and  T. A. Seliga, eds.  U.S.D.A. Forest
Service Tech. Rept. NE-2.

Schofield, C. L.  1980.  The  processes limiting fish  population in
acidified lakes.  In Atmospheric Sulfur Deposition, Eds. D. Shriner, C.
Richmond, S. Lindberg.  Ann Arbor Science,  pp. 345-356.

Schofield, C. L.  1981.  Historical  fisheries changes as related to
surface water pH changes in the  United States.  ^Proceedings of an
International Symposium on Acidic Precipitation and Fishery Impacts on
Northeastern North America.  Aug. 2-5, 1981. Cornell University, Amer.
Fish. Soc.,  Ithaca, NY.

Schofield, C. L. and J. R. Trojnar.   1980.  Aluminum  toxicity  to brook
trout (Salve!inus fontinalis) in acidified waters,  pp. 347-366.  In
Polluted  Rain.TorTbara,  T.  Y., M.  W. Miller,  M.  W., and P. E. Morrow,
eds.  Plenum Press, New York, NY.

Schofield, C. L., D. A. Webster, C.  A. Guthrie, and W. A. Flick.  1981.
Management of acidified waters.   Final Report  C-164480, New York State
Department of Environmental Conservation.

Shaffer,  P. W. and J. N. Galloway.  1983.   The  effect of acidic
deposition on the composition of streams  in Shenandoah  National Park.
Proceedings of Symposium on Acid Precipitation  from International
Symposium on Hydometerology.   Denver, CO.
                                  4-158

-------
Shilts, W. W.  1981.   Sensitivity  of bedrock to acid precipitation and
derived soils to acid precipitation.  Geol. Survey of Canada, Paper
81-84.

Siccama, T. G., W. H. Smith,  and D. L. Mader.  1980.  Changes in lead,
zinc, copper, dry weight,  and organic matter content of the forest floor
of white pine stands  in central Massachusetts over 16 years.  Environ.
Sci. Techno!. 14:54-56.

Skougstad, M. W., M.  J. Fishman, L. C. Friedman, D. E. Erdmann, And S.
S. Duncan, eds.  1979.  Methods for Determination of Inorganic
Substances in Water and Fluvial Sediments, Techniques of Water-Resources
Investigations of the United  States Geologic Survey, Book 5, Chapter Al,
U. S. Government Printing  Office,  Washington, D.C.

Smith, M. W.   1969.  Changes  in environment and biota of a natural lake
after firtilization.   J. Fish. Res. Bd.  Canada 26:3101-3132.

Smith, R. W. and J. D. Hem.   1972. Effect of aging on aluminum
hydroxide complexes in dilute aqueous solutions.  U.S. Geol. Survey
Water Supply Pap. 1827-D,  51p.

Stumm, W. and J. J. Morgan.   1970. Aquatic Chemistry.  Wiley
Interscience, New York, NY.   583 pp.

Stumm, W. and J. J. Morgan.   1980. Aquatic Chemistry, John Wiley and
Son  Inc., NY.

Stumm, W. and J. J. Morgan.   1981. Aquatic Chemistry.  John Wiley and
Sons, Inc., New York, NY.

Sverdrup, H. U.  1982.  Lake  liming.  Chemical Scripta  (in press) p. 30.

Sverdrup, H. and I. Bjerle.   1982. Dissolution of calcite and other
related minerals in acidic aqueous solution in a pH-stat.  Vatten.
38:59-73.

Swedish Ministry of Agriculture Environment Committee.  1982.
Acidification Today and Tomorrow.   Risbergs Tryckeri:  A. B. Uddevalla,
Sweden, p.231.

Thomas, J. F. J.  1960.  Water Survey Report  II.   Ind. Water Resour. of
Canada, Dept. of Mines and Tech.  Surv.  No. 864.

Thompson, M. E.  1982.  The cation denudation  rate  as  a quantitative
index of  sensitivity of eastern Canadian rivers  to  acidic atmospheric
precipitation.  Water, Air and Soil  Pollution, 18:215-226.

Thompson, M. E. and M. B.  Mutton.   1981.  Selected  Canadian
lake—cations, sulfate, pH,  PC02  relationships.  Environment Canada,
Canada Centre  for  Inland Waters,  31 pp.
                                  4-159

-------
Thompson, M. E., F. C. Elder,  A.  R.  Davis,  and  S.  Whitlow.   1980.
Evidence of acidification of rivers  of Eastern  Canada,  pp. 244-245.
In Ecological Impact of Acid Precipitation.   Proceedings  of  an
International Conference, Sandefjord,  Norway, March  11-14, 1980.
Drablj6s, D. and A. Tollan, eds.   SNSF-project,  Oslo-As, Norway.

Traaen, T. S.  1980.  Effects of acidity on decomposition of organic
matter in aquatic environments,  pp.  340-341.  In Ecological  Impact of
Acid Precipitation.  Proceedings  of  an International  Conference,
Sandefjord, Norway, March 11-14,  1980.  Drabltfs, D.  and A. Toll an,
eds. SNSF-project, Oslo-As, Norway.

Ugolini, F. C., R. Minden, H. Dawson and J.  Zachara.   1977.  An example
of soil processes in the Abies Amabilis zone  of Central Cascades,
Washington, Soil Science 724:291-302.

Underwood, J. K., H. H. Vaughar,  J.  G. Ogden, III, and C. G. Mann.
1982.  Acidification of Nova Scotia  Lakes II:   Ionic  balances in dilute
waters.  Report to Nova Scotia Department of the Environment.

U.S./Canada 1982.  Memorandum of Intent on Transboundary  Air Pollution.
1981.  Impact Assessment, Working Group I.   Phase  II. Interim Working
Paper.  October 1981.

van Breemen, N.  1973.  Dissolved aluminum in acid sulfate soils and in
acid mine waters.  Soil Sci. Soc. Am.  Proc. 37:694-697.

Vangenechten, J.H.D. and O.L.J. Vanderborght.  1980.   Acidification of
Belgian moorland pools by acid sulphur-rich rainwater. Jji Ecological
Impact of Acid Precipitation.  D. Drablos and A. Tollan,  eds. SNSF
project, pp. 246-247.

Vollenweider, R. A.  1968.  Scientific fundamentals  of the
eutrophication of lakes and flowing  waters, with particular  reference  to
nitrogen and phosphorus as factors in eutrophication.  O.E.C.D.
DAS/CSI/68.  27:159.

Watt, W. D., D. Scott, and S. Ray.  1979.  Acidification  and other
chemical changes in Halifax County lakes after  21  years.  Limnol.
Oceanogr. 24:1154-1161.

Watt, W. D., D. Scott, and W. J. White.  1983.   Evidence  of
acidification in some Nova Scotia rivers and of impact on Atlantic
salmon, Salmo salar. Can. J. Fish. Aquat. Sci.:  in  press.

Webb, T., III. and D. R. Clark.  1977.  Calibrating  micropaleontological
data  in climatic terms:  a critical  review.  Annals  of the New  York
Academy of Sciences 288:93-118.

Westcott, C. C.  1978.  pH Measurements.  Academic Press, New York, NY.
                                  4-160

-------
Wilson, D. E.  1979.   The influence of humic compounds on titrimetric
determination of total  inorganic carbon in freshwater.  Arch. Hydrobiol.
87:379-384.

Wiltshire, J. F. and J. R. Machell.   1981.  A  study of acidification in
sixteen lakes in mainland Nova Scotia and  southern New Brunswick.  A
preliminary report.  Environmental Protection  Service, Atlantic Region,
Halifax, Nova Scotia.

Wood, J. M.  1980.  The role  of pH and oxidation-reduction potentials in
the mobilization of heavy metals, pp. 223-236.  J^n Polluted Rain.
Toribara, T. Y., M. W.  Miller, and P. E. Morrow, eds.  Plenum Press, New
York, NY.

Wright, R. F.  1977.   Historical changes in the pH of 128 lakes in
southern Norway and 130 lakes in southern  Sweden over the period
1923-1976.  SNSF-project, TN34/77.  71 p.

Wright, R. F. and E.  T. Gjessing.  1976.   Acid precipitation:  Changes
in the chemical composition of lakes. Ambio 5:219-223.

Wright, R. F. and A.  Henriksen. 1978.  Chemistry of small Norwegian
lakes with special reference  to acid  precipitation.  Limnol. Oceanogr.
23:487-498.

Wright, R. F. and M.  Johannessen.  1980.   Input-output budgets of major
ions at gauged catchments in  Norway,  pp. 250-251.  Jji Ecological Impact
of Acid Precipitation.   Proceedings of an  International Conference,
Sandefjord, Norway, March 11-14, 1980.  Drabljos, D. and A. Tollan,
eds. SNSF-project, Oslo-As, Norway.

Wright, R. F. and E.  Snekvik.  1978.  Acid precipitation chemistry and
fish populations in 700 lakes in southernmost  Norway.  Verh.  Internat.
Verein. Limnol. 20:765-775.

Wright, R. F., T. Dale, A. Henriksen, G. R. Hendrey, E. T. Gjessing, M.
Johannessen, C. Lysholm and E. Storen.   1977.   Regional Surveys of Small
Norwegian Lakes.  SNSF project IR33/77.  Oslo, Norway.

Wright, R. F., N. Conroy, W.  T. Dickson, R. Harriman, A. Henriksen, and
C. L.  Schofield.  1980.  Acidified  lake  districts  of  the world:  a
comparison of water chemistry of lakes  in  Southern  Norway, Southern
Sweden, Southwestern Scotland, the  Adirondack  Mountains of New York, and
South  Eastern Ontario, pp. 377-379.   ^Ecological  Impact of  Acid
Precipitation.  Proceedings of an  International  Conference,  Sandefjord,
Norway, March 11-14. 1980.  Drabltfs,  D.  and A. Tollan, eds.
SNSF-project, Oslo-As, Norway.

Yan, N. D. and P. J. Dillon.   1981.   Studies  of Lakes and Watersheds
near Sudbury, Ontario.  Draft, Ontario  Ministry of the Environment
Technical Report, Water Resources  Branch,  Rexdale,  Ontario,  Canada.
                                  4-161

-------
Van, N. D. and C. Lafrance.   1982.   Experimental  fertilization of  lakes
near Sudbury, Ontario.   In Studies  of Lakes  and  Watersheds near  Sudbury,
Ontario:  Final Limnologjy~~Report of the Sudbury  Environmental Study.
Ont. Min. of Environ.

Van, N. D. and G. E. Miller.   1982.  Characterization  of lakes near
Sudbury, Ontario.  In Studies of Lakes and Watersheds  Near Sudbury,
Ontario: Final LimnoTogy Report of  the Sudbury Environmental Study.
Ont. Min. of Environ.

Van, N. D. and W. A. Scheider, and  P. J. Dillon.   1977.  Chemical  and
bilogical changes in Nelson Lake, Ontario following  experimental
elevation of lake pH.  ^ni Water Pollution Research Canada, Proceedings
of the 12th Canadian Symposium.

Yuen, S., S. J. Charlson, H.  H. Harrison, and M.  B.  Baker.  1979.  The
global and regional  tropospheric sulfur cycle.   Symp.  Budget and Cycles
of Trace Gases and Aerosols in the  Atmosphere.   Fourth International
Conference of the Commission  on Atmospheric  Chemistry  and Global
Pollution.  Boulder, CO.  August.

Zeman, L. J.  1973.   Chemistry and  troposphere fallout and stream  flow
in a small mountainous  watershed stream in Vancouver,  B. C.  Ph.D.
Thesis, Univ. of British Columbia.

Zimmerman, A. P. and H. H. Harvey.   1979. Sensitivity to acidification
of waters of Ontario and neighboring states.  Final  Report to Ontario
Hydro, University of Toronto, Toronto, Ontario.   118 p.
                                  4-162

-------
            THE ACIDIC DEPOSITION PHENOMENON  AND  ITS  EFFECTS
                   E-5.   EFFECTS ON  AQUATIC BIOLOGY

5.1  INTRODUCTION (J.  J.  Magnuson)

     The loss of fish  populations from seemingly  pristine oligotrophic
waters was the first and most obvious indication  that atmospheric
deposition was affecting  aquatic ecosystems (Dannevig 1959, Beamish and
Harvey 1972, Cowling 1980).   Changes in water chemistry, particularly
increases in acidity,  were found to  be associated with these local fish
extinctions.  Later studies have included the effects of acidification
on other aquatic organisms,  such as  those associated  with bottom
substrates (the benthos), tiny plants and animals floating  freely in the
water column (the plankton), and rooted aquatic plants (macrophytes).
The resultant literature is large, widely scattered,  and varies
considerably in its scientific merit.  The purpose of this  chapter is to
review and evaluate this literature  critically, and to summarize the
effects of acidification  on  aquatic  organisms.

     The chapter begins  with a section on naturally acidic  waters,
including a discussion of what organisms occur in such habitats and how
their distributions relate to distributions in habitats recently
acidified by man's activities.  Subsequent sections critically evaluate
the literature regarding  the response of benthic  organisms, macrophytes
and wetland plants, plankton, fishes and other aquatic biota to
acidification.  These  are followed by a discussion of ecosystem-level
responses to acidification and a section on mitigative options.  The
final section summarizes  the known effects of acidification on aquatic
biota and indicates potential effects that need to be addressed.

     It should be kept in mind that  acidification of  freshwaters is a
complex process that involves more than merely increases in acidity.
Other well-documented  changes include increased concentrations of metal
ions, increased water  clarity, the accumulation of periphyton
(microflora attached to bottom substrates) and detritus, and changes in
trophic interactions (e.g.,  loss of  fish as top predators).  The
response of aquatic systems to acidic deposition  must be viewed in terms
of all these changes that together constitute the acidification process.

     Evidence linking  changes in aquatic communities  to acidification
can be divided into three types.  The first type  consists of field
observations, which are  1) descriptions of conditions before and after
                                  5-1

-------
acidification is suspected to have occurred or 2)  contemporary
comparisons of water bodies thought to exhibit different degrees  of
acidification.  Problems exist with this type of correlation approach.
For example, before and after studies may be difficult to interpret if
methodologies have changed in the interim, or if other factors  such as
land-use practices have also changed.  In comparative studies,  pH is
frequently correlated with other limnological parameters (e.g., lake
size, nutrient concentrations), making it difficult to attribute
inter-lake biotic differences solely to differences in pH.  Despite
these problems, field observations provide the earliest indications of
changes in biotic communities and provide a basis for forming hypotheses
that can be further evaluated when consistent trends are observed in
repeated studies.

     The second type of evidence consists of field experiments, which
range from modifying the conditions of enclosures in a lake (Muller
1980) to intentionally acidifying an entire lake or stream (Schindler et
al. 1980b; Hall et al. 1980).  These studies generally minimize the
problem of confounding factors, which plague field observation  studies,
and have contributed much to our understanding of how organisms are
affected by the acidification process.  However, experimental
manipulations that focus on one variable may miss effects which are due
to the interaction of several variables.  For example, acidifying an
entire lake may not reveal a major reason for fish kills in waters
acidified by acidic precipitation, namely aluminum released when the
surrounding watershed is also acidified.  A great difference also exists
between the time scale of experimental acidifications (which typically
occur over a period of months or a few years) and of regional
acidification  (which occurs over many years).

     The third type of evidence consists of laboratory experiments,
whereby the effect of a particular stress (low pH, aluminum) is
evaluated after all other variables are carefully controlled.  These
experiments typically consist of bioassays involving one species and one
or a small number of stresses.  Most of our understanding of the
physiological effects of low pH on aquatic organisms is due to such
studies.  As with field experiments, these studies are time consuming,
expensive and have yielded data on only a few species.  Predicting
community-level changes from laboratory bioassays on a few species is
difficult.  A species may experience reduced growth or reproduction in
the laboratory at a low pH, but may prosper in an acidified lake at the
same pH if its competitors suffer even greater reductions in growth and
reproduction.

     It is obvious that all three types of evidence provide certain
kinds of information yet have certain drawbacks.  The strongest
conclusions regarding the effects of acidification on aquatic organisms
will be reached when all three types of evidence yield consistent
results.  Examples of such cases are given in the conclusions section
(Section 5.10.1).
                                  5-2

-------
     The significance of changes in species abundances or community
composition lies in how these changes affect important ecosystem
processes.  These processes include primary production (the production
of new plant tissue through photosynthesis), nutrient recycling (re-use
of nutrients released through decomposition of organic material), and
trophic interactions (transfer of energy from plants to herbivores to
carnivores).  A schematic presentation of these processes and how they
may be affected by acidification is given in Section 5.8 (Figure 5-17).
While direct toxic effects of acidification on organisms have been
relatively easy to document, assessing effects on ecosystem processes
has proven more difficult.  We know, for example, that certain species
of algae become dominant under acidic conditions, yet how this affects
the food supply to higher trophic levels, or how total primary
productivity is affected has not been well  studied.  The growth of algal
mats in acidified lakes has been observed,  yet how this seal over the
bottom sediments will affect nutrient cycling has not been measured.
Most effort to date has involved describing responses of various taxa to
the acidification process.  Future work will need to consider how these
changes affect ecosystem processes.

5.2  BIOTA OF NATURALLY ACIDIC WATERS (J. J. Magnuson and F. J. Rahel)

     Naturally acidic lakes and streams occur throughout the world and
have been known in the United States since at least the 1860's
(Hutchinson 1957, Patrick et al. 1981).  These naturally acidic waters
provide insight into the pH range normally tolerated by aquatic
organisms.  Such information is useful  in assessing how recent pH
declines attributed to cultural acidification might affect aquatic life.
This chapter's purpose is to summarize the literature on naturally
acidic waters and to examine the influence of low pH on plants and
animals found in such habitats.  North American waters are emphasized,
but reference to other geographic areas is  made when cosmopolitan taxa
are involved.  Methods for distinguishing between naturally acidic and
culturally acidified waters are discussed in Chapter E-4, Section 4.4.3.

5.2.1   Types of Naturally Acidic Waters

     Naturally occurring acidic waters  fall  into three groups.  In the
first group are inorganic acidotrophic  waters associated with geothermal
areas or lignite burns, where pH values between 2.0 and 3.0 are not
uncommon (Waring 1965, Brock 1978, Hutchinson et al. 1978).  Among the
most extreme values recorded are pH 0.9 for Mount Ruapehu Crater Lake,
New Zealand (Bayly and Williams 1973),  pH 1.7 from Kata-numa, a volcanic
lake in Japan (Hutchinson 1957), and pH's below 2.0 for several springs
in Wyoming (Brock 1978).  The high acidity is due to sulfuric acid,
which arises from the oxidation of sulfides such as hydrogen sulfide
(H2$)  and pyrite (FeS?).  In addition to being extremely acidic,
these waters frequently contain elevated metal  concentrations and are
often heated.  Assessing the biological effects of low pH under these
conditions is difficult, but such sites have provided insight into the
lower pH limit for various taxa (Brock  1973, 1978).  This type of
naturally acidic aquatic habitat occurs in  North America mainly in the
                                  5-3

-------
west, and has been most extensively  studied in the Yellowstone Park
region of Wyoming (Van Everdingen 1970,  Brock  1978).

     The second group of naturally acidic  waters  consists  of  brownwater
lakes and streams associated with peatlands,  cypress  swamps,  or
rainforests, depending on latitude (Janzen 1974,  Moore and Bellamy
1974).  Their acidity is derived from organic  acids leached from  decayed
plant material and from hydrogen ions released by plants such as
Sphagnum mosses in exchange for nutrient ions  (Clymo  1967).  These
waters commonly have pH's in the range of 3.5  to  5.0  and owe  their dark
color to large amounts of dissolved organic matter.  As with  acidic
geothermal waters, brownwaters have  other  qualities besides low pH that
may limit aquatic life.  For example, they are characterized  by low
concentrations of many of the inorganic ions  necessary for plant  growth
and osmotic balance in animals (Clymo 1967).   There is some evidence
that the dissolved humic compounds may be  toxic to amphibians, even at
neutral pH (Gosner and Black 1957, Saber and Dunson 1978). Low oxygen
and high carbon dioxide concentrations are also present in some
brownwater habitats (Welch 1952, Kramer et al. 1978).  Finally, the low
primary productivity of brownwaters  may mean  that even physiologically
tolerant species may be excluded due to food unavailability (Janzen
1974, Bricker and Gannon 1976).  Brownwater habitats  in North America
are associated with either northern peatlands  (Jewell and  Brown 1929,
Cole 1979, Johnson 1981) or with southeastern  swamplands (Beck et al.
1974, Forman 1979, Kirk 1979).

     The third type of naturally acidic habitat consists of ultra-
oligotrophic waters.  They are especially  common  where glaciation has
removed younger calcareous deposits and exposed weather-resistant
granitic and siliceous bedrock.  The absence of carbonate  rocks in the
drainage basin results in lakes with little carbonate-bicarbonate
buffering capacity; hence such lakes are very  vulnerable to pH changes.
They often have pH's in the 5.5 to 6.5 range,  and most of  the acidity
appears due to carbonic acid (HeCOa).  These lakes tend to be small
and have low concentrations of dissolved ions (Chapter E-4, Section
4.3.2).  In North America, this type of naturally acidic lake occurs  in
large areas of eastern Canada and the northeastern United  States, as
well as in sections of western United States and  northern  Florida
(Shannon and Brezonik 1972, Galloway and Cowling  1978). Many of  the
lakes which have been, or will be, affected by acidic precipitation
belong in this category (see Chapter E-4,  Section 4.3.2).

5.2.2   Biota of Inorganic Acidotrophic Waters

     In North America, the most extensively studied inorganic
acidotrophic waters are those of the Yellowstone  Park region  in Wyoming.
Certain species of eucaryotic algae, fungi, and bacteria have
demonstrated remarkable adaptation to this acidic environment and often
form extensive mats (Brock 1978).  For example, the alga Cynanidi urn
caldarium was found at pH 0.05, while the bacterium Sulfolobus
acidocaldarius thrived in a thermal  spring at pH  0.9  and 60 C.  Lower pH
limits for other taxa in this environment are summarized by Brock (1978)
                                  5-4

-------
and include a pH near 0.0 for fungi, pH 3.0 for Sphagnum mosses, and pH
2.5 to 3.0 for vascular plants such as sedges (Carex and Eleocharis
spp.) and ericacid shrubs (blueberries, cranberries).   Al though
generally considered eurytropic, blue-green algae are  conspicuously
absent from these acidic environments.  Brock (1973, 1978)  has assembled
data showing that these algae are intolerant of pH's below 4.0.  The
inability to survive under acidic conditions may be due to their lack of
membrane-bound chloroplasts that, in eucaryotic algae, prevent the
acid-labile chlorophyll from being decomposed at low pH.

     In ponds exposed to sulfur fumigations from burning bituminous
shales, the euglenoid Euglena mutablis was present at  pH 1.8 (Hutchinson
et al. 1978, Havas and Hutchinson 1982).  The red chironomid,  Chironpmus
riparius, and the rotifer, Brachionus urceolaris, were abundant at pH
2.8, but no copepods or cladocerans were present.

     Among the few insects reported from acidic thermal waters is the
ephydrid fly Ephydra thermpphila (Brock 1978).  This fly breeds in
streams at pH 2.0 and is the basis of a food chain involving several
invertebrate predators.  Extensive surveys of invertebrates in the
acidic geothermal waters of North America have not been done,  but it
seems reasonable that other invertebrate taxa might tolerate such low
pH.  For example, in streams polluted by acidic mine wastes, species of
rotifers, midges, alderflies and dytisscids have been  found at pH's near
3.0 (Roback 1974, Harp and Campbell 1967, Parsons 1968).

     Vertebrates such as amphibians and fish appear unable to survive in
inorganic acidotrophic habitats, but again no extensive surveys have
been undertaken.  Surprisingly, waterfowl do not avoid these lakes, and
Canadian geese have been reported to nest on Turbid Lake in Yellowstone
Park (pH ~ 3.0) (Brock 1978).

     Another group of inorganic acidotrophic lakes that have been well
studied are the volcanic lakes of Japan (Ueno 1958).  Some of the
organisms present in these lakes belong to cosmopolitan genera and hence
provide insight into the lowest pH which may be tolerated  by Morth
American genera.  Aquatic mosses (e.g., Rhynchostegium aplozia) dominate
the plant community, although reeds (Phragmites) occur" along the margins
of most lakes, even at pH's below 3.01Diatoms (Pinnularia) and
rotifers (Rotaria)  have been observed at pH 2.7.  A small  caldera lake
filled with water at pH 3.0 but fertile enough to support moderate
phytoplankton production contained several genera of Crustacea
(Simocephalus, Cnydorus, Macrocyclops) and a rotifer (Braehionus).  The
teleost Tribolodon hakonensis from Lake Osoresan-ko (pH 3.5) occurs at
the lowest pH reported for any fish species (Mashiko et al.  1973).

     While the work done on inorganic acidotrophic waters  has revealed
some outstanding examples of extreme pH tolerance, in  general, these
waters have very low species diversity and monocultures of tolerant
species are common.
                                  5-5

-------
5.2.3  Biota jn Acidic Brownwater Habitats

     Brownwater habitats do not experience  the extremes  of  temperature,
pH, and metal concentrations common to inorganic  acidotrophic  waters;
consequently they contain a greater diversity of  organisms.  They  are,
however, characterized by low ion concentrations, reduced light
penetration and, frequently, low dissolved  oxygen concentrations.   These
variables interact with the acidic pH (3.5  to 5.0)  to  determine species
richness and biological production.

     Among the genera of macrophytes reported from acidic brownwater
lakes are Alternanthera, Ceratpphyllum,  Ispetes,  Juncus, Limnobiuiri,
hluphar, Potamogeton and Utricularia (Jewell  and Brown  1924,  Griffiths
1973, Stoneburner and Smock 1980).  Many brownwater lakes,  however,  are
characterized by the absence of macrophytes,  which is  generally
attributed to the stained water and the  lack  of a firm substrate on  the
lake bottom (Welch 1952, McLachlan and McLachlan  1975, Marshall  1979).
The shoreline plant community has been well  described  for northern bogs
and includes sedges (Carex) , ericacid shrubs  (Vaccim'um  chamaedaphe) and
mosses (Sphagnum) (Gates 1942, Heinselman 1970, Vitt and Slack 1975).
The characteristic tree along the shore of  southeastern  brownwater lakes
is the cypress (Taxodium) (King et al . 1981).

     Phytoplankton have classically been described as  present  at low
densities (Birge and Juday 1927, Welch 1952,  Stoneburner and Smock
1980).  Recent work has emphasized the predominance of small -bodied
algae (the nannoplankton) in these waters (Bricker and Gannon  1976).
Although species from most phy topi ank ton phyla have been reported,
certain genera of desmids (Xanthidium, Euastrum,  Hyalotheca) and diatoms
(As ten' one! la, Eunotia, Ac tin ell a, Anomoeoneis, Pinnularia,  Melosira)
are especially characteristic (Woelkerling  and Gough 1976,  Marshall
1979, Patrick et al . 1979, Stoneburner and  Smock  1980).  As  with the
phy topi ank ton, the zooplankton in acidic dystrophic lakes are  frequently
dominated by small -bodied forms, particularly rotifers (Brachionus,
K era tell a, Monostyla, Polyarthra) and copepods (Diaptomus,  Cyclops)
(Welch 1952, Smith 1957, Bricker and Gannon 1976, Marshall  1979).
Relatively few cladocerans have adapted to this environment although
species from the following genera have been reported:  Alona,  Bosmina,
Chydorus, Daphm'a, Diaphanpsoma, Eubosmina, Leptodora, and  Pleurpxus
(Marshall 1979, Von Ende 1979, Stoneburner and Smock 1980).   In lakes
where fish are absent or where darkly stained water and  low hypolimnetic
oxygen offer some protection from fish predation, dipteran  larvae  of the
genus Chaoborus are an important part of the zooplankton community (Von
Ende
     A peculiar phenomenon in many acidic brownwater lakes is the large
standing crop of zooplankton relative to phytoplankton.   This paradox
has lead to suggestions that bacteria and suspended organic matter
(tripton) may be important food sources for zooplankton  in these lakes
(Bayly 1964, Bricker and Gannon 1976, Stoneburner and Smock 1980).
                                  5-6

-------
     The benthic community in acidic dystrophic  lakes  is  typically
impoverished.  This is particularly true of small  bogs where  a  deep
layer of decaying peat obliterates any sand or gravel  substrate and
prevents macrophyte growth.   Such lakes have dipteran  larvae
(Chaoboridae and Chironomidae),  dragonflies and  damsel flies (Odonata),
and alderflies (Sialidae)  as their main benthic  invertebrates (Welch
1952, McLachlan and McLachlan 1975).  Even habitats with  more diverse
substrates still  have few  benthic species although caddisflies
(Trichoptera), whirligig beetles (Gyrinidae), and cranefly larvae
(Tipulidae) are sometimes  present (Smith 1961, Patrick et al.  1979).
Jewell and Brown (1929)  described an interesting invertebrate community
living in pools in the sphagnum  mat of a Michigan bog  at  pH 3.5 to 4.0.
Air-breathing forms like beetles (Dytisicidae, Haliplidae, Helodidae,
Hydrophilidae) and mosquito  larvae (Culex)  predominated in these
low-oxygen pools, although several dragonfly species (Odonata)  and the
cladoceran, Acantholebris  curvirostri, were also present.

     Notably absent from acidic  bog waters are mayflies (Ephemeroptera);
crustaceans such as amphipods, ostrocods and crayfish;  molluscs (snails,
clams); sponges;  and annelids (oligochaetes, leeches)  (Pennak  1953,
Wetzel 1975).  The absence of organisms that have a calcified
exoskeleton is not unexpected in brownwater habitats due  to the low pH
and the extremely low concentration of calcium.   An exception  to this
generalization is the occurrence of the fingernail  clam (Pisidium) in
bog lakes at pH's below  5.0  (Griffiths 1973).

     Summaries of fish species distribution in relation to pH  exist for
both northern and southern brownwater habitats (Frey 1951, Hastings
1979, Rahel and Magnuson 1983).   Slow growth and low species  diversity
characterize the fish assemblages in these waters (Smith  1957,  Garton
and Ball 1969).  In northern midwestern lakes where ice cover occurs,
winter anoxia interacts  with pH  to determine the structure of fish
assemblages (Rahel 1982).  Lakes with adequate winter  oxygen
concentrations are dominated by  yellow perch (Perca flavescens), sunfish
(family Centrarchidae),  and  bullheads (Ictalurus spp.), even  down to pH
4.5.  If winter oxygen concentrations are low enough to exclude
predators, minnows (family Cyprinidae) dominate  the fish  fauna,  but only
if the pH is above 5.2 to  5.4.  Lakes that are both very  acidic (pH
below 5.2) and experience  winter anoxia contain  only yellow perch and
the central mudminnow (Umbra limi).  Other species that can survive in
acidic northern brownwaters  but  are probably excluded  because  suitable
habitat or spawning areas  are missing are the northern  pike (Esox
lucius), and brook trout (Salvelinus fontinalis)  (Jewell  and  Brown 1924,
Smith 1961, Dunson and Martin 1973).

     Southeastern brownwater lakes and streams (pH 4.0 to 5.0)  have a
more diverse fish fauna  than do  similar northern waters (Wiener and
Giesy 1979, Frey 1951, Laerm et. al 1980).   Among the  more common taxa
are various species of sunfish,  pickerel  (family  Esocidae), catfish
(family Ictaluridae), and  killifish (family Cyprinidontidae), along with
the American eel  (Anguilla rostrata), lake chubsucker  (Erimyzon
                                  5-7

-------
sucetta). eastern mudminnow (Umbra pygmaea),  pirate perch (Aphredoderus
sayanus), and the yellow perdu

     With the exception of the golden shiner  (Notemigonus crysoleucas),
ironcolor shiner (Notropis chalybaeus),  and the swamp darter (Etheostoma
fusi forme), minnows and darters are conspicuously absent from acidic
brownwaters, even though they may be abundant in nearby neutral  waters
(Frey 1951, Laerm et al. 1980, Rahel and Magnuson 1983).  Predation from
bass and pike may exclude these small-bodied  fishes from many habitats,
but even when predators are absent, minnows and darters are rarely found
below pH 5.2.  Other acid sensitive species are the smallmouth bass
(Mi cropterus do!omieui), and walleye (Stizostedion vitreum).

5.2.4  Biota in Ultra-Oligotrophic Waters

    The third category of naturally acidic waters consists of ultra-
oligotrophic lakes and streams.   Hydrogen ion concentrations fluctuate
in these waters as a function of photosynthetic activity and carbon
dioxide concentrations, with pH typically varying between 5.5 and 7.0.
Low nutrient concentrations result in low biological  productivity at all
trophic levels.  Most aquatic taxa are able to tolerate the hydrogen ion
concentration of these lakes and thus other physical/chemical factors
(e.g., thermal conditions) or biotic interactions (predation and
competition) are important in determining species composition.

     A great diversity of taxa has been  reported from ultra-
oligotrophic lakes, but certain groups are characteristic of this lake
type.  In the phytopiankton, for example, crysophytes and diatoms
(Chrysophyta) along with desmids and other green algae (Chlorophyta) are
diagnostic of oligotrophic conditions (Hutchinson 1967).  Numerous other
algae are usually present at low densities (Schindler and Holmgren 1971,
Baker and Magnuson 1976).

     Copepods appear to dominate the zooplankton community, but numerous
other taxa have been recorded in surveys of oligotrophic waters (Ratalas
1971, Torke 1979).  Factors like lake depth and size, thermal regimes,
phytoplankton abundance, and fish predation appear to be more important
than pH in determining zooplankton community structure in these lakes
(Anderson 1974, Green and Vascotto 1978).

     Benthic communities are diverse, although certain genera of midge
larvae (Tanytarsus, Chaoborus) along with fingernail  clams (Pisidium),
the amphipod Pontoporeia, and" the mysid My sis relicta have classically
been associated with oligotrophy (Hamilton 1971, Brinkhurst 1974, Wetzel
1975).  In acidic streams (pH less than 5.7), mayflies (Ephemeroptera),
molluscs, some caddisfly genera (Hydropsyche), and the amphipod
(Gammarus) are rare, even though they are abundant in downstream
sections having a higher pH (Sutcliffe and Carrick 1973).  These taxa
are also missing from streams affected by acidic mine drainage (Roback
1974).  Shell-forming molluscs and crustaceans may be excluded from
oligotrophic waters because of low calcium concentrations, even though
the pH is circumneutral.  Crayfish, for example, were absent from
softwater Wisconsin lakes having calcium concentrations below 2 mg
£-! regardless of lake pH (Capelli 1975).


                                  5-8

-------
     Aquatic macrophytes typical of oligotrophic waters have been
summarized by Hutchinson (1967) and Seddon (1972).  Among the
representative genera are Bidens, Elatine, Eriocaulpn, Ispetes, Juncus,
Lobelia, and Sparganium.  Most of these have a distinct physical  form,
consisting of stiff leaves placed in a close rosette or on short,
unbranched stems as opposed to the long-stemmed, branched leaf typical
of hardwater macrophytes (Fasset 1930).  Species occurring in
oligotrophic waters are probably not restricted to the low nutrient
conditions present there but are likely excluded from more fertile
waters by competition from other macrophyte species (Hutchinson 1967).

     Identifying fish assemblages typical  of oligotrophic waters is
complicated by human activities that affect community composition, such
as stocking, over-exploitation, and eutrophication (Regier and Applegate
1972).  Many high-elevation Palearctic lakes were probably barren of
fish following deglaciation, although the very long and poorly
documented history of fish introductions by humans makes it impossible
to know what percent were fishless (Nilsson 1972, Donald et al. 1980).
These coldwater lakes today are dominated by salmonids (trout and
salmon) and coregonids (whitefish and ciscoes).  Oligotrophic lakes with
slightly warmer thermal  regimes (because they are shallower or are
located at lower altitudes or farther south than the salmonid lakes) are
dominated by percids (yellow perch)  and certain centrarchids (typically
the smallmouth bass, Micropterus dolpmieui) and rock bass (Ambloplites
rupestris) (Adams and Olver 1977, Rahel and Magnuson 1983).

     As with the other faunal groups, the low productivity and biotic
interactions (predation/competition)  of these lakes probably have a
bigger influence on the species composition than pH per se.  For
example, many small-bodied fish species (e.g., minnows and darters) are
commonly absent from oligotrophic lakes even though they can tolerate
the pH's typical  of these waters (Rahel and Magnuson 1983).
Competition, or more likely predation by larger species, may exclude
these fish from biologically unproductive lakes where there are few
macrophytes to provide refuges.  Another example involves yellow perch
and whitefish (Coregonus spp.)  which  only successfully coexist in large,
cold lakes where the pelagic whitefish can avoid competition from the
more littoral-based yellow perch (Svardson 1976).

5.2.5  Summary

     Naturally acidic  waters provide  insight into the lowest pH
tolerated by various groups of aquatic organisms (Table 5.1).  While
life has been found in the most acidic environments sampled, the general
observation is that species diversity declines as pH decreases.  The
most tolerant organisms  are from the  lower trophic levels, with some
bacteria and algae able  to flourish at pH's below 1.0.  Invertebrates
are rarely found  below pH 3.0,  and fish are generally limited to pH's
above 4.0.  Some  organisms (especially certain genera of bacteria) are
true acidophiles,  unable to grow and  reproduce at neutral  pH (Brock
1978).   However,  most organisms occurring in acidic environments survive
                                  5-9

-------
    TABLE 5-1.   LOWER pH  LIMITS  FOR  DIFFERENT GROUPS OF ORGANISMS IN
                              NATURALLY ACIDIC WATERS

Group


Bacteria


Approx.
Lower
PH
limit
0.8

2-3

Examples of Species
Occurring at Lower pH Limit

Thiobacillus thiooxidans,
Suifolobus acidocaldarius
Bacillus, Streptomyces

Reference


Brock 1978

Brock 1978
Plants
  Fungi
  Eucaryotic
   algae
 Blue-green
  algae
 Vascular
  plants
  Mosses

Animals
  Protozoa
  Rotifers

  Cl adocera
  Cope pods

  Insects
  Amphipods

  Clams

  Snails

  Fish
0

0
1-2
4.0

2.5-3


3.0
2.0
3.0
3.5
3.0
3.0
3.6
2.0
3.0
5.8

5.8

4.5
6.0
5.8
6.2
3.5
4.0
4.5
Aconti urn velatum

Cyanidium caldarium
Euglena mutabl1 i s,
Chlamydpmonas acfdophila.
Chi orel la
Eleocharis, Carex,
Ericacean plants,
Phragmites
Sphagnum
Amoebae, Heliozoans
Brachionus, Lecane, Bdelloid
Col1otheca, Ptygura
SimocephaTus, Chydorus
Mac rocy clops
Cyclops
Ephydra thermophila
Chironomus riparius
Mayflies

Gammarus

Pi si di urn
Most other species
Amnicola
Most other species
Tribolodon hakonensis
Umbra limi
Sunfishes (Centrarchi dae)
Brock 1978

Brock 1978
Brock 1978
Mastigocladus, Synechococcus   Brock 1978
Brock 1978
Hargreaves et al. 1975
Ueno 1958
Brock 1978
Brock 1978
Hutchinson et al. 1978
Edmondson 1944
Ueno 1958
Ueno 1958
Hutchinson et al. 1978
Brock 1978
Hutchinson et al. 1978
Sutcliffe and Carrick
 1973
Sutcliffe and Carrick
 1973
Griffiths 1973
Pennak 1978
Pennak 1978
Pennak 1978
Mashiko et al.  1973
Rahel and Magnuson 1983
Rahel and Magnuson 1983
                                  5-10

-------
quite well at neutral pH but are excluded from such environments by
competitively superior species.

     Species distributions in natural pH gradients provide a means of
assessing the long-term effects of low pH exposure, integrated over all
life history stages and all physiological functions.  Such information
is  seldom obtained in laboratory bioassays,  which are generally short-
term, focused on one or two physiological responses, and ignore the
potential for genetic adaptation to acid stress.  Species' acid
sensitivity inferred from distributions in naturally acidic waters may
be  useful in selecting species to monitor in waters undergoing cultural
acidification.  For example, acid tolerance  rankings of fish species,
based on distributions among naturally acidic Wisconsin lakes (Figure
5-1), were correlated with acid tolerance rankings from culturally
acidified Canadian lakes (Figure 5-2).  This allowed predictions of
which fish species should be monitored in Wisconsin lakes susceptible  to
acidification (Rahel  and Magnuson 1983).

     Studies of species distributions relative to pH are subject to
misinterpretation if other correlated factors are not adequately
considered.  Among the factors that can interact to influence species
distributions are pH, metal concentrations and temperature in geothermal
waters; pH, oxygen concentrations,  and substrate composition in
dystrophic waters; and pH, low nutrient concentrations, and predation  in
ultra-oligotrophic waters. The problem of separating out the effects of
confounded factors is illustrated by work on the distribution of
rotifers in Wisconsin lakes.  Alkaline waters (above pH 7.0)  contained
relatively few species of rotifers  but large numbers of individuals.   In
contrast, acidic waters (below pH 7.0) contained large numbers of
species but few individuals (Pennak 1978).   Hence, rotifer species
diversity increased with decreasing pH.   However,  this was probably
because competitive interactions were influenced by factors correlated
with pH, not because most species of rotifers could not tolerate neutral
pH.  In another example,  Weiner and Hanneman (1982)  failed to find a
relationship between reduced fish growth and low pH in a set of
naturally acidic Wisconsin lakes, even though growth reductions at low
pH are consistently observed in laboratory bioassays (Section
5.6.4.1.3).   They attributed the lack of correlation between  fish growth
and pH to the overriding effects of population density.

     Experimental manipulations offer potential  for separating the
effects of these confounding factors from the effects of pH.   A good
example is the alkalinization of an acidic brownwater lake (Smith 1957).
When the pH  was raised by adding lime, several  stocked fish species
reproduced successfully  for the first time.   However,  as the  pH returned
to its former level,  reproduction stopped, indicating that hydrogen ion
concentration was the limiting factor.

     In some cases,  naturally acidic environments are free of the
confounding  stresses  associated with culturally  acidified environments.
This is especially true  of metal  toxicants,  which are common  in waters
affected by acidic mine  drainage or acidic precipitation (Parsons 1977,


                                 5-11

-------
      COMMON NAME

      CENTRAL MUDMINNOW
      YELLOW PERCH
      BLACK BULLHEAD
      BLUEGILL
      LARGEMOUTH BASS
      WHITE SUCKER
      YELLOW BULLHEAD
      PUMRKINSEED
      GOLDEN SHINER
      NORTHERN REDBELLY
      BROOK STICKLEBACK
      NORTHERN PIKE
      WALLEYE
      ROCK BASS
      MOTTLED SCULPIN
      SMALLMOUTH BASS
      MUSKEILUNGE
      BLACK CRAPPIE
      BURBOT
      CREEK CHUB
      CISCO
      IOWA DARTER
      JOHNNY DARTER
      REDHORSE
      COMMON SHINER
      MIMIC SHINER
      TROUT- PERCH
      BLUNTNOSE  MINNOW
       LOGPERCH
       BLACKNOSE  SHINER
       FATHEAD MINNOW
NUMBER OF LAKES IN A GIVEN pH RANGE
        (50  lakes >  70)
FAMILY 7.


P

Ce
Ce
Ca

Ce
Cy
s*
E
P
Ce
Co
Ce
E
Ce
Ga
iy
P
P
Ca
Cy
Cy
Pe
V
Cy
Cy
pH RANGE NUMBER OF
0 6.0 5.0 4.0 LAKES
.ii.



























, 1 . 1 L
1 1 1 ._ 1 . 1 ' -
50
114
50
84
80
83
35
78
63
13
10
65
51
56
29
44
40
59
25
13
16
24
37
23
31
22
10
45
15
15
7

23 ' 17  17  15
 Figure 5-1.  The distribution of 31 fish species in relation to pH for
             138 northern Wisconsin lakes.  Family names are abbreviated
             as follows:  Catostomidae (Ca), Centrarchidae (Ce), Cyprinidae
             (Cy), Esocidae (E), Gadidae (Ga), Gasterosteidae (G),
             Ictaluridae (I), Percidae (P), Percopsidae (Pe), Salmonidae
             (S), Umbridae (U).  Adapted from Rahel and Magnuson (1983).
                                  5-12

-------
             V
            _*
             <0
            XJ
             o>
             
-------
Cronan and Schofield 1979)  but rare in acidic brownwater and  ultra-
oligotrophic lakes.   As a result of organic  complexation,  comparison  of
fish species distributions relative to pH  in these different  water types
has helped to identify aluminum toxicity,  not pH,  as  the major  reason
for spring fish kills in lakes affected by acidic  precipitation (Muniz
and Leivestad 1980a).

     Data on the biota of naturally acidic environments  will  continue to
be instructive in studies of culturally acidified  waters and  should be
especially useful in evaluating the long-term effects of chronic  acid
stress.

This section is summarized as follows:

1.   Naturally acidic lakes fall into three major  groups:

    0 inorganic acidotrophic waters {pH commonly less than 4.0)

    0 dystrophic waters (pH commonly 3.5 to 5.0)

    0 ultra-oligotrophic waters (pH commonly 5.5 to 7.0)

2.   In naturally acidic waters, hydrogen ion concentration can be
     strongly implicated as limiting the occurrence of:

    *> invertebrates with calcified exoskeletons below pH 5.5
        (mayflies, Gammarus, snails, clams)

    0 blue-green algae below pH 4.0

    0 some species of minnows (Cyprinidae) and darters (Percindae)
        below pH 6.0

    o several species of sunfish (Centrarchidae) below pH 4.5

These pH limits for survival and reproduction are  similar to  those
observed in culturally acidified waters.

3.   Lower safe pH limits inferred from a species  distribution  among
     naturally acidic waters may not always be valid for culturally
     acidified waters.  For example, these limits  may be:

    0 too low if other stresses (e.g., aluminum) are present  in
      culturally acidified lakes, or

    o too high if species are absent from naturally acidic lakes
      because of factors other than low pH:  e.g., high temperature
      or metals in inorganic acidotrophic waters;  low sodium, and
      calcium concentrations or unsuitable habitat in dystrophic
      waters.
                                  5-14

-------
5.3  BENTHIC ORGANISMS (R. Singer)

5.3.1  Importance of the Benthic Community

     The term benthos refers to the community of organisms which  live in
and on bottom sediments of lakes and streams.  The following groups  are
important components of the benthos:  microbes, periphyton,
macroinvertebrates, Crustacea, Insecta,  Mollusca, and Annelida (Table
5-3).  These organisms interact with biological and chemical components
of the water column by processing detritus,  recycling inorganic
nutrients, mixing sediments, and serving as  a principal  food source  for
fish, waterfowl, and riparian mammals.   Most of the energy and nutrients
in lakes and streams ultimately passes through the benthos,  so any
alteration of this community is likely  to affect plankton, fish,  and
water chemistry.  Studies of the effects of  acidic precipitation  on  this
community have begun only recently (Singer 1981a), and not all  benthic
components have received equal treatment.

      Microbes rapidly colonize the surfaces of leaf litter and other
organic debris.  Many benthic macroinvertebrates then process the
debris, further facilitating its decomposition by microorganisms.
Macroinvertebrate "shredders" rip and chew leaves, vastly increasing
surface area, and partially digest material  as it passes through  their
guts.  Without these invertebrates, organic  detritus decomposes very
slowly (Brinkhurst 1974).

     After the macroinvertebrates have broken up the detritus,  fungi,
bacteria, and protozoans complete the digestion and release inorganic
nutrients Into the water.  The pH of the water in part controls the
solubility equilibria of these inorganic constitutents and largely
determines whether they will be available for recycling by plants.   In
addition, the rate of decay depends on  the metabolic efficiency of this
microbial community, which is also pH dependent (Laake 1976, Gahnstrom
et al. 1980).

     Macroinvertebrates aerate sediments by  their burrowing movements.
The top few centimeters of sediments generally demonstrate large
gradients of pH, Eh (oxidation-reduction potential--the concentration of
free electrons), dissolved 02, and other constituents (Hutchinson
1957).  Losses or alterations of plant and animal communities have
profound effects on the chemistry of this top layer of sediments
(Mortimer's "oxidized microzone" 1941,  1942), yet little work has
centered on this habitat in acidified lakes.  Mitchell et al. (1981b)
found that the presence of burrowing mayflies (Hexagenia) affected
sulfur dynamics in sediment cores taken from acidic lakes.
Sediment/water column biological and chemical interactions are difficult
to study because events occur across strong  chemical gradients over
short distances (Mitchell et al. 1981b).  These gradients are easily
perturbed by experimental procedures, including in situ measurements.
Despite these procedural  difficulties,  it is important to determine  the
influence of pH-related alterations of the sediment community on  the
chemistry and biota of the water column.
                                  5-15

-------
     Benthic animals are at the base of most food chains that lead to
game fish.  It has been suggested  that  eliminating the amphipod Gammarus
lacustris (Section 5.3.2.4)  and most molluscs (Section 5.4.2.6) might
reduce trout production by 10 to 30  percent  (0kland and 0kland
1980); however this prediction has not  been  verified.  Rosseland et al.
(1980) reported that trout in acidified waters  shifted their diet from
acid-sensitive invertebrates like  mayflies and  bivalves to acid-
tolerant forms such as corixid bugs  and beetles.  Although decline of
fish populations due to alteration of the benthic community has not been
studied, stress on fish populations  as  a result of nutrient changes
should be considered.  Fish fry, which  are more dependent on smaller
invertebrate prey than are adults, might be  more sensitive to changes in
the benthic community.  These effects have not  been considered
experimentally, however.

      Finally, changes in the benthic plant  community (Section 5.5)
affect macroinvertebrate distribution.   The  littoral habitat is an
important area for benthos,  and alterations  in  plant community
structures are likely to affect all  other trophic levels.  These
interactions remain to be investigated,  but  Eriksson et al. (1980b) have
suggested that many of the observed  changes  in  water chemistry and
plankton communities are due to biological alterations, not direct
chemical toxicology.  They reported  an  increase in clarity, alteration
of planktonic communities, and even  a drop in pH (by 0.5 units) when
fish were eliminated from a neutral  lake by  poisoning.  The results
extend and verify similar work reported by Stenson et al. (1978).

     Sources of energy to benthos  include primary production by higher
plants (macrophytes) and attached  algae (periphyton), and energy derived
from detritus raining from the water column  above (autochthonous inputs)
and from detritus washed into the  basin (allochthonous inputs).  Lakes
(lentic systems) receive most of their  energy from autochthonous
sources, whereas streams (lotic systems) derive their energy from
primarily outside, allochthonous sources (e.g., Wetzel 1975).
Consequently, shredding and scraping benthic insects and crustaceans are
relatively more important in streams than lakes, while detritus-
consuming worms and midges are more  abundant in lakes.

5.3.2  Effects of Acidification on Components of the Benthos

     The diversity of benthic organisms is often confusing to non-
specialists.  It must be emphasized  that the loss of fish populations is
one of the last biological effects of acidification, and alterations in
the benthic community integrate annual  loadings at levels of stress
which are not observable in fish populations.   The ultra-oligotrophic
lakes characteristic of sensitive  areas harbor  ecosystems which are
unique.  These ecosystems may be damaged at  levels of acidification (pH
< 6.5) that may not affect fish at all.  The concept of an endangered
ecosystem is as viable as the more generally accepted view of the
endangered species.
                                 5-16

-------
     Using historical collections and known  water  quality  requirements
of organisms allows specialists to generalize  about past water chemistry
parameters.  Moreover, the low mobility and  long life cycles of many
benthic organisms allow one to make conclusions about the  extremes of
water quality fluctuations in past years.

5.3.2.1  Microbial  Community—Studies of the effects of acidification on
benthic protozoans have not been conducted.  Other members of this
community include bacteria and fungi.  It was  reported that
acidification of lakes causes bacterial  decomposers to be  replaced by
fungi (Hendrey et al. 1976, Hendrey and Barvenik 1978) and proposed
(Grahn 1976, 1977;  Hultberg and Grahn 1976)  that the shift to fungi
accounts for the observed (Leivestad et al.  1976)  accumulation of
detritus in acidic lakes.   Liming of lakes to  increase the pH brings a
rapid restoration of normal microbial activity (Scheider et al. 1975,
1976; Gahnstrom et al. 1980).

     Traaen (1976,  1977) showed that leaf packs in lakes were processed
much more slowly at lower pH (5.0) than at higher  pH (6.0) values, but
he also cautioned (1977) that many other factors besides acidity can
affect leaf processing.  Burton (1982)  has confirmed the impact of low
pH on processing of organic matter.  Friberg et al. (1980) reported an
increased accumulation of detritus and a reduction in numbers of
scrapping insects in an acidic (pH 4.3 to 5.9) as  compared to a neutral
(pH 6.5 to 7.3) stream.  Hall et al. (1980)  and Hall and Likens
(1980a,b) artificially acidified a stream in Hubbard Brook, NH, and
showed that scrapers were largely lost.   In  addition, they reported that
insects that feed by collecting debris were  inhibited.

     Hall et al. (1980) observed a growth of basidiomycete fungus on
birch leaves in an artificially acidified portion  of a stream; such
fungal growth was lacking in the non-acidified control section.
Hultberg and Grahn (1976)  and Grahn et al. (1974)  described an
accumulation of a "fungal  mat" on the bottom of many acidified
Scandinavian lakes.  It is now understood that this coarse particulate
material is a mixture of detritus, some fungi, and mostly  algae (Stokes
1981) (Section 5.3.2.2).  The original  description of this layer of
material as a "fungal mat" (Hendrey et al. 1976) was erroneous (Hendrey
and Verticci 1980)  due to the senescent,  colorless state of the common
blue-green algal  (Phormidium spp.) component of the mat.

     Some controversy exists regarding the effects on microbial
metabolism brought about by acidification (Baath et al. 1979).  The
accumulation of detritus in acidic lakes suggests  a reduction in
decomposition by bacteria (Leivestad et al.  1976).  The reduction of
oxygen utilization by acidified cores (Hendrey et  al. 1976) supports
this view.  Furthermore, liming increased oxygen consumption of
previously acidic cores (Gahnstrom et al.  1980).   At pH levels below
5.0, oxygen consumption, ammonia oxidation,  peptone decomposition, and
total bacterial numbers all declined (Bick and Drews 1973).  In
contrast, Schindler (1980) reported no change  in decomposition rates in
an artificially acidified lake, and Traaen (1978)  observed no clear


                                  5-17

-------
changes in the planktonic bacterial  populations  from  seven  lakes of pH <
5.0 as compared to seven lakes  of pH >  5.0.   Traaen argued  that acidic
inputs should affect the plankton populations prior to  affecting benthic
algae.  His results showed that the  distribution of bacterial
populations was more strongly influenced by  organic inputs  and temporal
and spatial (depth) patchiness  than  by  pH.   Gahnstrom et  al.  (1980)
reported that inhibition of oxygen uptake by sediments  increased in
acidic lakes as compared to reference lakes  only in the littoral
sediments.  They argued that the inhibition  of microbial  activity  in  the
littoral  zone might be due to the inflow of  acidic runoff,  which is
restricted to the epilimnion during  snowmelt and autumn rains (Hendrey
et al. 1980a).  All these studies demonstrate that decomposition of
organic material is inhibited below  pH  5.0 but not necessarily by  a
reduction in standing crop of bacteria.   The resulting  accumulation of
organic matter undoubtedly affects water chemistry, fish  habitats,
nutrient cycling, and primary productivity.

     Microbial effects on other trophic  systems  probably  involve
alterations of sulfur, nitrogen,  and phosphorus  dynamics.   Methylation
of mercury (Tomlinson 1978, Jernelov 1980) and other heavy  metals  may
have profound effects on higher trophic  levels (Galloway  and Likens
1979; refer also to Chapter E-6). The  release of aluminum  from
sediments below pH 5.0 (Driscoll  1980)  is another potentially serious
impact that has not been adequately  studied.

5.3.2.2  Periphyton--The periphytic  community of algae  lives attached to
macrophytes and directly on sediments and makes  important contributions
to primary production and nutrient cycling,  particularly  in lotic
(stream)  systems.  Changes in the species composition of  this community
reflect changes in the chemistry of  both the water column and the
sediments.  These algae are an  important food source  for  the grazing
macroinvertebrates which are a  principal  source  of food for fish.  Algal
seasonal  growth and decomposition store  and  periodically  release
nutrients and other ions.

5.3.2.2.1  Field surveys.  Acidic lakes  develop  periphytic  communities
dominated by species known to prefer acidic  water, and  dramatic
decreases in species diversity  below pH  5.5  have been observed (Aimer et
al. 1974; see Section 5.2).  One of  the most striking aspects of many
acidified lakes is the presence of a thick mat of algae which overlies
the substrate.  This mat overgrows all  the rooted plants  and, to a large
degree, physically and chemically isolates the lake bottom  from the
overlying water.  The mat varies in  shape, texture, and species
composition from lake to lake,  seemingly irrespective of  water chemistry
parameters.  Three types of mats were described  by Stokes (1981):

    1)  Cyanophycean mats, dominated by  the  blue-green  algae
        pscillatoria sp., Lyngbya sp.,  and Pseudoanabaena sp. in
        Sweden at pH 4.3 to 4./ Uazarek 1980) and Phormidium sp.
        in New York at pH 4.8 to 5.1 (Hendrey and Vertucci  1980).
        These mats are dark blue-green  with  occasional  flecks of
        orange-colored carotene-rich material.  They are  thick,


                                  5-18

-------
        felt-like,  and  encrusting.  Stokes reported cyanophycean
        mats at depths  of 2 to 3 m, but they have been observed as
        deep as 5 m in  an acidic (< pH 4.9) Adirondack lake (Singer
        et  al. 1983).

    2)   Chlorophycean mats, dominated by green algae like Mougeotia
        sp. and Pleurodiscus  sp. at pH 3.9 to 5.0 in Canadian lakes
        (Stokes 1981).  These mats are coarser than cyanophycean
        mats.  They tend to be loosely packed, green to reddish
        purple, and may extend to 4 m deep.  Unlike cyanophycean
        mats,  chlorophycean mats are not compacted and do not
        retain their structural integrity when lifted.  A
        chlorophycean mat developed after the experimental
        acidification of a whole lake was completed (Schindler and
        Turner 1982).

    3)   Chlorophycean epiphytic or periphytic algae dominated by
        green  algae such as Spirogyra sp., Zygnema sp.,
        Pleurodiscus sp., and Mougeotla sp., Oedogonium, and
        BulbochaeteT This community appears as bright grass-green
        clouds hanging  from macrophytes and resting lightly on the
        bottom.   They appear  around pH 5.0 and have been reported
        in  Canada (Stokes 1981), the Adirondacks  (Hendrey and
        Vertucci  1980), and Sweden (Lazarek 1982).  They also
        appeared  in artificially acidified channels (Hendrey 1976),
        artificially acidified cylinders (Muller  1980, Van and
        Stokes 1978), and in  an artificially acidified lake at pH
        5.6 (Schindler  1980).

     I  have observed all three types of mat communities in a survey of
five Adirondack  lakes below pH 4.9.  These lakes  were all about the same
size (~ 30  ha), low in  nutrients, located near each other, and similar
in morphometry.   Why one community dominates one  lake but is not found
in another  is  unknown.  Part  of the explanation may be that the three
types of mats  may represent stages in a pattern of seasonal succession.
Lazarek (1982) has  reported seasonal succession among epiphytes from one
acidic (pH  4.3 to 4.7)  Swedish lake.  As these mats are the most
conspicuously  visible characteristics of acidified lakes, their
significance  and  effects on other physical and chemical components
deserve more attention.

5.3.2.2.2  Temporal trends.   The shells of diatoms (Bacillariophyceae)
are made of Si02  and are very resistant to weathering.  Deposition of
planktonic  and benthic  diatoms to sediments produces  a record of the
past populations  in the lake  once the cores are dated by  radioactive
decay (Norton  and Hess  1980). The pH tolerance of many diatoms has been
tabulated elsewhere (e.g., Lowe 1974).  Thus the  ancestral pH may be
inferred from  the stratigraphic record.  This technique is subject to
variances caused by macroinvertebrate mixing, local changes in  pH
sensitivity of species, and the numerous other factors besides  pH that
determine the  distribution of species  (Norton et  al.  1981).
Nonetheless,  inferred  pH generates a value that reflects  the real water


                                 5-19

-------
column value to within <  1.0  pH  unit,  which  often lies within the range
of normal  seasonal  pH variation.   The  method's accuracy is even better
for comparing groups of lakes with similar current pH values along a
temporal  gradient of past pH  levels  by regression analysis (Norton et
al. 1981).  The inferred pH values calculated  from diatom stratigraphy
related very well  to the  values  estimated  from using the shells of
cladoceran remains (Norton et al.  1981).

     Berge (1976)  compared the diatom  assemblages in sediments from
seven Norwegian sites with the communities from the same sites as
reported in 1949 and found no quantitative change in the diatoms in the
26-year period.  However, he  noted a marked  shift towards species that
required or preferred low pH.  In  an even  longer period (ca. 1920-1978)
Aimer et al. (1974)  reported  a reduction in  diatoms from cores taken
from Scandinavian lakes which have become  more acidic.  Dam et al.
(1980) reported a more obvious shift towards acid-tolerant diatoms in
sediments from acidic Dutch lakes.

     Three hundred years of diatom deposition  in sediments was used to
calculate pH values in two Norwegian lakes (Davis and Berge 1980).  The
pH tolerance of diatoms was determined from  present-day distributions,
and the pH in the past was inferred  from the species composition in the
dated sediment layers.  One lake has remained  constant at ~ pH 5.0
while the other went from pH  5.1 to  4.4 since  1918 (Davis et al. 1983).

     More recently (Davis et  al.  1983), results of sediment core
analyses from nine Norwegian  lakes and six New England lakes were
compared.   The range of pH tolerance of the  diatoms was determined by
studying current distributions in 36 Norwegian and 31 New England lakes.
The three Norwegian Lakes which  are  currently  acidic (pH < 5.0) have
decreased in pH by 0.6 to 0.8 units  since  1890-1927.  The lakes
currently above pH 5.0 have decreased  0 to 0.3 pH units since 1850.  All
six of the New England lakes  decreased 0.2 to 0.4 units and some of
these changes might be due to land use changes  (reforestation) which are
in the historical  record. Another anomaly was the record of heavy metal
pollutants in the sediments several  decades  prior to the changes in the
diatom communities.  This was ascribed to  the  buffering of the
watershed, which released metals while retaining protons for many years,
thus keeping the lake pH stable, or alternatively, to the former high
emissions of neutralizing particulates like  fly ash.

     An interesting change in the diatom community structure is also
apparent from an analysis of  the data  (Berge 1976, Dam et al. 1980,
Davis and Berge 1980, Norton  et  al.  1981,  Davis et al. 1983).  The
species of diatoms which indicate acidic  (pH < 5.0) conditions are
primarily benthic, whereas those from  circum-neutral (pH 6.0 to 7.2) are
planktonic.  This implies that the diatom  community shifts to benthic
production in acidic lakes.   Diatoms are common but not dominant members
of the algal mats of present-day acidic lakes  (Stokes 1981).

     Del Prete and Schofield  (1981)  used  sediment cores to study the
succession of diatom species  in  three  Adirondack lakes.  They observed


                                  5-20

-------
an increase in dominance by acid-tolerant species  in  the most
acid-impacted lakes.   A trend  towards  species  tolerant of low nutrient
waters was also reported.  label!aria  fenestrata and  Cyclotella
stelligera increased  in numbers most directly  with increasing acidity,
although some of the  results were equivocal.

     Coesel et al. (1978) have compared the desmid populations from  a
group of lakes in the Netherlands  with community compositions reported
in studies done in 1916-25, 1950-55, and with  their own  survey in  1977.
Many of the species from the rich  flora in the earliest  survey were  lost
due to cultural eutrophication.  In the most recent survey,  those  ponds
that were not impacted by nutrient additions were  affected by acidic
deposition, as reflected by the paucity of desmid  species.   These  ponds
appeared to have undergone oligotrophication.   The eutrophic ponds
remained well-buffered and unchanged.   Thus, the effects on  community
composition brought on by cultural  eutrophication  can be separated from
the changes caused by acidification.

     These studies of temporal trends  demonstrate  that many  acidic lakes
have become acidic in historic times,  but they do  not prove  that this
acidification is universally a consequence of  atmospheric  deposition.
Deforestation, followed by eutrophication and  reforestation, can cause
the pH of a lake to rise and then fall.  Even  so,  some lakes have  fallen
about 0.5 units in locally unperturbed watersheds  in  historic times.

5.3.2.2.3  Experimental studies.  Muller (1980) studied  the  succession
of periphyton in artificially acidified chambers  held in situ in Lake
223, Experimental Lakes Area, in northwestern  Ontario (Schindler et  al.
1980b).  At the control pH of 6.25, a  succession  occurred  in the
chambers from dominance by diatoms in  the spring  to dominance by green
algae (Chlorophyta) in mid-July.  In enclosures at pH <  6.0, Chlorophyta
dominated the periphyton throughout the sampling  period.   Blue-green
algae (Cyanophyta) were reduced and almost eliminated under  the most
acidic conditions.  Muller observed no trend with  respect  to changes in
biomass but noted a sharp decrease in  species diversity  (as  measured by
Hill's index] in the acidified (pH 4.0) chambers.   Changes in primary
production (14C) showed no trend with  pH.  The dominance of  the
periphyton by Chlorophyta in the acidified samples was  due almost
entirely to the growth of Mougeotia sp., which by  June represented 96
percent of the biomass and cell numbers at pH 4.0.  This taxon was
responsible for less than 4 percent of the biomass and cell  numbers  in
the natural lake water.  Interestingly, during May, the  blue-green alga,
Anabaena sp., rose from 3.4 percent of the biomass in the  lake water (pH
6.2) to 4.3 percent at pH 4.0, but this species was almost absent  by
June.  In  spite of its low biomass this alga accounted for 25 percent
and 41 percent of the total cell numbers in these two samples.  Muller's
(1980) work demonstrates the need to consider natural seasonal  patterns
of  succession when we superimpose the effects of acidification on
aquatic ecosystems.  The only  other report of seasonal  changes  in
periphyton (Lazarek 1982) dealt with algae living attached to Lobelia
dortmanna  and verified the succession from diatoms to green  algae
(Mougeotia spp.) during the growing season.


                                  5-21

-------
    Higher standing crops but lower  rates  of C-fixation per unit
chlorophyll  occurred in  periphyton growing in artificial stream channels
at reduced pH (Hendrey 1976).  The total rate of 14C-uptake was
similar over a wide range of  [H+].   Increased standing crop was
attributed to a combination of three mechanisms:   1) enhanced growth by
acid-tolerant taxa, 2) reduction  in  grazing by  the reduced
macroinvertebrate population,  and 3)  inhibition of microbial
decomposition (Hendrey 1976).

     In an artificially  acidified section  of a  softwater stream in New
Hampshire, Hall et al. (1980)  reported an  increase in periphyton numbers
and substrate chlorophyll £ concentration. They did not perform a
taxonomic analysis of the periphyton community.
                                                                  *
     Periphyton communities respond  to acidification by alterations in
species composition, increases in the standing  crop, decreases in the
amount of growth per unit of  biomass, and  formation of atypical mats
which cover the substrate. These changes  produce  dramatic, visually
obvious changes in lakes and  streams at pH < 5.0.

5.3.2.3  Mi'croinvertebrates--The  responses of several minor groups of
invertebrates to acidification have  been studied.  The Nematoda and
Gastrotricha are both common  but  poorly studied inhabitants of
interstitial water in sediments (meiofauna).  They feed on detritus and
other organic material lying  between the grains of sand in sediments.
The ubiquitous meiobenthic gastrotrich, Lepidodermella squammata, was
almost totally eliminated under laboratory conditions below pH 6.4
(Faucon and Hummon 1976).  Unfortunately,  the pH gradient was achieved
by mixing unpolluted creek water  with water from a stream receiving
acidic strip mine drainage, so it is not easy to generalize to streams
receiving acidic deposition.   Hummon and Hummon (1979) added CaC03 to
the acidic mine drainage and  showed  that at the same pH, water with more
carbonate (COa?-) ameliorated the deleterious effects of acid
stress.  The extreme sensitivity  of  these  animals  to some component of
the acidic water, possibly low 003?- or high concentrations of metal
ions, bears further investigation.   Roundworms  (Nematoda) normally have
a ubiquitous distribution (Ferris et al. 1976). However, in an
extensive survey of Norwegian lakes, sub-littoral  sediments of acidic
lakes had a scarcity of roundworms when compared to  shallow sediments
from the same lakes (Raddum 1976).   No other mention is made of the
Nematoda in the literature pertaining to the acidification of aquatic
systems.

     Freshwater sponges (Porifera) are epifaunal and directly exposed to
changes in water chemistry alterations. However,  their response to
acidic deposition has not been studied. Jewell (1939)  studied the
distribution of Spongillidae  from 63 lakes, bogs and rivers in Wisconsin
with various levels of hardness and  pH. She found that most of the
species did have limited ranges  of Ca2+ concentrations  in which they
flourished.  Six common species were exposed to chemically modified
water, and growth was observed.   The lowest pH  in  this  experiment was
5.9, but there were indications that the most  important parameter was
                                  5-22

-------
the availability of Ca(HC03)2.   As filter feeders,  sponges  are
important reprocessers of suspended organic  matter  and  are  particularly
useful indicators of water quality because of the large volume  of water
which passes through their tissues.

     Aquatic mites (Acarina)  are not generally collected in  surveys of
benthic fauna, but Raddum (1976) noted that  mites occurred  in great
abundance in the shallow water  of an acid-impacted  lak-e.  At a  depth of
0.5 m, mites were third in abundance after nematodes  and midges
(Chironomidae).  At depths >  2  m almost no mites  were observed.  The
shallow mites probably receive  their nutrition from the shore or the
water surface, rather than the  lake substrate.  In  contrast, Wiederholm
and Eriksson (1977) observed mites in deep water  (> 10  m) in an acidic
lake in Sweden, and Collins et  al. (1981)  reported  no differences
between the distribution of mites in acidic  and control  lakes.  Clearly,
much work needs to be performed on the distribution of  this  group to
obtain a more complete understanding of how  acidic  precipitation affects
their distribution.

5.3.2.4  Crustacea--Benthic crustaceans include familiar large  forms
like crayfish (Decapoda), sow bugs (Isopoda), and scuds (Amphipoda), but
also smaller forms such as benthic copepods, mysids,  cladocerans, and
other branchiopods (e.g., Lepldurus).  All these  forms,  whether large or
small, contribute to the ecosystem dynamics  by feeding  on detritus or on
smaller detritivores and thus converting the organic  material into a
form palatable to fish and other carnivores.

     The distribution and characteristics  of habitats containing the
isopod Asellus aquaticus (aquatic sow bug) and the  amphi pod  Gammarus
lacustris (scud) were summarized by K.  0kland (1979a, 1980a)~Both of
these species are important as  food for fishes and  as detritus
processors.   A_. aquaticus populations were reduced  below pH  5.2 and
absent below pH of 4.8.While  G. lacustris  was able  to out-compete A.
aquaticus at pH 7.0, Asellus  ouT-competed  Gammarus  at sites  stressedTby
either acidic inputs or organic enrichment.   A. aquaticus was widely
distributed in acid-stressed  lakes at pH 5.0 TK.  0kland 1980b)  but £.
lacustris was inhibited below pH 6.0 (K. 0kland 1980c)  probably due to
the low calcium concentration in the acidic  water.

     In the laboratory, Gammarus pulex demonstrated no  avoidance of pH
6.4 to 9.6 (Costa 1967).  However, within  12 to 15  minutes  after the pH
was lowered to 6.2 in one part  of the tank,  the amphipods began to stay
near the alkaline side.  Immature Gammarus performed  this avoidance
behavior faster than did adults.

     Sutcliffe and Carrick (1973) verified that in  England £. pulex is
not normally found below pH 6.0, but they  pointed out that  it was found
in France at pH 4.5 to 6.0.  They suggested  that  the  avoidance  response
(Costa 1967) might explain its  limitation  to near-neutral water, instead
of direct mortality due to low  pH.  Laboratory studies  (Borgstrom and
Hendrey 1976) suggest, however, that direct  mortality is important at pH
<_ 5.0.  G_. lacustris achieved 96 hr Tl-50 at  pH 7.26 in  Montana, but


                                  5-23

-------
populations from Utah withstood pH  5.7  in similar laboratory  bioassays
in hard (135 mg £-1 CaCOa in  Montana, 200 mg &-1  CaC03  in  Utah)  water
(Gaufin 1973).  A different species, G.  fossarum, from  Germany,  showed
no mortality at pH 6.0,  and had a 96 Tfr  TL-5Q of  ~ 4.7.   At pH 5.0,
30 percent of the laboratory  population  survived  for 10 days  (Matthias
1982).  K. 0kland (1980a) ascribed  these differences to the variable
sensitivity of different populations.

     Steigen and Raddum (1981)  noted  that A^. aquaticus  responded to
acidification by leaving the  water, so  they confined some  of  the animals
in wire-enclosed tubes.   The confined  individuals resorted to
cannibalism, but the increased  energetic demands  Steigen and  Raddum
measured caused by the H+ stress resulted in losses  of  total  caloric
value in the confined animals.   The unconfined specimens left the water
but returned to feed, sometimes cannibalistically, and  the survivors
gained in caloric content.  This behavioral  response may be the
mechanism by which Asellus can  tolerate  more acidity than  can Gammarus.

     The opossum shrimp, Mysis  relicta,  is a bottom-dwelling  crustacean
characteristic of deep water.  It enters the water column  at  night to
feed on plankton and, in turn,  provides  food for  fish (Pennak 1978).
When Experimental Lake 223 was  artificially acidified from pH 6.6 to
5.3, Mysis populations were eliminated  at ~ pH 5.9 (Schindler and
Turner 1982).

      Eggs of the tadpole shrimp, Lepidurus arcticus (Eubranchiopoda,
Notostraca) took longer to hatch and  the larvae matured more  slowly than
normal at pH < 5.5 than at pH values >  5.5 (Borgstrom and  Hendrey 1976).
At pH < 4.5, larvae of j^. arcticus  died  in two days  and eggs  never
hatched.  A survey from Sweden  (Borgstrom et al.  1976)  reported  that L_.
arcticus was not found below pH 6.1.

     Laboratory bioassays of the crustaceans Daphnia middendorffiana,
Diaptomus arcticus, Lepidurus arcticus and Branchinecta paUidpsa have
provided additional evidence (Havas and  Hutchinson 1982) of the
sensitivity of crustaceans to acid  stress.  Animals  collected from an
alkaline (pH 8.2) pond were exposed to  naturally  acidic water (pH 2.8)
from a nearby pond which received aerial deposition  from the  Smoking
Hills of the Canadian North West Territories. The acidic  water  was
amended with NaOH to provide  a  range of  pH treatments.   A  critical pH
was 4.5, at which mortality drastically  increased for all  of  the
individuals.  Mortality did not occur  in control  water  lacking heavy
metal contamination (Al, Mi,  Zn).  These authors  suggested that  their
critical pH of 4.5 was lower than that  reported  in other studies because
the water in the Smoking Hills  area is  higher in  total  conductivity (1.3
mho cm~l than that of other acidic  clear water systems  (Havas and
Hutchinson 1982).

     An increased abundance of benthic  cladocerans has  been reported
(Collins et al. 1981) from two  of three  acidic lakes studied  in
Ontario.
                                  5-24

-------
       Crayfish are very important components of the benthos as detrital
  processors and as food for larger game fish.  Species  of crayfish  show
  some variation in sensitivity to pH.  Malley (1980)  indicated that
  Orconectes virilis, in softwater of ~ 22 ymhos cm-1  conductivity
  and Ca2+ of 2.8 mg JT*, was stressed by pH < 5.5.   However,
  Cambarus sp. was reported (Warner 1971) in a stream  receiving acidic
  mine drainage at pH 4.6, Ca2+ of 12 mg £-1, and conductivity of 96
  ymhos cm-1.  Cambarus bartoni was found in three acidic  lakes (pH
  4.6 to 4.9, - 3 mg £-1 Ca^"1") and Orconectes propinquis  was
  collected in one of three acidic lakes (Collins et ai. 1981).  I have
  seen Orconectes spp. in two lakes of pH 4.8 and 5.0  in the Adirondacks.

       This apparent discrepancy in pH tolerances of various crayfish may
  not be entirely due to interspecific or inter-population differences.
  The crayfish Orconectes viri1is has difficulty recalcifying its
  exoskeleton after molting at pH < 5.5.  Uptake of 45ca2+ by crayfish
  stopped at pH 4.0 and was inhibited at pH 5.7 (Malley  1980).
  Infestation of this species by the parasitic protozoan Thelohonia  sp  and
  reduction in recruitment of young at pH 5.7 was also reported (Schindler
  and Turner 1982).  Hence the tolerance of Cambarus to  pH 4.6 from  an
  acidic mine drainage stream may be due to the higher Ca2+
  concentration in the stream compared to habitats affected by acidic
  deposition.  The ameliorative effect of cations is suggested by the
  inability of the crayfish Astacus pallipes to transport  22Na+ below
  pH 5.5 (Shaw 1960).  Stress Is a function of both low  pH levels and low
  calcium levels, and the responses to these stresses  undoubtedly vary
  between life cycle stages and species.

  5.3.2.5  Insecta--The importance of insects in lakes and streams is
  discussed in Section 5.2.  These animals are important ecologically but
  also, because their tolerance to various stresses is well known, they
  are important as water quality indicators.

       Studies of benthic insects exposed to acid stress include surveys,
  mostly from Europe and Canada, and some experimental manipulations.
  Survey work involves presence-absence data from which  tolerances have
  been assumed.   The general conclusion drawn from surveys of lakes  and
  streams (Sutcliffe and Carrick 1973; Conroy et al. 1976; Wright et al.
  1975, 1976; Hendrey and Wright 1976; Leivestad et al.  1976;  Wiederholm
  and Eriksson 1977; Raddum 1979; Friberg et al. 1980; Overrein et al.
  1980) is that species richness, diversity, and biomass are reduced with
  increasing acidity.  Because predation by fish is eliminated in some
  water and food should be abundant due to the accumulation of detritus
  (Grahn et al.  1974), one might suppose that insect biomass would
  increase.  However, acidity imposes stresses that are as severe as
  predation (Henrikson et al. 1980b), and the lack of  bacterial
  decomposition of detritus (Traaen 1976, 1977) may render the detritus
  unpalatable to insects (Hendrey 1976, Hendrey et al. 1976).

  5.3.2.5.1  Sensitivity of different groups.  The sensitivity of benthic
  insects to pH stress varies considerably among taxa  and  among different
                                    5-25
409-262 0-83-13

-------
life cycle stages (Gaufin 1973,  Raddum  and  Steigen 1981).  Responses are
physiological  and behavioral.

     Mayflies  seem to be particularly sensitive to acidic conditions.
Female mayfly  adults (Baetis)  did  not lay eggs on otherwise suitable
substrates in  water pH < 6.0,  although  three  different species were
found within 200 to 300 m in neutral brooks with similar substrates
(Sutcliffe and Carrick 1973).   The adult presumably can detect high
levels of acidity by dipping her abdomen into the water as she flies.
Besides Baetis, the common mayflies Ephemerella ignita, and Heptagem'a
lateral is were absent only from the acidic  region of the River Duddon,
England (Sutcliffe and Carrick 1973).   A Swedish survey (Nilssen 1980)
also found mayflies to be sensitive to  pH stress.  A plot of the number
of mayfly species vs pH of 35  lakes and 25  rivers indicated that the
number of species decreased logarithmically with decreasing pH.  Species
were lost in two groups;  one  group did not appear below pH 6.5, and
another decline in species numbers occurred below pH 4.5 (Borgstrom et
al. 1976, Leivestad et al. 1976).  In another  survey (Fiance 1978) the
distributional pattern of the  mayfly, Ephemerella funeral is, was studied
in the Hubbard Brook, NH,  watershed during  a  2-year period.  Nymphs were
absent from waters of pH < 5.5.  The 2-year life cycle of this mayfly
makes it particularly sensitive to irregular  episodic stresses, because
a single drop  in pH may eliminate  the insects for several years. In an
experimentally acidified section of a New Hampshire stream (pH 4.0),
mayfly (Epeorus) emergence was inhibited and  drift of nymphs increased
(Hall et al. 1980; Hall and Likens 1980a,b; Pratt and Hall 1981).  These
responses suggest that mayflies exhibit both  behavioral and
physiological  responses to acidity.

     Laboratory bioassays verified that mayflies were the most
acid-sensitive order of insects (Bell and Nebeker 1969, Bell 1971,
Harriman and Morrison 1980; Table  5-2). Exposing caged transplanted
insects to acidified river water showed that  mayflies could not survive
and would try  to leave in the  drift (Raddum 1979).

     In contrast, dragonflies  and  damselflies (Odonata) (Table 5-2) are
much more resistant to low pH  (Bell and Nebeker 1969, Bell 1971,
Borgstrom et al. 1976).  The dragonfly  nymph  Libel!ula pulchella
tolerated pH 1.0 for several hours (Stickney  1922).Dragonfly nymphs
(Anisoptera; Odonata) may be able  to endure episodic acidic stress by
closing their  anus, through which  they  respire, but this behavior has
not been investigated.  Dragonflies burrow  into sand and mud, turning
over material  and changing the structure of the habitat.  They are also
major predators on oligochaete worms, midges  (Chironomidae), and small
insects; they  are even known to feed on tadpoles and small fish (Needham
and Lloyd 1916).

     Tolerance to acidification within  the  Plecoptera (stoneflies) is
variable according to surveys  (Sutcliffe and  Carrick 1973, Leivestad et
al. 1976), field manipulations (Raddum  1979;  Hall and Likens 1980a,b)
and laboratory studies (Bell and Nebeker 1969, Bell 1971).  Stoneflies
and mayflies are preferred trout food in streams, as evidenced by the


                                  5-26

-------
  TABLE 5-2.  RESULTS OF LABORATORY STUDIES ON pH TOLERANCE  OF SELECTED
     INSECT NYMPHS.   TLso IS THE pH WHICH IS LETHAL  TO  50% OF THE
   ORGANISMS.  RESULTS OF DIFFERENT STUDIES ARE REPORTED  HERE AS THE
     NEGATIVE LOGARITHM OF THE AVERAGE HYDROGEN ION  CONCENTRATIONS
     Organisms
                               96 hr
                 PH
        Long-term  50% successful
          TL50       emergence   References3
Ephemeroptera
   Baetls sp.                   4.5
   Clnygniula par                6.11
   EphemereTIa doddsl           4.10
   Ephemere]fa grandls          3.6      5.8(48)
   EphemerelTa subvarfa         4.65     5.38(30)
   Heptagenia~sp.               6.17
   Hexagenla Umbata            5.66     5.5(33)b
   LeptppMebia sp.             5.20
   RhlthrogenaTrobusta          4.60
   StenonemaTubrurn             3.32

Odonata
   Boyerla vlnosa               3.25
   Ophtogomphus rtplnsulensis   3.50
Plecoptera
   Acroneun'a lycorias
   Acroneuri'a paclfi'ca
   Arcynopteryx parallel a
   Isogenus aestivails
   Isogenus frontalts
   Isoperla fulva
   Nemoura cinerea
   PterorTarcella badla
   Pteronarcys caHfornica
   Pteronarcy? dorsata
   Taem'opteryx maura
5.
3.
3.32
3.8
4.37
 .08
 .68
4.5
2.6
3.92
4.44
4.25
3.25
Trichoptera
   Hydropsyche betteni          3.15
   Hydropsycffe sp.              3.28
   Arctopsycfie" grandls          3.4
   Ltmnephtlus ornatu?          2.82
   BrachycenTrus americanus     1.50
   Brachycentrus occldentails
   Cheumatopsyche sp.
         4.42(30)
         4.30(30)
         3.85(30)
         5.8(90)
4.50(30)
         4.52(90)
         4.95(90)
         5.00(30)
         3.71(30)
         3.38(30)
         2.45(30)
         4.3(90)
         4.52(90)
                       5.9
                       5.2
                       5.2
              5.0
6.6
              5.8
              4.0
                       4.7
                       4.0
                                     M
                                     G
                                     G
                                     G
                                     B, BN
                                     G
                                     G
                                     G
                                     G
                                     B
                            B, BN
                            B, BN
B,BN
G
G
G
B, BN
G
M
G
G
B, BN
B, BN
                            B, BN
                            G
                            G
                            G
                            B, BN
                            G
                            G
                                 5-27

-------
                         TABLE 5-2.  CONTINUED
Organl sms
Diptera
Atherix variegata
Holorusia sp.
Simulium vlttatum

96 hr
TL50
2.8
3.63
pH
Long-term 50% successful
TLso emergence
4.2(68)
References3
G
G
G
aReferences:   B = Bell  1971,  BN  =  Bell and Nebeker 1969, G = Gaufin
 1973, M = Matthias 1982.
^Seventy of 90 survived.
                                  5-28

-------
attempts of fishermen to mimic these body  forms with  their flies
(Schweibert 1974).   Plecoptera are  ecologically very  important
components of streams,  where smaller forms  cling  to rocks, feeding on
the drift of detritus,  and algae and larger forms seek smaller
invertebrates as prey.   Critical sensitivity of this  group begins
between pH 4.5 to 5.5,  and their distribution generally  follows that of
mayflies, except for some tolerant  forms like Taeniopteryx, Nemoura,
Nemurella, and Protonemura (Raddum  1979).

     Caddisflies (Trichoptera) include  burrowers, sprawlers, filter
feeders, predators, detritivores, and forms found specifically in
running or standing water.  They occupy many niches and  are difficult to
lump into generalizations.  Most of the larvae live in cases made from
local materials.  Caddisflies have  been found near pH 4.5 in field
surveys (Sutcliffe and Carrick 1973, Leivestad et al. 1976, Raddum 1976)
but not at pH 4.0 (Raddum 1979;  Hall and Likens 1980a,b).  Raddum (1979)
observed that the running water caddisflies Rhyacophila  nubila,
Hydropsyche sp., Polycentropus flavomaculus, and  Plecforenemia conspersa
a 11 survived pH 4.0 in the laboratory,  but only P. conspersa did well in
situ at pH 4.8.  Raddum explained the loss  of Rhyacophila and
Hydropsyche in the field by alterations in  their  food supply.  P_.
navomacuTatus became cannibalistic at  pH  4.0, which  may explain its
absence in the stream but its survival  when isolated  during laboratory
experiments.  The problem of cannibalism points out the  difficulties in
relating laboratory studies to field observations.  Another caddisfly,
Limnephilus pal lens, was collected  from an  alkaline (pH  8.2) pond and
subjected to more acidic water both in  the laboratory and in situ (Havas
and Hutchinson 1982).  The larvae survived  in pH  3.5  water, and actually
did better in metal contaminated sulfate-fumigated water.  This acidic
water was near the alkaline pond from which the caddisflies were
collected, but no larvae lived in the acid pond.  Possible explanations
for the absence of the caddisflies  from water in  which they could
survive were:  1) absence of suitable food, 2) sensitivity to the
acidity during emergence, 3) absence of suitable  case building material
in the acidic pond.

      Most other insects are largely unaffected or slightly favored  in
acidic lakes and streams.  The alderfly, Sialis (Megaloptera), increased
its emergence rates in an artificially  acidified  stream  (Hall and Likens
1980a,b).  It was found commonly in shallow water in  an  acidic (pH 3.9
to 4.6) Swedish lake (Wiederholm and Eriksson 1977) and  in a highly
variable (pH 6.2 to 4.2) Norwegian  lake (Hagen and Langeland 1973).

     Several true flies (Diptera) increase in relative abundance at  low
pH (Hagen and Langeland 1973, Wiederholm  and Eriksson 1977, Raddum 1979,
Collins et al. 1981, Raddum and Saether 1981). The most successful
dipterans are the midges (Chironomidae),  the predacious  phantom midge
(Chaoborus, Chaoboridae) and in streams, the black fly  (Simulidae).
Black fly adults are notorious as biting  pests when they emerge  in the
spring.  Often, the principal insects in acidic lakes are the midges
(Chironomidae) Chironomus riparius  (Havas  and Hutchinson 1982)
Procladius sp.. LlmnochTronomous sp., Sergentia coracina,


                                  5-29

-------
Stlchtochironomus sp. and phantom midges (Chaoborus)  (Leivestad  et  al.
1976, Raddum and Saether 1981).   These insects  comprised  56  and  41
percent of the benthos of a Swedish acidic lake (pH  3.9 to 4.6)
(Wiederholm and  Eriksson 1977).   Chironomids appear to be preadapted
for acidification, because the same species are found in  clearwater
acidic lakes as in humic acid lakes (Raddum and Saether 1981).   Uutala
(1981) reported that the chironomid fauna of two acidic Adirondack  lakes
were reduced in biomass as compared to fauna in nearby control lakes.
The different life cycle stages  have variable responses to pH  stress,
but the molting period is the most sensitive (Bell 1970).

     The dominance of the benthos of acidic lakes by midge larvae is not
surprising, as these insects are abundant in almost  all lakes, but  the
observed shift in dominant species does suggest that benthic community
structure is altered.  Direct toxicity is probably not the explanation
for the absence of certain species.  For example, some Orthocladius
consobrinus tolerate pH 2.8 in the laboratory,  but this species  was not
found in acidic pools (pH 2.8) in the Smoking Hills,  even though it was
found in nearby alkaline (pH 8.2) pools (Havas  and Hutchinson  1982).

     Other insects abundant in acidic waters are the true bugs
(Henri ptera) like water striders  (Gerridae), backswimmers  (Notonectidae),
and water boatmen (Corixidae), and beetles (Coleoptera) of the families
Dytiscidae and Gyrinidae (Raddum 1976, Raddum et al.  1979, Nilssen
1980).  These insects prey on other insects and small crustaceans,  both
benthic and planktonic.  They are metabolically very  active and  receive
most of their 03 from the atmosphere, thus reducing  the amount of soft
body tissue exposed directly to  the water, in contrast to gilled insects
and crustaceans.

5.3.2.5.2  Sensitivity of insects from different microhabitats.
Important generalizations are better made by analyzing the data  after
grouping the taxa by functional  guilds and microhabitats  rather  than by
phylogenetic associations (Merritt and Cummins  1978).  Collins et al.
(1981) compared three acidic softwater lakes (4.6 to 4.9) with 11
neutral  softwater lakes in central  Ontario and  reported no significant
decreases in populations of animals living in sediments (infauna).
Observations of epifauna by scuba divers concurred with the general
observation that acidic lakes have depauperitic mollusc and  insect
populations.

     It is hardly surprising that infaunal  communities, which  are
protected by the buffering capacity of the substrate, are less affected
than epifaunal  communities.  Still, few studies have  organized data in
such a manner as to verify that  epifaunal  insects are indeed the targets
of acid stress.  Also, a perusal  of the data presented above suggests
that it is epifaunal forms with  filamentous gills that are most
sensitive to low pH.  Air-breathing beetles and bugs  survive low pH
stress well as do infaunal forms  with filamentous gills,  such  as the
burrowing mayfly, Hexagenia.   Metabolic and physical  actions of
Hexagenia nymphs increased the Eh,  NHa,  inorganic S,  SOa, and
decreased the pH as compared to  control  microcosms lacking nymphs or
                                  5-30

-------
with dead nymphs (Mitchell  et al.  1981b).   Thus,  not only  does  the
chemistry affect the biota, but conversely  the  biota alters the
chemistry.

5.3.2.5.3  Acid sensitivity of insects  based  on food sources.   Total
invertebrate biomass in an  acidic  (pH 4.3 to  5.9)  stream was -2.6
times less than that of a neutral  stream  (pH  6.5  to 7.3) 6 km away  in
southern Sweden (Friberg et al. 1980).  Organizing species lists  into
guilds based on eating methods shows that in  the acidic water,  shredders
increased in relative abundance at the  expense  of scrapers.  These  data
differ from those reported by Hall and  Likens (1980a,b) from an
artificially acidified stream in New Hampshire, where  shredders and
predators were not affected.  The tolerance of  predators,  mostly
predacious diving beetles (Dytiscidae), water striders (Gerridae) and
water boatmen (Corixidae) has been noted  in numerous corroborated
surveys (Leivestad et al. 1976, Raddum  et al. 1979).   Shifts in the
activities of these different functional  guilds affect detritus
processing and may be either a cause or a result of the inhibition  of
microbial detritus processing (Section  5.3.2.1).

5.3.2.5.4  Mechanisms of effects and trophic  interactions.   It  is likely
that other factors besides H"1" concentration stress organisms in
acidified waters. (Overrein et al. 1980).   Malley's (1980) work  (see
Section 5.4.2.4) suggests that reduced  calcium  deposition  may limit
insects as well as crustaceans.  Havas  (1981) suggested that Na+
transport may be affected.   Effects of  increased Al concentrations  on
invertebrates have not been studied as  intensively as  they have with
fish (Baker and Schofield 1980).  Other metals, such as  Hg (Tomlinson
1978) may also be important.  Nutrient  depletion, inefficient microbial
digestion, substrate alteration, dissolved  oxygen stress,  and changes  in
other populations (e.g., fish predation)  all  may act on insect
populations.  The water boatman, Glaenpcorisa propinqua propinqua a
predator on zooplankton and other small invertebrates, is  tolerant  of
acidity and is common in acidic lakes.   The addition of perch to
one-half of a lake divided by a net vastly  reduced numbers of
Glaenocorisa on the side with fish.  The  only change in  water chemistry
was a decrease in total phosphorus from 3-8 to 2-6 yg  &-1  when
fish were added (Henrikson and Oscarson 1978).   Different  taxa  respond
in various ways.  Some may make behavioral  adaptations;  others, like  the
water boatmen (Corixidae) can alter rates of  Na+ pumping  (Vangenechten
et al. 1979, Vangenechten and Vanderborght  1980).

     For reasons which  are not clear, a shift towards  larger  species
within a higher taxon occurs (Raddum 1980).  This may  be  due  to reduced
predation pressure on larger insects in the absence  of fish or  because
larger species have less surface/volume and can cope better with
chemical and osmotic stress.   Increased abundance of  insect predators
may be due to the opening of this niche as  a  result of fish loss
(Henriksen et al. 1980b) or due to the larger  size of  these predators.
Community alterations, and even modifications of water column chemistry,
have been traced to fish removal  (Stenson et al. 1978) independent  of  pH
changes.  Thus, it is dangerously simplistic  to ascribe changes in


                                  5-31

-------
community composition to merely the physical-chemical  alterations  of
acidification without also considering the varied biological
interactions.

     Alterations in insect populations are likely to affect fish
populations (0kland and 0kland 1980).   Rosseland et al.  (1980)
reported that corixids composed 15 percent of the gut content,  by
volume, of trout from neutral  waters but 44 percent in a  declining
population from an acidic (pH < 5.5) stream.   However, no causal
relationship between shifts in diet and population decline can  be  made
at this time.

5.3.2.6  Mollusca--Molluscs provide food for vertebrates  (fish, ducks,
muskrats, etc.).  Clams are filter feeders and are important bio-
indicators of water quality conditions.  Snails scrape the substrate  and
the surfaces of aquatic plants, controlling the periphyton in waters  in
which they live.  The impact of acidity on molTuscan populations is
dramatic.  The calcareous shell of these animals is highly soluble at pH
< 7.0 and acidic conditions require that the  animals precipitate fresh
CaC03 faster than it can dissolve.

     The only thorough survey of clams and snails in acid-impacted
waters was done in Norway (J.  flkland 1969a,b,  1976, 1979a,b, 1980; K.
0kland 1971, 1979b,c,  1980b;  0kland and 0kland 1978,  1980;
0kland and Kuiper 1980).  About 1500 localities,  mostly lakes in
Norway, were surveyed between 1953 and 1973.   Fingernail  clams
(Sphaeriidae)  and snails (Gastropoda)  were sampled.  Sphaeriidae live in
sediments (infaunal)  and no surveys of the more epifauna!  unionid
mussels have been conducted.   Norway has 17 species of Pisidium and
three of Sphaerium.   None of these clams normally occurred below pH 5.0.
The six most common sphaeriids were eliminated below pH 6.0.  These
common species were found in lakes with low alkalinities  but with  pH
values  ~ 6.0.   Thus,  their absence from these  poorly buffered lakes
serves as an indication of acidification,  not  just low CaCOa stress
(0kland and Kuiper 1980).

     Freshwater snails (Gastropoda) were reported to be stressed much
like the clams from the Norwegian survey.   Of  the 27 species of snails
reported in Norway, only five were found below pH 6.0  (J.  0kland
1980).  Snails could  tolerate  higher H+ concentrations if the total
hardness were higher,  indicating that pH may  stress snails by reducing
the CaCOs availability (J.  0kland 1979b).   These authors  (0kland
and 0kland 1980) estimated that the crustacean Gammarus lacustris  and
the molluscs accounted for 45  percent  of the caloric  input of trout, and
they predicted that trout production could be  reduced by  10 to 30
percent below pH 6.0  due to the loss of food resources.   This prediction
has not been supported by the fish surveys (Section 5.6.2).

     Some additional  distributional data,  which corroborate the
0klands' conclusions  cited above,  have been reported  from Sweden
(Wiederholm and Eriksson 1977), Norway (Hagen  and Langeland 1973,
Nilssen 1980)  and from a river in England (Sutcliffe and  Carrick 1973).
                                  5-32

-------
These later authors emphasized the absence  of the  freshwater limpet
snail Ancylus fluviatilis as an indicator of  pH levels which frequently
fall below 5.7~They also concluded that pH  served  to limit the
distribution of  molluscs by reducing the availability of CaC03» as
measured by water hardness.

     The physiological  response of molluscs to pH  stress was studied by
Singer (1981b).   In Anodonta grandis (Unionidae) from six lakes ir New
York and Ontario with various levels of pH  and hardness, marked
differences in shell  morphometry and ultrastructure  were observed.  The
clams from alkaline lakes (pH > 7.2) had thick shells with  fine layers
of organic conchiolin interspersed.  The clams from  softwater neutral
lakes had thinner shells, with relatively thick prismatic layers.  Clams
from a slightly  acidic  lake (pH 6.6) had thin shells with heavy plates
of organic material substituting for the normal CaCOs matrix.  Using
unionid shells from museum collections  as indicators of pre-acidifica-
tion water quality was  suggested.

5.3.2.7  Anne!ida--Aquatic worms have been  used extensively as
indicators of organic (Goodnight 1973,  Brinkhurst  1974) and inorganic
(Hart and Fuller 1974)  pollution.  With an  increase  in organic detritus
and a decrease in fy concentrations, the benthic community  is
typically dominated by  Tubifex spp. and Limnodrilus  hoffmeisteri
(Brinkhurst 1965, Howmiller 1977).   Considering their tolerance of other
stresses and the abundance of detritus, it  is surprising that
oligochaetes are reduced in biomass in  acidic lakes.  Raddum (1976,
1980) found few  oligochaetes in water deeper  than  20 m in 18 acidic
lakes (pH < 5.5) and normal fauna in 16 other more neutral  Norwegian
oligotrophic lakes.  The acidic lakes,  however, had  more oligochaetes
than the non-acidic lake at a depth of  0.5  m. The difference in numbers
at greater depths was more pronounced in the  spring  and autumn.  Neutral
lakes had three  to four times the total number of  oligochaetes per
square meter.  Raddum (1980) attributed the reduction in numbers of
oligochaetes in  acidic  lakes to pollutants  associated with  acidic
deposition (e.g., heavy metals and aluminum). These worms, however, are
routinely collected in  vast numbers directly  below sewage and industrial
effluents with far greater concentrations of  pollutants (Hart and Fuller
1974, Chapman et al. 1980).  An alternative explanation for their
reduction in numbers might be the unpalatability of  their detrital food
due to the slower decomposition rates in acidic lakes (Traaen 1977).
Oligochaetes are not normally abundant  in nutrient-poor waters, and
their low numbers in acidic lakes may be as much a function of the low
nutrient level characteristic of acidic lakes as of  pH.

     One study that mentioned the distribution of  leeches (Hirudinea) in
acidic lakes (Nilssen 1980) reported that these worms disappeared below
pH 5.5.  Leeches characteristic of eutrophic  waters  (Hirudo medicinal is,
Glossiphpnia heteroclita) were absent from  even mildly acidic lakes.
Raddum (1980) reported  that Hirudinea were  restricted to waters above pH
5.5, largely because of the loss of prey below this  pH, even though many
leeches are detritivores and scavengers, not  obligate carnivores (Pennak
1978).  These anecdotal observations should be viewed with  caution
                                  5-33

-------
because leeches are not always  common  in  neutral oligotrophic lakes, and
I saw an unidentified leech  on  the  bottom of acidic Woods Lake (Herkimer
Co., NY) while diving in 6 m of water.

5.3.2.8  Summary of Effects  of  Acidification on Benthos—Table 5-3
summarizes some of the expected consequences of acidifying a lake or
stream to pH 4.5.  The following generalizations may also be made, based
on the best available current evidence.

1.   Bacterial decomposition of litter in bags in situ and debris in
     vitro is reduced significantly (p <  .001), as measured by
     respiratory rates and weight loss, between pH 6.0 and 4.0.
     Planktonic bacterial  standing  crops  do not change significantly,
     although metabolic rates are depressed.  Insects and crustaceans
     responsible for shredding  and  processing detritus are almost
     completely eliminated between  pH  6.0 and 4.0.

2.   In most acidified lakes below  pH  5.0, a mat of algae covers most of
     the substrate from ~ 1  to  5 m  to  the limit of light penetration.
     These mats are of 3 types:   a)  an  encrusting, felt-like, black to
     blue-green mat composed of blue-green algae (Cyanophyta) 0.5 to 2
     cm thick; b) coarse,  loosely compacted dark green mats composed of
     green algae (Chlorophyta)  1 to 4  cm  thick; c) cloud-like layers of
     green filamentous algae (Chlorophyta) which rest on the bottom in
     depths as thick as 1.5  m.   All  three types of mats include debris,
     diatoms, fungi and bacteria.   These  mats are often the most visible
     aspect of acidified lakes.   They  may have profound effects on fish
     spawning habitats, nutrient cycling, and sediment chemistry, but
     their origin, differentiation  into types, and chemical interactions
     have not been studied.   They have  been extensively noted in field
     surveys and have developed in  artificially acidified chambers and
     stream channels below pH 5.0.

 3.  Many invertebrates are  very sensitive to pH.  Amphipods, which are
     an important fish food  in  rivers  and some lakes, cannot tolerate pH
     < 6.0, based on field observations,  laboratory bioassays, and field
     enclosure experiments.   Snail  populations are stressed below pH 6.0
     and absent from the field  below pH 5.2.  Large mussels cannot
     survive below pH 6.6, but  fingernail clams can survive in sediments
     with overlying water with  pH values  as low as 4.8.  The crustacean
     water louse (Isopoda) and  many species of stoneflies (Plecoptera),
     mayflies (Ephemeroptera),  and  caddisflies (Trichoptera) die at pH <
     5.0, as determined by field observations and laboratory bioassays.
     Insects are often limited  by mechanisms not related to direct
     toxicity.  Some dragonflies and many predacious beetles
     (Coleoptera), and true  bugs (Hemiptera) occur commonly in acidified
     (pH < 5.0) lakes.  They fill the  niche normally occupied by
     planktivorous fish and  represent  a major alteration of food chains.
     Most of these active  predaceous insects receive their air supply
     from the surface.
                                  5-34

-------
                             TABLE 5-3.   SUMMARY OF EFFECTS  OF  pH  4.5 WATER ON BENTHOS
            Taxon
                     Common name
                                  Microhabitats
                      Sensitivity to acid
                        (pH 4.5) stress
                       References
            Bacteria
            Periphyton
                   Algae
                     Blue-greens (Cyanophyta)
                     Greens (Chlorophyta)
                     Diatoms (Bacillariophyceae)
                     Dinoflagellates (Pyrrophyta)
oo
en
Crustacea
  Decapoda
Crayf 1 sh
              Isopoda
              Amphipoda
                   Aquatic sowbug
                   Scud
                                                 Al 1 substrates.
                               On plants  (epiphytic)
                               On rocks  (epipl ithic)
                               On mud (epipelic)
                               On "encrusting  mat".
Burrowers deposit
feeders and grazers
lakes and streams.
                               Deposit feeder in
                               lakes and streams,
                               under rocks and in
                               littoral  vegetation.

                               Detritivore in lakes
                               and slow-flowinq
                               areas of streams.
                               Found among pi ant
                               stems.
                                                     Growth rate or Og
                                                     uptake inhibited.
                      Increasing standing
                      crop in lakes and
                      streams.
                                                                                    Development of dis-
                                                                                    tinct types of peri-
                                                                                    phyton communities.
Sensitivity variable
and highly species
specific.  Effects
vary depending on
other cation concen-
trations.

Asellus aquaticus
tolerant to - pH
5.0.
                      Not generally found
                      below pH 6.0.  Sen-
                      sitive differences
                      between species have
                      been described.
                     Bick and Drews
                     1973, Baath et al.
                     1979, Gahnstrom et
                     al.  1980

                     Hendrey 1976, Hall
                     et al. 1980,  Van
                     and  Stokes 1978,
                     Hendrey and
                     Vertucci 1980,
                     Muller 1980

                     Singer et al. In
                     press, Stokes 1981
Mai ley 1980,
Collins et al.
1981, Shaw 1960
                                            K. Okland 1978a,
                                            1980b.
                      K.  0kland
                      1980a,b,c;
                      Sutcliffe and
                      Car rick  1973;

-------
                                                       TABLE  5-3.   CONTINUED
              Taxon
Common name
Microhabitats
Sensitivity to acid
  (pH 4.5) stress
References
                Eubranchiopoda   Tadpole  shrimp
                            Small, often tenpo-
                            rary, ponds or back-
                            waters of streams;
                            often only abundant
                            seasonally.
                    Not  found  in  Sweden
                    pH 6.1.  Growth  re-
                    duction  and hatching
                    failure  below pH
                    5.0.
                      Borgstrom  et  al.
                      1976,  Borgtrom  and
                      Hendrey  1976
              Insecta
                Ephemeroptera    Mayflies
                            Include burrowing and  Sensitivity  varies
en
 i
co
CD
                Odonata           Dragonflies  (Anisoptera)
                                 Damselflies  (Zygoptera)
                Plecoptera        StonefHes
                Trichoptera      Caddlsflies
                                                                surface  dwelling
                                                                forms.   Found  in
                                                                lakes and  streams.
                                                                Predators,  detriti-
                                                                vores, herbivores.
                            Predators in mud,
                            littoral debris,  and
                            rock substrates in
                            lakes and streams.
                            Predators, detriti-
                            vores and herbivores
                            in flowing streams.
                            All benthic habitats.
                                                   between  groups  but
                                                   generally  not
                                                   tolerant
                    Tolerant  to  pro-
                    longed  severe acid
                    stress.
                    Most  genera are
                    sensitive  but  some
                    are tolerant (see
                    text).
                    Some  genera  are very
                    tolerant,  hut others
                    are sensitive (see
                    text)
                      Sutcliffe  and
                      Carrick  1973,
                      Nilssen  1980,
                      Pratt  and  Hall
                      1981,  Leivestad
                      et  al. 1976,
                      Borgstrom  et al.
                      1976,  Raddum 1976

                      Stickney 1922,
                      Borgstrom  et al.
                      1976,  Bell  and
                      Nebeker  1969, Bell
                      1971

                      Sutcliffe  and
                      Carrick  1973;
                      Leivestad  et al.
                      1976;  Hall  and
                      Likens 1980a,b;
                      Raddum 1979

                      Sutcliffe  and
                      Carrick  1973;
                      Leivestad  et al.
                      1976;  Hall  and
                      Likens 1980a,b;
                      Raddum 1976, 1979

-------
                                                     TABLE 5-3.   CONTINUED
             Taxon
   Common name
   Microhabitats
Sensitivity to acid
  (pH 4.5) stress
                                                                                                             References
               Diptera
01
 i
GO
               Hemiptera
               Coleoptera
             Mollusca
               Pel ecypoda
True flies
  Midges (Chironomidae)
                                 Phantom ghost midge
                                  (Chaobotidae)
                                Black  flies (Simulidae)
True bugs
  Water striders  (Gerridae)
  Backswimmer (Notonectidae)
  Water boatman  (Corixidae)

Beetles
  Predacious diving  beetle
    (Dytiscidae)
  Whirligig beetle (Gyrinidae)
Clams
Major detritivores in
lakes and enriched
streams, living in
mud or on substrates
in tubes.

Predator on sub-
strates and in water
column of lakes
Predatory in streams
on rock substrates

Predators on water
surface, in water
column, and over sub-
strates.

Predators on water
surface, in water
column, and over sub-
strates.
Filter feeders in
substrates, detriti-
vores in lakes and
streams.
Most reports show
increase in numbers,
but some report
decreases.
                                                      Tolerant of acid
                                                      stress.
Tolerant of acid
stress

Tolerant of acid
stress.
Tolerant of acid
stress.
All mollusca are
highly sensitive to
pH stress.  The most
tolerant are finger-
nail clans which are
rarely found as low
as pH 5.0.
Raddum and Saether
1981, Uutala 1981
Wi ederholm and
Eriksson 1977,
Leivestad et al.
1976

Leivestad et al.
1976

Raddum 1976,
Raddum et al.
1979, Nilssen 1980
Raddum 1976,
Raddum et al.
1979, Nilssen
                                                                                                                         1980
J. (Bkland 1976,
1979b, 1980; K.
0kland 1971,
1979b,c, 1980b;
(Jkland and
flkland 1978,
1980;  0kland
and Kuiper 1980;
Singer 1981b

-------
en
 i
to
CO
                                                      TABLE  5-3.    CONTINUED
             Taxon
   Common name
Mlcrohabitats
                      Sensitivity to acid
                         (pH 4.5) stress
                                                                                                            References
             Annelida
               01igochaeta
               Hirudinea
Aquatic earthworms
Leeches
Detritivores  in  lakes  Standing crops low
and streams with soft  in acidic waters
substrates.
                                                              Predators, detriti-
                                                              vores.
                                                     Anecdotal  observa-
                                                     tions report  no
                                                     leeches below pH
                                                     5.5.
                                         Raddum 1976,  1980
                                         Nilssen  1980,
                                         Raddum 1980

-------
 4.  Forms which live cm  the substrate  (snails, stoneflies, mussels,
     etc.) are more sensitive to  pH  drops than those which live In the
     substrate (e.g., fingernail  clams, midge larvae burrowing may-
     flies).   In those groups that have been studied in the laboratory
     (crayfish, backswimmers, molluscs), high calcium concentrations (>
     2 mg £-1 can ameliorate the  effects of low pH.

     Fish shift their food  to available prey, but the nutritional
effects of switching from a diet  of  largely amphipods,. mayflies, and
stoneflies to one of water  boatmen,  beetles, and water striders are not
known.  Effects on different age  classes of fish are likely to vary.
Changes in the rates of detrital  processing and decomposition rates
affect primary productivity and hence the whole ecosystem.

5.4  MACROPHYTES AND WETLAND PLANTS  (J. H. Peverly)

5.4.1  Introduction

     The softwater, low alkalinity,  oligotrophic lakes in temperate
regions susceptible to acidic deposition support a  flora characterized
by the isoetid or rosette plants.  This contrasts with hard waters which
support vittate species,  having elongated stems with leaf nodes.
Plants commonly observed  in softwater lakes are listed in Table 5-4.

     In general, emergent plants  in  these lakes grow only in a narrow
band along the shore.  The  submerged, three-inch high isoetids extend
from shore to the 3 to 4  m  depth  and coexist with some lilies and
bladderwort.   Beyond 4 m, Mi tell a spp., bladderwort, and mosses
dominate.

     Life in the water depends on the presence and  growth of aquatic
plants as well as other inputs from  the basin (Section 5.3.1).
Macrophytes stabilize the sediments; clear, cool and oxygenate the
water; and provide colonization sites for insects,  small plants and
animals, and bacteria. These in  turn are a major food source for the
larger aquatic animals, such as fishes, amphibians, aquatic mammals, and
waterfowl.  Thus, aquatic plants  fill an important  role in the entire
aquatic ecosystem.

     Macrophyte growth in softwater  lakes can be a  major part of total
lake production and is largely attributable to growth by isoetids
(Hutchinson 1975, Hendrey et al.  1980b).  Because isoetids are perennial
and evergreen, they can continue  to  photosynthesize and produce oxygen
under winter ice cover, and provide  a stable, constant source of grazing
material.  Standing crop  varies from <  5 to 500 g dry wt m-2 in
August, but annual productivity is only about 50 percent of standing
crops (Moeller 1978, Sand-Jensen  and Sondergaard 1979).

     Plant productivity in softwater lakes  is not high because the
carbon dioxide (003) level  in the water is low (.02 mM C02 at pH
5.0) and major nutrient minerals  such as P, K, N, and Ca are in limited
supply (Hutchinson 1975).  However,  these aquatic macrophytes have


                                 5-39

-------
  TABLE 5-4.  PLANTS COMMONLY OBSERVED IN SOFTWATER (LOW ALKALINITY)
                         OLIGOTROPHIC LAKES
    Species
                     Common Name    Type
                   Response to
                   Acidification
Pontederia
cordata L.
Pickerel
weed
Emergent
Unknown
Juncus sp.
Spargam'um spp.

Brasenia
  schreberi Gmel

Nuphar
  advena Ait.

Nymphaea
  odorata Ait.

Isoetes spp.
Lobelia
  dortmanna L.
Eriocaulon
  septangulare With
Myriophylliim
      Tk
tenel1umBigel
Potamogeton spp.
                     Rush
                     Burreed

                     Water
                       shield

                     Yellow
                       lily

                     White
                       lily

                     Quillwort
                     Pipewort
                     Pond  weeds
Emergent
Emergent

Floating leaves,
  rooted

Floating leaves,
  rooted

Floating leaves,
  rooted

Submerged,
  rooted
  (iosetids)

Submerged,
  rooted
  (iosetids)
Submerged,
  rooted
  (iosetids)

Submerged,
  rooted
  (iosetids)

Submerged,
  rooted
Stimulated
growth
(Hultberg and
Grahn 1976)

Unknown

Unknown


Unknown


Unknown
Overgrown
(Hultberg and
Grahn 1976)

Overgrown
(Hultberg and
Grahn 1976)
Oxygen evolu-
tion falls
(Laake 1976)

Unknown
                                                      Unknown
Decreased
growth
(Roberts et
al. 1982)
                                  5-40

-------
                       TABLE 5-4. (CONTINUED)
   Species
Common Name    Type
                   Response to
                   Acidification
Eleocharis spp.


Utricu'iaria spp.


Sphagnum sp.
Nitella spp.
Spike rush    Submerged,
                rooted

Bladderwort   Submerged,
                unrooted
Moss
Drepanocladus spp.     Moss


Fontinalis spp.         Moss
Submerged,
  attached
              Submerged,
                attached

              Submerged,
                attached
Stonewort     Submerged,
                attached
Unknown


Unknown


Stimulated
growth (Grahn
1977)

Unknown


Unknown


Unknown
                                  5-41

-------
several means of overcoming such  difficulties and producing enough
tissue to support an aquatic  animal community.  First, aquatic
macrophytes recapture up to 50  percent  of  their own respiratory C02
and store it in an internal  gas chamber systan for reuse in
photosynthesis (Sondergaard 1979).  Secondly, the isoetids are able to
exist and grow in oligotrophic  water, where other aquatic macrophytes
cannot, by more efficient use of  nutrients in the sediments.  This is
accomplished by the root systems, which are efficient sites for
absorption of carbon, nitrogen, phosphorus, and potassium.  The relative
root-to-shoot ratio is large  in these plants  (0.5 to 0.6, Sondergaard
and Sand-Jensen 1979), indicative of the greater role of roots in
nutrient absorption.  In addition,  water in the sediments where the
roots grow often has a carbon level of  1 to 5 mM (Wiurn-Andersen and
Andersen 1972), 50 to 100 times that in the overlying water column.
Vittate plants, which depend  more on leaf  absorption for carbon supply,
cannot grow in these low carbon waters.

     The accumulation of nutients in plant tissues, acquired through the
roots from the sediments, recirculates  sediment nutrients back into the
overlying water, where they can be  used by other organisms.  For
instance, in a 200 g dry wt m~2 crop of Eriocaulon septangulare, there
would be about 50 g carbon, 2 g nitrogen,  0.1 g phosphorus, and 1.5 g
potassium.  About 0.24 g carbon,  0.2 g  nitrogen, 0.01 g phosphorus, and
0.4 g potassium (Moeller 1975)  would be dissolved in water 1 m deep over
this meter square area.  Clearly, nutrient release from such plant beds
could increase the concentration  of available nutrients in the water
column.

     Lilies and emergent plants can also obtain carbon by absorption of
CO^ from the atmosphere and translocation  to  carbon reserves in
rhizomes under the water surface.   Mosses  and algae are not as involved
in processes that transfer nutrients from  sediments or air into the
water column.

     In addition to major nutrients, rooted aquatic macrophytes
(including isoetids) are exposed  to elevated  levels of metals in the
sediments (e.g. iron, manganese,  copper, zinc, aluminum).  These
elements can also be absorbed by  roots  and transported to the shoots,
where they are able to enter biological  cycles slowly as the plants
senesce and decay.  However,  concentration differences between sediment
and water column levels of metals available for absorption are not
always as great as for the major  nutrients.   This is especially the case
where rooted plant activity is  high, as oxygen release at the root
surfaces (Wiurn-Andersen and Andersen 1972) raises the redox potential.
This tends to precipitate iron  and  manganese  compounds (Tessenow and
Baynes 1978) and remove phosphorus  from solution.  Metals not affected
by redox potential, like aluminum,  would remain in solution in the
rhizosphere, and still be available for uptake by the roots.  Indeed,
the aluminum contents of plant  tissues  (0.4 to 22 g kg-1) from both
neutral and acidified lakes (Al 0.03 to 0.2 mg £-M in the
Adirondacks and Ontario were elevated above Hutchinson's (1975) mean
value of 0.36 g Kg-1 (Best and  Peverly  1981,  Miller et al. 1982).
                                  5-42

-------
     Lilies interact much more with sediments than with water and
generally tend to accumulate less of the above metals.   The  mosses and
algae interact almost exclusivey with the water column  and accumulate
metals (Ca, K, Fe, Al) under certain water conditions.   Aquatic
macrophytes can recycle Fe, Mn, Cu, Zn,  and Al  metals  from sediments,
but they can also restrict exchange of Fe and Mn between water and
sediment by oxidizing the top 15 to 20 cm of sediments  (Tessenow and
Baynes 1978).

     Mosses and algae that grow close to the bottom not only absorb
metals metabolically, but also physically adsorb then  onto tissue
surfaces.  Sphagnum spp. are known to have especially  high adsorption
capacities for metals, including calcium, iron,  aluminum, and potassium
(Clymo 1963, Hendrey and Vertucci 1980).  Metals adsorbed in this
fashion are effectively removed from biological  cycles  for long periods,
as the elements remain bound to dead tissues, which often persist for
years.  Mats of Sphagnum spp. and algae have formed on  the bottom of
some softwater lakes.Hultberg and Grahn (1976) suggested that mats of
this nature decrease productivity by restricting exchange of nutrients
between sediments and water.

     The tissues produced by growing plants eventually  die,  releasing
nutrient elements and metals back to the water by a variety  of decay
processes.  Carbon dioxide is produced by plankton and  microorganisms
from this dead plant material, along with dissolved phosphorus,
potassium, ammonia and calcium.  The metals are released, often in a
form complexed with organic acids that keeps them in solution, thus
readily available for uptake.

5.4.2  Effects of Acidification on Aquatic Macrophytes

     Direct effects of acidification on  aquatic macrophytes  have not
been well-documented.  However, in two reports of laboratory results,
oxygen evolution was reduced up to 75 percent by a pH  decrease from 7.0
to 4.0 in both softwater (Laake 1976) and hardwater plants (Roberts et
al. 1982).  In the field, nutrient ions  and metals (such as  calcium,
magnesium, sodium, potassium, manganese, and iron) may  be leached out of
the tissues, especially during the episodic pH drops associated with
snow melt.  This could have a negative effect on plants in the spring
when new growth is quite susceptible to  nutrient imbalances.

     Most effects of acidification on aquatic plant distribution and
growth are indirect.  Specifically, these would include decreased carbon
supply for photosynthesis, nutrient depletion, increased metal
concentrations, and decreased rates of nutrient recycling (Grahn et al.
1974, Andersson et al. 1978b, Schindler  et al. 1980a).   The  dominance of
isoetid species in soft water lakes of pH 5.5 to 6.5 is a response in
part to low carbon and major nutrient availability in  the water column.
As acidic deposition causes the pH to decline, these factors become even
more limiting.  For instance, Lobelia dortmanna rooted  in sediment cores
showed a 75 percent reduction in oxygen  production at  pH 4.0 compared to
the control (pH 4.3 to 5.5), and the period of flowering was delayed 10


                                  5-43

-------
days at the low pH (Laake 1976).   As  a  result, species more tolerant of
low nutrient supplies and higher  metal  concentrations may become
dominant.

     Measurements over 15 years  in one  acidified Swedish lake with a pH
drop of 0.8 units between 1967 and 1973 showed that isoetid species were
replaced by Sphagnum sp.  and blue-green filamentous algae, which grew
over the bottom in that time span, smothering the low-growing isoetids
(Grahn 1977).  This is viewed as  detrimental to overall lake quality
because Sphagnum beds are not a good  habitat for most aquatic animals.
In addition, Sphagnum tends  to perpetuate  the conditions that exclude
other species by exchanging  metabolically  produced hydrogen ions for
nutrients and metals in the  water via adsorption processes.  Thus,
acidification and oligotrophication continue.  As the Sphagnum grows, it
forms a mat of increasing area.   The  dead  stems decay slowly and
continue to hold adsorbed elements.   As a  result of this mat barrier and
because Sphagnum has no roots to  exploit the sediment, interchange of
dissolved nutrients between  overlying water and sediments is minimized
following Sphagnum invasion.  With the  exception of dense Sphagnum beds
observed in Lake Golden (pH  4.9)  in the Adirondack Mountains of New York
State (Hendrey and Vertucci  1980), large expanding beds have not been
observed in acidified waters of the northeast United States or Canada
(Best and Peverly 1981, Wile 1981).

     The effect of acidification  on nutrient availability is unclear.
Generally, slower breakdown  of organic  matter (including Sphagnum
tissues) in acidic waters (see Section  5.8.2.1) would tend to decrease
the amount of major nutrients available for plant growth.  In addition,
softwater lakes are inherently low in nutrients.  In the Adirondacks,
plant tissue concentrations  of the major nutrients indicated that
phosphorus was limiting in both  acidic  and nonacidic lakes (Best and
Peverly 1981).

     Other possible indirect effects  of acidity on macrophytes are those
associated with increased metal  (aluminum, cadmium, iron, manganese,
copper, lead, zinc) concentrations in water and sediments.  Tissue
analysis of isoetid plants from  both  acidified and nonacidified lakes in
the Adirondacks and Ontario  have  shown  elevated levels of aluminum,
copper, iron, and lead in roots  and shoots from acidic waters (Best and
Peverly 1981, Miller et al.  1982). Concentrations of manganese,
cadmium, and zinc were lower in  plants  from acidic waters, corresponding
to one report of lower measured  metal levels in sediment of an acidified
lake (Troutman and Peters 1982).

     Toxic tissue levels of metals discussed above are not presently
known.  Effects of increased metal accumulation on isoetid productivity
are not clear, but these metals  have  been  shown to be toxic to aquatic
plants.  Concentrations of Al, Zn, and  Cu  in sediments measured by
Stanley (1974) produced 50 percent reduction in Myriophyllum spicatum
root weight.  However, these concentrations were greater than those
reported to occur in acidified lake sediments, at least  for Adirondack
lakes (Best and Peverly 1981).


                                  5-44

-------
     If metal  concentrations Increase  in  tissues, but do not inhibit
growth, there  is a potential  for  increased cycling of metals.  However,
Sphagnum spp.  growth may be a positive factor,  removing metals from the
water by adsorption (Clymo 1963)  and by barrier formation between the
sediments and  water.

      Acidification of brown waters that  contain organic acids causes
clearing of the water column by organic precipitation with metals,
especially aluminum (Aimer et al.  1978).  The  result is increased light
penetration to greater depths, with plant growth perhaps increased over
a larger area.  This could lead to a larger  food base for aquatic
animals and could be a positive factor if the  increased growth is not
represented solely by Sphagnum spp. and blue-green algae.

5.4.3  Summary

    0   There  is currently no trend towards  dominance of macrophyte
        communities by Sphagnum sp. in 50 oligotrophic, softwater lakes
        surveyed in North America.  In fact, dominant species are the
        same in both acidified (pH less than 5.6) and non-acidified (pH
        5.6 to 7.5) lakes.

    0   With continued acidification,  shifts to Sphagnum spp.-dominated
        macrophyte communities have been  documented in six Swedish lakes
        acidified for at least 15 years.  This does not seem to be a
        general property of acidified  lakes.

    o   Standing crops of macrophytes  vary widely (5 to 500 g dry wt
        m-2) in softwater, oligotrophic lakes, and acidification
        produces no definite trend in  standing crop changes.  Based on
        one report, annual productivity is equal to one-half the summer
        standing crop in a nonacidified lake.   Oxygen production was
        reduced 75 percent at pH  4.0 versus  pH 4.3 to 5.5 in one
        flow-through experiment.

    o   The only known effect of  acidification on macrophytes in the
        field is that ofincreased metal content in the tissues,
        especially Al.  In acidified lakes,  mean aluminum concentration
        in plant tissue (dry wt basis) is 3.0  to 5.0 g kg-1 (about ten
        times  higher than normal) while mean manganese concentration is
        0.02 to 4.0 g kg"1 (about one-fifth  of normal).  In general,
        concentrations of iron, lead and  copper are higher, while
        cadimium and zinc are lower  in the tissues of plants from
        acidified lakes.

5.5  PLANKTON  (J. P. Baker)

5.5.1  Introduction

     The term plankton refers to  organisms that live suspended within
the water column, are generally  small  to  microscopic in size, have
limited or no powers of locomotion, and are  more or less subject to


                                  5-45

-------
distribution by water movements  (Wetzel 1975).  The plankton community
consists of animals (zoo plank ton),  plants  (phy to plank ton), and microbes.
Effects of acidification  on  zooplankton and phy to plank ton will be
considered within this section;  effects of acidification on the
microbial community were  included  in Section 5.3.  Discussions focus on
plankton communities within  the  open-water zone.  Interactions with
populations in littoral  and  benthic regions are important, but poorly
understood with regard to potential effects of acidification.

     Zooplankton and phytoplankton  communites are usually quite complex,
composed of a large number of  species, and subject to significant
spatial and temporal  variations.   These variations in occurrence and
importance of species of phytoplankton and zooplankton make it difficult
to obtain a representative sampling of the plankton community.  Attempts
at relating differences in plankton communities between lakes or within
a given lake to acidity or other environmental parameters are hindered
by this natural diversity and  variability.

     Six phyla of algae typically  contribute to phytoplankton
communities of freshwater ecosystems:  Cyanophyta (blue-green algae),
Chlorophyta (green algae), Pyrrophyta  (primarily dinoflagellates),
Chrysophyta (yellow-green algae; includes  the chrysomonads and diatoms),
Euglenophyta (euglenoids), and Cryptophyta (primarily cryptomonads).
Photosynthesis by phytoplankton  plays a significant role in the
metabolism of lakes (Schindler et  al. 1971, Jordan and Likens 1975,
Wetzel 1975), and in determining the quantity of secondary or tertiary
(e.g., fish) production within a lake  (Smith and Swingle 1939, Hall et
al. 1970, Makarewicz and Likens  1979).

     The animal components of  freshwater  plankton communities also
constitute a diverse collection  of organisms from many phyla.  The most
important taxonomic groups are protists (Phylum Protozoa), rotifers
(Phylum Aschelminthes, Class Rotifera, or  as a  separate Phylum
Rotifera), insects (Phylum Arthropoda, Class Insecta), and two subgroups
of the class Crustacea (Phylum Arthropoda), the Subclass Copepoda and
the Order Cladocera  (Subclass Branchiopoda)  (Edmondson 1959).  A large
number of trophic levels are also  represented—herbivores, omnivores,
and carnivores.  Thus, both  the  structure  (variety in types of organisms
represented) and function (energetic  interactions among individual
organisms) of the plankton community are complex.

     Data on acidification and effects on  plankton communities are
limited  almost entirely to field observations  and correlations.
Experiments designed to elucidate  causal mechanisms  for observed changes
are,  for the most part, lacking, at both  the physiological and
ecological level.  The large number of interacting factors potentially
involved in the reaction of plankton  to acidification makes a critical
analysis of currently available  data  very  difficult.  In  some cases,
results  appear contradictory.  With an increased understanding of causal
mechanisms, many of these apparent contradictions should be resolved.
                                  5-46

-------
5.5.2  Effects of Acidification  on  Phytoplankton

5.5.2.1  Changes in Species  Composition—In extensive surveys of acidic
lakes in Norway, Sweden,  eastern Canada, and the United States, altered
species composition and reduced  species  richness (number of species) in
the phytoplankton community  were consistently correlated with low pH
levels.  Results from 18 field studies that support  this conclusion are
summarized in Table 5-5.   Decreases in species  richness appear most
rapidly in the pH interval 5.0 to 6.0  (Aimer et al.  1974, 1978;
Leivestad et al. 1976; Kwiatkowski  and Roff 1976).   For example, in a
survey of lakes in the west  coast region of Sweden,  lakes with pH values
of 6.0 to 8.0 generally contained 30 to  80 species of phytoplankton per
100 ml sample.  Lakes with pH levels below 5.0  had only about a dozen
species.  In some very acidic lakes (pH  4.0), only three species were
collected (Aimer et al 1978).

     In general, species are lost from all classes of algae as pH
declines.  However, proportionally  larger losses occur within some
groups than in others.  As a result, the dominant algae in acidic lakes
are often different from those characteristic of circumneutral lakes.

     In six out of nine investigations  (Table 5-5),  dinoflagellates
(Phylum Pyrrophyta), and often the  same  species of dinoflagellates, were
reported to dominate in acidic lakes.  Aimer et al.  (1974, 1978)
reported that the dominant species  in  acidic waters  sampled in the west
coast region of Sweden were  Peri dim'urn  inconspicuum  and Gymnodinium cf.
uberrimum (both dinoflagellatesT^Stokes  (1980) and Van (1979) noted
that, in lakes in the Sudbury Region of  Ontario with pH values below
5.0, up to 50 percent of the biomass consisted  of dinoflagellates,
especially Peridinium limbatum and  Peridinium inconspicuum.  In Carlyle
Lake (pH 4.8 to 5.1) near Sudbury,  acidification experiments within
limnocorrals resulted in the proliferation of Peridinium limbatum (a 75
percent increase in biomass). At pH 4.0  this single  species accounted
for 60 percent of the total  phytoplankton biomass  (Van and Stokes 1978).
Hendrey (1980) investigated  3 lakes in  the Adirondack Region of New York
State.  In the most acidic lake (pH 4.9), Peridinium inconspicuum
comprised a significant fraction of the  biomass in the ice-free season.
Species of chrysophyceans (Phylum Chrysophyta)  were  also important.  The
dominance of dinoflagellates in many acidic waters has not been
adequately explained (National Research  Council Canada 1981).

     Dinoflagellates are not always reported as the  dominant algal group
in acidic environments.  In  a survey of  Florida lakes, Crisman et al.
(1980) reported that in the  most acidic  lakes  (pH  4.5 to 5.0) green
algae  (Phylum Chlorophyta) accounted for about  60  percent of the total
phytoplankton abundance.  However,  the  genus Peridinium was also
reported as a dominant taxon in these  lakes.   In Wavy Lake  (pH 4.3 to
4.8) near Sudbury, Ontario,  Conroy  et  al.  (1976) noted that
chrysophyceans (Phylum Chrysophyta) of the genus Dinobryon dominated.
Together chrysophyceans and  green algae  constituted  an average of 90
percent of the standing crop.  In two  non-acidic lakes sampled, these
algae accounted for only 21  and 23  percent of the  standing crop.  On the


                                  5-47

-------
                  TABLE  5-5.    SUMMARY OF  OBSERVATIONS  RELATING  SPECIES  DIVERSITY  AND  SPECIES  COMPOSITION
                                                  OF  THE PHYTOPLANKTON  COMMUNITY  TO  ACIDITY
               Location
              (reference)
    Reductions in
  species diversity
      Dominant species
       in acid water
      Species mi sslng
       1n acid water
     General
     comments
                 Swedish  West
                 Coast  (Aimer et
                 al.  1974 and
                 1978)
Numbers of species per 100 ml
  sample:
  pH 6-8:  30  to 80 species
  pH  < 5: about 12
  pH  < 4: 3
en

CO
 In most acid waters:
  dlnoflaqellates  (Pyrrophyta)
  Peridinium inconspicuum
  Gymnodinium cf.  uberrimum

 In a few lakes  with pH about 4:
  green algae (Chlorophyta) -
  Ankistrodesmus convolutus
  pocystis submarina
  Oocystf? lacustrTs

Other common species:
  chrysophyceans (Chrysophyta)
  Dinohryon crenulatum
  Dinohryon sertularia

  qreen algae (Chlorophyta)
  Chlamydomonas sp.
The classes Chlorophyceae
  (Chlorophyta)  and
  Chrysophyceae  (Chrysophyta)
  had greatly  reduced numbers
  of species

Absence of diatoms (class
  Bacillarlophyceae, Phylum
  Chrysophyta) and bluegreen
  alqae (Cyanophyta) at pH <5:
  Chroococcus  limneticus
  Heri snipped ia~tenu1ssima

Species common 1n oligotrophlc
  lakes, but absent at pH <6:
  bluegreen algae (Cyanophyta) -
    (lOmphpsphaeria lacustrls
  green algae  (Chlorophyta) -
    Scenedesmus  serratus
  chrysophyceans (Chrysophyta) -
    Dinobryon divergens
    Dinobryon bavarlcum
    DinobryoF borgel
    Dinobryori sucecicum
    Kephyripn spirale
    Stichogloea doederlelnll
  diatoms  (Chrysophyta) -
    Rhizosolenia lonqiseta
    CyclotellaTodanica
  cryptophytes (Cryptophyta) -
    Rhodomonas mi nuta
  dinoflagellates (Pyrrophyta) -
    Ceratium hirundinella
One stop survey of
  115 lakes in  August
  1972 and 60 lakes  1n

Greatest change in
  species composition
  occurred in the pH
  Interval 5 to 6
             2.
                 Swedish West
                 Coast  (Hultberg
                 and Andersson
                 1982)
Following  Hming, number of
  species  generally increased
Dominant species in acid,
  ollqotrophic lakes:
  dinoflagellates {Pyrrophyta) -
    Perldinlum Inconspicuum
    Gymnodinium sp.

Following Umlnq, the Importance
  of genus Perldinlum declined

Prevalent (30 to 40% of the
  blomass) 1n humlc lake:
  bluegreen alqae (Cyanophyta) -
    Merismopedia sp.
  green  alqae (Chlorophyta) -
    Oocystls sp.
Diatoms  Insignificant In all
  acid lakes

Following  liming, the Importance
  of species of green algae
  (Chlorophyta) and
  chrysophyceans (Chrysophyta),
  and, 1n  some cases, diatoms
  (Chrysophyta) Increased
pre- and post-1imlnq
  studies; long-term
  monitoring of four
  lakes

-------
                                                     TABLE  5-5.    CONTINUED
Location
(reference)
Reductions 1n
species diversity
dominant species
1n add water
Species missing
1n acid water
General
comments
    Southern Norway   Number of species Identified
    (Hendrey and       per lake:
    Wriqht 1976,       pH > 4.5:  10 to 25 species
    Leivestad et       pH  4 to 4.5: < in
    al. 1976)
                                                                                 Decrease  1n  importance  of
                                                                                   species  of  green  algae (Class
                                                                                   Chlorophyceae, Phylum
                                                                                   Chlorophyta)

                                                                                 No consistent  trend relating pH
                                                                                   to numbers  of  species of
                                                                                   diatoms  (Chrysophyta) or
                                                                                   bluegreen  algae  (Cyanophyta)
                                                                                                  One-stop  survey  of 55
                                                                                                    lakes  in  October 1974
4.
    Southern Norway
    (Raddum et al.
     1980)
                The number of algal  species
                  collected at any one time
                  was generally lower 1n clear
                  water acid lakes
                                                                                                  Periodic  sampling  of  13
                                                                                                    lakes throughout an
                                                                                                    entire  growing  season
5.
    Canadian
    Shield-Sudbury
    Ontario (Stokes
    1980)
                 Indices of both diversity and
                  species richness declined
                  with decreasing pH level
                              At pH < 5, up to 50% of the
                                blomass consisted of
                                dlnoflagellates (Pyrrophyta)
                                especially-
                                  Peridinium limbatum
                                  Peridinium InconspTcuum

                              However, this was not the case
                                1n a naturally acidic
                                dystrophlc lake
                                  In oligotrophic lakes with pH
                                  < 5, importance of species of
                                   green algae (Chlorophyta) and
                                   chrysophyceans  (Chrysophyta)
                                   decreased
                                 9  lakes  (pH  3.9 to 7.0)
                                    sampled  at monthly
                                    intervals  for 2 summer
                                    seasons

                                 Acidic lakes near
                                    Sudbury, Ontario have
                                    high concentrations of
                                    metals that may
                                    influence  phytoplankton
6.
Sudbury Region
of Ontario
(Van 1979)
Number of  taxa observed in
  acidic lakes was less than
  in non-acidic lakes
Biomass 1n acid lakes  dominated
  by a dinofl agellate
  (Pyrrophyta)  Peridinium
  Inconspicuum

Proportion of the biomass
  contributed by dlnoflagellates
  was correlated with  hydrogen
  1on activity, hut not with
  phosphorus concentration

Most common genera in  acid
  lakes:
  dinoflaqcllates (Pyrrophyta) -
    Peridinium
  cryptophytes  (Cryptophyta) -
    Cryptomonas
  chrysophyceans (Chrysophyta) -
    Ilinohryon
  green algae (Chlorophyta)  -
  Chlamydomonas Oocystis
Non-acidic oligotrophic  lakes
  typically dominated  by
  chrysophyceans (Chrysophyta)
  and diatoms (Chrysophyta),
  but In acidic lakes  sampled
  a dinoqlagellate (Pyrrophyta)
  dominated
Comparison of 4 acidic
  lakes with 10 non-acidic
  lakes.   Intensive
  sampling.   Samples
  collected  at a weekly  or
  bi-weekly  frequency at
  2 m depth  intervals at
  the deepest spot in each
  lake for one or two
  summer  seasons

The change 1n community
  structure apparently
  occurs  over a pH range
  of 4.7  to 5.6

-------
                                                                TABLE  5-5.    CONTINUED
Location Reductions in
(reference) species diversity
7. Sudbury Region
of Ontario
(Dillon et al
1979)
Dominant species Species missing
in acid water in acid water
Following liming, dominance
shifted from dinofl agel lates
(Pyrrophyta) and cryptophytes
(Cryptophyta) to the
chrysophyceans (Chrysophyta)
more typically observed in
circumneutral waters
neneral
comments
Three of the acidic lakes
sampled by Yan (1979)
were limed 1973-1975;
pH levels were raised
from <4.7 to above 6
               Sudbury  Region
               of  Ontario
               (Conroy  et al.
               1976)
(Jl
CD
In the two acidic lakes, a few
  qenera usually dominated the
  bicmass, resulting \n a low
  diversity index.   In the
  non-acidic lakes,  the
  blomass was more evenly
  distributed throughout a
  large number of qenera
In acidic Wavy  Lake,  the
  dominant genus  was  a
  chrysophycean (Chrysophyta)
  Dinohryon.  Most  of the
  species identified  in  Wavy
  Lake belonged to  the green
  algae (Chlorophyta) and
  chrysophyceans  (Chrysophyta).
  Together these  two  groups
  represented on  the  average
  90% of the  standing crop.
  In  the 2 non-acidic lakes,
  these 2 groups  accounted for
  only 217. and  23%  of the
  standing crop

In acidic Florence  Lake, a
  considerable  biomass of the
  bluegreen algae (Cyanophyta)
  Merismopedia  sp.  developed in
  August
Few or no diatoms (Chrysophyta)
  present in acidic waters while
  they dominated in non-acidic
  Millerd Lake and were
  significant in non-acidic
  Flack Lake

Acidic Wavy Lake had few blue-
  green algae (Cyanophyta)

In acidic Florence Lake,
  however, a considerable bloom
  of the bluegreen algae
  Merismopedia sp. developed in
  August

Both of the non-acidic lakes
  also had substantial
  populations of bluegreen
  algae, although of different
  species
               LaCloche
               Mountain Region
               of Ontario
               (Kwlatkowski
               and Roff 1976)
Strong relationship  between
  diversity of phytoplankton
  and pH level, with the
  diversity Index  dropping
  off sharply below  pH 5.6

All of the major groups of
  phytoplankton decreased
  markedly 1n their  numbers of
  species with Increasing
  add conditions.   Comparing
  the highest pH lake sampled
  (pH about 6.7) with the
  most add lake (pH about
  4.4), the numbers  of species
  of green algae (Chlorophyta)
  were reduced from  26 to 5;
  Chrysophyta (from  22 to 5;
  bluegreen algae  (Cyanophyta)
  from 22 to in.  Numbers of
  species of diatoms
The important  species  1n  each  lake  shifted according to pH level.
  In  the  more  neutral  lakes, the  green algae  (Chlorophyta)
  comprised  between  40 and  50% of the total algal  flora, with
  bluegreen  algae  (Cyanophyta)  accounting for only 30%.  In acidic
  lakes,  however,  bluegreen algae constituted about 60%, and green
  algae only about 25% of the  algal  flora.
                                                              Species common 1n acidic waters:
                                                                Bluegreen algae (Cyanophyta)  -
                                                                  Aphanocapsa sp.
                                                                  Chroococcus Prescottii
                                                                  fiscillatorla sp.
                                                                  Rhabdodgrma 1Ineare
                                                                Green algae (Chlorophyta)  -
                                                                  Carteria sp.
                                                                  Chiamydomonas sp.
                                                                  Chlorella eTTlpsoidea
                                                                  Closterium sp.
                                   Aphanocapsa sp.
                                   Chroococcus dispersus
                                   Chroococcus 1imneticus
                                   Osci1latoria  sp.

                                   Ankistrodesmus sp.
                                   Carteria sp.
                                   Chiamydomonas sp.
                                   Oocystis sp.
                                   Scenedesmus sp.
                                  6 la,kes sampled weekly
                                    for 2 months  in 1972
                                    and 1973 (lake pH
                                    range of 4.4  to 6.7)

                                  Lakes with similar pH
                                    values had similar
                                    species  composition
                                    as evaluated  by the
                                    coefficient of
                                    community and
                                    percentage similarlHy
                                    of community.  Thus,
                                    community structure
                                    in these lakes
                                    reflected the pH
                                    gradient

-------
                                                                TABLE  5-5.   CONTINUED
Location
(reference)
Reductions 1n
species diversity
Dominant species
1n acid water
Species missing
1n acid water
General
comments
            9.   Cont.
  (Chrysophyta) collected 1n
  samples were also greatly
  reduced 1n the two most
  acidic lakes relative to the
  other lakes sampled
                                                                 Cryptomonads (Cryptophyta)  -
                                                                 (considered by the  authors  as  1n the Phylum Pyrrophyta)
                                                                   Cryptomonas erosa
                                                                   Cryptomonas ovata
                                                               Dinoflaqellates (Pyrrophyta)
                                                                 species of the genera Peridinium
                                                                 and Glenodinlum,  although present
                                                                 1n some lakes, never reached
                                                                 significant proportions  1n  either
                                                                 acidic or non-acidic lakes
                                                                  Many of the species  of  diatoms
                                                                    (Chrysophyta)  common  to the
                                                                    more neutral  lakes were
                                                                    absent from acidic lakes
            10.
Ul
                LdCloche
                Mountain Region
                of Ontario
                (Van and Stokes
                1978)
                               Phytoplankton community
                                dominated by Peridinium
                                limbatum  (a dinoglagellate,
                                Phylum Pyrrophyta), and
                                Cryptomonas ovata
                                (a cryptomonad, Phylurn
                                Cryptophyta, but considered by
                                the  authors in the Phylum
                                Pyrrophyta)

                               These  2 groups formed between
                                50-90% of the biomass in all
                                collections
                                  Intensively  sampled one
                                    acid  lake, Carlyle
                                    Lake  (pH about 5.0),
                                    also  studied by
                                    Kwiatkowksi and Roff,
                                    1976.   Samples
                                    collected  at weekly
                                    Intervals  late June to
                                    late  July, 1974
            11. Ontario,  North
                of Lake Huron
                (Johnson  et al.
                1970)
Species  diversity lower 1n 2
  acid contaminated lakes than
  in the circumneutral lake
Many species of the Class
  Chrysophyceae (Chrysophyta),
  the class Myzophyceae
  (Cyanophyta; bluegreen algae),
  and diatoms (Class Bacillario-
  phyceae, Phylum Chrysophyta)
  developed in the circumneutral
  lake that were absent or
  occurred in only small  numbers
  in the 2 acidic lakes
Three lakes -  one
  circumneutral  and  two
  acidic lakes,
  acidified as a result
  of contamination by
  acid leachate  from
  processing of  local
  urani um ores

Associated with  low  pH
  levels were  high
  levels of calcium,
  sulfate, and nitrate,
  and, to a lesser
  extent, elevated heavy
 metals concentrations

-------
                                                                TABLE  5-5.    CONTINUED
Location
(reference)
Reductions In
species diversity
Homlnant species
1n acid water
Species missing
In acid water
General
comments
            12.  Adirondack
                Region of  New
                York  State
                (Hendrey 1980,
                Hendrey et al.
                1980b)
            13.  Adirondack
                Region  of Hew
                York  State
                (Charles 1982)
Total  number of  species
  identified 1n  each lake
  decreased with increasing
  acidity:
    clrcumneutral lake - 64
    intermediate      - 38
    acidic             - 27
Species of  the Class
  Chrysophyceae  (Chrysophyta)
  dominated the  bioraass of the
  most acidic lake, although
  dinoflaqellates  (Pyrrophyta),
  especially Peridinium
  inconspicuum.  comprised a
  significant fraction of the
  biomass  In the ice-free
  season
cn
en
ro
Numbers of species  of  green
  algae (Chlorophyta)  and blue-
  green alqae (Cyanophyta)
  decreased most markedly

Dinobryon spp. (Chrysophyta)  are
  the typical dominant phyto-
  plankters during  the summer 1n
  Adirondack lakes
                                                                All lakes with pH > about 5.8
                                                                  had euplanktonic diatoms
                                                                  (class Badllarlophyceae,
                                                                  Phylum Chrysophyta)  present in
                                                                  their surface sediments.
                                                                  Lakes with a lower pH had
                                                                  none.
Intensive sampling of
  three lakes -  one
  acidic (pH about 4.9),
  one intermediate (pH
  about 5.5), and  one
  circumneutral  (pH
  about 7.0)
                                                                   Survey of sediment
                                                                     diatom assemblages  and
                                                                     lake water
                                                                     characteristics  for
                                                                     39 lakes
            14.  Florida
               (Crisnan  et  al.
               1980)
Mean number of  taxa in acidic
  lakes was 10.8  vs. 16.5 for
  non-acidic lakes
In most acidic  lakes  (pH  4.5 to
  5.0), green algae  (Chloro-
  phyta)  accounted for 60% of
  the total  phytoplankton
  abundance; blue-green algae
  (Cyanophyta)  only  25%.
  Opposite pattern In circum-
  neutral lakes.

Highly acidic lakes  were
  dominated by:
  Green algae (Chlorophyta) -
    Scenedesmus
    Ankistrodesmus
    Staurastrum
                                                                   and several  species of small
                                                                   coccoid green  algae
                                                                 Rlnoflagellates  (Pyrrophyta) -
                                                                   Peridinium
In lakes of pH 6.5 to 7.0,
  bluegreen algae (Cyanophyta)
  made up 63% of total  phyto-
  plankton abundance, while
  green algae (Chlorophyta) were
  responsible for only 31%.
  Opposite pattern in acidic
  lakes
Survey of 13 poorly
  buffered lakes  1n
  northern Florida  with
  pH levels below 5.6,
  and 7 comparable  lakes
  in southern Florida
  but with pH levels
  above 5.6
            15. Missouri  (Lind
                and Campbell
                1970)
Reduced species diversity In
  add lake
                                                                   Study of a  very  acid
                                                                     lake (pH  3.2 to  4.1)
                                                                     affected  by  strip
                                                                     mining)

-------
                                                               TABLE 5-5.    CONTINUED
Location
(reference)
Reductions 1n
species diversity
Dominant species
In acid water
Species ml sslng
In acid water
General
comments
            16. England
                (Hargreaves et
                al. 1975)
                    Number of alga)  specfes
                      present per water was
                      negatively correlated with
                      total acidity
                                                                                                 Study of 15 waters  with
                                                                                                   pH levels of  3.0  or
                                                                                                   less; most affected  by
                                                                                                   strip mining
                                                                                                   activities
cn
 i
en
CO
17. Smoking Hills
    Region, North-
    west Terr.,
    Canada
    (Hutchinson et
    al.  1978)
In these very acidic ponds,
  phytoplankton populations
  were depleted
Even at these  extremely low pH
  levels,  some species of algae
  still commonly occurred:
  Euglenoids  (Euglerrophyta) -
    Euglena mutablis
  Diatom (Chrysophyta) -
    Nitzshia sp.
  Rinoflagellate (Pyrrophyta) -
    Gymnodinium ordinatura
Ponds affected by
  spontaneous burning
  of bituminous shale
  deposits.   pH values
  as low as  1.8
            18.
                New Zealand
                (Brock and
                Brock 1970)
                                                                                    Lower pH limit below which blue-
                                                                                      green algae (Cyanophyta) were
                                                                                      unable to grow is about 4.8 to
                                                                                      5.0.  However, at lower
                                                                                      temperatures (<56 C)  then In
                                                                                      the study waters, bluegreen
                                                                                      algae may be able to  tolerate
                                                                                      more acid pH values
                                                                                                 Analysis of algal
                                                                                                   populations along  the
                                                                                                   pH gradient as acidic
                                                                                                   (pH about 3.8) thermal
                                                                                                   waters and alkaline
                                                                                                   (pH 8.2 to 8.7)  hot
                                                                                                   springs flow into  a
                                                                                                   lake, Waimangu
                                                                                                   Cauldron

-------
other hand,  during the experimental acidification of Lake 223, Ontario,
from pH 6.7  to 7.0 in  1976  to  pH 5.4  in 1980, the importance of
chrysophyceans gradually  decreased, with a corresponding increase in
green algae  (Phylum Chlorophyta) (Schindler and Turner 1982).  Blooms of
Chlorella, a green alga,  within the hypolimnion (associated with
increased water clarity)  for the most part accounted for the increase in
importance of green algae.

     A dominance of blue-green algae  in acidic waters has also been
reported.  Conroy et al.  (1976) observed a bloom of blue-green algae
(Merismopedia sp.) in  acidic Florence Lake (pH 4.4 to 4.9) in Ontario.
Hultherg and Andersson (1982)  noted that blue-greens (again Merismopedia
sp.) were prevalent in humic acid  lakes in Sweden.  Stokes (1980) noted
that the typical dominance  of  dinoflagellates in acidic waters near
Sudbury, Ontario did not  apply to  naturally acidic, dystrophic lakes.
Thus, various circumstances, such  as  the presence of high concentrations
of humic organic materials  in  the  water, may be conducive to developing
populations  of blue-green algae under acidic conditions.

     Another approach  to  assessing the effect of acidification on
phytoplankton is to determine  which taxa common in circumneutral lakes
are missing  or reduced in waters at low pH levels.  Again, it is
difficult to generalize.  Of 11 papers dealing with this question (Table
5-5), in seven papers, diatoms (Class Bacillariophyceae, Phylum
Chrysophyta) were reported  to  be reduced in importance in acidic waters;
green algae  (Phylum Chlorophyta) in six papers, blue-green algae (Phylum
Cyanophyta)  in five papers, and chrysophyceans (Class Chrysophyceae,
Phylum Chrysophyta) in four papers.   In many cases, shifts in acidity
were also associated with a shift  in  major species within a given group
of algae.

     The observation that different species of algae are characteristic
of waters with different pH levels has also been used to  predict an
approximate lake pH level based upon  the composition of the algal flora
within the lake.  Because the  siliceous cell walls of diatoms (both
planktonic and benthic) are well  preserved in lake sediments, this group
of algae has most frequently been  used in  these analyses.  Use of this
technique for estimation  of historic  changes in pH is discussed  in
greater detail  in Section 5.3.2.2.2.

5.5.2.2  Changes in Phytoplankton  Biomass  and Productivity--Available
data on acidification  and primary  productivity in acidic  lakes yield no
clear correlation between pH level and algal biomass or  productivity.
Relative to primary productivity  and/or  phytoplankton biomass in
circumneutral lakes, levels in acidic lakes  in  some cases are reduced,
in others unchanged or even increased (Table 5-6).

     Field correlations must be  interpreted with care.  For  example,
lakes low in nutrients may  be  particularly  sensitive to  acidification.
At the same time, low  nutrient levels limit  primary  productivity.  As  a
result, any correlation between  lake  pH  level  and  phytoplankton  biomass
or productivity may reflect only  their common  association with  nutrient


                                  5-54

-------
                      TABLE 5-6.    THE  RELATIONSHIP  BETWEEN  LAKE ACIDITY  AND  PHYTOPLANKTON  BIOMASS  AND/OR
                                               PRODUCTIVITY—OBSERVED RESPONSE TO  LOW  pH
CJl
CJ1
Significant  Decrease

In six lakes near Sudbury, Ontario,
concentrations  of chlorophyll £ were
positively correlated (p < 0.01) with pH;
primary productivity (on a volumetric
basis) was lowest in the most  acidic lake
(KwiaUowski  and Roff 1976).

In three Adirondack lakes, the  most acidic
lake (pH 4.7 to 5.1) had the  lowest level
of chlorophyll  £; the least acidic lake had
the highest  level of primary  productivity
(on an areal  basis) (Hendrey  1980).

In a survey  of  Florida lakes,  mean
chlorophyll  £ concentrations  were
siqnficantly lower in acidic  lakes (1.88
nig m"3) than in non-acidic lakes (7.53 mg
ra-3) (Crlsnan et al. 1980).
Significant  Increase

In 58 lakes  along the west coast of Sweden,
the larqest  biomass of phytoplankton occurred
in the most  acidic lakes (pH 4.5),  and  the
lowest biomass at Intermediate pH levels
(pH 5.1 to 5.6)  (Aimer et al. 1978).

In acidification experiments within Hmno-
corrals 1n Carlyle Lake (pH 4.8 to  5.1), near
Sudbury, Ontario, after 28 days the biomass
of phytoplankton was highest at the lowest pH
tested (pH 4.0), and lowest at pH 6.0 and 6.5
(Van and Stokes  1978).

During experimental acidification of Lake 223
(Experimental Lakes Area In western Ontario),
the pH decreased gradually from pH  6.7  to 7.0
1n 1976 to pH 5.4 1n 1980.  Over that time
period, chlorophyll and algal bionass Increased
significantly, associated with hypoHmnetlc
algal blooms of  Chlorella, and apparently in
response to  increased water clarity (Schlndler
and Turner 1982).
The National Research Council  of Canada  (1981) collated
measurements of algal biomass  and productivity for
oligotrophic lakes in the Canadian Shield Region of
Ontario.   Neither biomass nor  production were significantly
correlated with pH.  Algal biomass was significantly (p <
0.01)  correlated with total phosphorus concentration.

In the fall of 1973, the pH of one Ontario  lake, Middle
Lake (pH  about 4.4) was raised to around 7.0 by additions
of base.   Total phosphorus levels did not increase, no,r did
phytoplankton biomass (Van 1979).  Experimental increases
in phosphorus levels 1n acidic lakes (with  or without
neutralization) have, however, induced significant
increases in phytoplankton biomass (Dillon  et al. 197R,
Hendrey et al. 1980b).

Within eight plastic enclosures in Eunice lake, an
oligotrophic lake with pH 6.5-in British Columbia, add
addition  (minimum pH 5.5) resulted 1n no significant change
in chlorophyll content.  Additions of acid  plus nutrients
(minimum  pH 5.0) increased algal biomass (Marmorek 1983).

In three  Swedish lakes sampled from 13 to 15 May 1975,
rates  of  phytoplankton production per volume of water were
somewhat  lower In the most acidic lake  (pH  4.6).  However,
because of greater water transparency in this acidic lake,
measurable primary productivity was maintained to a greater
depth. Levels of primary productivity on an areal basis,
per square meter of lake surface, were similar 1n all three
lakes  (Aimer et al. 1978).

In 13  lakes in southern Norway, chlorophyll ^ content was
not significantly correlated with lake pH  (Raddum et al.
1980).

-------
status and not a cause-and-effect relationship  between  pH and
phytoplankton response.

     Three investigators have reported  lower  levels of  phytoplankton
biomass and/or productivity in acidic lakes than  in circumneutral lakes,
based on measurements from six lakes near  Sudbury, Ontario  (Kwiatkowski
and Roff 1976), three lakes in the Adirondacks, New York (Hendrey 1980),
and a survey of Florida  lakes (Crisman  et  al. 1980).  None  of these
studies included a simultaneous analysis of nutrient availability. In
addition, careful  examination of data on primary  productivity collected
by Kwiatkowski and Roff  (1976) indicates that,  with the exception of two
lakes, no clear relationship exists between productivity and lake pH.
The productivity reported for the most  acidic lake (pH  4.0  to 4.6, about
3 mg C m~3 hr"l) is well  within the range  normally observed in
non-acidic lakes in the  region (0.3 to  6.9 mg m-3 hr"1) (National
Research Council Canada  1981).  Values  Kwiatkowski and  Roff measured in
the five remaining lakes were well  above the  norm.  Thus, no conclusive
data are available to support the hypothesis that acidification results
in decreased algal biomass and productivity.

     In contrast,  three  field surveys and  four  field experiments suggest
that acidification causes no change, or perhaps even an increase, in
phytoplankton biomass (Table 5-6).  Surveys in  Ontario  (compiled in
National  Research  Council  Canada 1981)  and Norway (Raddum et al. 1980)
found no correlation between lake pH and algal  biomass; in  Sweden (Aimer
et alI. 1978), the  largest biomass occurred in the most  acidic lakes.
Acidification experiments within limnocorrols yielded no change
(Marmorek 1983) or an increase (Van and Stokes  1978) in algal biomass.
Experimental acidification of an entire lake  (Lake 223  in the
Experimental Lakes Area,  Ontario)  also  was associated with  a significant
increase in phytoplankton biomass (Schindler  and Turner 1982).

     Increased accumulations of phytoplankton in  acidic waters may
reflect either an  associated increase in the rate of production or a
decrease in the rate of  loss (e.g., decreased predation).   No studies
report an increase in phytoplankton productivity with acidification or
in acidic lakes, although data are not  abundant.  Two field surveys
suggest no relationship  between lake pH and primary productivity (Table
5-6).   Predator-prey interactions within the plankton community are
complex.   Detailed studies related  to effects of acidification on
phytoplankton mortality  are not available.  Potential changes, based on
ecological  theory, are discussed in Section 5.5.4.

      In a number  of laboratory studies, primary productivity in algal
cultures has been  shown  to be a function of pH  level (e.g., Hopkins and
Wann 1926,  Bold 1942, Sorokin 1962, Brock  1973, Loefer  1973, Moss 1973,
Cassin 1974).  For each  species, growth responses to pH form an inverted
U-curve,  with an optimum pH level  for maximum growth, and significantly
lower growth rates at lower and higher  pH  levels.  The  optimum pH for
growth varies significantly between species.  Moss (1973) found a lower
limit for growth of most algal species  at  pH levels above 4.5 to 5.1.
However,  three of  33 species tested grew well at  pH levels  below 4.0.


                                  5-56

-------
 Sixteen of 33 species were capable of significant growth  below  pH  5.0.
 No distinct differences were found between groups or  types  of algae  with
 regard to minimum pH tolerated (Moss 1973).  Blue-green algae in general
 (Phylum Cyanophyta), however, may be less tolerant of pH  levels below
 5.0  (Bold 1942, Brock 1973, Moss 1973).

      The presence of an alga at a low pH level  does not necessarily
 imply a preference for acidic conditions or that photosynthesis and
 growth are optimal (Hendrey et al. 1980b).  The proliferation of
 Perl dim'urn species at pH levels 4.0 to 5.0 does not mean  that these
 organisms do best at pH levels 4.0 to 5.0, only that its  competitors do
 less well.

      The growth of algae in acidic waters indicates a physiological
 ability to tolerate low pH levels, and conditions associated with  low
 pH,  e.g., a shift in the form and availability  of aqueous inorganic
 carbon and other necessary plant nutrients, and increased concentrations
 of some metals, especially aluminum (Chapter E-4, Section 4.6).
 Research has not yet clearly defined physiological responses of algae to
 acidic conditions, or why some species can tolerate higher acidity than
 others.

 5.5.3  Effects of Acidification on Zooplankton

      Results from 14 field surveys of zooplankton communites are
 summarized in Table 5-7.  In each study, acidic lakes had fewer
 zooplankton species (e.g., Figure 5-3).  In Norway, clearwater  lakes
 with pH levels below 5.0 contained 7.1 species  on the average as
 compared to 16.1 species in less acid lakes (pH > 5.5)  (Overrein et  al.
 1980).  Sprules (1975a,b) found nine to 16 species of crustacean
 zooplankton in lakes with pH levels above 5.0 in the LaCloche Mountain
 Region of Ontario, but only one to seven species in acidic lakes,  pH <
 5.0. In the northeastern United States, lakes with pH below 5.0
 contained three to four species of planktonic crustaceans;  lakes with pH
 above 5.5 contained six to 10 species (Confer et al.  1983). The
 greatest change in species number and types of dominant species occurred
 between pH 5.0 to 5.3 (Sprules 1975a, Roff and Kwiatkowski  1977).

      Likewise, experimental acidification of Lake 223, Ontario, from pH
 6.7  to 7.0 in 1976 to pH 5.4 in 1980, resulted in a decline in  the
 number of zooplankton species present in the lake.  A decrease  in  the
 mean epilimnetic pH from 6.1 to 5.8 was associated with the
 disappearance of one species; decrease to pH 5.6 led to the loss of  two
 more species (Mailey et al. 1982).

      For the most part, species dominant in acidic lakes  are also
 important components of zooplankton communities in non-acidic lakes  in
 the  same region.  There is no invasion of new species.

      Certain species of planktonic rotifers of the genera Keratella,
 Kellicottia, and Polyarthra tolerate acidic conditions and can  be  found
 in the pH range 4.4 to 7.9.  In Scandinavia, species common in  acidic


                                   5-57
409-262 0-83-14

-------
               TABLE 5-7.   SUMMARY OF OBSERVATIONS RELATING SPECIES  COMPOSITION,  SPECIES  DIVERSITY,  AND
                                        BIOMASS OF THE ZOOPLANKTON  COMMUNITY TO ACIDITY
en
en
00
Changes in species composition and abundance of:
Location
(Reference)
1. Southern
Sweden
(Aimer et
al. 1974
and 1978)
General
Observations
Number of species
lower add lakes
In acid lakes.' often
just a few species
occur but the number
of Individuals can be
rather great
In highly acidic
lakes (pH <5)
Polyarthra remata
Bosmina coreqoni , and
niaptomus qracilis
often dominate
Rotifers
Polyarthra remata,
Polyarthra vulqarls,
Keratella cochlearis, and
Kellicottia lonqispina
common at most pH levels,
4.4 to 7.9
Polyarthra remata dominant
in several lakes with pH
< 5.5
Conochllus unlcornls
present in many lakes but
less prevalent in acid
waters
Many of the other rotifers
Cladocerans
Bosmina coregonl common
and occurred at all pH
levels
All Daphnia species
were sensitive to low
pH levels. Only a few
individuals found at
pH < 6
niaphanosoma
brachyurum, Holopedium
qlbberum, and Leptodora
kindti common but
mainly at pH > 4.9
Bythotrephes longimanus
Copepods
Dlaptomus gracllis and
Cyclops spp. common at all
pH levels
Heteroeope appendlculata
occurred mostly at pH>5.5
Others Comments
One-stop survey
of 84 lakes In
August 1971
Samples collected
with 75 11 mesh
net
                                         appear to have preferences
                                         above 5.5
found more frequently
in lakes with pH < 5.4.
At higher pH levels,
fish predation may keep
the population at low
levels

Common 1n non-act die
lakes:
 Dlaphanosoma

2. Southern
Sweden
(Hultberg
and
Anderson
1982)

In acidic lakes,
zooplankton community
dominated by a few
species

Dominants 1n acid lakes:
Polyarthra spp.
Keratella cochlearis
Kelllcottia longisplna
Common after liming:
Polyarthra spp.
Keratella cochlearis
Asplanchna prlodonta
Conochilus mincornis

Holopedium
Daphnia cri stata
Bosmina
Dominant 1n acid lakes:
Bosmina coreqoni
Common after liming:
Bosmina coreqoni
niaphanosoma sp.
Daphnia cristata
Limnosida froutosa
Holopedium qibberum
Ceriodaphnia
quaaranqula

Dominants in add lakes:
Eudlaptomus gracllis
Cyclops spp.
Common after liming:
Eudiaptomus qracilis

Pre- and post-
liming studies.
Effects of lining
on zooplankton
are difficult to
evaluate due to
simultaneous
rotenone treat-
ments

-------
                                                              TABLE  5-7.    CONTINUED
Changes 1n species composition and abundance of:
Location
(reference)
3. Southern
Norway
(Hendrey
and Wright
1976)
General
observations
Total number of
species collected
decreased with
decreasinq pH
Rotifers Cladocerans
Daphnia galeata absent
at pH < 6.9
Eubosmina lonqispina
common at al 1 pH
levels, 4.1 to 7.7
Hoi oped 1 urn gibberum
occurred frequently at
Copepods Others
Eudiaptomus qracilis
common over wide ranqe of
pH, 4.1 to 6.6 most
frequently dominant at low
pH levels; rarely dominant
at pH>5.5
Heterocope sal lens
occurred pH 4.1 to 6.6
Comments
One-stop survey
of 57 lakes
durlno fall 1974
Samples collected
with single
vertlcle haul of
a 75 v mesh
net
01

01
UD
pH levels 4.2 to  7.2

Daphnia longispina
appeared in samples pH
4.6 to 6.8
Ancanthodiaptomus
denticornis and
Mixodiaptomus lacinlatus
did not occur at pH < 5

Cyclops scutifer appeared
at pH 4.6 to 7.7
4. Norway
(Raddum et
al. 1980;
Raddum,
1978;
Hobaek and
Raddum
1980)
Number of species
lower in add lakes.
In southern Norway,
clear-water lakes with
pH < 5 held on the
average 7.1 species;
equally acid humic
lakes, 11.7 species;
less acid (pH > 5.5)
clearwater lakes,
16.1 species on
average
All major groups
contributed to the
lowered number of
species, but
cladocerans
apparently most
affected
Species occurring with
equal frequency in acid
and non-acid clearwater
lakes:
Kellicottia lonqispina
Keratella serrulata
Species more frequent in
non-acidic lakes:
Conochllus spp.
Asplanchna sp.
Keratel 1 a
cochlearis
Keratella
hlemalis
Species more frequent in
acidic lakes:
Polyarthra spp.
Species occurinq with
equal frequency in acid
(pH < 5) and non-acid
(pH > 5.5) clearwater
lakes:
Bosmina (Eubosmina)
lonqispina
Species more frequent
1n non-acid lakes:
Holopedium gibberum
Diaphanosoma
TTrachyurum
Ceriodaphnia
nuadranqula
Daphnia lonqispina
Daphnia galeata
Bythotrephes
lonqimanus
Polyphemus pediculus
Leptodora kindti
Species occurrlnq with Chaeborus
equal frequency in add flavicans
and non-acid lakes: absent 1n
Eudiaptomus qracilis clearwater
Heterocope saliens acid lakes
Species more frequent in
non-acid lakes:
Cyclops scutifer
Cyclops abyssorum
Mesocyclops leuckartl

Survey of 27
lakes; sampled
(3 vertical net
hauls, with 90
u mesh net , per
visit) from June
to September 1977
- 1979

-------
                                                                TABLE 5-7.    CONTINUED
Location
(reference)

General
observations
Changes
Rotifers
in species composition and abundance of:
Cladocerans Copepods Others Comments
en
 i
4. cont.        Some species tolerate
               acid conditions  1n
               the presence of
               humus, but are ahsent
               from add clearwater
               lakes

               The species number of
               filter-feeders
               reduced in clear-
               water acid lakes.
               Changes for
               rapturial species
               not as obvious
5. Southern Lower abundance of
Norway Daphnia longlsplna and
(Nilssen Daphnia lonqlremus at
1980) pH<5.5
Bosmina longlsplna a
dominant species at all
pH levels, 4.5 to 7.0
Leptodora klndtl absent
at pH levels 5.0 to 5.5
In sediment cores from
lake with pH 4.2 to
5.0, Daphnia Ignqispina
present below 3 to 4
cm, hut absent in
surface sediments and
In the lake
Eudiaptomus qracilis
a dominant species at
all pH levels, 4.5 to 7.0
Cyclops scutifer
common at all pH levels,
4.5 to 7.0
Heterocope sal lens
increased 1n abundance
1n more acid lakes
(pH < 5.0)
Chaoborus
flavicans
absent at pH
levels <
4.5 to 5.0
In sediment
cores from
lake with pH
4.2 to 5.0,
Chaoborus
flavicans
remnants
common below
3 to 4 cm,
but absent 1n
surface
sediments and
1n lakes
Samples collected
with 90 v mesh
net June -
October 1973 and
1974 at fourth
nlqht intervals

-------
                                            TABLE 5-7.   CONTINUED
en
en
Changes In species composition and abundance of:
Location General
(reference) ohservations
6. LaCloche Signifacant reduction
Mountain in number of species
Region of and numbers of
Ontario individuals at lower
(Roff and pH levels (pH 4.4 to
Kwiatkow- 4.8)
ski, 1977)
Diversity index
declined sharply
below pH 5.3
Mean size of
crustacean
zooplankters
identical in acid vs.
non-acid lakes
Rotifers
Standing crop of rotifers
reduced at pH levels 4.4
to 4.8
In all lakes with pH>5.8,
rotifers represented by a
variety of species with no
one species being dominant
In highly acidic waters
(pH about 4.4), Keratella
turocephala dominated. As
the pH increased,
Keratella cochlearis.
Kellicottia bostoniem's,
and Kellicottia longispina
increased in occurrence
Polyarthra euryptera and
Polyarthra dolichoptera
rare at pH 4.7 to 5.0;
absent pH<4.4
In highly acidic waters
(pH about 4.4), Keratella
taurocephala dominated.
As the pH increased,
Keratella cochlearis,
Kellicottia bostoniensis,
and Kellicottia longispina
Increased in occurrence.
Polyarthra euryptera and
Polyarthra dolichoptera
rare at pH 4.7 to 5.0;
absent pH < 4.4
Cladocerans
Standing crop of
Cladocerans reduced at
pH levels below 5;
maximum at pH 5 to 6
Leptodora klndti found
only at pH>5.0
Daphnia galeata
mendotae'Daphnia
retrocurva, and
Diaphanosoma
leuchtenbergianum found
in all lakes but rare
at pH 4.4 to 4.8
Bosmina longirostris,
Eubosmia tubicen, and
Holopedium gibberum
common in all lakes
Copepods Others
Standing crops of
cyclopoid copepods but not
calanoid copepods reduced
at pH levels 4.4 to 4.8
Diaptomus minutus occured
abundantly in all lakes at
all pH levels, 4.4 to 6.0
Diaptomus oregonensis and
Epischura lacustris only
encountered in lakes with
pH>5.6
Cyclops bicuspidatus
thomasi and Hesocyclops
edax found in all lakes,
pRT.4 to 6.8
Comments
Six lakes with
pll levels 4.0 to
7.1 samples at
weekly intervals
June and August
197? and Hay and
July 1973
Vertical haul
with 60 |i mesh
net; and
Schindler-Patalas
trap at various
depths.

-------
                                                                TABLE  5-7.   CONTINUED
en
 i
01
Changes in species composition and abundance of:
Location
(reference)
7. LaCloche
Mountain
Region of
Ontario
(Sprules
1975a,
1975t>)
General
observations
Above pH 5.0,
communities with 9-16
species, 3-4
dominants in lakes
with pH < 5.0, 1 to 7
species with only 1
or 2 dominants
Discontinuity in
species distribution
at pH 5.0 to 5.2.
64% of all species
identified occurred
never or rarely at
pH < 5.0.
Rotifers Cladocerans
Tolerant species
distributed Independent
of pH:
Bosmlna
Diaphanosoma
leuchtenhergianum
Holopedium gibberum
Never occur pH < 5.0:
leptodora klndtl
Daphnia galcata
mendotae
Daphnia retrocurva
Daphnia ambiqua
Daphnia lonqiremis
Copepods Others
Tolerant species
distributed independent of
pH:
Mesocyclops edax
Cyclops bicuspidatus
thomasi
Diaptomus mlnutus
Never occur pH < 5.0:
Tropocyclops prasinus
mexicanus
Epischura lacustris
Diaptomus oregonesis
Comments
One-time sampling
of 47 lakes f'-om
July to early
September 1972 -
1973 .
Vertical hauls
with either
75 v or 110 u
mesh net
pH ranged from
3.8 to 7.0
                         In some lakes, only
                         Diaptomus  minutus
                         remains.   Above pH
                         5.0, pH had  little
                         effect on  tolerant
                         species and  only a
                         slight effect on the
                         total number of
                         species

                         In regression
                         analyses,  pH alone
                         accounted  for 53% of
                         the variance in
                         number of  species
Occur primarily in
lakes with  pH < 6.0:
  Polypemus pediculus
  Daphnia~catawba
  Daphnia pulicaria
Oiaptomus rcinutus dominant
in most lakes pH < 5.0; In
some cases  the only
species present

-------
                                                      TABLE 5-7.   CONTINUED
GO
Changes
Location General Rotifers
.(reference) observations
8. Sudbury Numbers of species
Region of reduced 1n acid lakes
Ontario (pH 4.1 to 4.4) with
(Yan and an average of only
Strus 1980) 3.7 species per
sample vs 10.6 in
non-acid lakes
Total community biomass
lower In acid lakes
than in nonacid lakes.
Decreased biomass
resul ted from both a
decrease In numbers
(except in one lake)
and the small size of
the community domi-
nants (primarily
Bosmina longirostris)
In acid lakes
The greatest reductions
were observed in the
lake with the highest
metal concentrations
Contamination with
copper and nickel
appeared to have some
In species composition and abundance of:
Cl adocerans
Major species 1n non-
acid lakes:
Bosmina longirostris
Holopedium qibberum
Diaphanosoma
leuchtenbergi anum
Daphnia galeata
mendotae
In acid waters, Bosmina
longirostris accounted
for an average of 79%
of the total crustacean
biomass vs 3% in
non-acid lakes
In acidic Clearwater
Lake zooplankton
community characterized
by the importance of
Bosmina longirostris,
and the absence of
Daphnia sp. and the
other common
cl adocerans Holopedium
gibberum and
Diaphanosoma
leuchtenbergianum
Copepods Others
Major species 1n non-acid
lakes:
Cyclops bicuspidatus
thomasi
Tropocyclops praslnus
mexicanus
Diaptomus minutus
Copepods contributed an
average of 65% of the
total biomass and 85% of
the total individuals 1n
non-acid lakes
Diaptomuro minutus formed
between 44 and 73% of all
crustacean zooplankton,
and dominant in all
non-acid lakes
In acidic Clearwater Lake,
zooplankton community
characterized by the
absence of Tropocyclops
prasinus mexicanus and
Mesocyclops edax and by
the scarcity of Cyclops
bicuspidatus thomasi and
Diaptomus minutus
Comments
Sampled 4 acidic
lakes (pH 4.1 to
4.4) and one less
acidic lake (pH
5.7) in the
vicinity of
Sudbury plus 6
non-acidic lakes
(pH 5.7 to 6.6)
in Muskoka-Halib
urton Reglor of
Ontario
Acidic lakes also
have high levels
of copper and
nickel which nay
adversely effect
zoo-plankton
Samples collected
summers 1973-1977
as vertical hauls
with 80 u mesh
tow net and at 2-
to 3-n Intervals
with a plastic
trap
                     plankton community over
                     and above effects of
                     low pH

-------
                                                                 TABLE  5-7.    CONTINUED
en
 i
OY
Changes In species composition and abundance of:
Location
(reference)
9. Sudbury
Region of
Ontario
(Van et
al. 1982)
10. Georgian
Bay
Region of
Ontario
(Carter
1971)
General Rotifers
observations
In the non-add lake, Rotifers generally form
collections Included only about IX of total
7 species on the zooplankton biomass In
average vs 3.7 from non-acidic oligotrophic
the add lake lakes in the Sudbury area
Standing crop
generally greater 1n
very acid (pH 4.7 to
5.2) than in slighly
acid or alkaline
ponds
About 14 species
Cladocerans
Cladocerans unimportant
1n non-acid lake,
forming <5% of the
average biomass
In the add lake,
Bosmina longirostris
comprised 403 of the
crustacean zooplankton
biomass
Species occurring in
all ponds independent
of pH:
Bosmina longirostris
Ceriodaphnia
quadrangula
Diaphanosoma
leuchtenbergianum
Copepods
Diaptomus minutus major
contributor to total
zooplankton biomass 1n the
non-acid lake. Cyclops
scutlfer, Mesocyclops
edax, and Tropocyclops
prasinus mexicanus also
important
In the acid lake,
Diaptomus- minutus
comprised 321 of the
crustacean zooplankton
biomass. Chydorus
sphaericus and Cyclops
vernal is also common

Species occuring in all
ponds independent of pH:
Diaptomus reighardl
Cyclops vernal is
tycTops bicuspidatus
thomasi
Mesocyclops edax

Others
2 individuals
of Chaoborus
flavicans
collected in
non-add
lake.
In the acidic
lake,
Chaoborus
flavicans,
Chaoborus
albatus, and
Chaoborus
americanus
occurred

Commons
Pre- and post-
fertll Ization
study of one
acidic (pH about
4.6) and one non-
acidic (pH about
6.0) lake only
pre-fertll Izotion
data included
here
Samples collected
in plexiglass
trap at 1-, 4-,
and 7-m; 76 u
mesh net
32 ponds sampled
to 10 times ove-r a
3-year period.
Samples collected
with a Clarke-
Dumpus or trans-
parent zooplankton
trap
                          present 1n non-add
                          waters were absent
                          from the very  acid
                          lakes
Species  occurring only
in less  acid  and
alkaline ponds:

  Leptodora kindta
  Daphnia ambiqua~
  Daphnla retrocurva
  Ceriod
    Tac"u
  jipsmina coregonl
    coreqom
  Holope'jrum  gibberum
Species occuring only in
less acid and alkaline
ponds:

  Epischura  lacustrls
  Diaptomus  minutus
  Diaptomus  oregonensls
  Tropocycfops prasinus
    mexicanus
The acidity  in
these waters Is
attributed mainly
to large amounts
of organic  (humlc)
acids

-------
                                                                TABLE  5-7.    CONTINUED
             Location
            (reference)
                                                            Changes  In species composition and abundance of:
  General
observations
                                                      Rotifers
                                                                              Cladocerans
Copepods
                                                                                                                             Others
                                                                                                                                        Comments
          10.  cont.
tn
                                           Bosmlna Ipnglrostrls
                                           was the most conslst-
                                           ently abundant crusta-
                                           cean In all ponds.  Its
                                           greatest numbers  were
                                           usually found In  the
                                           very add ponds.
11. Smoking
Hills area
in North-
west Terr.,
Canada
(Hutchlnson
et ai. 1978)
12. Adirondack
Region of
New York
State and
White
Mountain
region of
New Hamp-
shire
(Confer et
al. 1983)
The only zooplankton
present In these very acid
waters (pH 2.8 to 3.6)
were rotifers Branchlonus
urceolarls the dominant
form
number of zooplankton
species and
zooplankton biomass
sharply related to pH
(p < 0.01). For each
unit decrease on pH
lakes contained on
the average 2,4 fewer
species and 22.6 mg
dry wt m2- less
zooplankton biomass

Identified pH range for
distribution of
species:
Bosmuna longlrcotrls,
5.2-5.1
Bosmina corrgoni ,
4.5-7.2
Daphnia catawba,
S.Z-6./
Daphnia ambigua,
4.5-6.6
Holopedium gibberum,
4.5-7.2
Diaphanosomean
leuchtenberglanum,
4.7-6.6
Polyphemus pedlcutus,
'4.7-7.2 	
Leptodora kindtH
6.3-6.4

Diaptomus minutus domi-
nated at pH < 5,
Identified pH range for
dlstlbutors of species:
Diaptomus minutus,
4.5-7.2
Cyclops scutlfer.
5.4-7.2
Mesocyclops edax,
4.5-6.7
Tropocyclops praslmus,
5.3-7.2
Epischura lacustris,
5.3-7.2

Chaoborus sp. Two stop survey
occurred (July and August
throughout 1979) of 20 lakes
the pH range (10 in NY; 13 'n
4.5 to .6.7 NH)
Samples collected
with Van Oorn
sampler at 3-5
depths per lake
All headwater
lakes; pH ranged
from 4.5 to 7.2

-------
                                              TABLE 5-7.   CONTINUED
tn
 i
CTl
Changes In species composition and abundance of:
Location General Rotifers Cladocerans
(reference) observations
13. Great Identified pH range for
Britain distribution of
(Lowndes species:
1952) Diaphanosoma
brachyurum,
4.3-9.iJ
Daphnia pulex,
5.8-9.2
Daphnia longispina,
HP? .2
Ceriodaphnia
retlculata, 6.2-9.2
Ceriodaphnia
quadrangula,
3.2-9.2
Bosmina longirostrls,
6.9-9.2
Polyphemus pedlculus,
1.6-9.2
Bythotrephes
longimanus, 6.7-7.2
Leptodora kindtl ,
6.7-8.4
14. Great Number of species Fond 1n waters with
Britain lower In low pH pH < 5.0:
(Fryer waters In pH range 3 Diaphanosoma
1980) to 7. brachyurum
Ceriodaphnia
quadranguTa
Bosmina coregoni
Polyphemus pedTculus
Copepods
Identified pH range for
distribution of species:
Diaptomus gracills,
"'4.7-9.2 	
Cyclops abyssorum,
6.2-7.3
Cyclops vernalis,
4.4-9.2
Cyclops bicuspidatus,
4.1-9.?
Found in waters with
pH < 5.0:
Cyclops abyssorum
Tropocyclops prasinus
Others Comments

One-time sampling
of 70 w.itcr
bodies
Acidity attrib-
utable primarily
to high levels of
organic (humic)
acids

-------
                                                     NUMBER  OF  SPECIES  PER  COLLECTION
en







3=>
Q.
Q;
c+
CD
Q.
-+>
o

z
PQ
O
o

(— >

oo
1— >

































o
-h

<
QJ
-s
—I.
O
c
— 1
QJ
fD
to

__i
O
QJ
fD
Q.
O

rh
•^p
__>
fD
O
01
3
O.

QJ
3

oo
3-
_ i*
fD
_j
Q.
V

0
3
QJ

_i.
O








-o
QJ
3
rt-
O
3
O
O
-s
to
ft*
QJ
O
fD
QJ
3
Q.
fD
3

.—i,
t
n
fD
Q-
_j.
3
fD
D->
0
3-

o
o
—t
._— 1
fD
O
rt-
_j.
O
3
cu
M*
3
Q.
-a
re
o -> oo r>o en
~n
IQ
-s
-P"
V "o
co
fD

Q)
.^.
0
3 en
V) .
3- o
•a
cr
fD
{-^
^
fD
fD
3

r+
T cn
•§: b
Qi
<
fD
-s
IQ
fD
3
C
i-
fD .
"5 O

0
-h

VI
~&
fD
O
—j*
fD
CO

3, 00

O


1 1(1111 1

D -h X D
c_ m r-
D 2 ^ "^
D Of D D K P
D rt o
D ° D D co o
- D OD $ m
D OD D _
D D 5 o
a°S ° S 1
D D D O D *•
a 5
D D D °°
a


X
o
a o
t>
a a
a XCD c> a
a
x a a
a

o &&
+ X t> 4- 0 +
OX O
o
x x
X D X D







.







•^


c
1

0 O
oo :r
c.: 3> r~
o r— m
CO CD CD
f~* f~— |-rj
5 5 0
> i

>














4*























-------
lakes were Keratella  cochlearls, Keratella  serrulata, Kellicottia
longispina, Polyarthra  remata, and Polya'rthra vulgaris.  Species reduced
in abundance with acidification  indtided Asplanchna priodonta,
Conochilus unicprm's, Conochilus mincorm's, and KeFatella hTemalis
(Aimer et al. 1974,  1978;  Raddum 1978;  Hultberg and Andersson 1982).  In
Ontario, species of Keratella and Kell icottia were also important in
acidic lakes (Keratella taurocephala,  Keratella cochlearls, Kellicottia
bostoniensis,  Kellicottia longisp'Tinia) (Roff and KwiatkowskT 19/7).
Experimental acidification of Lake 223, to  pH 5.4, resulted in increased
numbers of Polyarthra yulgaris,  Polyarthra  remata, Keratella
taurocephala, and Kell icottia longi'spina  (Malley et al. 1982).

     Among crustacean zooplankters,  several  species of cladocerans
appear sensitive to acidity.  In particular, field surveys (Table 5-7)
indicate that many species of the genus Daphnia are absent or uncommon
below pH 5.5 to 7.0 (Lowndes 1952, Carter 1971, Aimer et al. 1974,
Sprules 1975a, Hendrey  and Wright 1976, Hobaek and Raddum 1980, Van and
Strus 1980, Nilssen 1980).  In addition,  in  laboratory experiments with
Daphnia magna and Daphnia  pal ex, reductions in survival and
reproduction, and physiological  imbalances  occurred at pH levels below
5.0 to 6.0  (Davis and Ozburn 1969, Potts  and Fryer 1979).

     Counterbalancing the  scarcity of daphnids in acidic lakes is an
increase in the abundance  of species  of the cladoceran genus Bosmina.
In Scandinavia, Bosmina coregoni and  Bosmina longispina were common at
all pH levels greater than 4.1 to 4.5 (Aimer et al. 1974, Hendrey and
Wright 1976, Raddum 1978,  Hultberg and Andersson 1982).  In Ontario,
Bosmina longirostris accounted  for a  large  fraction of the zooplankton
biomass in  acidic lakes (pH < 5)  (Carter  1971, Roff and Kwiatkowski
1977, Van and Strus 1980,  Van et al.  1982).  Other cladocerans common in
temperate,  oligotrophic lakes  (e.g.,  Diaphanosoma brachyurum,
Diaphanosoma leuchtenbergianum,  Leptodora kindti, Holopedium gibberum,
Polyphemus  pediculus, Ceriodaphnia quadrTngula. and Bythotrephes
Tongimariu?) often are less abundant  in waters with  pH 1 eve!s below 4.7
to 5.0 (Table 5-7).  Acidification of Lake  223 down to pH 5.4, however,
resulted in no consistent  trends in  the numbers of Bosmina longirostris.
Daphnia galeata mendotae,  and Diaphanosoma  brachyurum, and a  possible
increase in the numbers of Holopedium gibberum  (Malley et al. 1982).

     Copepods prevalent in acidic waters  (pH 4.1 to 5.0) are Diaptomus
gracilis in Scandinavia and Diaptomus minutus  in North America  (Table
5-7).Tn addition,  frequently  reported as  common in acidic waters are
Heterocope  saliens in Scandinavia,  and Cyclops vernal is, Cyclops
bicuspidatus thomasi, and  Mesocyclops edax  in North America"!  S~pecies
noted as being more  frequent in  non-acidic  lakes  include Epischura
lacustris,  Diaptomus oregonensis, Tropocyclops  prasinus mexicanus,
Heterocope  appendiculata,  Ancanthodiaptomus denticorni's, and
MIxodiapTblmus laclniatus'.   Similarly, experimental  acidification of Lake
223 to  pH 5.4 resulted in  no consistent change  in  populations of
Diaptomus minutus, Cyclops bicuspitatus thomasi, and Mesocyclops edax,
but a dec!ine in numbers of Tropocyclops~prasinus mexicanus,  and
                                  5-68

-------
extinction of Epischura lacustris  below  pH  5.8 and Diaptomus sicilis
below pH 6.1 (Mailey et al.  1982).

     Experimental  acidification  of Lake  223, Ontario also resulted in
the extinction of the opposum  shrimp, Mysis relicta, an important
planktonic predator, below  about pH 5.6  (Mailey et al. 1982).

     Of the insects, midge  larvae  Chaoborus spp. are important
zooplankters in many lakes.   Little is known about effects of acidifi-
cation on Chaoborus. although  it appears to persist in some acidic
environments down to pH 4.2  to 4.5 (Scheider et al. 1975, Van et al.
1982, Confer et al.  1983, Marmorek 1983).   On the other hand, Hobaek and
Raddum (1980) observed that Chaoborus flavicans was absent in
clear-water acid lakes (pH  < 5.0).   fTilssen (1980) reported the
extinction of Chaoborus in  an  acidic lake  (pH 4.2 to 5.0), where
carapace remnants in bottom  sediments verified its presence in earlier
years.

     No data on impacts of  acidification on the productivity of the
zooplankton community are available. Studies on changes in community
biomass are also limited.   Thus, the functional response of the
zooplankton community to increasing levels  of acidity is still largely
unknown.

     Three surveys of abundance  of zooplankton in acidic lakes have been
conducted, involving lakes  near  Sudbury, Ontario contaminated with both
acid and metals (Van and Strus 1980), lakes in the LaCloche Mountain
Region of Ontario (Roff and Kwiatkowski  1977), and headwater lakes in
the Adirondacks, New York,  and White Mountain Region of New Hampshire
(Confer et al. 1983).  In each case, the biomass and/or numbers of
zooplankton in acidic lakes were reduced relative to that in
circumneutral lakes  in the  same  region.   Confer et al. (1983) reported
an average decrease  of 22.6  mg dry wt m-2  per unit drop in pH.  Roff
and Kwiatkowksi (1977) concluded that standing crops of rotifers,
cladocerans, and cyclopoid  copepods (but not calanoid copepods) were
reduced at pH levels below  5.6.  The mean  size of crustacean
zooplankters was, however,  identical in  acidic vs. non-acidic waters.
Van and Strus (1980) found  total community  biomass to be markedly lower
(by almost 80 percent, on the average)  in acidic lakes (pH 4.1 to 4.4)
than in non-acidic lakes (pH > 5.7). Decreased biomass resulted from
both a decrease in numbers  of individuals (except in one acidic lake)
and the small size of the dominant species  (primarily Bosmina
longirostris).

     In contrast, experimental acidification of Lake 223, Ontario, and
1imno-corrals within Lake Eunice,  British  Columbia, resulted in no
change, or even a slight increase, in zooplankton standing crops (Malley
et al. 1982, Marmorek 1983).  The  lowest pH level attained in both these
cases, however, was  pH 5.4.

     Although more data are necessary,  particularly for regions outside
Ontario, the tentative conclusion  is that acidification to pH < 5.0


                                  5-69

-------
results in not only fewer species  but also  decreased biomass of
zooplankton.

5.5.4  Explanations and Significance

5.5.4.1  Changes in Species Composition--The most discrete and
identifiable changes that occur in plankton communities with
acidification are a decline in  the number of species and a shift in
species composition.  It is possible  to  speculate on why these changes
occur and what they may mean to the system.

     The species that predominate  in  an  environment are those best
adapted to survive and reproduce in that environment.  Acidification
changes the environment; thus,  it  is  not surprising that the composition
of the plankton community also  changes.

     Adaptation to acidic conditions, however,  involves more than just
an ability to tolerate low pH levels. Numerous other chemical,
physical, and biological changes associated with acidification require
organisms to make adjustments.   Chemical changes associated with low pH
include elevated concentrations of metals and  alterations in the form
and availability of plant nutrients,  particularly inorganic carbon and
phosphorus (Chapter E-4, Section 4.6.3.5).  With increased acidity, lake
transparency typically increases (Chapter E-4,  Section 4.6.3.4),
potentially altering physical mixing  and thermal regimes.  Finally, as
the increased acidity directly  and indirectly  affects other organisms in
the water, predator-prey and competitive interactions will shift.  All
these factors influence which (and how many) species will be important
within an ecosystem.  Unfortunately,  at  this time we do not know enough
about tolerances and preferences of species for pH levels,
concentrations of metals, etc.  to elucidate which factors result in
observed changes in species composition.

     One factor that has received some attention is the possible
importance of predator-prey interactions.   Acidification results in a
decline in abundance of fish (Section 5.6), important zooplankton
predators.  Changes in plankton communities in response to changes in
fish populations have been clearly demonstrated in numerous studies
(e.g., Brooks and Dodson 1965,  Hall et al.  1970, Nilssen and Pejler
1973, Zaret and Kerfoot 1975, Andersson  et  al.  1978a, Lynch 1979,
McCauley and Briand 1979, Henrikson et al.  1980a, b, and Lynch and
Shapiro 1981).  In general, in  the absence  of  planktivorous fish, the
zooplankton community is typically dominated by large-bodied species.
Fish prey preferentially on larger, more-visible zooplankton (O'Brien
1979).  With the elimination of fish, increased populations of
relatively large-bodied carnivorous and  omnivorous  zooplankton (e.g.,
Chaoborus spp., Leptodora kindti,  and, Epischura lacustris and Mysis
relicta)~consume smaller zooplankton  species and reduce standing crops
of small-bodied zooplankton to  low levels  (Dodson 1974).  Often, as a
result of increased zooplankton grazing  on  phytoplankton, inedible algal
species constitute a greater proportion  of  the total phytoplankton
biomass.
                                  5-70

-------
     In acidic waters, however, the species of zooplankton  that
frequently dominate are relatively small.   Bosmina  coregoni, Bosmina
longispina, and Bosmina longirostris are all  small  (maximum length  about
0.5 to 0.7 mm) compared to other species of cladocerans  common  in non-
acidic, temperate, oligotrophic lakes,  e.g.,  Daphina  longispina (2.2
mm), Daphm'a galeata mendotae (2.3 mm),  Daphnia ambigua  (1.7 mm),
HolopedTum gibberum (1.2 mm), Diaphanosoma brachyurum (1.1  mm), and
Ceriodapffnia quad^angula (0.9 mm) (Nilssen and Pejler 1973,  Makarewicz
and Likens 1979, Lynch 1980).  Diaptomus minutus, a common  copepod  in
acidic lakes in North America, has a maximum length of about 1.0 mm as
compared to 1.2 mm for Cyclops scutifer and Mesocyclops  edax (Makarewicz
and Likens 1979).

     Lynch (1979), in an experimental  investigation of predator-prey
relationships in a Minnesota pond, concluded that zooplankton community
structure was controlled not only by the abundance  of vertebrate
predators, but also by the abundance of invertebrate  predators  and  the
relative competitive abilities of herbivorous zooplankters.  Small-
bodied zooplankton are presumably less  susceptible  to vertebrate
predators but also more susceptible to  invertebrate predators.   Small-
bodied zooplankton (including Bosmina Ipngirostris) dominated in
vertebrate-free environments when invertebrate predators (e.g.,
Chaoborus) were rare and the competitive dominant was of intermediate or
small size.  Janicki and DeCosta (1979)  suggested that Bosmina
longirostris dominates in acidic Cheat Lake (impacted by acid mine
drainage) because of its high reproductive potential  and the intolerance
of its major predator, Mesocyclpps edax, to acidic  conditions.
Populations of a number of crustacean pianktonic predators  including
Epischura lacustris, Leptodora kindti,  and Mysis relicta do seem to be
reduced in acidic lakes, (Nilssen 1980,  Schindler and Turner 1982;  Table
5-7).  Data on abundance of Chaoborus are scarce and  somewhat
contradictory (Section 5.5.3TIThe characteristic  abundance of
small-bodied zooplankton in acidic lakes may, however, be related to a
reduced abundance of invertebrate predators.   Data  are insufficient for
a detailed analysis of this hypothesis.

     The elimination of fish and the reduced importance  of  predaceous
zooplankton in acidic lakes are probably direct consequences of acidi-
fication.  Changes in these populations  occur while their food  supplies
are still abundant (National  Research Council  Canada  1981,  Malley et al.
1982).  The persistence of small-bodied  herbivores  is indicative of
their tolerance of low pH and elevated metal  concentrations.  The
dominance of small-bodied herbivores may,  however,  be the result of a
complex interaction between declining fish populations,  reduced
invertebrate predation, increased water  clarity, and  the relative
survival, growth, and reproductive capabilities of  zooplankton  species
in acidic environments.

     In addition to changes in zooplankton communities,  associated  with
acidic conditions are marked shifts in the species  composition  of the
pnytoplankton community (Section 5.5.2), an important food  source for
zooplankton.   Some algae are more edible than others  (Porter 1977).  A


                                  5-71

-------
high proportion of the phytoplankton  in  many  acidic lakes are
dinoflagellates, a relatively  large phytoplankter that may be less
readily consumed and digested  by  many herbivorous zooplankters.  Van and
Strus (1980)  found that the  average diameter  of the alga Peri dim'urn
inconspicuum, the dominant phytoplankter in acidic Clear-water Lake, was
14 ym.  Yet,  the maximum size  of  a particle likely to be ingested by
Bosmina longirostris, the dominant zooplankter, was 10 to 14 ym, with
85 percent of the particles  ingested  usually  less than 5 ym in
diameter.  The dominant phytoplankter, comprising almost one-half of the
phytoplankton biomass in Clearwater Lake,  may therefore be relatively
unavailable as an energy source to the dominant zooplankter in the lake.

     It is possible that the dominance of dinoflagellates in acidic
waters reflects primarily the  change  in  zooplankton community structure.
The abundance of relatively  small-bodied,  herbivorous zooplankton may
result in selective removal  of edible algal taxa, and the subsequent
dominance of the phytoplankton by larger,  inedible forms.  Van and Strus
(1980), however, discount this hypothesis.  Based on observations in
acidic Clearwater Lake, Ontario,  Van  and Strus (1980) calculated that
the filtering rate for zooplankton in this acidic lake was 5 to 18 times
lower than estimated rates for non-acidic oligothrophic lakes in the
same region.   Assuming these calculations are correct, herbivore grazing
should exert little control  over phytoplankton community structure.

     Alternatively the shift in the phytoplankton community may reflect
relative tolerance to low pH and  elevated metal levels.  If the tolerant
species of algae are also less edible, then transfer of energy from
phytoplankton to herbivorous zooplankton may  be reduced.  This may occur
even though the total biomass  and productivity of these primary
producers are comparable to  those in  circumneutral waters.
Repercussions at higher trophic levels (e.g., fish) are possible, but
the current level of understanding suggests that changes in
phytoplankton community structure are relatively insignificant for the
ecosystem as a whole compared  to  other documented ecological changes
associated with acidification.

5.5.4.2  Changes in Productivity—Available data on acidification and
primary productivity in acidic lakes  yield no clear correlation between
pH level and algal biomass or  productivity.   Primary productivity and/or
phytoplankton biomass in a few cases  were lower in acidic lakes relative
to circumneutral waters; in other cases  equal or even greater (Section
5.5.2.2, Table 5-6).

     Changes in phytoplankton  community  biomass and productivity with
increased acidity may reflect  a balance  between positive and negative
factors.  Differences in the importance  of these factors between systems
may account for inconsistencies in the response of different aquatic
systems to acidic deposition.

     The biomass of phytoplankton at  any given time is a function of its
rate of production vs. its rate of loss.  In  some acidic systans
phytoplankton biomass accumulates (Aimer et al. 1978, Van and Stokes


                                 5-72

-------
1978), suggesting either an increase in primary  productivity per unit
biomass or a decrease in the loss  function.   No  studies  have indicated
increased productivity per unit biomass with increased acidity  (Section
5.5.2.2).  Thus, most authors (Hendrey 1976, Hall  et al.  1980)  have
concluded that any accumulation of algal  biomass in acidic waters
results from a decreased rate of loss or depletion, i.e., decreased
grazing or decreased decomposition.   Lower zooplankton biomass  or  shifts
in zooplankton community structure (Section  5.5.3) may decrease grazing
pressure on phytopl ankton.  Such a conclusion,  however,  is purely
speculative.

     As common as increased standing crops of phytoplankton are
observations of decreased biomass  associated with acidic conditions.
Conclusions that phytoplankton biomass decreased with increasing acidity
imply that either rates of production have decreased or  rates of loss
have increased, or that both have  occurred.   Although good evidence for
lower primary productivity in acidic waters  is  lacking,  there is a
theoretical basis suggesting that  a  number of changes associated with
acidification and acidic deposition  could reduce productivity.  Factors
that could decrease primary productivity with declining  pH levels
include: (1) a shift in pH level below that optimal for  algal growth;
(2) an increase in metal concentrations above those optimal for growth;
(3) decreased nutrient availability; and (4) a  shift in  species
composition within the phytoplankton community  to species with  lower
photosynthetic efficiencies.

     Three primary mechanisms have been proposed whereby nutrient
availability may be reduced in acidic environments:  inhibition of
nutrient recycling, decreased availability of inorganic  carbon, and/or
decreased availability of phosphorus.  Grahn et al. (1974) suggested
that a decreased rate of decomposition and the accumulation of  coarse
detritus, benthic algae, and macrophytes (especially Sphagnum)  on the
bottom of acidic lakes decreased recycling of nutrients  and prevented
exchange of nutrients and other ions between sediments and the  overlying
water (Sections 5.3 and 5.4).  A reduction in these processes could
significantly reduce quantities of nutrients available to primary
producers (e.g., Kortmann 1980) and  induce what Grahn et al. (1974)
termed oligotrophication of the lake system.  No data are available
however to confirm this hypothesis.

     Besides this decrease in nutrient cycling  resulting from a
biological perturbation, increased acidity may  also decrease nutrient
availability via chemical interactions.  Potential effects on inorganic
carbon and phosphorus have received  the most attention.

     At lower pH levels, the total  quantity  of  inorganic carbon
available for algal  uptake is reduced and a  greater proportion  of  it
occurs as aqueous C02 rather than  as bicarbonates or carbonates.  The
National Research Council Canada (1981) calculated that  for a typical
soft-water lake at pH 4.2 in equilibrium with the atmosphere, the
quantity of inorganic carbon consumed by phytoplankton per day  amounted
to about 14 percent of the total dissolved inorganic carbon available in


                                  5-73

-------
the lake.  Thus,  it is possible that during  periods of peak
photosynthesis,  phytoplankton  may  take  up  inorganic carbon from the
water at a rate faster than  it can be replaced  by  diffusion from the
atmosphere.  Phytoplankton productivity at these times nay be carbon
limited.  The significance of  these occasional  limitations during
periods of peak  photosynthesis to  annual levels of production has not
been evaluated.

     In oligotrophic lakes,  phosphorus  availability often limits primary
production (Wetzel  1975,  Schindler 1975).  Chemical interactions between
aluminum and phosphorus (Chapter E-4, Section 4.6.3.5) in acidic waters
or within watersheds receiving acidic depositions, may decrease
phosphorus availability with decreasing pH level and, as a result,
decrease primary productivity.  Despite considerable research on the
chemical nature of aluminum-phosphorus  interactions, no field studies
regarding acidification of surface waters  have  been completed to confirm
or reject this hypothesis.

     Shifts in species composition within  the phytoplankon community
with increased acidity were  discussed in preceding sections.  It is
possible that species of algae predominating in acidic waters have
inherently lower levels of photosynthetic  efficiency than do species
dominant in similar but non-acidic waters.  In  this case, a reduced
level of primary productivity  may be an indirect effect of the shift in
species composition.  Following removal of the  fish population from an
oligotrophic, circumneutral  lake in Sweden,  not only did the species
composition and-diversity of the phytoplankton  community change, but
limnetic primary production  was reduced (Henrikson et al. 1980a,b). It
was hypothesized that, with  the removal of fish, increased grazing
pressure by zooplankton selected for relatively inedible forms of algae,
and the inedible forms were  less productive  and less efficient users of
available nutrients, in part because of their larger size.  None of
these hypotheses has been tested.  Andersson et al. (1978a) also found
decreased primary productivity in the absence of fish, and Redfield
(1980) varied zooplankton grazing intensity  and determined that
concentrated grazing decreased algal productivity.

     Despite these apparently  good reasons for  why acidification should
decrease primary productivity, the available evidence  suggestss  that
there is no consistent decrease.  In part, this may reflect
counterbalancing factors working to increase productivity with
acidification, e.g., increased lake transparency or, to  a lesser extent,
increased nutrient availability resulting  from  plant nutrients
associated with acidic deposition.

     A notable feature of many acidic lakes  is  their remarkable  clarity.
Water chemistry changes with acidification that may contribute  to
increased water clarity are  discussed in Chapter E-4,  Section 4.6.3.4.
As the absorption and scattering of light in the water decreases with
acidification:  (1) a greater amount of light may  be available  for
photosynthesis; (2) light may penetrate to greater depths  increasing  the
size of the euphotic zone;  and (3) adequate light  for  photosynthesis may
                                  5-74

-------
extend down into the thermocline and hypolimnion where  nutrient levels
are generally higher (Johnson  et al.  1970).  Thus, photosynthesis per
unit area of lake surface may  increase.

     Associated with acidic deposition  are relatively large  inputs of
sulfate and nitrate (Chapter E-4,  Section  4.4.1).  Both are  nutrients
required for plant growth.   Productivity  in most oligotrophic lakes,
however, is phosphorus-limited.   Thus,  nutrients associated  with acidic
deposition probably stimulate  primary productivity very little.  In the
few lakes that are nitrogen-limited,  the  response may be more
significant, but no studies are  available  to confirm this.

     It is obvious that transformations in the  structure and function of
the plankton community with increased acidity are the result of a
complex series of reactions.   There  is  no  simple explanation for why
observed differences or changes  occur,  nor is there any reason to expect
responses to be identical  in different  aquatic  systems.  Photosynthesis
by phytoplankton plays a significant role  in driving and controlling the
metabolism of lakes (Section 5.5.1.2).  Any decrease in productivity
could have repercussions at all  trophic levels, including reduced fish
production.  The limited evidence available (Section 5.6), however,
indicates that direct effects  of acidification  on fish  appear more
important than indirect food chain effects.  Thus, although
acidification affects the quality  and may, to a lesser  extent, affect
the quantity of plankton production,  the  significance of these changes
to the aquatic ecosystem as a  whole  has yet to  be established.

5.5.5  Summary

       °   Acidification results in  a marked shift in the structure of
           the plankton community.  For both phytoplankton and
           zqolankton, the  total  number of species represented decreases
           with increasing  acidity.   For  zooplankton. the greatest
           change in species composition  occurs in trie  pH range 5.0 to
           5.3; for phytoplankton, in the  pH interval 5.0 to 6.0.

       0   Zooplankton communities in acidic lakes are  simplifications
           of communities typical  of circumneutral lakes in  the region.
           Species dominant in acidic lakes are also important
           components of zooplankton  communities in non-acidic lakes.
           In Scandinavia,  acidic  lakes (pH < 5.0) are  characterized by
           the prevalence of Diaptomus  gracilis, and Bosmina coregoni or
           Bosmina longispina"In North  America the typical dominant
           association in acidic waters is Diaptomus minutus and/or
           Bosmina longirostris.

       0   Generalizations  about changes  in community structure for
           phytoplankton populations  with  acidification are  more
           difficult to make.  In  many  acidic waters (but certainly not
           all), dinoflagellates (Phylum  Pyrrophyta) predominate.
           Dinoflagellate species  Peri dim'urn inconspicuum and Perl dim'urn
           11mbaturn in particular  are reported  as dominants, often


                                  5-75

-------
           constituting large proportions  of  the  total biomass.
           Dinoflagellates  also  occur  in circumneutral lakes.  Their
           abundance in acidic lakes is often counterbalanced by the
           absence of most  planktonic  species of  diatoms and some common
           species of green algae,  blue-green algae, and chrysophyceans.

       0   Despite the altered structure of the plankton community,
           productivity may remain  unaffected.  Relative to levels of
           primary (phytoplankton)  productivity in circumneutral lakes,
           primary productivity  in  acidic  lakes in some cases islower,
           in others equal.  A cause-and-effect relationship between
           primary productivity  and acidification has not yet been
           established.  In two  field  experiments, increased acidity
           resulted in increased phytoplankton biomass.  In one field
           experiment, acidification had no effect on phytoplankton
           biomass.

       °   Data on zooplankton productivity in acidic lakes are
           non-existent.  In three  lake surveys,  zooplankton biomass was
           lower in acidic  lakes than  in circumneutral lakes in the same
           region.  In contrast,  in two field acidification experiments,
           zooplankton standing  crop was unchanged, or even slightly
           increased.

       0   Shifts in the structure  and function of the plankton
           community with acidification may represent both direct and
           indirect reactions to the decrease in  pH level.  Associated
           with the increased acidity  are  modifications in a large
           number of other  chemical, biological,  and physical aspects of
           the environment  whose changes may  affect the plankton
           community.  Because of the  complexities of these
           interactions,  little  is  known about what controls potential
           changes in phytoplankton and zooplankton communities, why
           responses differ in different lakes, and the significance of
           these changes to other trophic  levels.  Loss of fish
           populations seems to  occur  independently of effects of
           acidification on lower trophic  levels.  However,
           phytoplankton and zooplankton do play  significant roles in
           nutrient and energy cycling.

5.6   FISHES (J. P.  Baker)

5.6.1   Introduction

     The clearest evidence  for impacts of  acidification on aquatic biota
is the documentation of adverse  effects on fish populations.  The
literature is extensive and varied. Available data on effects of
acidification on fish are of at  least  seven types:

     1)  historic records of declining fish populations in lakes and
         rivers, coupled with historic records of increasing acidity;
                                  5-76

-------
     2)   historic records  of  declining  fish populations  in lakes and
         rivers  currently  acidic  but with no historic records on levels
         of acidity;

     3)   regional  lake survey data  and  correlations of present-day fish
         status  with  present-day  acidity levels in lakes and rivers;

     4)   data on success/failure  of fish stocking efforts related to
         acidity of the surface water;

     5)   experimental  acidification of  aquatic ecosystems and
         observations of biological responses;

     6)   results of in situ exposures of fish to acidic  waters; and

     7)   laboratory bioassay  data on survival, growth, behavior and
         physiological responses  of fish to low pH, elevated aluminum
         concentrations, and  other  water quality conditions associated
         with acidification.

Each of these data sets is reviewed: numbers (1) through (4) in Section
5.6.2, Field Observations; numbers  (5)  and  (6) in Section 5.6.3, Field
Experiments; and number (7) in Section  5.6.4, Laboratory Experiments.
Combined, they provide strong evidence  that acidification of surface
waters has adverse effects on fish, in  some cases resulting in the
gradual  extinction of fish populations  from acidified lakes and rivers.

     Loss of fish populations from  acidified surface waters is not,
however, a simple process  and cannot be accurately  summarized as "X" pH
results in the disappearance  of "Y" species of fish.  At the very least,
biological and chemical variation within and between aquatic ecosystems
must be taken into account.   For  example, tolerance of fish to acidic
conditions varies markedly, not only between different species but also
between different strains  or  populations of the same species and among
individuals within the same population. In addition, the water
chemistry within an acidified aquatic system typically undergoes
substantial temporal  and spatial  fluctuations.  The survival of a
population of fish may be  more closely  keyed to the timing and duration
of acid episodes in relation  to the presence of particularly sensitive
life history stages,  or to the availability of "refuge areas" during
acid episodes, or to the availability of spawning areas  with suitable
water quality, than to any expression of the annual average water
quality.  Because of these complexities, summary of effects of
acidification on fish in one  or a few simple concluding  tables can be
misleading.  In addition,  our understanding of functional relationships
between acidification and  fish responses is still incomplete.

5.6.2  Field Observations

     By themselves, field  observations  often fail to establish
cause-and-effect responses definitively.  Most extensive field
observations are simply correlations between acidity of  surface waters


                                 5-77

-------
and absence of various fish species.   Unfortunately, only  In a few
Instances are historic records  available  that provide concurrent
documentation of the decline of the fish  population and the gradual
increase in water acidity.   Clear  demonstration that the absence of fish
resulted from high acidity  requires supporting evidence from experiments
conducted in the field or laboratory.   A  review of observed fish
population changes apparently related  to  acidification does, however,
serve to establish the nature and  extent  of the potential  impact of
acidification on fish.

5.6.2.1   Loss of Populations

5.6.2.1.1   United States

     5.6.2.1.1.1  Adirondack Region of New York State.  The Adirondack
region of New York State is the largest sensitive  (low alkalinity) lake
district in the eastern United States  where extensive acidification has
been reported (Chapter E-4, Section 4.4.3.1.2.3).  The region
encompasses approximately 2877  individual  lakes and ponds  (114,000
surface ha) (Pfeiffer and Festa 1980), and an estimated 9350 km  (6700
ha) of significant fishing  streams {Colquhoun et al. 1981).  Twenty-two
fish species are native to  the region, including brook trout (Salvelinus
fontinalis), lake trout (Salvelinus namaycush), brown bullhead
(Ictalurus nebulosus), white sucker TCatostomus commersoni), creek chub
(Semotnus atromaculatus),  lake chub (Couesiou? plurebeus), and  canmar
shiner (Notropis comutus)  (Greeley and Bishop 1932).  Tn~  addition, a
variety of other species (e.g., smallmouth bass, Micropterus dolqmieui;
yellow perch, Perca flavescens) have been introduced into  Adirondack
waters, especially into the larger, more  accessible lakes.  Brook trout
are frequently the only game fish  species resident  in the  many small
headwater ponds that are located at high  elevations and are particularly
susceptible to acidification (Pfeiffer and Festa 1980).  Although native
to the Adirondacks, in some waters brook  trout populations were
introduced and must be maintained  by stocking due  to a lack of suitable
spawning area.

     Information relevant to effects of acidification on Adirondack fish
populations evolves primarily from three  sources:   (1) a comprehensive
survey of water quality and fish populations in many Adirondack  surface
waters conducted by the New York State Conservation Department in the
1920's and 1930's (Greeley  and Bishop  1932), followed by sporadic
sampling of lakes and rivers up until  the 1970's  (data maintained on
file by the State); (2) in  1975, a complete survey  of all  lakes  (214)
located above an elevation  of 610  m (Schofield 1976b); and (3) from 1978
to the present, accelerated sampling by the New York State Department of
Environmental Conservation  (DEC) of low alkalinity  lakes or lakes that
contain particularly valuable fisheries resources  (Pfeiffer and  Festa
1980).  In addition, a preliminary survey of fish  populations and water
quality for 42 Adirondack streams  was  completed by  the DEC in 1980
(Colquhoun et al. 1981).  None of  these efforts has involved Intensive
studies of individual aquatic systems.
                                  5-78

-------
     Evaluations of Adirondack  data to date are limited to correlations
of present-day fish status with present-day pH levels and, for a limited
number of lakes, a comparison of current data with historic data on pH
and fish population status.  Each of the studies concluded that the
geographic distribution  of fish is strongly correlated with pH level,
and that the disappearance of fish populations appears to have been
associated with declines in  pH.  Indices of fish populations in
Adirondack streams were  statistically (p < 0.05) correlated with pH
measurements (taken in the spring 1980) (Colquhoun et al. 1981).
Schofield (1976b,  1981,  1982) noted fewer fish species in lakes with pH
levels below 5.0 (Figure 5-4).  Schofield and Trojnar (1980) also
observed that poor stocking  success for brook trout stocked into 53
Adirondack lakes was significantly (p < 0.01) correlated with low pH and
elevated aluminum levels.

     In many of the acid waters surveyed in the 1970's, no fish species
were found.   In high elevation  lakes, about 50 percent of the lakes had
pH less than 5.0 and 82  percent of these acidic lakes were devoid of
fish.  Thus, of the total lakes surveyed, 48 percent had no fish.  High
elevation lakes, however, constitute a particularly sensitive subset of
Adirondack lakes,  and these  percentages do not apply to the entire
Adirondack region.  Unfortunately, neither a complete survey nor a
random subsampling of all Adirondack lakes and streams has yet been
attempted.

     All lakes now devoid of fish need not, however, have lost their
fish populations as a result of acidification or acidic deposition.  A
portion of these lakes never sustained fish populations.  In addition,
if earlier fish populations  have disappeared, it must be demonstrated
that acidification was the cause.

     For 40 of the 214 high  elevation lakes, historic records are
available from the 1930's (Chapter E-4, Figure 4-18) (Schofield 1976b).
In 1975, 19 of these 40  lakes had pH levels below 5.0, and also had no
fish.  An additional two lakes with pH 5.0 to 5.5 also had no fish.
Thus, 52 percent had no  fish.   In the 1930's, only three lakes had pH
levels below 5.0 and, again, none of these had fish at that time.  One
additional lake with a pH 6.0 to 6.5 also had no fish.  Thus, in the
1930's, only 10 percent  of the 40 lakes had no fish.  This implies that
17 lakes (or 42 percent)  have lost their fish populations over the
40-year period.  If this holds  true for high elevation lakes in general,
then 39 percent (83 lakes) of the high elevation Adirondack lakes may
have actually lost fish  populations.  However, this assumes that the
subset of 40 lakes represents an unbiased subsample of the 214 high
elevation Adirondack lakes.

     For Adirondack lakes in general, the DEC reports that about 180
lakes (6 percent of the  total), representing some 2900 ha (3 percent of
the total),  have lost their  fish populations (Pfeiffer and Festa 1980,
Schofield 1981).  The basis  for this estimation has not, however, been
clearly delineated.  Presumably, there are 180 lakes for which recent
(1970*s) fish sampling efforts  have yielded no fish and for which


                                 5-79

-------
  a:
        40


        30


        20
        50


        40


        30


        20


        10


          0
               NORWAY
            WRIGHT et al.
                 1975
                                     LA  CLOCHE MOUNTAINS,  ONTARIO
                                              HARVEY 1975
ADIRONDACK MOUNTAINS, NEW YORK
        SCHOFIELD 1976
            LEGEND
       0  NO  FISH PRESENT

       D  FISH  PRESENT
               4.0    4.5   5.0   5.5   6.0   6.5   7.0    7.5

                                      pH

Figure 5-4.   Distribution of fish  in  relation  to lake  pH.

                                   5-80

-------
historic records of fish surveys (1930's to 1960's)  are available that
indicate the presence of fish in earlier years.  All are listed as
former brook trout ponds (Pfeiffer and Festa 1980).  Because the names
of these lakes have not been published and  the  data  are available only
in DEC files, this important conclusion cannot  be critiqued or
validated.

     It is also necessary to demonstrate that the loss of  fish from
Adirondack lakes has occurred as a result of acidic  deposition and/or
acidification of surface waters.

     Retzsch et al. (1982)  argued that "although precipitation acidity
cannot be excluded as a possible cause,  it  represents only one of a
number of factors that may  alter fish  populations in the Adirondacks."
They consider that loss of fish populations in  the Adirondacks may also
be a result of (1)  natural  acidification with the development of
naturally acidic wetlands adjacent to  lakes (see Chapter E-4, Section
4.4.3.3.); (2) declines in  the number  of fish stocked into Adirondack
lakes and changes in management practices;  (3)  introductions of
non-native fish species; (4) increased recreational  use and fishing
pressure; and (5) construction of dams (manmade or beaver) and
manipulations of lake levels and stream flow.

     All  of these reasons sound feasible, yet the DEC argues in return
that loss of fish has occurred in the  absence of alternative
explanations other than acidification  of surface waters (N.Y. DEC 1982).
For example, inadvertent introductions of non-native fish  species occur
primarily in accessible low elevation  waters that are generally not, at
present,  impacted critically by acidification.  Nongame fish species,
not subject to stocking, management, or fishing pressure, have also been
reduced or eliminated.  In  addition, numerous waters located in the
immediate proximity of high-use public campgrounds in the Adirondacks
have maintained excellent trout populations throughout the years despite
heavy fishing pressure (N.Y. DEC 1982).   Dean et al. (1979) evaluated
the impact of black fly larvacide on 42 Adirondack stream  fish
populations and found no significant differences in occurrence and
density of fish in treated  versus untreated streams.  By default,
acidification has been implicated as a factor causing the loss of fish
in a number of lakes and streams.

     A detailed analysis of the raw data set has not, however, been
published that examines, for individual  lakes,  evidence for loss of fish
populations and potential explanations for  these losses, including
acidification.  Still, the  data set in total  is sufficient to conclude
that loss of fish in the Adirondacks,  at least  for some surface waters,
was associated with acidification.  The  number  of fish populations
adversely impacted, and the significance of these losses relative to the
total  resource available in the Adirondacks is, however, inadequately
quantified at the present time.

     5.6.2.1.1.2  Other regions of the eastern  United States.  Schofield
(1982)  summarized available data relating water acidity and fish
                                  5-81

-------
population status for areas  in  the  eastern United States with waters
potentially acidified by  acidic  deposition (Chapter E-4, Section
4.4.3.1.2.3).   Very  few of these studies, with  the exception of studies
in the Adirondack region,  included  comprehensive inventories of fish
populations or historic changes  in  fish  population status with time.
Davis et al. (1978)  noted that  in Maine  lakes biological effects have
not yet been detected. Haines  (1981a) discussed the potential for
adverse effects of acidification on Atlantic salmon (Salmo salar) rivers
of the eastern United States.   Although  the rivers were defined as
"vulnerable,"  no discernable effect on salmon returns was reported.
Crisman et al. (1980) sampled gamefish populations in the two most
acidic lakes (pH 4.7 and  4.9) in the Trail Ridge area of northern
Florida.  Populations of  largemouth bass (Micropterus salmoides) and
bluegull sunfish (Lepoim' s macrochirus) exhibited no clear evidence of
stress directly related to low  pH values or elevated aluminum
concentrations.  In Pennsylvania,  some  fish species have disappeared
from a few headwater stream  systems (Arnold et  al. 1980), but no
consistent trends in the  data set conclusively  demonstrated
acidification impacts (Schofield 1982).   Section 5.2 reviews the
distribution of fish in naturally acidic waters of the United States.

     In regions of the United States, other than the Adirondack Mountain
area of New York State, no adverse  effects of acidic deposition and/or
acidification on fish have been  definitely identified.  Discussions
generally refer only to "potential  impact."

5.6.2.1.2  Canada

     5.6.2.1.2.1  LaClpche Mountain Region of Ontario.   Information
collected on fish populations in the LaCloche Mountain region of Ontario
provides some of the best evidence  of adverse effects of acidification
on fish.  The principal source  of acid entering the LaCloche area is
sulfur dioxide emitted from  the Sudbury  smelters  located about 65 km
northeast (Beamish 1976). Large acidic  inputs  have resulted in
relatively rapid acidification  of many of the region's  lakes--
acidification rapid enough that fish population declines, and in some
cases extinctions, have occurred over the course  of the  15 years that
the lakes have been monitored by researchers from  the University of
Toronto (H. Harvey, R. Beamish,  and other associates).

     Metal concentrations measured  in acidic waters in  the LaCloche  area
ranged from 2 to 5 yg Cu  £-1, 8 to  12 yg Ni  jr1,  24 to  36
yg Zn £-1, and 1 to 4 yg Pb  JT1 (Beamish 1976).  Because of
atmospheric transport of  metals from the relatively nearby Sudbury
smelters,  these values may be slightly  greater  than levels typical  of
acidic waters in other regions discussed in  Section 5.6.2.

     The LaCloche Mountains  cover 1300 km? along  the  north shore of
Lake Huron.   Contained within this  area  are 212 lakes,  approximately 150
of which have been surveyed  for chemical characteristics; 63  for fish
populations.  Fish populations in several of the  lakes  have  been studied
in detail  since the late 1960's and early 1970's  (Beamish and Harvey
                                  5-82

-------
1972, Beamish 1974a,b;  Beamish et al.  1975,  Harvey 1975).  Major sport
fishes common in these  lakes  include lake  trout,  small mouth bass, and
walleye (Stizostedion vitreum).   Other fish  occurinq very frequently are
yellow perch, pumpkinseed sunfish (Lepomis gibbpsus), rock bass
(Ambloplites  rupestris), brown  bullhead,  lake  herring  (Coregorus
artedii), and white sucker.   LaCloche  Mountain  lakes in general have
waters with low ionic content and are  quite  clear, indicative of low
organic acid content (Harvey  1975).  Of 150  lakes surveyed in 1971, 22
percent had pH levels below 4.5  and 25 percent  were  in  the pH range of
4.5 to 5.5 (Beamish and Harvey 1972).

     Harvey (1975) noted that the number of  species  of  fish in 68
LaCloche Mountain lakes was  significantly  (p <  0.005) correlated with
lake pH (Figure 5-4).  In addition,  however, number  of  species of fish
per lake was also significantly  correlated with lake area and other
physical features.  Because small  lakes tend to have low pH values, the
effects of these two independent variables on fish may  be confounded.  A
covariate analysis based on data presented in Harvey (1975) indicated,
however, that the correlation with lake pH was  still significant
(p < 0.005) even after  adjustment for  differences in lake area.  Of the
31 lakes with pH < 5.0, 14 had no fish. Fourteen lakes had pH values of
6.0 or greater, and all of these had at least one species of fish
present with usually seven or more species occurring.

      For the 68 LaCloche Mountain lakes surveyed during 1972-73, 38
lakes are known or are  suspected to have had reductions in fish species
composition (Harvey 1975). Based on historic fisheries information,
some 54 fish populations are  known to  have been lost, including lake
trout populations from 17 lakes, smallmouth  bass from 12 lakes,
largemouth bass from four lakes, wallyeye  from  four  lakes, and yellow
perch and rock bass from two  lakes each.   Assuming that lakes with
current pH < 6.0 originally contained  the  same  number of species as
lakes with an equal surface  area and pH >  6.0,  an estimated 388 fish
populations have been lost from the 50 lakes surveyed with pH < 6.0
(Harvey and Lee 1982).

     The gradual disappearance of fish populations with time and with
increased acidity has been described in detail  for Lumsden Lake, George
Lake, and O.S.A. Lake (Table  5-8; Beamish  and Harvey 1972, Beamish
1974b, Beamish et al. 1975).   Lake pH  levels measured in 1961 by Hellige
color comparator were 6.8, 6.5,  and 5.5 in Lumsden,  George, and O.S.A
lakes, respectively.  In 1971-73, pH levels  measured in the three lakes
with a portable pH meter were 4.4, 4.8 to  5.3,  and 4.4  to 4.9,
respectively.   In the 1950's, eight species of fish were reported in
Lumsden Lake.  Over the period 1961-71, a  drastic decline in the
abundance of both game and non game fish occurred.   In  George Lake,
during the interval 1961-73,  lake trout, walleye, burbot, and  smallmouth
bass disappeared from the lake,  and from 1967 to 1972 the white sucker
population decreased in number by 75 percent and in  biomass by 90
percent.  For O.S.A. Lake in  1961, local  residents reported good catches
of lake trout and smallmouth  bass.  In 1972, intensive  fish sampling
                                  5-83

-------
       TABLE 5-8.   LOSS OF  FISH  SPECIES FROM LUMSDEN LAKE AND GEORGE
     LAKE,  ONTARIO  (FROM BEAMISH AND HARVEY 1972, BEAMISH ET AL. 1975,
                         HARVEY AND LEE 1982)
Date
       Species information
Lumsden Lake
  1950's
  1960

 1960-65  •
  1967

  1968
  1969

  1969
  1970
Eight species present
Last report of yellow perch
Last report of burbot
Sport fishery fails
Last capture of lake trout
Last capture of slimy sculpin
White sucker suddenly rare
Last capture of trout-perch
Last capture of lake herring
Last capture of white sucker
Last capture of lake chub
George Lake
  1961
  1965
  1966

  1970

  1971

  1972

  1973
Last spawning of walleye
Last capture of smallmouth bass
Last spawning of lake trout
Last capture of trout-perch
Last capture of burbot
Most white suckers fail  to spawn
Last capture of walleye
Brown bullhead fail to spawn
Northern pike, pumpkinseed sunfish,  rockbass,
  brown bullhead, and white sucker fail  to spawn
Last capture of lake white fish
Lake trout rare
                                  5-84

-------
                         TABLE 5-8.   CONTINUED
Date                            Species  information


George Lake (continued)


  1974                   Northern  pike and  pumpkinseed  sunfish rare

  1978                   Few age classes of white  suckers remain

  1979                   Brook  trout and muskellunge rare
                         White  sucker, brown bullhead,  rock bass, lake
                           herring,  and  yellow  perch present
                                  5-85

-------
yielded only four yellow perch,  two rock  bass,  and eight lake  herring.
By 1980, no fish remained (Harvey and Lee 1980).

     Harvey (1979) summarized the apparent tolerance  of fish in  the
LaCloche Mountain region to pH,  based on  the occurrence of species in
lake surveys and their disappearance with acidification (Figure  5-5).
Beamish (1976)  concluded that increased acidity was the principal  factor
resulting in the loss of fish populations.

      5.6.2.1.2.2  Other areas of Ontario.   Harvey (1980)  estimated  that
approximately 200 lakes in Ontario have lost their fish populations.
For the most part, however, these lakes are in  the vicinity of Sudbury,
Ontario.  Studies that suggest fish loss  in response  to acidification
for other areas of Ontario are very limited.  Although  the
Muskoka-Haliburton region of Ontario receives large inputs of  acidic
deposition, and decreases in alkalinity have been  suggested for  some
lakes (Chapter E-4, Section 4.4.3.1.2.2), no adverse  effects on  fish
populations have been documented; pH values apparently  have not
decreased to levels harmful to fish.

     5.6.2.1.2.3  Nova Scotia.  In Novia  Scotia,  rivers with pH  < 5.4
occur only in areas underlain by granitic and metamorphic  rock;  all  flow
in a southerly  direction to the Atlantic  Coast  (Watt  et al. 1983).
Thirty-seven rivers within this  region have historic  records indicating
that they sustained anadromous runs of Atlantic salmon.   For 27  of these
rivers (Table 5-9), almost complete angling catch  records  are  available
from annual reports of Federal Fishery Offices  for the  period  1936 to
1980.  Of these 27, five rivers  have undergone  major  alterations since
1936 that potentially could have impacted salmon stocks.   For  the 22
remaining rivers, 12 presently have pH >  5.  Statistical  analysis of
angling catch from 1936 to 1980  indicated that  only one of these 12
rivers had experienced a significant (p < 0.01)  decline in salmon catch
since 1936, one river a significant (p <  0.05)  increase,  and 10  no
significant trend in angling catch with time.   In  contrast, of the 10
rivers with current pH < 5.0, nine have had significant (p < 0.02)
declines in success since 1936,  and one,  no significant trend.

     Salmon angling records for  rivers with pH  <  5.0  vs pH > 5.0 are
compared in Figure 5-6.  From 1936 through the  early  1950's, angling
catch in the two groups of rivers were similar.   After  the 1950  s,
angling catch in rivers with pH  < 5.0 declined,  while salmon catch in
rivers with pH  > 5.0 continued to show no significant trend with time.

     Year-to-year variations in  salmon catch are considerable,
reflecting the many factors affecting angling success and reporting
accuracy.  Between the two groups of rivers (pH <  5.0 and  pH > 5.0],
however, occurrence of high and low success years, generally correspond.
Both groups of rivers are well distributed along the  500 km Atlantic
coastline of Nova Scotia.  Tag return data suggest that salmon stocks in
this area all  share a common marine migratory pattern.   Biological and
physical factors leading to greater or lesser angler  success (e.g.,  sea
survival, river discharge rates, or juvenile year-class survival)


                                  5-86

-------
     SPECIES
   YELLOW PERCH
   PUMPKINSEED
   ROCK BASS
   WHITE SUCKER
   NORTHERN PIKE
   LAKE HERRING
   BLUEGILL
   LAKE WHITEFISH
   SMALLMOUTH BASS
   LARGEMOUTH BASS
   LAKE TROUT
   BROWN BULLHEAD
   GOLDEN  SHINER
   IOWA DARTER
   JOHNNY  DARTER
   COMMON  SHINER
   BLUNTNOSE  MINNO
                                                    NUMBER OF LAKES
                                                    CONTAINING SPECIES

I—
1 	 __ 	





h"
1 	 — 	







« ' ' 	 1 	 > 	
40
37
29
25
20
23
6
6
19
7
g
7
10
20
11
6
q

                  4.0   4.5    5.0
                   ) 11   |  19  I
         5.5
         pH
19  I  13  |  8   I  10
  NUMBER OF LAKES
6.0   6.5   7.0
Figure 5-5.   Frequency of occurrence  of fish  species  in  six  or  more
             La Cloche Mountain lakes in relation  to  pH.   Vertical bar,
             lowest pH recorded; dashed line, stressed populations,  e.g.,
             missing year classes;  solid line, populations which  appear
             unaffected (Harvey 1979).
                                  5-87

-------
TABLE 5-9.  MAJOR RIVERS IN NOVA SCOTIA ON THE ATLANTIC  COAST,
       pH LEVELS AND STATUS OF ATLANTIC SALMON STOCKS
Mean3
PH
River 1980-81
Musquodoboit6
St. Mary's
LeHave*
Ecum Secum
Petit
Ship Harbour
Gold
Salmon (Digby)
East Ship Harbour
West Ship Harbour
Moser
Quoddy
Kirby
Medway6
Salmon (Port Dufferin)
Gaspereau
Mersey6
Middle
Li scomb
Ingram
Tangier
East
Tusket
Issacs Harbour
Nine Mile
Salmon (Lawrencetown)
Clyde
Barrington
Jordan
Sable
Broad
Roseway6
Larry 'sf
6.7
6.1

5.7
5.6
5.6
5.5



5.4
5.4
5.4
5.4
5.3
5.2

5.0
5.0
5.0
4.9
4.8
4.8
4.8
4.8
4.7
4.6






Recorded
Presence (+)
Rangeb or Absence (-) Regression of
pH of Salmonc Angling Catch
1979-80 <1960 1980-82 on Year3
6.6-6.9 +
6.1-6.8 +
6.0-6.1
+
+
5.6-5.9 +
5.6-6.0 +
5.1-5.7
5.3-5.4
5.0-5.4
5.5-6.2 +
+
+
5.2-5.8 +
+
+
4.9-5.4
+
5.0-5.3 +
5.0-5.5 +
+
4.9-5.1 +
4.5-4.8 +
+
+
+
4.6-4.6 +
4.5-4.7 +
4.4-4.6 +
4.3-4.6 +
4.3-4.5 +
4.3-4.5 +


+ NS
D
NS
NS
NS

D
D
D
NS
NS
NS
+ NS
NS
NS
D
+
NS
+
— _
+
_
_
_
_ _
.. _
_
_
_
_
-

                        5-1

-------
                              TABLE  5-9.  CONTINUED


 aWatt et al.  1983,  Rivers  with  1980-81 mean  pH recorded have angling
  data available over the past 45 years and are represented in Figure
  5-11.

 bFarmer et al.  1981; pH  range from  three pH  measurements per
  river—April  or May 1979, September  or November 1979, and February or
  March  1980.

 cWatt et al.  1983;  <1960 Presence/Absence based on catch records;
  1980-82 based on electrofishing  for  juvenile salmon and/or catch data.

 dWatt et al.  1983;  27 rivers  with angling records 1936 to 1980—no
  significant  trend  (NS), significant  increase in catch with time (+)
  decrease in  catch  with  time  (-), major disturbance in watershed (D).

 Historical  pH records available.

 fpH level reported  as <  4.7 in  Watt et al. 1983.
                                   5-89
409-262 0-83-15

-------
                    200
                3
                CO
                Of.
                o
100


 80



 60




 40
en

«3
O
                Z
                UJ
                o
                a:
                o


                o
 20
                     10


                      8
      —Q—MEAN FOR 12 RIVERS WITH pH >5.0 (1980)


    	O—MEAN FOR 10 RIVERS WITH pH<5.6 (1980)
                      1935     1940     1945     1950     1955      1960

                                                               YEAR
                                                        1965
1970
1975
1980
   Figure 5-6.  Average angling success for Atlantic salmon  in 22  Nova  Scotia  rivers  since 1936.   Data
                were collected from reports of federal fishery offices  and  normalized by expressing each
                river's angling catch as a percentage of the average  catch  in  that river during the first 5
                years of record (1936-40) (Watt et al. 1983).

-------
probably act uniformly over the entire area (Watt et  al.  1983).
Decreases in salmon catch over time are,  on the  other hand, clearly
correlated with present-day pH values 5.0 and below.

     Watt et al. (1983), concluded that at present in Nova  Scotia, seven
former salmon rivers with mean annual  pH < 4.7 no longer  support salmon
runs (Table 5-9).  An electrofishing survey in the summer of  1980 failed
to find any signs of Atlantic salmon reproduction in  any  of these seven
rivers.  Farmer et al. (1980), however, observed that for the most part
these rivers are all also naturally somewhat acidic (highly colored
waters, indicating the presence of organic acids), and historically  had
relatively low fish production.  Peat deposits and bogs are common to
much of this area.  Inputs from these materials  probably  contribute  to
the low pH levels and have some impact on salmon production.  Historical
records of pH for a few rivers within this area  (Chapter  E-4, Section
4.4.3.1.2.2) do, however, indicate that acidity  has increased in recent
years.   Acidic conditions and acidification,  therefore, probably
contribute to the loss of Atlantic salmon populations in  Nova Scotia.

     The estimated lost (rivers with pH < 4.7) or threatened  (rivers
with pH 4.7 to 5.0) Atlantic salmon production potential  represents 30
percent of the Nova Scotia resource, but only 2  percent of  the total
Canadian potential.  Atlantic salmon rivers salmon in New Brunswick,
Prince Edward Island, and other areas of Novia Scotia generally have pH
levels above 5.4, and are not under any immediate acid threat (Watt
1981).

5.6.2.1.3   Scandinavia and Great Britian

     5.6.2.1.3.1   Norway.  Extensive information on  acidification and
loss of fish populations in Norwegian waters has been collected under
the auspices of the joint research project SNSF--"Acid Precipitation-
Effects on Forest and Fish," 1972-1980.  Documentation of the effects of
acidification on fish is derived principally from (1) yearly  records of
catch of Atlantic salmon in 75 Norwegian rivers  from  1876 to  the
present; (2) a survey of water chemistry and fish population  status  in
700 small lakes in southern Norway in 1974-75; (3) collation  of
information on fish population status (current and historic)  for some
5000 lakes in southern Norway, validated with testfishing in  93 lakes
during 1976-79; and (4) detailed analyses of historic changes in fish
population status related to land use changes with time in  selected
watersheds.  Together these data provide strong  evidence  that
acidification has had profound impacts on fish.

     Statistical data for the yearly salmon catch from major  salmon
rivers in Norway have been recorded since 1876 (Figure 5-7) (Jensen  and
Snekvik 1972, Leivestad et al. 1976, Muniz 1981).  While  catch in all
rivers declined slightly from 1900 until  the 1940's,  in 68  northern
rivers the decline was followed by a marked increase, and catch in the
1970's equalled or exceeded that around 1900. In contrast, in seven
southern rivers, annual catch dropped sharply over the years  1910-17,
has declined steadily since then, and is now near zero.   This decrease


                                  5-91

-------
                                                                        i/o
                                                                        o
    1900
1980
Figure 5-7.  Yearly yield for Atlantic  Salmon  fisheries  in  seven  rivers
             from the southermost part  of  Norway  (bottom curve) compared
             with 68 rivers from the rest  of the  country (top curve).
             (Leivestad et al. 1976).
                                  5-92

-------
is reflected in all seven rivers and cannot be explained by  known
changes in exploitation practices.   Massive fish  kills  of Atlantic
salmon (Section 5.6.2.4) were reported in these rivers  as early  as  1911.
Efforts over the last 50 years to restock with hatchery-reared fry  and
fingerlings have been unsuccessful.   In the seven southern rivers,  pH
levels averaged 5.12 in 1975, as compared to an average pH of 6.57  for
20 of the 68 northern rivers.  Leivestad et al. (1976)  reported  that
acidity in southern rivers has been  steadily increasing from 1966 to
1976 hydrogen ion concentration increased by 99 percent.

     In 1974-75, the SNSF project completed a synoptic  (nonrandom)
survey of water chemistry and fish  population status  in 700  small to
medium-sized lakes in Stfrlandet (the four southernmost  counties  of
Norway) (Wright and Snekvik 1978).   Based on interviews with local
residents, fish populations in lakes were classified  as barren,  sparse
population, good population, and overpopulated.  The  principal species
of fish was brown trout (Salmo trutta).  Other important species were
perch (Perca fluviatilis), char (Salvelinus alpinus), pike (Esox
lucius). rainbow trout (Salmo gairdTieri). and brook trout.   "ATxMTt 40
percent of the 700 lakes were reported as barren  of fish,  and an
additional 40 percent had sparse populations.  Fish status was clearly
related to water chemistry; most low pH,  low conductivity  lakes  were
either barren or had only sparse populations.  Above  pH 5.5, few lakes
were barren.  A stepwise multiple regression of fish  status  against
chemical variables pH, N03-, S042-   C1-,  Na+, K+,  Ca2+,  Mg2+, Al3+,
and HCOs" indicated that pH and Ca2* were the two most  important chemical
variables (r = 0.53).

     The original data base on fish  populations in Stfrlandet collected
by Jensen and Snekvik (1972) and Wright and Snekvik (1978) has gradually
been extended to the whole country.   By 1980, data on fish in more  than
5000 lakes in the southern half of  Norway had been collected by
interviewing fisheries authorities,  local landowners, local  fishermen's
associations, and other local experts (Sevaldrud  et al.  1980, Overrein
et al. 1980, Muniz and Leivestad 1980a).   Interview data ware validated
for 93 lakes by comparison with results from a standardized  testfishing
program.  Interview data provided an accurate assessment of  actual  fish
stocks for over 90 percent of the lakes (Rosseland et al.  1980).

     At present, fish population damage has apparently  occurred  in  an
area of 33,000 km2 in southern Norway.   Twenty-two percent of the
lakes at low elevations below 200 m  have  lost their brown  trout
populations; 68 percent of the trout populations  in high  altitude lakes
above 800 m are now extinct.  In 13,000 km2 of this area,  fish
populations in all lakes are extinct, or near extinction.  Water
chemistry data are available for a  subset of these 5000 lakes, and  again
fish population status is clearly correlated with pH  (Figure 5-8).

     Besides information on the current status of fish  populations  in
these 5000 lakes, the SNSF project  has  also compiled  available historic
information on changes in fish populations with time.   For almost 3000
lakes in Sjfrlandet, the population  status of brown trout has been
                                  5-93

-------
    Kc-=10yS * cnTin « 203,
1UU
80
60

40
20

n
»
•
-
.
•
•
.

r
'/,
X
X
\
X
/
/
X,







1
7
^
/
X
^
^
j
/







~j-
•^
/
/
^
/
^
•^










r
/
/
/
/
/
r
/








Y
K













17
-







               Kc 10-20 ytS • on"1n - 320

                                                 X
                                                            1,
                          284
1UU
80
60
40
20
n

•
•
-
•
cri
^
/
\



]
/
XXXXXXI
-




p

















g
•™






g




        CO    t-i
         »     .
        *f    in

        to    CT*
co
in
                   CM
                   in
                Kc>30yiS  • cm"1 = 227
•
.
»
X
X
vXXXXXXN


11
-






















PT;



in co «-J ** r^ oo
^ «*">*« in in
V ^j en CNJ J,
                                                          pH
     LEGEND

   LOST POPULATIONS

   TROUT  PRESENT
Figure 5-8.  Status of brown trout populations from the affected areas
             in the four southermost counties (Rogaland, Vest-Agder,
             Aust-Agder, and Telemark)  in Norway grouped according to
             lake pH and conductivity.   The data are given as percentage
             of lakes with or without trout within each class of pH and
             conductivity (Muniz and Leivestad 1980a).
                                    5-94

-------
 recorded by local fishermen since about 1940.  The time trend for loss
 of populations is diagrammed in Figure 5-9.  The rate of disappearance
 of brown trout from lakes in Stfrlandet has been particularly rapid
 since 1960.  Today, more than 50 percent of the original  populations
 have been  lost, and approximately 60 percent of the remaining are in
 rapid decline (Sevaldrud et al. 1980).  Attempts at restocking acidified
 lakes containing reduced populations have largely failed (Overrein et
 al. 1980).

     A relationship between water acidity and fish population status or
 even water acidification and concurrent loss of fish populations does
 not necessarily implicate acidic deposition as the primary  cause for
 adverse effects on fish.  Evidence for the association between acidic
 deposition and acidification of surface waters is considered in Chapter
 E-4.  However, several studies have been completed in Norway that
 examine alternate explanations for acidification, e.g.,  changes in land
 use, specifically as they relate to historic changes in fish populations
 (Drabljte and Sevaldrud 1980, Drablsfs et al. 1980).  In each of three
 study areas, no correlation between shifts in land use and  human
 activities and changes in fish status was found.  Areas that have
 experienced changes in land use (e.g., abandonment of pasture farms  or
 discontinuance of lichen harvests) do not have any higher proportion of
 lakes with declines in fish population than do areas without such land
 use changes.  In contrast,  fish population  declines are  correlated with
 inputs of acidic deposition.

     5.6.2.1.3.2   Sweden.   Sweden h.as about 90,000 lakes,  many of which
 have low alkalinity and are potentially sensitive to acidic deposition.
 Extensive surveys of acidification and fish population status,  have  not,
 however, been completed.  In southern Sweden, 100 lakes  with pH 4.3  to
 7.5 were sampled in the 1970's (Aimer et al.  1978).  Apparently as a
 result of acidification (i.e., disappearance of fish was  associated  with
 current low pH levels in lakes), 43 percent of the minnow (Phoxinus
 phoxinus) populations, 32 percent of the roach (Rutilus  rut11 us),  19
 percent of the artic char,  and 14 percent of the brown trout populations
 had been lost.  In a study  of six lakes in  southern Sweden,  Grahn  et al.
 (1974)  cited historic pH data suggesting a  pH decline of 1.4 to 1.7
 units since the 1930's-40's and the simultaneous elimination of minnows,
 roach,  pike and brown trout from two or more of these six lakes.
 Disappearances of populations of roach in lakes in southwestern Sweden
 were recorded as early as the 1920's and 1930's (although not definitely
 correlated with acidification) (Dickson 1975).   In eastern  Sweden, loss
 of roach from Lake Arsjon near Stockholm occurred in association with a
 decrease in pH readings:   pH 5.1 to 5.3 in  1974 as compared to  pH  6.0
 measured colometrically in  the 1940's (Milbrink and Johansson 1975).

     5.6.2.1.3.3   Scotland.  Investigations in Scotland  (Harriman and
Morrison 1980, 1982)  indicated that intensive afforestation  can result
 in acidification of streams and subsequent  reduction and  loss of fish
populations.   The role of acidic deposition in  this acidification
 process has not yet been clearly established.   In a study of 12 streams
draining forested and nonforested catchments,  an electrofishing survey


                                  5-95

-------
     3000
     2500
     2000
  o
  i—i
  
-------
failed to yield any trout In most streams draining forested catchments
(mean pH 4.34), while moorland streams (mean pH 5.40)  Invarlbly had
resident trout populations.

5.6.2.2  Population Structure--The well-being of a population can be
judged In part by examination of Its age composition (National  Research
Council Canada 1981).  Theoretically, age one fish should be more
numerous than age two fish; age two fish more numerous than age three
fish; age three fish more numerous than age four fish, etc.  Two factors
commonly alter this theoretical distribution:  year selectivity and
large natural variations in year class strength.  Almost all  procedures
for sampling fish populations are size selective.   Often, small, young
fish are poorly sampled.  In addition, relative numbers of fish in each
age group may fluctuate greatly from year to year  as a consequence of
natural environmental and biological factors (e.g., year-to-year
temperature variations, competition between age groups).   The frequent
absence of one or several age groups within a population may, however,
be indicative of a population under stress or undergoing change.
Studies of fish populations in acidic waters frequently reveal  reduced
or missing agre groups.

     Deviations from the expected age class distribution in acidic lakes
result in some cases from the absence of young fish; in others  from the
absence of older fish.   A population with only fairly  large,  fairly old
individuals, suggests that recruitment and/or reproduction have failed.
A population with only  young fish may imply the occurrence of a
mortality factor acting only on fish after a certain age (e.g.,  after
sexual  maturity).  Both types of distributions have been  observed in
acidic waters, although the absence of young fish  occurs much more
frequently.   Decreased  recruitment of young fish has been cited as a
primary factor leading  to the gradual extinction of fish  populations in
acidic waters (Schofield 1976a, Overrein et al. 1980,  Haines  1981b).

     Studies of lakes in the LaCloche Mountain region  of Ontario by
Beamish,  Harvey, and others provide detailed observations of  the
structure of fish populations 1n acidic and acidifying lakes.   White
suckers were last reported in Lumsden Lake in 1969 (Table 5-8)  at a pH
of 5.0 to 5.2 (Beamish  and Harvey 1972) (Section 5.6.2.1.2.1).
Intensive sampling 1n 1967 yielded no young-of-the-year and very few age
one fish, suggesting poor recruitment of white suckers in both  1967 and
1966.  In contrast,  in  George Lake examination of  the  age distribution
of white  suckers in  1972 indicated no obviously missing year  classes  and
thus no major reproductive failures prior to 1972  (pH  4.8 to  5.3)
(Beamish  et al.  1975).   Although reduced in number,  white suckers  were
still present in George Lake in 1979 (Harvey and Lee 1980).   In  1972,
O.S.A.  Lake had a pH of about 4.5.   Intensive sampling yielded  only a
small number of very old fish—eight lake herring  aged 6  to 8 years,
four yellow perch aged  8 years, and two rock bass  aged 13 years  (Beamish
1974b).  By 1980, no fish remained in O.S.A. Lake  (Section
5.6.2.1.2.1).
                                  5-97

-------
     In addition to these intensive  studies  of  individual lakes in the
LaCloche Mountain region, Ryan  and Harvey  (1977,  1980) surveyed (through
rotenone applications)  the age  distribution  of  populations of yellow
perch and rock bass in  32 and 20  LaCloche  Mountain lakes, respectively.
For both species, lakes with lower pH  levels had  a higher frequency of
populations missing the age 0 group  (young-of-the-year).  The most
acidic lake yielding young-of-the-year yellow perch was characterized by
a pH of 4.4, for rock bass by a pH of  4.8.

     Absence of young age groups  in  fish populations from acidic and
acidifying lakes has also been  documented  for a few lakes in the
Adirondack region and in Scandinavia.   In  South Lake in the Adirondacks,
white suckers netted in 1957-68 (pH  5.3 in 1968)  ranged in length from
15 to 51 cm, suggesting a wide  range of age  classes.  By 1973-75 (pH 4.9
in 1975), however, recruitment  of young fish appears to have ceased.
White suckers collected ranged  from  30 to  49 cm in length.  Five suckers
captured in 1975 were aged 6 to 8 years (Schofield 1976a, Baker 1981).
In Lake Skarsjon in Sweden, prior to lake  liming  (pH 4.5-5.5) only very
large, old perch remained in the  lake  (Figure 5-10).  One year after
liming (pH  ~6.0), reproduction was  reestablished and two size classes
of perch were present,  both very  large, old  fish  and a new group of
small, one-year-old perch (Muniz  and Leivestad  1980a).

     Recruitment failure may result  either from acid-induced mortality
of fish eggs and/or larvae or because  of a reduction in numbers of eggs
spawned.  Beamish and Harvey (1972)  attributed  the lack of reproduction
in fish populations in  LaCloche Mountain Lakes  to a failure of adult
fish to spawn.  In Lumsden Lake in 1967, no  spawning activity was
observed in the lake or in the  inlet or outlet  streams during the normal
spawning period.  Mature female white  suckers were found to be resorbing
their eggs in June.  In George  Lake, in 1972 and  1973 about 65 to 75
percent of the population of female  white  suckers failed to release
their ova to be fertilized.  In 1973,  most brown  bullheads, rock bass,
pumpkinseed sunfish, and northern pike had also not spawned when
examined after their normal spawning period  (Beamish et al. 1975).
Biochemical analyses of fish from George Lake indicated that females
exhibited abnormally low levels of serum calcium  during the period of
ovarian maturation.  Lockhart and Lutz (1977) hypothesized that a
disruption in normal calcium metabolism, induced  by low pH, affected
female reproductive physiology.  In  these  lakes,  therefore, failure of
female fish to spawn was an important  contributing factor to
reproductive failures.

     This response, failure of  female  fish to spawn, has not, however,
been reported elsewhere.  From  a  survey of 88 lakes in Norway, Rosseland
et al. (1980) noted that female fish remaining  in acidic lakes had
normal gonads, and indications  of unshed or  residual eggs were rare.
Studies conducted in Scandinavia  and the United States (Schofield 1976a,
Muniz and Leivestad 1980a) suggest that increased mortality of eggs and
larvae in acidic waters is the  primary cause of recruitment failures.
                                  5-98

-------
                 PERCH POPULATION   LAKE ST. SKARSJ0N  1976
           CO
o
6

4
2
0


-
•
*
•
i

RECRUITMENT FAILURE

Approximately
15 years old

, r

















1

ONE YEAR AFTER LIMING
8r
6

4

2
n
•
.
F

: ''<
i /
1975 Reproduction
\
/TI 1 year old
M

//,i
r.'i,
/ / / 0








t











1
10         20

   LENGTH (cm)
                                                    30
Figure 5-10.   Liming of Lake St.  Skarsjon, Sweden,  in 1975 reestablised
              reproduction of perch population (Muniz and Leivestad
              1980a).
                                  5-99

-------
In Norway, total mortality of naturally  spawned trout  eggs was observed
in an acidic stream a few weeks after spawning  (Leivestad et al. 1976).

     In addition to the lack  of young fish  in a population, associated
with recruitment failure as described above,  in some cases loss of older
fish has been observed in acidic waters.  Three lakes  in the Tovdal
River, Norway, were test fished from 1976 to  1979  (Figure 5-11)
(Rosseland et al.  1980).  Before 1975, brown  trout populations in these
lakes were stunted and grew to 8 to 10 years  of age.   In 1975, the
Tovdal River had a severe fish kill.   Since 1976,  no post-spawning brown
trout (age 5 and up)  have been found, and the population is dominated by
young fish.  Test fishing in  autumn indicated the  presence of maturing
recruit-spawners.   By each subsequent year, however, this age group had
disappeared while their offspring survived.   Researchers speculated that
stress associated with spawning activities, coupled with acid-induced
stress, resulted in significant post-spawning mortality (Muniz and
Leivestad 1980a).

     Harvey (1980) proposed that loss of older  fish with acidification
was also occurring in George  Lake (LaCloche Mountain region)
(colorimetric pH 6.5  in 1960;  pH 5.4  in 1979).   In 1967, white suckers
up to 14 years of age occurred in the lake.   By 1972,  few fish were
older than 6 years.  Sampling in 1979 revealed  a population with 90
percent of the white  suckers  aged 2 and 3 years.

     It is unlikely that loss of older fish in  either  of these cases
resulted from over-fishing.

5.6.2.3  Growth—Observations on fish growth  in acidic waters and
changes in growth rate over time with acidification suggest that
indirect effects of acidification,  via changes  in  food availability, are
generally insignificant for adult fish.  In very few cases have reduced
growth rates been reported.  For the most part,  fish in acidic and/or
acidified waters grow at the  same rate or faster than  fish in
circumneutral waters  in the same region.

     Decreases in fish growth rate associated with acidification have
been documented only  for acidic lakes in the  LaCloche  Mountain region,
Ontario.  In 1975, Beamish et al. (1975) reported  that growth rates for
white suckers in acidic George Lake (pH 4.8 to  5.3, 1972-73) had
declined over the period 1967 to 1973, and  this was apparently
associated with lake  acidification.  In more  recent surveys, however,
this trend appears to have reversed.   Fish  collected in 1978 and 1979
were larger (at a given age)  than fish in 1972, and similar in size to
fish collected in 1967 to 1968 (Harvey and  Lee  1980).  Therefore, even
in this instance,  consistent decreases in growth over  time with
increased water acidity have not occurred.

     On the other hand, several studies  suggest increased fish growth in
acidic waters and/or  with acidification.  For two  acidic lakes in the
Adirondacks sampled in the 1950's and 1970's, numbers  of brook trout
caught decreased over the 20-year period, and significant increases in
                                  5-100

-------
CATCH YEAR |l979ll978|1977il976|l975
1
80
1975 60
40
20
1976
1UU
80
60
40
20
	 iuu
80
i—
1977 | 60
£ 40
20
1978
	 1UU
80
1979 60
40
20
1UU
80
60
40
20
p—O
|1979


1978
• u




I






///t

%





/xx


1974
•


///r

W7,



1973


XXX!
y//


D
0

1972




•
JUVEf
RECRl
POST

1971


W



1970





1969


Vxx



1968 1967

xW







LEGEND
ULE FISH(I-II)
JIT SPANNERS (III-IV)
SPAWNERS (VII- )




1967
Figure 5-11.   Age distribution of brown  trout in  Lake  Tveitvatn,
              Tovdal, Norway (Rosseland  et al.  1980)
                                  5-101

-------
fish growth were observed (Schofield 1981).   Roach  in  acidic lakes  (pH
4.6 to 5.5) in Sweden grew at substantially  faster  rates than roach in
circumneutral lakes (pH 6.3 to 6.8)  (Aimer et al. 1974, 1978).  Growth
of rock bass in 25 LaCloche Mountain lakes was also significantly (p <
0.05) faster in lakes with greater acidity,  even after adjustment for
effects of lake depth on fish growth (Ryan and Harvey  1977, 1981).
Jensen and Snekvik (1972) described a common pattern of change  in lakes
in Stfrlandet, Norway over the last 50 years.  Densities of fish in
lakes declined, presumably associated with acidification and the onset
of increased recruitment failure.   Simultaneously,  fishing quality
increased, with a greater number of large  trout available.  Eventually,
however, with continued recruitment failures,  in many  lakes populations
disappeared entirely.

     Rosseland et al. (1980), on the other hand, in a  survey of 88 lakes
in southern Norway, found no obvious tendency  for increase in growth in
sparse populations in acidic lakes,  despite  the fact that fish  from
acidic lakes had higher proportions  of full  stomachs and were in better
condition (i.e., weighed more for a given  length).   Ryan and Harvey
(1980, 1981) observed that yellow perch in 39  LaCloche Mountain lakes
grew more quickly in more acidic waters up to  age 3 years, but
thereafter grew more slowly.  In addition, yellow perch collected from
George Lake 1n 1973 and 1974 (pH 4.6)  at age 1 to 4 years were
significantly larger than perch of the same  age collected 1n 1966 (pH
5.8); this trend was reversed for age groups 5 years and older.  Up to
age 4, yellow perch feed primarily on  plankton and  benthic
invertebrates.  Large perch feed preferentially on  small fish.

     Fish growth reponse to acidification  may  be a  complex function of
two factors:  acid-induced metabolic stress  and food availability.
Reduced growth in acidic waters as a result  of physiological stress has
been noted frequently in laboratory  experiments (Section 5.6.4.1.3).
Presumably, similar responses occur  1n acidic  lakes and streams.
Observations of increased or unchanged growth  1n acidified surface
waters, however, suggest that adverse effects  of acidity on fish
metabolism and physiology are counterbalanced, in part or totally, by
changes in food availability.

     Acidification is associated with substantial changes in the
structure and, in some cases, the function of  lower trophic levels
(Sections 5.3 and 5.5).  Despite the fact  that some Important prey
organisms are sensitive to acidic conditions and, as a result, fish may
be required to shift their predation patterns, still In most acidic
lakes food does not seem to be a significant limiting  factor for adult
fish (Beamish et al. 1975, Hendrey and Wright  1976).   Possibly, with
decreased fish density resulting from recruitment failures or fish
kills, decreased interspecific and/or Intraspecific competition for food
supplies may lead to increased food  availability for the fish remaining.
Increased food availability may balance any  negative effects of
acid-Induced metabolic stress.
                                  5-102

-------
     Detailed studies of effects of food availability on  fish, at all
life history stages,  in acidic  waters  are not,  however, available.
Therefore, the conclusion that  shifts  in food availability with
acidification have no adverse effects  on fish survival or production is
preliminary.  The growth response for  any particular species may depend
on its sensitivity to acidic  conditions  relative  to the sensitivity of
desirable prey items.  As a group,  aquatic invertebrates  appear more
tolerant than fish.  Therefore, fish that feed  primarily  on
invertebrates often experience  increases in growth with acidification.
However, fish that require or prefer prey intolerant of acidification
may be adversely affected by  reduced food supplies.

5.6.2.4  Episodic Fish Kills—Observations of dead and dying fish in
acidifying waters are not common.   Mechanisms of  population extinction
(e.g., recruitment failure)  are often  too subtle  to be easily detected.
However, instances of massive acute mortalities of adult  and young fish
have occurred, typically associated with rapid  decreases  in pH resulting
from large influxes of acid into the system during spring snowmelt or
heavy autumn rains.  Chemical characteristics and occurrence of these
short-term acid episodes are  described in Chapter E-4, Section 4.4.2.
In general, organisms are less  tolerant  of rapid  increases in toxic
substances than they are of chronic exposure and  gradual  changes in
concentration.  As a result,  the rapid fluctuations in acidity associ-
ated with short-term acidification  (defined in  Chapter E-4, Section
4.2.3) may be particularly lethal  to fish and may play an important role
in the disappearance of fish  from acidified lakes and streams.

     Fish kills apparently associated  with acid episodes  have been
reported numerous times in the  streams and rivers of southern Norway
(Jensen and Snekvik 1972, Muniz 1981).  The first records of mass
mortality of Atlantic salmon date from 1911 and 1914, and coincide
closely with the sharp drop in  salmon  catch recorded for  rivers in
southern Norway over the years  1910-17 (Figure 5-7).  Additional obser-
vations of mass mortality were  reported  in 1920,  1922, 1925, 1948, and
1969, in each case following either heavy autumn  rains or rapid
snowmelt, particularly in May to June.  In 1948,  a massive mortality of
salmon and sea trout (Salmo trutta) occurred in the Frafjord River.  At
least 200 dead salmon and sea trout were collected, some  of the salmon
weighing more than 20 kg.  pH measurements (colormetric)  taken when dead
fish first appeared were 3.9  to 4.2.  One month later the pH was 4.7 to
4.8.

     A similar episode occurred in the Tovdal River (Norway) in the
spring of 1975 (Leivestad et al. 1976).   Dead fish were first observed
at the end of March.   During the first weeks of April thousands of dead
trout covered a 30 km stretch of the river.  The  Tovdal River valley is
sparsely populated and has no industry.   Veterinary tests failed to find
signs of any known fish diseases.   The pH of the  river was about 5.0.
In March, at two stations downstream,  a  drop in water pH  was recorded
apparently associated with a  period of snowmelt at altitudes below 400
m.  At higher altitudes, no dead fish  were found, and temperatures
probably never rose above freezing.


                                  5-103

-------
     Leivestad and Mum'z (1976)  observed the physiological  response of
fish to this acid episode in the Tovdal  River.   Trout  surviving within
the affected 30 km area of river had substantially  lower  levels of
plasma chloride and plasma sodium than  did fish  from apparently
unimpacted reaches of the river.  In the upper reaches of the river,  the
snow started to melt on April  21 and continued at a moderate rate until
May 6.  The pH dropped from 5.2  to a minimum of  4.65.   Blood samples
from fish collected in this area on May 15 had significantly lower
plasma sodium and/or chloride compared  to samples from fish from the
same area taken before and after snowmelting.  Leivestad  and Muniz
(1976) proposed that-increased acidity  interferred  with osmoregulation
via perhaps impairment of the active transport mechanism  for sodium
and/or chloride ions through the gill epithelium.   Additional evidence
for the adverse effects of acidity on ionic balance in fish is
available from laboratory bioassays (Section 5.6.4.1.5).

     Fish kills attributed to short-term acidification have been
reported for only one water outside of  Norway.   During each spring 1978
to 1981, coincident with spring  run-off, dead and dying fish, especially
pumpkinseed sunfish, were observed in Plastic Lake, LaCloche Mountain
region, Ontario (Harvey 1979, Harvey and Lee 1982). Measured pH levels
were 5.5 at the lake surface and 3.8 in the major inlet.   Field
experiments to verify these toxic conditions in  Plastic lake were
completed in"1981 and are described in  Section 5.6.3.3.
                                                    •
     In addition to these observations  of mass mortalities of fish
attributed to acid episodes under natural  field  conditions, several
instances of unusually heavy fish mortality have been  reported within
fish hatcheries receiving water  directly from lakes or rivers.  In
Norway, poor survival of eggs and newly-hatched  larvae of Atlantic
salmon, attributed to water acidity, were reported  as  early as 1926 in
hatcheries on rivers in Stfrlandet (Muniz 1981).  In Nova  Scotia, 19 to
38 percent mortality of Atlantic salmon fry occurred in 1975 to 1978  at
the Mersey River hatchery (Farmer et al. 1981).  In Norway and Nova
Scotia, neutralization of the water by  passage through limestone
alleviated the problem.  In the  Adirondacks, adult, yearling, and larval
brook trout, which had been maintained  without incident over the winter
1976-77 in water from Little Moose Lake, experienced distress and
mortality during the first major winter thaw in  early  March (Schofield
and Trojnar 1980). The minimum pH measured was 5.9  on  March 13 (with
0.39 mg Al £-!). Mortalities occurred over a 5-day  period March 13
to 17.  Deaths included three adult brook trout, 25 yearlings (132 to
167 mm), and an undetermined number of  recently  hatched fry.  Eyed brook
trout eggs exposed to the same water did not experience significant
mortality.

     All of the above observations of fish kills were  associated with
episodic increases in acidity.  Grahn (1980), however,  recorded fish
kills in two lakes in Sweden associated with decreases in acidity.  In
June 1978 in Lake Ransjon and in June 1979 in Lake  Amten, large numbers
of dead ciscoe (Coregonus albula) were  discovered.  A  weather pattern of
heavy rainfall, decreasing pH levels, and increasing aluminum


                                  5-104

-------
concentrations in the lakes,  followed by  a long  period of  dry,  sunny
weather preceded fish kills in both lakes.  pH levels in the lake
epilimnion increased from approximately 4.9 and  5.4  to 5.4 and  6.0,
respectively.  Grahn (1980) hypothesized  that the  increase in pH level
precipitated aluminum hydroxide,  and that ciscoe,  migrating into the
epilimnion to feed,  were exposed  to these lethal conditions.  Laboratory
experiments (Section 5.6.4.2)  have also noted that aluminum is
particularly toxic to fish as  it  precipitates out  of solution.  Dlckson
(1978) reported that acidic lake  waters immediately  after  liming (pH
values increased to 5.5 and above), were  toxic to  trout.   Concentrations
of aluminum were still high and,  presumably,  aluminum would be  actively
precipitating out of solution.

5.6.2.5  Accumulation of Metals in Fish--An indirect result of
acidification of surface waters may be accumulation  of metals in fish.
Evidence for this relationship is derived from correlations between
metal concentrations In fish and  lake and stream pH  levels, and
evaluations of metal chemistry and availability  in oligotrophic, acidic
waters.  Data are presented in Chapter E-6, Section  6.2.3.  Elevated
levels of mercury in fish from acidic waters  have  been measured in
Sweden, Ontario, and the Adirondack region of New  York.  (Aimer et al.
1978, SchofieW 1978, Bloomfield  et al. 1980, Hakanson 1980,
Jernelov 1980, Suns et al. 1980).  There  is no evidence that this
bioaccumulation has adverse effects on the fish} although  it may
represent a hazard for human health.   Other metals in addition  to
mercury occur at elevated concentrations  in acidified waters and
potentially may accumulate in  fish and other  biota.  Data  on these
accumulations and their effects on fish are,  however, very limited.

5.6.3   Field Experiments

     Correlations between fish population status and acidity of surface
waters, and field observations of declines in fish populations
concurrent with acidification  of  a lake,  river,  or stream, strongly
imply that acidification has serious detrimental effects on fish.  Such
observations, however, rarely  prove cause-and-effect.  In  experiments,
one variable is changed, and the  response to  that  change is recorded.
Thus, the cause and its effect are clearly delineated.

     Whole-ecosystem acidification experiments have  been carried out  at
two locations:  Lake 223 in the Experimental  Lakes Area, Ontario and
Morris Brook in the Hubbard Brook Experimental Forest, New Hampshire.
In both cases, acid was added  directly to the water  and pH levels were
held fairly constant.  Despite these deviations  from the process of
acidification 1n nature, results  from these two  experiments demonstrate
important biological changes associated with  increased water acidity.

5.6.3.1   Experimental acidification of Lake  223,  Ontario—Lake 223 is a
small, oligotrophic lake on the Precambrlan Shield of western Ontario.
Prior to acidification, surface waters had an average alkalinity of
about'80 yeq sr1 and pH of 6.5 to 6.9.  Five  species of fish were
present: lake trout, white sucker, fathead minnow  (P1mephales promelas),
                                  5-105

-------
pearl dace (Semotolus  margarita)  and  slimy  sculpin  (Cottus cognatus).
Beginning in 1976,  additions  of  sulfuric  acid  to the lake epilimnion
gradually reduced lake pH.   Early in  each ice-free  season, lake pH was
decreased to a predetermined  value and  then  maintained at that value
through the following  spring,  at which  time  pH was  again reduced.  Mean
pH values were 6.8  in  1976,  6.1  in 1977,  5.8 in 1978, 5.6 in 1979, 5.4
in 1980, and 5.1 in 1981.   Biological responses to  this acidification
have been described in Schindler et al. 1980b, Schindler 1980, Malley et
al. 1982, Schindler and Turner 1982,  Mills  1982, NRCC 1981, and
U.S./Canada MOI 1982,  and  are summarized  in  Table 5-10.

     A number of important biological changes  occurred at pH values of
5.8 to 6.0, notably the disappearance of  the opossum shrimp (Mysis
relicta), a henthic/planktonic crustacean (Section  5.5.3), and the
collapse of the fathead minnow population.   Although both these species
were important prey for lake trout in the'lake, no  effects on trout
populations were detected.   Lake trout  density and  population structure
remained stable, and year-class  recruitment failures were not detected
until 1981 at a pH of  5.1.   At the onset  of acidification (1976),
fathead minnows were abundant while pearl dace were rare.  With the
collapse and eventual  extinction of the fathead minnow population as the
pH declined to 5.5, pearl  dace abundance  increased  dramatically (perhaps
in response to the loss of its closest  competitor). The increased
abundance of pearl  dace and a succession  of strong  year classes of white
suckers in 1978 to 1980 apparently provided adequate food alternatives
for the lake trout.

     Despite many changes  in lower trophic  levels,  lake trout and white
sucker populations showed  no definite indications of stress until 1981,
pH about 5.1, when reproductive  failures  occurred.  During the early
years of acidification, population numbers  of  both  species increased and
growth rates were relatively unchanged.  The primary food source for
white suckers, benthic dipterans, increased in abundance.  Although
types of prey available to lake  trout changed  dramatically, suitable
food remained abundant.  Both species spawned  successfully all years of
study prior to 1981, and there were no  indications  of egg resorption or
skeletal malformations.

     The population of bottom-dwelling  slimy sculpin gradually declined
throughout the acidification 1976 to  1981.   Potential reasons for the
decline include direct adverse effects  of increased acidity and/or
increased trout predation, associated with  an  increase in water clarity.

     Among the fish, fathead minnow seemed  to  be most sensitive to
acidification.  Fathead minnows  are ubiquitous in  lakes  in northern
North America and form an  important part  of aquatic food chains. The
population in Lake 223 disappeared extremely quickly, probably as a
result of two factors:  its particular  sensitivity  to acidity and its
short life span.  Recruitment failure occurred initially at pH 5.8  in
1978.  Prior to acidification, fathead  minnow  in  Lake 223 typically
lived only three years.  Natural mortality  rates  during  their  second and
third years of life were extremely high,  over  50  percent per year,


                                  5-106

-------
      TABLE 5-10.  BIOLOGICAL CHANGES IN LAKE 223 IN RESPONSE TO
   EXPERIMENTAL ACIDIFICATION (MILLS 1982, SCHINDLER AND  TURNER  1982)
pH                        Recorded change


Below 6.5   Increased bacterial  sulfate reduction partially neutralize
              acid additions
            Increased abundance  of Chlorophyta (green  algae)
            Decreased abundance  of Chrysophyceans (golden brown  algae)
            Increased abundance  of rotifers
            Increased dlpteran emergence

5.8-6.0     Disappearance of the opossum shrimp (Mysis relicta)
            Reproductive Impairment of the fathead minnow (Pimephales
              promelas)
            Possible Increased embryonic mortality of  lake trout
              (Salyelinus namaycush)
            Inhibition of calcification of exoskeleton of crayfish
              (Orconectes vlrlUs)
            Disappearance of the copepod Dlaptomus sicills

5.3-5.8     Increased hypollmnetlc primary production
            Development of Mougeotea  algal mats along  shoreline
            Increased infestation of  crayfish  with a parasite Thelohania
              sp.
            Collapse of the fathead minnow population
            Increased abundance  of the pearl dace minnow  (Semotilus
              margarita)
            Decreased abundance  of the slimy sculpin (Cottus cognatus)
            Decreased abundance  of crayfish           	
            Increased abundance  of white sucker (Catostomus commersonl)
            Increased abundance  of lake trout
            Disappearance of copepod  Epischura lacystris
            First appearance of  the cladoceran DaphnlaTcatawba x
              schoedleri

Below 5.3   Recruitment failure  of lake trout
            Recruitment failure  of white sucker
                                  5-107

-------
presumably as a result of heavy  trout predation.  Few individuals
remained after the second year of life.  Year-class failure in 1978,
therefore, left few spawning  adults  (age 2  and 3) the following year.
Successive year-class failures in 1978 and  1979 assured the rapid
disappearance of this species from Lake 223.

     In summary, experimental  acidification of Lake 223 resulted in
severe changes in fish populations at pH values as high as 5.8 to 6.0.
Adverse effects on fish and loss of  populations occurred  primarily as a
result of recruitment failures rather than  as a result of increased
mortality of adult fish or reductions in food supplies.

5.6.3.2  Experimental Acidification  of Morris Brook, New  Hampshire--
Norn's Brook, a third order stream in the Hubbard Brook Experimental
Forest, New Hampshire, was experimentally acidified to pH 4.0 from April
to September 1977 (Hall  et al. 1980, Hall and Likens 1980a,b).  Brook
trout were observed in the study section before and after acid addition.
Small numbers of trout confined  in the study section during low water in
June, July, and August were exposed  continuously to water at pH 4.0 to
5.0 and total aluminum levels up to  about 0.23 mg srl.  Trout
captured at pH 4.0, 5.0 and 6.4  in August showed no evidence of
pathological changes in gill  structure.  Most of the trout, however,
moved downstream to areas of  higher  pH at the onset of acid addition in
the spring.  No mortality was observed, only a general avoidance
reaction.  Potential effects  on  young-of-the-year trout and reproductive
success were not included in  this study.

5.6.3.3  Exposure of Fish to  Acidic  Surface Waters—In addition to the
above field experiments involving acidification of an entire ecosystem,
smaller scale field experiments  have been conducted involving the
transfer of fish into acidic  lakes and streams.  It is important to
distinguish these small-scale field  experiments from similar exposures
of fish to acid waters in laboratory experiments for two  reasons:   (1)
water quality conditions in field experiments may undergo substantial
natural fluctuations while conditions are usually held rather constant
in laboratory experiments, and (2) many laboratory experiments create
acidic water by diluting strong  acids (H^SO^, HN03, HC )  into
nonacidic background water.  These artificially acidic waters may not
precisely mimic acidified surface waters, and,  as a result, fish
responses recorded in laboratory bioassays  may  not always accurately
represent what would occur in the  field.   In this section,  in situ
exposures of fish to acidic surface  waters  are  reviewed in  addition to
experiments that, although conducted in  a laboratory or hatchery, used
unmodified acidic water taken directly from an  acidic lake(s) and/or
stream(s).

     Excessive mortality of adult  fish has  been observed  in a number of
in situ experiments with fish held  in cages in  acidic waters.  Following
observation of fish kills in  Plastic lake  (LaCloche Mountain  region,
Section 5.6.2.4) in 1979 and  1980,  during the  spring of 1981  rainbow
trout  (Salmo gairdneri) were  held  in cages  at four  locations  in Plastic
Lake and at four locations in a  control,  nonacidic  lake  (Harvey et  al.
                                  5-108

-------
1982).  No mortality occurred at any of the cage sites  in the control
lake (pH 6.09 to 7.34).  In Plastic Lake,  however,  mortality  ranged  from
12 percent at the lake outlet (pH 5.0 to 5.85)  to 100 percent at  the
inlet (pH 4.03 to 4.09).  At the inlet, mortalities commenced on  the
first day and all fish were dead within 48 hr.   Aluminum  accumulated
rapidly on the gills of fish tested in Plastic  Lake.

     During the winter (December to April) 1971-72, Hultberg  (1977)
placed seatrout and minnows (Phoxinus phoxinus), both with a  mean length
of 6.5 cm, at ten test stations ranging in pH from 4.3  to 6.0 within the
watershed of Lake Alevatten, Sweden.  At all  but three  of the test
stations native minnow populations had disappeared  within the ten years
preceding the experiment.   Fifty-three percent  of the seatrout and 91
percent of the minnows died during the four-month test.   Most of  the
mortalities (68 percent of the seatrout total mortality;  59 percent  for
minnows) coincided with periodic drops in  pH  level.

     Several Norwegian laboratory experiments with  adult  fish have used
acidic stream waters (Leivestad et al. 1976,  Grande et  al.  1978).
During simultaneous exposure to water from an acidic brook, pH 4.4 to
4.7, all yearling rainbow  trout, Atlantic  salmon, and brown trout died
within 32 days.  Brook trout were more tolerant, with 30  percent
survival of one-year-old trout after 80 days.  Similarly, in  tests with
finger!ing age 0+ fish in  acidic stream water,  rainbow  trout  and
Atlantic salmon were least tolerant (all dead within 12 days),  brown
trout intermediate (all dead within 32 days), and brook trout
substantially more tolerant (50 percent survival after 42 days).   By
comparison, in stocking experiments at Lake Langtjern, Norway (mean  pH
4.95), 24 and 61 percent (age 0+ and age 1+ fish, respectively) of brook
trout stocked were recaptured, as compared to 0.6 and 19  percent  of  the
brown trout and none of the rainbow trout (Grande et al.  1978).
Long-term exposure of brook trout to acidic stream  water  (mean pH 4.6,
range 4.2 to 5.0) resulted in decreased growth  and  reductions in  plasma
sodium and chloride levels.

     A number of studies have also examined survival of fish  eggs
incubated in waters from acidic lakes and  streams (Table  5-11).
Hatching success and egg survival  of brook trout ova decreased sharply
between pH levels 5.0 and  4.6.  For brown  trout, hatching was near 100
percent at pH levels 6.2 and 6.5,  but 0 percent at  pH 4.8 and 5.1.   The
critical pH for hatching of Atlantic salmon eggs appears  to be 5.0 to
5.6; for walleye about pH  5.4; for roach,  something above pH  5.7.

     In three studies, results from in situ incubation experiments were
compared with concurrent surveys of occurrence  of fish  species within
the same waters.  Leivestad et al. (1976)  reported  that no brown  trout
eggs hatched and few trout fry were found  (by electrofishing)  in  an
acidic tributary (pH 4.8),  formerly an important spawning ground.  By
contrast, in a second tributary with inferior spawning conditions but pH
6.2, numerous trout fry were collected. Harriman and Morrison (1982)
reported no survival of Atlantic salmon eggs  incubated  in acidic  streams
(pH 4.2 to 4.4) draining forested catchments  in Scotland  and  the  absence
                                  5-109

-------
         TABLE 5-11.  SUMMARY OF FIELD EXPERIMENTS WITH FISH EGGS
                     EXPOSED TO ACIDIC SURFACE WATERS
Species
Brook trout9








Brown troutd


Brown troutf


Atlantic
salmon*1
Location
Hatchery with
water from
Honnedaga Lake
plus 6 tribu-
tary streams




In situ in 2
Norwegian
streams
In situ in 2
Norwegian
streams
In situ in
acidic Mandal
PH
4.5
4.6
5.0
5.1
5.3
5.4
5.6


4.8
~7

5.13
6.55

4.9
~7
% Survival
25
60
90
95
80
85
85


0
-100

0
90

< 1
80
Comments Reference
0.10 mg Zn £-1
0.05
0.002
0.002
0.04
0.03
0.02
Exposure from
eyed stage



Spawning observed
in acidic brook



g








d


f


d

Atlantic
  sal monk
Atlantic
  salmon3
River and a
near-neutral
tributary,
Norway

In situ at
several rivers
in Stfrlandet
Norway

In situ in
streams,
Scotland
 5.0
 5.5
4.2
4.4
4.9
5.8
 0
 0
54
30
       Critical  pH
       for hatching
Comparison of
forested vs non-
forested catchments
                                5-110

-------
                             TABLE  5-11.   CONTINUED
    Species
 Roache
 Walleyec
  Location
 pH   % Survival     Comments
                        Reference
Perche





In situ in
Lakes
Stensjon,
Trehorningen,
and Malaren,
Sweden
4.7
5.7
7.5



28
50
89



e





As above
In situ in
series of
small streams
in LaCloche
Mt. area,
Ontari o
4.7
5.7
7.5
4.6
6.7
  0
 14
100
       Hatching success
       significantly
       reduced at pH
References

aHarriman and Morrison 1982
bHendrey and Wright 1976;  Muniz  and Leivestad  1980a
cHulsman and Powles 1981
Leivestad et al.  1976
eMilbrink and Johansson 1975
fMuniz and Leivestad 1980a
gSchofield 1965
                                 5-111

-------
of fish from the same  streams  in  an electrofishing survey.  Finally,
Milbrink and Johansson (1975)  incubated perch (Perca fluviatilis) and
roach eggs in situ in  Lakes Malaren (pH 7.5), Stensjon (pH   5.7), and
Trehorningen (pH - 4.7)  in Sweden,  while some perch eggs hatched in
all three lakes (89, 50,  and 28 percent, respectively), very few or no
roach eggs hatched in  the two  acidic lakes  (14 percent in Lake Stensjon,
0 percent in Trehorningen).  Likewise, perch populations occurred in all
three lakes, although  extremely few perch were collected in the most
acidic lake, Trehorningen.  Roach, on the other hand, have apparently
disappeared from Lake  Trehorningen.  Roach  are still prevalent in both
Lake Stensjon and Malaren.

5.6.4  Laboratory Experiments  (J. P. Baker  and P. G. Daye)

     One of the best ways to prove cause and effect is to conduct
experiments in a carefully controlled environment, i.e., the laboratory.
Experimental conditions and fish  response can be clearly quantified and
dose-response relationships developed with  a minimum of time and effort.
Unfortunately, laboratory experiments have  several drawbacks.  For one,
the simplified, controlled environment of the laboratory may differ from
the natural environment in essential attributes.  Factors that cannot be
easily incorporated into laboratory experiments include:  (1) the
temporal and spatial  variability  in the field environment; and (2) the
potential for compensatory mortality, i.e., shifts in the efficacy of
natural mortality factors (e.g.,  predation, starvation) resulting from
the addition of acid-induced mortality and/or stress.  Consequently,
results from laboratory experiments cannot  be translated automatically
into an expected response in the  field.

     Serious gaps exist in the understanding of how to use laboratory
results in a quantitative assessment of field observations.  It  has
never been definitely  demonstrated that "X" conditions that yield "Y"
response in the laboratory  (e.g., 40 percent mortality) will also yield
"Y" response in the field.  Laboratory results are, however, useful in
firmly establishing cause-and-effect, that  increasing acidity has
adverse effects on fish, and a qualitative  estimate of the levels of
acidity of concern.

     The more closely  the laboratory environment simulates the field
experience, the more realistic the observed response.  Laboratory
bioassays conducted to date vary  substantially in their use of
conditions appropriate to the  problem of acidification of surface
waters.  Most laboratory experiments concerned with acidification have
focused on the effects of low  pH  on fish.   With acidification, however,
other factors also change in association with decreasing pH (Chapter
E-4, Section 4.6).  Increased  aluminum concentrations in acidic  waters,
in particular, have been shown to affect fish adversely (Section
5.6.4.2).   Unfortunately, most of the bioassay results to date  have
failed to include aluminum.  Thus, these results must be interpreted
with caution. In addition to aluminum concentration, other factors
change with acidification, e.g.,  increased  manganese and zinc
concentrations and perhaps  a decrease in dissolved organic carbon


                                 5-112

-------
(Chapter E-4, Section 4.6).  The importance of these other  changes  to
fish populations in acidified waters has yet to be  delineated  in either
laboratory or field experiments.

     Within the discussion of laboratory experiments,  Section  5.6.4.1
considers effects of low pH on fish.  Section 5.6.4.2  examines combined
effects of both low pH and elevated aluminum (and other  metals). Because
of the large number of experiments dealing with low pH,  Section 5.6.4.1
is subdivided into experiments dealing with survival,  reproduction,
growth, behavior, and physiological  responses.  Reproduction is
arbitrarily defined as including data on survival of fish larvae and fry
in acidic water.  Section 5.6.4.1.1  (Survival) therefore considers  only
data for fish approximately aged 4 months (finger!ings)  and older.
Questions related to acclimation to  acidic waters and  differences in
tolerances among fish strains, as related to possible  mitigation of
effects of acidification, are discussed in Section  5.9.  Interpretation
of laboratory results must also consider that fish  response in a
bioassay is a function of testing conditions (e.g.,  temperature,
flow-through or statij: water supply), background water quality (e.g.,
water hardness, concentrations of dissolved gases),  and  characteristics
of the fish tested (e.g., prior exposures and stress,  size, age,
condition).

5.6.4.1  Effects of Low pH —

5.6.4.1.1  Survival.  The majority of laboratory experiments designed to
determine the direct toxicity of elevated hydrogen  ion concentrations to
fish have been short-term, acute bioassays involving principally pH
levels 4.0 and below (Table 5-12).  If 2 days is arbitrarily selected as
the length of an acid episode, laboratory experiments  suggest  that  a 50
percent fish kill would occur at approximately pH 3.5  for brook trout,
pH 3.8 for brown trout, pH 3.8 to 3.9 for white suckers, and pH 4.0 for
rainbow trout.  In contrast,  field observations of  fish  kills  (Section
5.6.2.4 and 5.6.3.3) indicate mortality of:  (1) Atlantic salmon and
sea-run brown trout in Frafjord River,  Norway in 1948  at pH 3.9 to  4.2;
(2) brown trout in the Tovdal  River, Norway in 1975  at pH 5.0; (3)
rainbow trout in Plastic Lake, Ontario at pH 4.0 to 4.1; (4) brook  trout
in Little Moose hatchery, Adirondacks,  NY, at pH 5.9;  and (5)  brook
trout in Sinking Creek, PA, at pH 4.4 and below.

     A few experiments have considered survival of  fish  following
longer-term exposure to low pH levels (Table 5-13).  Apparently, adult
fish can survive quite low pH levels for fairly long time periods.  For
periods up to 11 days, brook  trout were able to withstand pH levels as
low as 4.2 with only small reductions in survival.   During  even longer
periods of exposure (65 to 150 days), however, a pH  level of 4.4 to 4.5
was severely toxic, and only  at pH levels of 5.0 and above  was brook
trout survival unaffected.  Long term experiments (> 100 days) with
adult rainbow trout, brown trout, arctic char, and Tathead  minnow
indicated no substantial  reductions  in  survival  at  the lowest  pH levels
tested, 5.0, 4.8 and 4.6, respectively.
                                  5-113

-------
                        TABLE 5-12.  MEDIAN SURVIVAL TIME (HR) FOR FISH EXPOSED TO pH LEVELS
en
i

Age/ 2.0- 2.6-
Species size 2.5 2.8
Brook trout * 10-60 g < 1
fngl*
2 9 1 2
90 g 1 4
60-130 g < 1
50 g
50-90 g

Rainbow trout * 1 g
* 130 g 1-4
* 200-300 g 2 2
2-5 g
* 2-5 g
* 5-15 cm
5-15 cm

Brown trout * 1-5 g
* 6 g 1-2
* 60-80 g 3 4
3
Arctic char * 100-170 g 3
White sucker 7 mo
Roach 7-13 cm
*Experiments using low alkalinity water.
tfngl = fingerling, age 0+, weight usually
References -
a. Daye and Garside 1975
b. Johnson 1975
c. Robinson et al. 1976
d. Packer and Dunson 1972
e. Swarts et al. 1978
f. Falk and Dunson 1977
g. Kwain 1975

3.0-
3.1

2-3
3-6
9
1




4
5
1
< 1
1
2


3-7
9

4
1
< 1

< 50 g.








PH level
3.2- 3.4- 3.6- 3.8- 4.0- 4.2- 4.4-
3.3 3.5 3.7 3.9 4.1 4.3 4.5
7
3-6 6-18 10-38 14-51 20-270
12-14 45
18 61-66 334
5-9
25 66-70
10-32
8 37
23
8 18

2 3 6 17 83 117 133
1 2 6 27 133
2 3 8 22 70
3 7 18 55
120
25 40 2-4




2 5 10 30-200 350 1000
1 3



h. Edwards and Hjeldnes 1977
1. McDonald et al. 1980
j. Lloyd and Jordan 1964
k. Brown 1981 with 0.1 mM Ca
1. Edwards and Gjedrem 1979
m. Beamish 1972


Reference

a
b
c
c
d
e
f
g

g
h
i
i
j
j
k

1
h

h
m
j











-------
  TABLE  5-13.   PERCENT SURVIVAL OF  FISH  FOLLOWING  CHRONIC  EXPOSURE TO  LOW  pH LEVELS
Species
Brook trout
*
*



Rainbow trout*

Brown trout*
Arctic char*
Fathead Minnow
Hagflsh*

Age/Size
100-300g
10-60g
5g
50g
150-360g

200-300g

60-80g
100-170g
1 yr
Mature
Adult
Length
of
Exposure
(days) 3.2 3.6
5
7 0 85
11
65
150

100

100
100
400
20

4.2- 4.5- 4.8-
4.4 4.6 5.0
60-90 100
100 100
100
0-36 0 75

93

94
90
86
80
36

5.2-
5.6

100




96

98
100
75
79

5.9- 6.5- 7.0-
6.2 6.8 7.5
100
100


75 100

97

95
100
85 75 85
100 93

Reference
a
b
c
d
e

f

f
f
9
h

•Experiments using low alkalinity water.
References

'Dively et al.  1977
bDaye  and Garslde 1975
C8aker 1981
dSwartz et al.  1978
eMenendez 1976
fEdwards and Hjeldnes 1977
SMount 1973
"Craig and Baksl 1977

-------
     An important objective of many of these experiments was  not  solely
to determine fish mortality at low pH  levels but  also  to evaluate
factors that influence fish tolerance  to low pH.   For  example, Lloyd and
Jordan (1964) and Kwain (1975)  concluded that as  fish  grow older they
became more acid tolerant.   Higher temperatures (5 to  20 C) tended to
decrease fish survival  at low pH (Kwain 1975,  Edwards  and Gjedrem 1979,
Robinson et al.  1976).   Water hardness also affected fish tolerance.
Lloyd and Jordan (1964) and McDonald et al. (1980) noted that at low pH
levels (pH^4.0), the resistance of rainbow trout to  acids increased
with increasing  hardness of water.   As a result,  experiments conducted
in high alkalinity, hard water (see Tables  5-12 and 5-13) are relatively
inappropriate for assessing effects of acidic  deposition on fish, a
phenomenon confined to dilute,  poorly  buffered surface waters.  Brown
(1981) suggested that higher calcium levels (more so than higher sodium,
potassium, or magnesium levels)  in harder water may be responsible for
the increase in  resistance.  Within even dilute,  low alkalinity waters,
small changes in calcium concentration (0 to 2 mg £-1) have been
shown to have a  significant influence  on survival  times of fish (Brown
1982).  Similarly, in the field (in Norway) the number of fish!ess lakes
was correlated with both pH level  and  calcium level, with the greatest
number of fishless lakes having both low pH and low calcium (Wright and
Snekvik 1978; Section 5.6.2.1.3.1). The sensitivity of fish to low pH
obviously interacts with a number of other  stress and  condition factors.

5.6.4.1.2  Reproduction.  As discussed in Section 5.6.2.2, loss of fish
populations with acidification  is in many lakes and rivers preceded by
successive recruitment failures.  These field observations suggest that
fish reproductive processes are particularly sensitive to acidic
conditions.  This conclusion is supported by laboratory experiments on
effects of low pH on spawning behavior, egg production, and egg and fry
survival.  Tolerance to low pH  varies  considerably among the early
development stages and reproductive processes. At the same time, many
fish reproduce during the spring season, a  period of large fluctuations
in water chemistry.  Information on the timing of these fluctuations in
water quality and the occurrence and sensitivity  of various reproductive
processes and stages has yet to be tied together  in an analysis of which
reproductive process(es) and/or stage(s) may play key  roles in the
success or failure of recruitment and  survival of the  population.

     Studies on  the effect of low pH on the entire reproductive cycle
have been completed only for brook trout (Menendez 1976), fathead minnow
(Mount 1973), flagfish (Jordanella floridae) (Craig and Baksi 1977), and
desert pupfish (Cyprinodon n. nevadensls)' (Lee and Gerking 1981)  (Figure
5-12).  pH level had some eTfect on all stages (processes) tested, with
the exception of number of eggs spawned by  brook  trout.  However,
sensitivity varied among both life history  stages (processes) and
species.  For brook trout,  survival of eggs and fry appeared  to be the
phase most sensitive to low pH levels, with survival significantly {p <
0.05) reduced at pH 6.1 and below.  For fathead minnow, flagfish  and
desert pupfish,  on the other hand, egg production appeared particularly
sensitive to low pH, with reductions in eggs produced  per female  at pH
levels between 6.0 and 7.0.  Lee and Gerking (1981) concluded that
                                  5-116

-------
Q  O
UJ  t-
o.  o
oo
   M-
OO  O
C5
O  &«
UJ
UJ  TJ

i  S-
rj     80
  co
  CJ
       60
  tt   40

  UJ

  CO

  2   20
                                                  0'   '   '   '   '
                                                        1   I   I
o
z  
-------
reduced egg production at low pH  levels  resulted  primarily  from
inhibition of oogenesis (rather than  interference with  normal spawning
activity).  Ruby et al. (1977, 1978)  also observed  retarded oocyte
growth (and reduced sperm production)  for flagfish  exposed  to pH 6.0
relative to the control of pH 6.8.

     Unfortunately three of these four experiments  (all  except Craig  and
Baksi 1977) were conducted in hard  water (alkalinity  >  500  yeq £"1
and two used fish species that do not occur in surface  waters sensitive
to acidic deposition.  Conclusions,  therefore, must be  interpreted
cautiously.  Results for brook trout (Menendez 1976), in particular,
differ markedly from results from other  researchers using low alkalinity
water (Figures 5-13 and 5-14) and/or naturally acidic surface waters
(Section 5.6.3.3).  Life cycle experiments with both  fish species and
conditions appropriate to acidification  of dilute surface waters are  not
yet available.  Thus, the relative  sensitivities  of reproductive stages
to low pH cannot be accurately assessed  at this time.

     Data on survival of fish embryos at low pH levels  in laboratory
experiments are summarized in Figure 5-13.  In each case, hatching  was
reduced at low pH levels.  Among  North American freshwater  species,
brook trout was the most tolerant.   Excluding results from  Menendez
(1976), numbers of brook trout embryos surviving  through hatching were
reduced substantially (< 50 percent hatching) only  at pH levels below
4.5.  Hatchability of white sucker eggs, on the other hand, dropped off
sharply at pH levels 5.0 to 5.2.   Number of fathead minnow  embryos
hatching declined at pH 5.9.  In  experiments conducted  in Scandinavia
and Great Britain, survival through hatching was  reduced below
approximately pH 4.4 for sea-run  brown trout and  below  pH 4.6 for roach.
Experiments with perch and Atlantic salmon yielded  inconsistent results.
These pH values for effects on egg survival are distinctly  higher than
values noted as acutely toxic to  adults  (pH 3.5 for brook trout; pH 3.8
to 3.9 for white suckers; pH 3.8  for brown trout)  (Section  5.6.4.1.1).

     A number of studies have noted that the hatching process  itself
appears pH sensitive (Runn et al. 1977;  Peterson  et al. 1980a,b; Baker
1981).  For eggs exposed to low pH either throughout  their  development
or just during hatching, a large  proportion of embryos  hatch
incompletely, with fry remaining  partially encapsulated for days
following hatching.  Delay or prevention of hatching  can be induced by
transfer of eggs into low pH water just prior to  hatching,  and  normal
hatching may occur if eggs are transferred just prior to hatching from
low pH water into control water.   Thus,  mechanisms  involved in  the
hatching process especially may be key factors limiting embryo  survival
in low pH water (disintegration of the chorion, facilitating mechanical
rupture of the chorion by embryo  trunk movements  at hatching;  Bell  et
al. 1969, Yamagami 1973, 1981).  Mechanisms proposed  involved:   (1) the
relationship between pH and activity of the hatching  enzyme (Yamagami
1973), (2) thicker, more rigid egg capsules at lower pH, with increased
resistance to degradation  (Runn et al. 1977, Peterson et al. 1980b),  and
(3) reduction in body movements inside eggs at low pH (Peterson  et  al.
1980b).


                                   5-118

-------
                                                               NORTH
                                                               AMERICAN
                                                               SPECIES
                                                                EUROPEAN
                                                                SPECIES
                                                                ATLANTIC
                                                                SALMON
                   4.0   4.5   5.0   5.5  6.0  6.5   7.0   7.5  8.0  8.5

                                         PH

                                        LEGEND
                           •  BROOK TROUT
                           o  FATHEAD MINNOW
                           *  PERCH
                           •  BROWN TROUT
a WHITE SUCKER
• ROACH
a ATLANTIC SALMON
Figure  5-13.  Effect of low pH on survival  of fish through hatching.

                                 References:
       a     Baker and Schofield   1982       g
       b     Swarts et al.          1978       h
       c     Trojnar                1977a      i
       d     Trojnar                1977b      j
       e     Johansson et al.      1977       k
       f     Mount                  1973
    Carrick                     1979
    Runn  et al. (in  1975)      1977
    Johansson and Mil brink     1976
    Peterson et al.             1980a
    Peterson et al.             1980b
                                     5-119

-------
100
80
60
40
20 -
 3.5

                   1
               I   /
               /   /

              I  /
              I   /
          LEGEND
                  /     •  BROOK TROUT
                  /     o  WHITE SUCKER
                                                             • ATLANTIC SALMON
                                                             °BROWN TROUT
                                                             xPIKE
                 pH
                                                        PH
 Figure 5-14.
Effect of low pH on survival  of fish as sac fry.  Solid
line, sac fry survival  through  swin-up following
development of eggs and hatching of larvae in low pH water
(expressed as percent  normal  hatch); Dashed line, sac fry
survival without previous  exposure to low pH.
        PART  (A)
          References
       a     Baker  and Schofield  1982
       b     Swarts et al.         1978
       c     Johansson et al.      1977
       d     Trojnar              19775
PART (B)

a    Daye and Garside           1975
b    Johansson and Kihlstrom   1975
c    Johansson et al.           1977
                                   5-120

-------
     Exposure of embryos to low pH levels during early stages of
development (particularly within the first day after fertilization  or
during water hardening) also adversely affected survival,  although  to  a
lesser extent than did exposure during hatching (Johansson et al. 1973,
Johansson and Milbrink 1976, Daye and Garside 1977,  Lee and Gerking
1981, Baker 1981).  For roach eggs exposed to pH 7.7 throughout their
development, 89 percent hatched successfully.  After exposure to pH 4.7
for the first 24 hr and then to pH 7.7 from 24 hr to hatch, 52 percent
hatched.  With exposure to pH 7.7 for 24 hr followed by pH 4.7 to hatch,
20 percent hatched.  Finally with exposure to pH 4.7 throughout
development, only 6 percent hatched successfully (Johansson and Milbrink
1976).

     The egg changes its character rapidly after being spawned.
Permeability decreases and the chorion hardens during the first few
hours after release, allowing the egg to become more resistant with time
(Lee and Gerking 1981).  Zotin (1965) noted that teleost eggs exchange
water with the surrounding solution primarily immediately after
fertilization and just before hatching.  Exchange of water and ions
between the egg and external medium during intermediate periods of
development occurs but is limited (Kalman 1959, Zotin 1965).

     Given the evidence that timing of exposure substantially affects
the sensitivity of embryos to low pH, it is obvious  that to determine
the impact of acidification on embryo survival, the  occurrence of
particularly susceptible stages must be evaluated in relation to the
timing of fluctuations in pH level in acidified surface waters.  As with
the toxicity of low pH to adult fish, the effect of  low pH on fish
embryos was also found to be a function of temperature (Kwain 1975).

     At intermediate pH levels, between those recorded to have no
consistent adverse effect on embryo survival and pH  levels that result
in near 100 percent mortality, some researchers (Mount 1973, Runn et al.
1977, Trojnar 1977b) have observed increased incidence of deformities  in
larvae after hatching.  Runn et al. (1977) suggest that these
malformations result, at least in part, from the prolongation of the
non-hatching period.  Peterson et al. (1980a), in contrast, reported no
increase in deformities of Atlantic salmon fry hatched at low pH levels
(5.5 to 4.5).

     Finally, pH may determine recruitment success for fish populations
in acidic waters by influencing the survival of young fish larvae (or
fry) after hatching.  The direct effect of low pH on fry survival has
been examined in laboratory experiments.  Fry survival in field
situations would also be strongly influenced by food availability,
predation, temperature, and a large number of other  environmental
factors.  In general, survival of fry in laboratory  bioassays decreased
below pH 4.0 to 4.5 for Atlantic salmon; pH 4.2 to 4.4 for brook trout;
pH 4.8 for brown trout; pH 5.0 to 5.5 for white suckers; and pH 5.2 for
pike (Figures 5-14 and 5-15).
                                  5-121
t09-262 0-83-16

-------
     100
 oo
°-    40 -
      20
                                       LEGEND

                                    • BROOK  TROUT
                                    O WHITE  SUCKER
Figure 5-15.
                                   pH

             Effect of pH on survival of fry exposed  for  14 days
             after swim-up and initiation of feeding.

             jjBaker and Schofield 1982
             bTrojnar 1977a; previous exposure during development at
              pH 8.'0 (o); previous exposure at pH 4.6 to 5.6 (•).
                                  5-122

-------
     Evaluations of the relative sensitivities of eggs, sac fry (fish
larvae after hatching but prior to initiation of feeding and swim-up),
and fry  (after initiation of feeding) have been inconsistent among
experiments, perhaps reflecting differences in species response.   Baker
and Schofield (1982) and Swarts et al. (1978) found in successive
experiments with brook trout and/or white sucker that sensitivity to low
pH decreased with age.  Also, a high proportion (> 75 percent)  of
embryos  alive at hatching survived through swim-up with continued
exposure to the same low pH level (Trojnar 1977a, Craig and Baksi 1977,
Baker and Schofield 1982).  Daye and Garside (1977, 1979),  on the other
hand, concluded that Atlantic salmon fry were more sensitive to low pH
than were eggs.  Likewise, Johansson et al. (1977) observed that
Atlantic salmon and brown trout (and to a lesser extent brook trout)
that survived through hatching at low pH levels (pH 4.1 to  5.0)
subsequently suffered substantial mortality (10 to 100 percent) during
the four weeks after hatching until  just prior to full  resorption of the
yolk sac.

     Therefore, while some researchers have concluded that  fry  are
relatively (as compared with fish eggs)  tolerant of low pH, other
researchers considered fry to be a particularly sensitive stage in the
reproductive cycle of fish.  Because as fry emerge from the nest,
"redd,"  or spawning tributary upon swim-up they may be subjected  to an
environment and water quality distinctly different from that to which
the eggs (and sac fry) were previously exposed, an understanding  of
these relative tolerances is important.

5.6.4.1.3  Growth.  The direct effect of low pH on fish growth  has been
examined in several  laboratory experiments.   Although field observations
of changes in growth with acidification  indicate a variable response  to
increased acidity (Section 5.6.2.3), reflecting the large number  of
variables determining growth in natural  situations, in the  laboratory
low pH has consistently resulted in  decreased growth.   These decreases
in growth often occur at pH levels above those producing substantial
fish mortality.  Edwards and Hjeldnes (1977)  observed a significant
(p < 0.001)  decrease in growth (relative to the control  at  pH 6.0)  of
yearling rainbow trout, brown trout, and arctic char held at pH 4.8 for
3.5 months;  mortality levels were less than 10 percent.   Jacobsen (1977)
found no significant decrease in growth of 18 month old brown trout
after 48 days, but tested pH levels  only down to 5.0.    Swarts  et al.
(1978)  and Baker (1981) noted delayed development of brook  trout  sac  fry
hatched at pH 4.6 and below.  For brook  trout embryos reared at pH 6.5,
6.0 and 5.5, fry were significantly  (p < 0.05)  shorter after 3  months
than were fry in control  water at pH 7.1 (Menendez 1976).   Likewise,
flagfish surviving through embryo development and 45 days after hatching
weighed significantly less at pH 6.0, 5.5, and 5.0 than  did fry at pH
6.8 (Craig and Baksi 1977)  and rainbow trout reared at pH 4.3 to  4.8
were shorter (p < 0.001)  than controls at pH  7.0 to 7.2  (Nelson 1982).

     The decrease in growth at low pH represents a sublethal  response to
elevated hydrogen ion concentrations and suggests that fish are
                                 5-123

-------
physiologically stressed at pH levels above those that produce acute or
chronic mortality.

5.6.4.1.4  Behavior.  Behavioral  responses of fish to low pH  probably
play an important role in determining the effect of surface water
acidification on fish populations.   Within a given aquatic system at any
time, water quality may vary substantially (Driscoll  1980).  If fish can
detect regions of low pH and by behavioral adaptation avoid exposure to
these toxic conditions, the impact of acidification may be, in part,
mitigated.  Mum'z and Leivestad (1980a)  reported observations of trout
concentrated into "refuge areas"  during  acid incidents".  In the acidic
river Gjor in Norway during spring  snow  melt, hundreds of brown trout
from the river crowded into a tiny  tributary with a higher pH.  If
experimentally restrained within the river, fish died within  a week.
Information on the presence of such "refuge" areas and the ability of
fish to detect and use these areas  are necessary for a complete
assessment of the impact of acidification.

     Unfortunately, laboratory (and field) data on behavioral responses
of fish to low pH are very limited.  Jones (1948) tested sticklebacks
(Gasterosteus aculeatus) in a sharp concentration gradient in a
laboratory apparatus.Fish were able to detect and avoid waters with pH
< 5.4, a value slightly above the lethal level of pH 5.0.  Hoglund
T1961) concluded that Atlantic salmon fingerlings avoid water at pH 5.3
and below, roach at pH 5.6 and below.  Johnson and Webster (1977)
investigated the effect of low pH on spawning site selection  of brook
trout.  Female trout clearly avoided areas of water upwelling at pH 4.0
and 4.5.  Discrimination was not evident at pH 5.0.  Preference by adult
brook trout for spawning in areas receiving neutral or alkaline aquifer
water may protect eggs and sac fry  from  adverse water quality
conditions.  Decreased spawning activity at low pH (discussed in Section
5.6.4.1.2) may therefore partially  reflect a behavioral response rather
than an adverse effect on reproductive physiology.

5.6.4.1.5  Physiological responses.  In  the laboratory a decrease in pH
level has been demonstrated to result in a wide diversity of
physiological responses in fish.   Some of these observed responses may
reflect only a general response of fish  to stress; others appear to be
specifically related to low pH.  The following does not represent a
complete review of the extensive and varied literature available on fish
responses to acidity; only major topics  are summarized.  Fromm (1980)
and Wood and McDonald (1982) have provided a thorough critique of the
literature on physiological and toxicological responses of freshwater
fish to acid stress.

     The best documented physiological response, and probably the
response most widely accepted as the physiological basis for the
toxicity of low pH, involves interference of elevated hydrogen ion
levels with osmoregulatory mechanisms and impaired body salt regulation.
Freshwater fish maintain a higher salt concentration in their tissues
than is in the water that surrounds them, and must actively take up ions
from the surrounding water through  the gill epithelium.  Sodium in the
                                  5-124

-------
water is exchanged for hydrogen ions or ammonium ions,  and chloride for
bicarbonate (Maetz 1973, Evans 1975).   Increased hydrogen  ion  activity
in the surrounding medium may impede the active uptake  of  sodium.   Brown
trout surviving in the Tovdal River, Norway,  collected  immediately
following a fish kill (apparently resulting from an acid episode),  had
significantly reduced plasma chloride and sodium levels (Leivestad  and
Muniz 1976, Section 5.6.2.4).  The plasma content of potassium,  calcium
and magnesium was not affected.  Therefore, impairment  of  the  active
transport mechanism for sodium and/or chloride ions through the  gill
epithelium was suggested as the primary cause of fish death.   Severe
Internal Ionic Imbalance would affect fundamental  physiological
processes such as nervous conductions and enzymatic reactions.

     Laboratory experiments have also found decreased plasma  (or whole
body) sodium and/or chloride levels as a result of exposure of organisms
to low pH levels (Packer and Dunson 1970, 1972;  Leivestad  and  Muniz
1976; McWilllams and Potts 1978; Jozuka and Adachi  1979; Leivestad  et
al. 1980; McWilliams et al. 1980;  McDonald et al.  1980;  McDonald and
Wood 1982;  Ultsch et al. 1981).  The exact mechanisms behind these
effects are not, however, fully understood.   A major influence on
branchial ion fluxes is the transepithelial  potential (TEP)  across  the
gills.  The TEP of brown trout has been shown to be strongly dependent
on the pH of the external medium,  being negative in neutral  solutions
but positive in acid solutions (McWilliams and Potts 1978).  At  near
neutral pH, the influx and efflux  of sodium were similar,  indicating
that trout were in sodium balance.  As the pH in the external medium
declined, sodium influx decreased and sodium  efflux increased  until, at
pH 4.0, the rate of loss of sodium amounted to about 1  percent of the
total body  sodium per hr.

     These processes are influenced by the content of dissolved  salts in
the water,  particularly calcium and sodium (McDonald et al.  1980; Brown
1981, 1982).  Calcium is essential in the maintenance of ionic balance
in freshwater fish, probably as a  result of its  influence  on the
permeability of gills to certain ions (McWilliams  and Potts 1978,
McWilliams  1980a).  Increased calcium concentrations (from near  zero to
about 40 mg £-1) decreased membrane permeability and thus  decreased
the rate of passive sodium efflux  from fish.   At the same  time,  calcium
appeared to have no significant effect on sodium influx  (McWilliams
1980a, 1982).   The result was a decrease in the  overall  rate of  sodium
loss from fish exposed to low pH in waters with  higher  calcium content.
Gill  permeability also varied between  species and  populations of fish
(McWilliams 1982), and sodium loss rates declined with  acclimation  of
fish to acid waters (McWilliams 1980b).  These results  help explain the
observed correlation between low calcium levels  and loss of fish
populations in Norwegian Lakes (Section 5.6.2.1.3;  Wright  and  Snekvik
1978) and imply that small changes in calcium availability in  natural
waters (e.g.,  during spring snowmelt,  see Chapter  E-4,  Section 4.3.2.6)
and previous exposure of fish to high  acidity are crucial  factors in
determining the response of fish exposed to sudden  acid  episodes.
                                 5-125

-------
     A decrease in blood pH levels (by  0.2  to 0.5  pH  units) is often
associated with the drop in plasma sodium levels in fish exposed to low
pH waters (Lloyd and Jordan 1964,  Packer and Dunson 1970, Packer 1979,
Jozuka and Adachi 1979, Neville 1979a,  McDonald et al.  1980, McDonald
and Wood 1982, Ultsch et al. 1981) and  is possibly a  result of flux of
hydrogen ions across gill  membranes into the blood.   McDonald et al.
(1980) noted that in moderately high alkalinity waters  (calcium 30 to 50
mg £•!), fish exposed to a pH of 4.3 developed a major  blood
acidosis (drop in blood pH) but exhibited only a minor  depression in
plasma ion levels.  In acidified,  low alkalinity water  (calcium 6 mg
r1), only a minor acidosis occurred, but plasma ion  levels fell and
mortality was substantially greater. Possibly the nature of the
mechanism of acid toxicity varies  with  the  nature  of  the ionic
environment.

     A drop in blood pH level  would affect  a large number of
pH-sensitive metabolic reactions.   The  oxygen carrying  capacity of fish
blood drops sharply below a blood  pH level  of 7.0  (Green and Root 1933,
Prossner and Brown 1961).   Decreased oxygen consumption by fish exposed
to acid waters has been found by Packer and Dunson (1970, 1972), Packer
(1979), and Ultsch (1978).  Carrick (1981), however,  observed no
significant differences in oxygen  uptake by brown  trout fry at pH 7.0 vs
pH 4.0.  Neville (1979b) concluded that an  observed increase in serum
erythrocyte concentration  offset the reduced capacity of the hemoglobin
to transport oxygen brought about by acidosis.  The increase in
hemoglobin level, maintenance of arterial oxygen tension (pOg). and
constancy of blood lactate levels  in rainbow trout exposed to pH 4.0
suggested that there was no oxygen stress despite  the acidosis.

     At critically low pH  levels (£ 3.5), where death occurs within
hours rather than days, a  failure of oxygen delivery  to the tissues may
be of primary importance.   A marked reduction in blood  oxygen capacity
due to massive acidosis, combined  with  impaired branchial oxygen
diffusion as a result of accumulation of mucous on the  gills and a
sloughing of gill epithelial tissue (e.g.,  Plonka  and Neff 1969, Daye
and Garside 1976, Ultsch and Gros  1979), may result in  eventual cellular
anoxia.  However, such low pH levels are rarely encountered by fish in
natural situations.  At more moderate pH levels, mucous accumulation on
the gills has not been observed and blood gas levels  remain normal,
indicating acid-base and/or ion regulatory  failure are  more probable
mechanisms of toxicity (McDonald et al. 1980, Frornm 1980).

     Finally, Nelson (1982) reported that ossification, amount of
calcium deposited in bone, in rainbow trout fry varied  significantly (p
< 0.005) as a function of pH of the medium  (pH 4.3, 4.8, and 7.3).
After calcium stores from  the yolk sac  are  exhausted, fry must
accumulate calcium from the surrounding water and  from  food intake.  A
decrease in skeletal ossification  at low pH level  could be partially
responsible for increased incidence of  skeletal deformities observed in
some laboratory bioassays  at low pH (e.g.,  Beamish 1972, Mount 1973,
Trojnar 1977b) and in white suckers from acidic George  Lake, LaCloche
                                  5-126

-------
Mountain region, Ontario (Beamish et al.  1975).   Nelson  (1982),  however,
noted no increase in deformities despite  decreased  ossification.

5.6.4.2  Effects of Aluminum and Other Metals  in  Acidic  Waters--
Increases in certain metal  concentrations can  be  associated  with
deecressing pH levels in acidified surface waters (Chapter E-4,  Section
4.6).  Declines in fish populations as a  result of  acidification may,
therefore, be a function of both low pH levels and  elevated
concentrations of some metals.  Critical  values for survival  of  fish
populations developed only  on the basis of pH  level  may  therefore be
misleading.

     Muniz and Leivestad (1980a) noted that naturally  acidified  water  is
generally more toxic to fish than are dilute sulfuric  acid solutions of
the same pH.  Brown trout exposed to soft waters  acidified by additions
of sulfuric acid (a pure hydrogen ion stress)  demonstrated physiological
stress (impaired regulation of body salts) only at  pH  levels below 4.6
(Leivestad et al. 1980).  When tests were performed in water from
acidified brooks and rivers in southern Norway, water  with a pH  of 4.6
resulted in significant physiological stress,  rapid salt depletion, and
mortality after 11 days (Leivestad et al. 1976; Section  5.6.3.3).  For
Atlantic salmon, Daye and Garside (1977)  found lower limits  for  survival
of fry to be around pH 4.3  and pH 3.9 for eggs exposed from
fertilization through hatching (Section 5.6.4.1.2). Bua and Snekvik
(1972), on the other hand,  used water from the acidic  Mandal  River,
Norway and found lower limits for survival to  be  pH 5.0  to 5.5.
Schofield observed stress and heavy mortality  among adult, yearling,
and sac fry of brook trout held in an Adirondack  hatchery receiving lake
water from Little Moose Lake at pH 5.9 during  spring snowmelt in 1977
(Schofield and Trojnar 1980; Section 5.6.2.4).  In  contrast, in
laboratory experiments (Sections 5.6.4.1.1 and 5.6.4.1.2) critical pH
levels for brook trout were generally between  pH  3.5 and 4.5. These and
other comparisons strongly imply that acidified  lake and river water
must contain toxic agents in addition to  hydrogen ions (Muniz and
Leivestad 1980a).

     Metals consistently exhibiting increased concentrations in  acidic
surface waters, apparently as a result of increased solubility with
decreasing pH level, are aluminum, manganese,  and zinc (Chapter  E-4,
Section 4.3.4.1).  In some regions, concentrations  of  cadmium, copper,
lead, nickel, and other metals are also elevated  in acidic  lakes.  High
concentrations of these metals, however,  probably result primarily from
increased atmospheric loading and deposition and  occur principally in
surface waters in close proximity to pollutant sources (e.g., Sudbury,
Ontario).  As such, they are not specifically a  result of acidic
deposition but may still interact additively or  synergistically  with
toxic effects of low pH, aluminum, manganese,  or zinc.  Unfortunately,
with the exception of aluminum, data are not sufficient  for  a thorough
evaluation of possible adverse effects of metals  on fish in  acidic
waters.  Spry et al. (1981) and Baker (1982) have reviewed  the available
literature.
                                  5-127

-------
     Total zinc concentrations  measured  in  acidic  surface waters in the
Adirondack region,  in  southern  Norway and in  southwestern Sweden ranged
up to 0.056 mg jr*  (Schofield 1976b, Henriksen  and Wright 1978,
Dickson 1980).  Although  laboratory bioassays examining effects of zinc
on fish are numerous ('Taylor et al. 1982),  none of these studies
considered low alkalinity water with pH  levels  below 6.0, and results
should not be automatically  extrapolated to conditions in acidified
surface waters.  For the  most part, however,  lethal concentrations of
zinc in bioassays are 10  times  zinc concentrations found in acidic
waters (Spry et al. 1981, Taylor et al.  1982).  Sinley et al. (1974)
estimated that the maximum acceptable toxicant  concentration (MATC) for
rainbow trout exposed to  zinc in low alkalinity circumneutral  water was
between 0.14 and 0.26 mg  £-!.   Benoit and Hoi combe (1978) determined
that the threshold level  for significant adverse effects on the most
sensitive life history stage of fathead  minnows was between 0.078 and
0.145 mg «,-!.  Taylor et  al. (1982) concluded from a review of the
available literature that concentrations of zinc that could be tolerated
by aquatic organisms lie  between 0.026 and  0.076 mg «,-!.

     Manganese has been considered a relatively nontoxic element; thus
toxicological data are very  limited.  Total manganese concentrations
measured in acidic surface waters ranged up to  0.13 mg jr1 in the
Adirondacks (Schofield 1976b) and up to  0.35  mg £-1 in southwestern
Sweden (Dickson 1975). Lewis  (1976) determined that manganese
concentrations up to 0.77 mg r^ had no  effect  on  survival of
rainbow trout in soft waters with pH levels of  6.9 to 7.6.

     Relationships between pH and levels of cadmium, copper, lead, and
nickel vary markedly between regions.  Excluding lakes within 50 km of
Sudbury, concentrations of cadmium, copper, lead,  and nickel measured in
acidic Ontario surface waters  ranged up  to  about 0.6, 9, 6, and 48 yg
Ni £-1, respectively (Spry et al. 1981).  In  southwestern Sweden,
concentrations of cadmium in acidic waters  were measured up to 0.3 yg
A-l; lead up to 5 yg t-1  (Dickson 1980).  Spry  et  al. (1981)
reviewed available bioassay data and noted  no significant adverse
effects on survival and reproduction at  concentrations up to 0.7 to 11
yg Cd £-1, 9.5 to 77 yg Cu A-l, 13 to 253 yg  Pb £-1, and 380 yg Ni £-1.

     In general, all of these  reported  toxic  concentrations and/or
maximum acceptable concentrations for zinc, manganese, cadmium, copper,
lead, and nickel are above the  highest  levels of these metals measured
in acidic surface waters  of Scandinavia  and eastern North America
(unless a local source of metal emissions  exists). However, the lack of
sufficient bioassay data  collected in low  alkalinity, acidic waters
makes this statement tentative.  In  addition, sublethal and additive or
synergistic effects with  other  toxic components in acidified surface
waters cannot be ruled out.

     Aluminum, on the other hand, has been found to be toxic to  fish at
concentrations as low as  0.1 to 0.2 mg  jr1  (Schofield and Trojnar
1980, Muniz and Leivestad 19805, Baker  and Schofield 1982), a  level
within the range of concentrations measured in  acidic surface  waters.


                                  5-128

-------
Total aluminum levels measured  ranged  up  to  1.4 mg a~l in the
Adirondack region,  New York  (Schofield 1976b), 0.76 mg i'1 in
southwestern Sweden (Dickson 1975,  Wenblad and Johansson 1980), 0.6 ing
JT1 in southern Norway (Wright  et al.  1980),  and 0.8 mg jr1 in
the Pine Barrens of New Jersey  (Budd et al.  1982).  In addition,
analysis of survival  of brook trout stocked  into 53 Adirondack lakes as
a function of 12 water quality  parameters indicated aluminum to be a
primary chemical factor controlling trout survival (Schofield and
Trojnar 1980; Section 5.6.2.1.2.1).

     Baker (1981, 1982), Baker  and  Schofield (1982), and Driscoll et al.
(1980) examined the effect of aluminum complexation on aluminum toxicity
to fish in laboratory experiments.   Complexation of aluminum with
organic chelates appeared to eliminate aluminum toxicity to fry, and
survival of brook trout and  white sucker  fry in acidic Adirondack waters
correlated most accurately with inorganic aluminum concentrations and
pH.  The toxicity of a given inorganic aluminum concentration varied at
different pH levels and with different life  history stages.  At low pH
levels (4.2 to 4.8),  the presence of aluminum was beneficial to egg
survival.  In contrast, in experiments with  sac fry and fry, aluminum
concentrations of 0.1 mg &~1 (for white suckers) or 0.2 mg &"1
(for brook trout) and greater resulted in measurable reductions in
survival and growth at all pH levels (Schofield and Trojnar 1980, Baker
and Schofield 1982, Muniz and Leivestad 1980a).

     The toxic action of aluminum seems to be a combined effect of
impaired ion exchange and respiratory  distress caused by mucous clogging
of the gills (Muniz and Leivestad 1980a).  Muniz and Leivestad (1980b)
observed rapid loss of sodium and chloride from the blood of brown trout
exposed to aluminum concentrations  as  low as 0.19 mg £~1 at pH 5.0.
Schofield and Trojnar (1980) noted  moderate  to severe gill damage at
aluminum levels of 0.5 and 1.0  mg £-1  at  pH  4.4 and higher.
Aluminum was particularly toxic in  solutions over-saturated with
aluminum at pH levels 5.2 to 5.4 (Baker and  Schofield 1982).

     The pH level in  acidic  waters  therefore affects fish survival both
as a direct toxicant and by  controlling the  concentration of inorganic
aluminum.

5.6.5  Summary

5.6.5.1  Extent of Impact

     Loss of fish populations associated  with acidification of surface
waters has been documented for  five areas—the Adirondack region of New
York State, the LaCloche Mountain region  of  Ontario, Nova Scotia,
southern Norway, and southern Sweden.   The following summarizes major
points from Section 5.6.2.1:

    o   The best evidence that  the  loss of fish has occurred in response
        to acidification is  derived from  observations of lakes in the
        LaCloche Mountain region, Ontario (Section 5.6.2.1.2.1).
                                  5-129

-------
Twenty-four percent of 68 lakes surveyed had no fish  present.
Fifty-six percent of the 68 lakes are known or suspected  to  have
had reductions in fish species composition (Harvey  1975).  Based
on historic fisheries information, 54 fish populations  are known
to have been lost, including lake trout populations from  17
lakes, smallmouth bass from 12 lakes, largemouth bass from four
lakes, walleye from four lakes, and yellow perch and  rock bass
from two lakes each (Harvey and Lee 1982).  The principal source
of atmospheric acidic inputs to the LaCloche area is  sulfur
dioxide emitted from the Sudbury smelters located about 65 km to
the northeast.  Large acidic inputs have resulted in  relatively
rapid acidification of many of the region's lakes.  For some
lakes the development of increased lake acidity and the
simultaneous decline of fish populations have been  followed and
recorded by a single group of researchers (Beamish  and  Harvey
1972, Beamish et al. 1975, Harvey and Lee 1982)  from  the
mid-19601 s to the present.

In Norway (Section 5.6.2.1.3.1), sharp drops in catch of
Atlantic salmon in southern rivers began in the early 1900's and
are associated with current low pH levels and a recorded
doubling of the hydrogen ion concentration in one of  these
rivers from 1966 to 1976 (Jensen and Snekvik 1972,  Leivestad et
al. 1976).  For almost 3000 lakes in Stfrlandet (southernmost
Norway) data on the status of brown trout has been  recorded
since about 1940 (Sevaldrud et al. 1980).   Today, more  than 50
percent of the original  populations have been lost, and
approximately 60 correlated with acidity,  acidification, and/or
inputs of acidic deposition (Wright and Snekvik 1978).

Extensive surveys of fish population status and acidity of
surface waters in Sweden have not been completed (Section
5.6.2.1.3.2).  However,  for 100 lakes in southern Sweden with
historic records on fish populations, loss of fish  was
correlated with present-day low pH levels in lakes  (Aimer et al.
1978).  Forty-three percent of the minnow populations,  32
percent of the roach,  19 percent of the char,  and 14  percent of
the brown trout populations had disappeared.

In Nova Scotia, records  of angling catch of Atlantic  salmon in
rivers date back, in some cases, to the early  1900's  (Section
5.6.2.1.2.3).  Of 10 rivers with current pH <  5.0 and historic
catch records, 9 have had significant declines in angling
success over the time  period 1936 to 1980.   For 12  rivers with
pH > 5.0,  only one experienced a significant decrease in salmon
catch.  Decrease in salmon catch over time is  correlated with
present-day pH values  5.0 and below.  In addition,  6  former
salmon rivers with current pH < 4.7 have no long-term catch
records, but surveys in  1980 indicated they no longer support
salmon runs.  Acidification of rivers in the area between 1954
and 1974 has been reported (Chapter E-4,  Section 4.4.3.1.2.2).
The high organic content in many of the low pH waters
                          5-130

-------
        (especially pH < 4.7)  suggests,  however, that these rivers are
        naturally somewhat acidic,  and  perhaps always had fairly low pH
        values and low fish production  (Farmer et  al. 1981).  The
        estimated lost or threatened  Atlantic salmon production
        potential represents 30 percent of the Novia Scotia resources
        but only 2 percent of the total  Canadian potential (Watt 1981).

    °   Finally, fish populations in  Adirondack lakes and streams have
        also declined over the last 40  to  50 years (Section
        5.6.2.1.1.1).  The New York State  Department of Environmental
        Conservation reports that about 180 lakes  (2900 ha) out of a
        total  of 2877 lakes (114,000  ha) in the Adirondacks have lost
        their fish populations (especially brook trout) (Pfeiffer and
        Festa 1980).  The absence of  fish  in Adirondack lakes and
        streams is clearly correlated with low pH  levels (Schofield
        1976b), although  several factors  may confound this
        relationship, e.g., lake size dystrophic conditions.  Records of
        pH and other information have not, however, been published to
        substantiate that loss of fish  in  these 180 lakes resulted from
        acidification.  For very few  individual lakes are historical
        data available that suggest both lake acidification and
        simultaneous loss of fish.  Acidification  probably contributed
        to the disappearance of fish  for at least  some surface waters,
        but exactly how many lakes  and  streams (perhaps substantially
        less than or more than 180) have been impacted cannot be
        satisfactorily evaluated at this time.

    o   In other regions of the world with low alkalinity waters and
        receiving acidic deposition (e.g., Muskoka-Haliburton area of
        Ontario and Maine) (Sections  5.6.2.1.1.1 and 5.6.2.1.2.2)
        acidification of surface waters does not appear to have
        progressed to levels clearly  detrimental to fish (Schofield
        1982).  No damage to fish populations has  been reported.

5.6.5.2  Mechanism of Effect—Three major  mechanisms for the
disappearance of fish populations with  acidification have been proposed:
(1) decreased food availability and/or  quality; (2) fish kills during
episodic acidification; and (3) recruitment failure.  Each probably
plays some role, although recruitment failure has  been hypothesized as
the most common cause of population loss (Schofield 1976a, Harvey 1980,
NRCC 1981, Overrein et al. 1980, Haines 1981b).  The following
summarizes major points from Sections 5.6.2.2 through 5.6.2.4, and
5.6.3.1:

    0   The influence of food chain effects on decreases in fish
        populations in acidified waters has received little attention to
        date,  but available information suggests it  plays a relatively
        minor role (Beamish 1974b,  Hendrey and Wright 1976, Muniz and
        Leivestad 1980a, Rosseland  et al.  1980).   With acidification, or
        in comparisons between acidic and  circumneutral lakes, fish
                                  5-131

-------
growth is often unaffected or increased with  increasing  acidity
(Section 5.6.2.3).   Some important prey organisms are sensitive
to acidic conditions and disappear with acidification yet  fish
seem capable of shifting to other suitable  prey.  During the
experimental acidification of Lake 223 (Section 5.6.3.1),  lake
trout production remained unchanged in spite  of the loss of
fathead minnows, a  major prey item prior to acidification  (Mills
1982).  Few studies, however, have examined the potential  effect
of reduced food quantity and/or quality on  survival of early
life history stages of fish or on fish production at pH  levels
above those that result in recruitment failures and reduced
population size.

Fish kills have been observed during episodic acidification of
surface waters (Section 5.6.2.4), and in certain instances may
play an important role in the disappearance of fish from
acidified surface waters.  For example, in  the Tovdal River,
Norway, in 1975 thousands of dead adult trout were observed in
association with the first major snow melt  in spring (Leivestad
et al. 1976).  Dead and dying fish are, however, seldom  reported
in acid-stressed waters relative to the large number of  lakes,
streams, and rivers with fish populations apparently impacted by
acidification.  In  contrast, a substantial  portion of fish
populations examined in acidified lakes lack  young fish  (Section
5.6.2.2) and apparently have experienced recruitment failure.

Recruitment failure may result either from  acid-induced
mortality of eggs and/or larvae or because  of a reduction  in
numbers of eggs spawned.  The number of eggs  spawned could be
reduced as a result of disruption of reproductive physiology and
ovarian maturation  or inhibition of spawning  behavior.   Evidence
exists that supports each one of these proposed mechanisms
(Sections 5.6.2.2,  5.6.4.1.2,,5.6.4.1.4, and  5.6.4.1.51.   For
example, Johnson and Webster (1977) demonstrated experimentally
that brook trout avoid spawning in waters with pH below  5.0.
Beamish and Harvey  (1972) observed that recruitment failure in
several acidic lakes in the LaCloche Mountain region, Ontario
was associated with a failure of the female fish to spawn.
Lockhart and Lutz (1977) hypothesized that  a  disruption  in
normal calcium metabolism, induced by low pH, had adversely
affected the reproductive physiology of female fish in these
lakes.  Adverse effects of low pH levels and  elevated aluminum
concentrations on survival of fish eggs and larvae have  been
demonstrated in numerous laboratory and field experiments
(Sections 5.6.3.3 and 5.6.4.1.2).  In Norway, total mortality of
naturally spawned trout eggs was observed in  an acidic stream a
few weeks after spawning (Leivestad et al.  1976).

It is likely that each one of these factors plays some role in
recruitment failure but the importance of each factor probably
varies substantially among aquatic systems, depending on the
particular circumstances.
                          5-132

-------
     0    More research is  necessary  to  define clearly the specific
         mechanism for population  decline  in a given water.  However,
         many studies  in the United  States and Scandinavia (Schofield
         1976a,  Muniz  and  Leivestad  1980a) emphasize increased mortality
         of eggs and larvae in  acidic waters as the primary cause of
         recruitment failures,  and recruitment failure as a common cause
         for the loss  of fish populations with acidification of surface
         waters.

 5.6.5.3   Relationship Between  Water Acidity and Fish Population
 Response—To assess the impact of acidification on fish resources
 quantitatively,  the functional relationship between acidification and
 fish population  response must  be  understood.  Unfortunately, loss of
 fish populations  from acidified surface waters is not a simple process
 and cannot be accurately summarized as "X" pH (or aluminum
 concentration) yields  "Y" response.  The mechanism by which fish are
 lost (Section 5.6.5.2) seems to vary between aquatic systems and
 probably  within a given system from year-to-year.

     The1  water chemistry within a given aquatic system is likewise
 extremely  variable  both spatially and temporally, and these variations
 are very  important  to  the survival or decline of fish populations.
 Lakes with  seemingly  identical water quality may show marked differences
 in fish response, perhaps reflecting, in part,  the existence or lack of
 water quality "refuge" areas for  fish survival  (Muniz and Leivestad
 1980a).  A  circumneutral  tributary or small  segment of a lake may
 provide an  area for successful  fish reproduction for a number of years
 following acidification of the main body of a lake.

     Fish species differ not only in their ability to tolerate acidic
 conditions  but also in their ability to exploit these chemical
 variations  in their environment (e.g.,  spawning time and location).
 Within a given fish species,  sensitivity to acidity  varies  with life
 history stage, age, condition,  previous exposures to acidity,  associated
 water quality conditions (e.g., aluminum and calcium concentrations,
 temperature), and other parameters.   In addition, for reasons  discussed
 at the beginning of Section 5.6.4, results from laboratory  experiments
 cannot be translated automatically into an expected  response in the
 field.   Serious gaps exist in the understanding of how  to use laboratory
 results in  a quantitative prediction of fish response in the field  and
 in the analysis of the complexity of the natural  environment and the
 significance of this complexity in determining  the impact of
 acidification.  It is therefore not surprising  that  development of  an
 accurate functional relationship  between acidification  and  fish response
 is impossible at this time.

     First  steps, however, in developing such a relationship are to:
examine in a semi-quantitative  manner all  of the  available  information
connecting acidity and fish (summarized in Table  5-14),  produce an
 initial  approximation  of  the  dose-response relationships (Figure 5-16),
and then assess patterns  and  reasons for deviations  from this  initial
                                  5-133

-------
TABLE 5-14.  SUMMARY OF FIELD OBSERVATIONS,  FIELD EXPERIMENTS,  AND LABORATORY EXPERIMENTS
                        RELATING WATER pH TO FISH RESPONSE

























tn
*^


+^


















FIELD OBSERVATIONS
Recruitment failure



Population loss



Fish kill

FIELD EXPERIMENTS
Recruitment Failure
Population extinction
Adult mortality


Embryo mortality


LABORATORY EXPERIMENTS
Adult mortality-
acute (2 day)
Adult mortality-
chronic (> 20 day)
Embryo mortality



Fry mortality




Reduced production
of viable eggs
Reduced growth

Avoidance

FIELD OBSERVATIONS
Recruitment failure

Population lots

Brook Lake Arctic
Trout Trout Char

5.0-5.5* S.2-5.50 5.2*
6.9«
5.7*

5.0« 5.0-5.5'
4.5-4.8J 4.4*


5.9"


5.1-5.34

4.8-5. jn
4.7-5.1*
4.4-4. 6l
4.S-4.6S



3.8«
3.500
4i4ff < 4.899
4.511
•4.1JJ
6.1"
4. add
<4.6PP
•4.4-4.9"
•4.5-4.8JJ
,;2dd
<4.4kk
4.0-5.0PP

5.1"
6.5" 4.899

4.5»»
Lake Herring Lake Smallmouth
Uhtteflsh Bass

4.5-4. 7b 5.0* 5.5-6.QO
5.0C
4.4C 4.4« 4.4<=

Brown Rainbow Atlantic Unite European Walleye Fathead
Trout Trout Salmon Sucker Perch Minnow

< 5 .IK 4.7-5.011 5.2* 4.4-4.9C 5. 5-6. Ob
4.7-5. if 4.7-5.2° 5.0-5.59 5.4«
5.0e <4.7h 6.5'

4.5-4.8J 5.5-6.0J S.I* 5.1«
4.5-5.0' 4.3* 5.2<
5. 1" 5.5<
4.9-5.1*
4.6J 3.9-4.2°
5.01 5.0"

5.1-S.jt 5.8-6.01
5.3-5.14
4.0-5.01"


4.5° 4.5-5.0' 4.7-5.7" S.4»
4.8* ' 5.0-5.5°
5.1" 4.9*

3.8y 4.0* 3.9** < 4.6"
3.8-4. 8«
4. 0-4 .Zee
< 4.899 < 5.099 5.09 S.91*
5.6"
4.1" 4.1k* +> s.&JJ
4.0-4.5"" S.snn s.joo
S.S"11*
4i544
4.4** 3.7-4.0T «5.4-S.63J < S.*"!"
5.0** 5.0-5.4«°



6.6hh

4.899 4.899
< 5.0" 4.3-4.8U"
S.ww
Largenouth Rock Pumpklnseed Blueglll Yellow Common Bluntnose
Bass Bass Sunflsh Perch Shiner Minnow

5. 1C 4.7-5.20 5.0<1 4.5-4.7° 5.5-6.01
5.011 S.O* 4.4' 5.7C
4.4C 4.8-5.01" 4.3<: 4.4« 4.4-5.0* 5.7°
6.0<1 4.3' 4.3C
Roach Northern SI Imy Brown
P1ke Sculpln Bullhead

5.5« 5.0e 4.7-5.20
5.1' 5.01 4.9«
4.4-4.9C 5.2'
4.7h

4.7' 4.3« 4J-5.0*
4.7e




5.3-5.84




5.7°





5.6"





4.2-5,2"






ww .
•

Lake Creek Trout Gadldae
Chub Chun Perch Burbot

4.5-4.7" 5.2-5.50 5.5-6.0>>

4.5-5.0* 5.0* 5.»d


-------
REFERENCES
 a  Schofield 1976c
 b  Beamish 1976
 c  Aimer et al.  1978
 d  Watt et al. 1983
 e  Harvey 1979
 f  Hultberg 1977
 g  Runn et al. 1977
 h  Grahn et al.  1974
 i  Beamish et al. 1975
 j  Grande et al. 1978
 k  Leivestad et  al.  1976
 1  Harriman and  Morrison 1982
 m  Overrein et al. 1980
 n  Schofield and Trojnar 1980
 o  Jensen and Snevik 1972
 p  Farmer et al. 1981
 q  Mills 1982
 r  Harvey et al. 1982
 s  Schofield 1965
 t  Dunson and Martin 1973
 u  Milbrink and  Johansson 1975
 v  Hul sman and Powles 1981 as
    reported in MO I 1982
 w  Mimiz and Leivestad 1980
 x  Johnson 1975
 y  Brown 1981
 z  Kwain 1975
aa  Beamish 1972
bb  Robinson et al.  1976
cc  McDonald et al.  1980
dd  Swarts et al.  1978
ee  Lloyd and Jordan 1964
ff  Swartz et al.  1978
gg  Edwards and Hjeldnes  1977
hh  Mount 1973
ii  Menendez 1976  (e from Table  5-22)
jj  Baker and Schofield 1982
kk  Johansson et al. 1977
11  Johansson and  Milbank 1976
mm  Carrick 1979
nn  Peterson et al.  1980a
oo  Trojnar 1977b
pp  Trojnar 1977b
qq  Peterson et al.  1980b (conf)
rr  Daye and Garside 1976
ss  Johansson and  Kihlstrom 1975
tt  Jacobsen 1977
uu  Nelson 1982
vv  Johnson and Webster 1977
ww  Hoglund 1961
xx  Ryan and Harvey  1977
yy  Ryan and Harvey  1980
*Refers to laboratory experiments taking into account both low pH  and  inorganic
 aluminum (at the expected concentration for that pH based on Driscoll  1980).
                                  5-135

-------
   SPECIES
   YELLOW PERCH
   NORTHERN PIKE
   ROCK BASS
   PUMPKINSEED SUNFISH
   LAKE HERRING
   LAKE WHITEFISH
   BLUEGILL
   LAKE CHUB
   EUROPEAN PERCH
   WHITE SUCKER
   LARGEMOUTH BASS
   BROOK TROUT
   BROWN TROUT
   SMALLMOUTH BASS
   BROWN BULLHEAD
   ATLANTIC SALMON
   ROACH
   LAKE TROUT
   CREEK CHUB
   RAINBOW TROUT
   ARCTIC CHAR
   SLIMY SCULPIN
   TROUT-PERCH
   BURBOT
   WALLEYE
   FATHEAD MINNOW
   COMMON SHINER
   BLUNTNOSE MINNOW
                            4.5        5.0        5.5
              LEGEND                             PH
                 pH RANGE OF SUCCESSFUL REPRODUCTION
                 pH LEVELS AT WHICH  POPULATIONS OCCUR
                 VARIATIONS IN OBSERVED LOWER pH LIMITS
                                            6.0
6.5
Figure  5-16.
Initial  estimates of relationship between acidity and
fish response,  based on  references  in Table  5-14.
                                          5-136

-------
approximation.   In large part,  the analysis  of deviations and variations
must be done on a lake-by-lake,  population-by-population basis, and is
the subject for further research.   Several points are, however, obvious.
Acidification adversely affects fish  populations.   Sensitivity of fish
to acidity is species-dependent,  and  determined by  aluminum and calcium
concentrations, in addition to  pH values.  Loss of  fish  populations need
not be associated with large declines in  annual  average  pH, but could
result from indirect effects on aluminum  chemistry  or episodic
acidification.

5.7  OTHER RELATED BIOTA (R. Singer and K. L.  Fischer)

5.7.1  Amphibians

     Direct effects of acidity  on vertebrates  have  been  demonstrated
only on fish (Section 5.6)  and  amphibians.   Amphibians are particularly
sensitive because many frogs, toads,  and  salamanders breed in vernal
pools filled by acidic snowmelt and spring rains.   The salamanders
Ambystoma maculatum and A_.  jeffersom'anum breed in  shallow hilltop  ponds
that have pH values 1.5 pH  units less than nearby permanent ponds in New
York State (Rough and Wilson 1977).  Spotted salamander  (A_. maculatum)
egg mortality increased to  > 60 percent in water less than pH 6.0,  a
substantial rise from the normal  mortality of < 1 percent at  pH 7.0.  In
contrast, the Jefferson salamander, A.  jef fersonianum, was most
successful at pH 5.0 to 6.0 (Rough 1"976TT  The preference for neutral
water by adult spotted salamanders may be a  result  of the absence of
their preferred prey, the stickleback (Eucalia), from acidic water
(Bishop 1941).   When a stretch  of stream  was artificially acidified to
pH 4.0, "salamanders" were  reported (Hall and Likens 1980a) to leave the
water.  Elsewhere in its range  in central Ontario,  the number of egg
masses of the spotted salamander positively  correlated with pH (Clark
and Euler 1980).  Adults are not as sensitive  to pH stress, but given a
choice, adult spotted salamanders (A. maculatum) preferred neutral
substrates (Mushinsky and Brodie 1975).

     The contribution of salamanders  to the  energy  flow  of a  forest
aquatic ecosystem is considerable.  In one  study (Burton and Likens
1975a), 20 percent of the energy available to  birds and  mammals passed
through salamanders, and these  amphibians represented twice as much
standing crop of biomass as did birds and an amount equal to that of
small mammals (Burton and Likens 1975b).  Most (94  percent) of the  sala-
manders were terrestrial, but all  salamanders  are aquatic as larvae. Not
only do amphibians provide  energy for birds  and mammals, but they repre-
sent the top predators in many  temporary  ponds (Orser and Shure 1972).

     The species specific tolerance of amphibians to low pH was
confirmed by a survey of newts 1n England (Cooke and Frazer 1976).
Smooth newts (Triturus vulgaris)  were rarely encountered 1n water with
pH < 6.0, but the palmate newt (T_. helveticus) was  routinely captured in
bogs at pH 4.0 to 3.8.  The distributions of these  species were
correlated most strongly with potassium and  calcium concentrations, both
of which co-varied with pH.  The variable sensitivity of newts to


                                  5-137

-------
acid-stress is demonstrated by  the american  red-spotted newt,
Notophthalamus  viridescens,  which one  of us (RS) has observed at 6 m in
acidic (pH 4.9) Woods Lake.  This same  species  has been reported at 13 m
in neutral (pH 7.4) Lake George,  also in  the Adirondacks  (George et al.
1977).

     Many anurans are sensitive to acidity,  too.  Calling densities (an
estimate of population size)  of spring  peepers  (Hyla crucifer) were
positively correlated with the  pH of water in whTcTTthey  occurred (Clark
and Euler 1980).  Bullfrogs (Rana catesbeiana)  (Clark and Euler 1980,
Cecil and Just 1979, Saber and  Dunson 1978), wood frogs (R.  sylvatica)
(Clark and Euler 1980) the common frog  (R_. temporaria) (rfagstrom 1977),
and the leopard frog (R_. pi pi ens) (Noble  1979)  have all been reported to
be sensitive to acidity below pH  5.0.   Evidence  from counts of dead and
moldy egg masses in the Netherlands (Strijbosch  1979) supports the
relationship between acidity  and  mortality of frogs.  The most serious
effects occur in the immature stages (Gosner and Black 1957).  Cricket
frog (Acris gryllus) and spring peeper  (Hyla crucifer) embryos exposed
to pH 4.0 for a few hours suffered 85 percent mortality.  Noble (1979)
reported embryonic mortality  in the leopard  frog (R. pi pi ens) at pH <
4.7, and Schlichter (1981) observed sub-lethal  reductions in sperm
mobility in this species below  pH 6.5 and some  embryonic mortality at pH
< 6.3.  In spite of the sensitivity of  R. pi pi ens to acidity in the
laboratory, one of us (RS) has  seen aduTt leopard frogs in an acidic (pH
4.8) Adirondack lake.  The larvae may have been  breeding  in  ponds that
provided refuge near the lake.   Reports of only  adult amphibians are of
questionable value because of the much  greater  sensitivity of the larval
forms.

     Toads, although terrestrial  as adults,  are  also sensitive to
acidity as larvae and embryos.   The common toad  (Bufo bufo) was not
reported below pH 4.2 in Sweden (Hastrom  1977),  aTuTthe natterjack toad
(Bufo calamita) was not found below pH  5.0 (Beebee and Griffin 1977) in
EngTaruT

     The mechanism by which acidity effects  amphibians is not known.
Huckabee et al. (1975) suggest  that the aluminum, manganese, and zinc
mobilized by low pH (Chapter  E-4, Section 4.6) may be toxic agents for
the shovel-nosed salamander (Leurognathus marmoratus) larvae in the
Great Smokey Mountains National  Park.   Another mechani sm may be the
inability to control ion fluxes across  membranes against  strong H+
gradients.  This has been indicated in  fish  (Section 5.6), invertebrates
(Sections 5.3 and 5.5), and frogs (Fromm  1981).

5.7.2  Birds

5.7.2.1  Food Chain Alterations--Direct effects  of acidity on birds are
not expected, but indirect effects by alterations in food resources and
bioaccumulation of toxic metals is possible. Waterfowl that feed on
fish are likely to avoid lakes  devoid of  prey.   Indeed, species richness
of fish-eating birds such as  mergansers,  loons,  and gulls is positively
correlated with pH (Aimer et  al.  1978,  Nilsson and Nilsson 1978).  The
                                  5-138

-------
diet of the common loon,  Gavia  Imrner,  is approximately 80 percent fish,
the remainder consisting  of crustaceans, molluscs, aquatic insects, and
leeches (Barr 1973).   The range of the loon includes the sensitive areas
of Canada's Precambrian Shield  (Godfrey 1966) and the Adirondack
Mountains.   Populations have declined  in the Adirondacks (Trivelpiece et
al. 1979),  but no causal  relationship  between acidification and
declining bird populations was  implied (Mclntyre 1979) but habitat
restriction unrelated to  acidification is important in.   In Quebec, the
common merganser (Mergus  merganser)  and the kingfisher (Megaceryle
alcyon) were observed only on those  lakes where the summer pH is higher
than 5.6 (DesGranges  and  Houde  1981).  The distribution of the black
duck (Anas rubripes)  has  been restricted in some .lakes in Maine because
of the lack of their  preferred  invertebrate prey (Reinecke 1979) but
habitat restriction unrelated to acidification is important in this
area. Some waterfowl  may  prefer acidic lakes if they can prey on the
large predatory insects which are often very common in these lakes
(Section 5.3.2.5). Goldeneye ducks  (Bucephala clangula) were shown to
favor acidic fishless lakes that had large insect populations (Eriksson
1979) and to feed in  larger numbers  around a lake after the fish were
experimentally removed (Eriksson et  al. 1980a).  Birds are opportunistic
feeders.  The alteration  of a food resource from a number of lakes may
reduce the population but not cause  a  total loss of population as the
birds switch to other resources and  to other lakes in the region.

     Birds such as swallows, flycatchers, and kingbirds that feed on the
aerial adult form of  aquatic insects are forced to find alternative food
sources if the insect populations upon which they normally feed are
depleted (Section 5.3.2.5).   In early  spring when many aquatic insects
emerge, acid runoff to lakes and ponds is at a peak.  It is also in
early spring that the birds depend heavily on a supply of food to
prepare for nesting and raising young. This may be the explanation for
the observation in southern Quebec,  where the tree swallow (Iridoproene
bicolor) was more common  during the  breeding season around the less acid
lakes studied (DesGranges and Houde  1981).  Blancher (1982) observed
that weight gain of kingbirds was related to insect emergence, not lake
pH.  Lake pH was not  correlated with densities of red-winged blackbirds
(Agelaius phom'ceus)  and  barn swallows (Hirundo rustica).

5.7.2.2  Heavy Metal  Accumulation—Al terations of food resources may not
be the only mechanism by  which  birds may be inhibited by acidity.  The
mobilization of metals at low pH (Chapter E-4, Section 4.6) may result
in increased body burdens in higher  trophic levels.  Studies by Nyholm
and Myhrberg (1977) and Nyholm  (1981)  have implicated aluminum in the
impaired breeding of  four species of passerines in Sweden.  Aluminum is
quite insoluble in the alkaline conditions characteristics of vertebrate
intestines, but it might  be actively transported across the intenstinal
barrier if calcium is in  short  supply. Effects were manifested by
reductions in breeding success; formation of thin, porous eggs; small
clutch size; and lower egg weight near acidic lakes.  The causal link
with lake pH was suggested to be the high aluminum content of the
insects near the acidic lakes (Nyholm  1981).  A laboratory study (Gilani
and Chatzinoff 1981)  proved that aluminum is toxic to bird embryos but
                                  5-139

-------
results from Al  injected into eggs  are not comparable  to  field  responses
to dietary Al.   Similar findings of decreased egg  size  and weight were
found for the eastern kingbird (Tyranus tyranus) in  Ontario  (Blancher
1982).  Mercury levels were found to be elevated in  the eggs  of
goldeneye ducks (Bucephala  clangula)  near acidic  Swedish lakes
(Eriksson et al.  1980a,b).  Across  eastern North America  where  extensive
pesticide use has occurred, the mobilization of pesticides and  heavy
metals by acidification may have even  more serious effects,  but these
considerations have not been researched.  This whole area concerning how
acidification may affect metal  and  pesticide toxicity  requires  more
research.

5.7.3  Mammals

     Mammals that feed on aquatic plants and animals,  such as muskrats,
minks, otters,  shrews, and raccoons, will  be affected  variously by
acidification,  depending on the sensitivity of their food organisms to
acidity and their ability to choose alternate food sources and  suitable
habitats in acidified areas.  While many species are not  directly
affected, they are likely to experience major changes  in  availability of
food and habitat quality.   An increase in the concentration  of  heavy
metals in the diet of certain species  of wildlife  may  occur  (Newman
(1979).  Raccoons (Procypn lotor) from the sensitive Muskoka  area of
Ontario contain mercury levels of 4.5  ppm in their livers, a  level five
times greater than in raccoon livers from an area  with non-acidic waters
(Wren et al. 1980).  Metal contamination of roe deer (Capreolus
capreolus) resulted in reduced weight and antler size  in  an
industrialized region in Poland (Sawicka-Kapusta 1978,  1979;  Jop 1979),
but this metal  deposition is not related to the long-range deposition
characteristic of North America.  In remote areas  of Sweden,  however,
cadmium accumulated in the body tissues of roe deer  and moose (Alces
alces) (Frank et al.  1981, Mattson  et al.  1981).

     The long-term effects of anthropogenic acidification on  caribou
(Rangifer tarandus caribou) are potentially severe.  The  primary source
of winter browse for caribou (Thompson and McCourt 1981), the lichen
Cladina stellaris is very sensitive to acidity (Chapter E-3,  Section
3.2.2).  Exposure of this lichen to simulated acidic rain at pH 4.0
reduced photosynthetic rates by about a quarter (Lechowicz 1982).
Recovery time from drying was also  impaired.  The  caribou/caribou lichen
relationship is very sensitive, as  the lichen normally  grows  only 6 mm
per year (Scotter 1963) and an adult caribou eats  5  kg of lichen per day
(Hanson et al.  1975).  Any impairment of lichen growth rate  may have a
serious impact on the winter range  of caribou, but it  will take years
for this effect to be noticeable as normal regeneration of lichen
communities requires in excess of 30 years.

     Acidic deposition may affect mammals by causing changes  in soil
chemistry that can sequester important nutrients  (Chapter E-2,  Section
2.3.3.3).  One mineral that is likely  to be made less  available to
herbivorous animals is selenium.  The solubility of  selenium  in soil
pore water declines with pH (Geering et al. 1968,  C. M. Johnson 1975),


                                  5-140

-------
and uptake by grasses is inhibited by SOX (Davies  and Watkinson  1966,
Gissel-Nielsen 1973), so concentrations of selenium  in forage  are
reduced in areas of sensitive soils receiving acidic deposition
(Gissel-Nielsen 1975, Shaw 1981).   Furthermore,  excess sulfur  in the
diet of animals can scavenge selenium from tissues (Harr 1978).  Dietary
deficiency of selenium leads to degeneration of  the  liver,  kidney, and
heart (Schwarz and Foltz 1957,  Harr 1978).  Selenium deficiency  leads to
muscular dystrophy ("white muscle  disease")  in sheep,  cattle,  swine, and
horses (Muth et al. 1958, Muth  and All away 1963, Hidiroglou et al. 1965,
Harr 1978).  Many soils in eastern North America are naturally low in
selenium and produce forage with concentrations  below the 0.1  ppm level
recognized as essential (Kubota et al.  1967, Levesque 1974, Winter
and Gupta 1979).  Incidence of  white muscle disease  has been related to
the use of sulfur-containing fertilizers in areas  naturally deficient in
selenium (Davies and Watkinson  1966, Allaway and Hodgson 1964, Allaway
1970).  Effects on the availability of  other essential  minerals, like
molybdenum (Chapter E-2, Section 2.3.3.3), may be  equally important but
have not yet been considered.

5.7.4  Summary

     Effects of acidification on vertebrate animals, not including fish
(Section 5.6) are still largely speculative.  The  potential effects are
diverse and research is at an early stage.  These  data are  summartized
in Table 5-15.  Many of the effects are expected to  take many  years to
appear; therefore long term monitoring  will  be essential.   The following
tentative conclusions can be drawn:

     0   Direct effects are most severe on the embryos and  larvae of
         amphibians, including  salamanders,  newts, frogs, and  toads.
         Sensitivity to acidity varies  widely within closely related
         taxa, but total amphibian biomass may decline in areas  exposed
         to acidic rainfall and snowmelt.

     0   Fish-eating birds (e.g.,  loons, mergansers) will be unable to
         rear young in areas where fish populations  are limited,
         resulting in smaller population sizes for portions of the
         breeding range.

     o   Some insectivorous bird populations may be  limited by the
         reduced availability of preferred prey  (flycatchers,  swallows,
         kingbirds) around acidic  lakes, but others  (goldeneye ducks)
         seek out the species of aquatic insects found in acidic lakes
         and may actually prosper in impacted areas.

     o   Mammals that feed on plants and animals in  acidic  lakes may
         accumulate higher than normal  body burdens  of heavy metals, but
         population losses have not yet been demonstrated.

     o   The large North American  herds of caribou may be affected in
         the long-term due to the sensitivity of the caribou lichen upon
         which they depend for  winter browse.


                                  5-141

-------
                         TABLE 5-15.  SUMMARY OF EFFECTS OF ACIDITY ON NON-FISH VERTEBRATES
en
i
ro
Taxa
AMPHIBIA
Ambystoma maculatum
A. jeffersonianum
Trituris vulgaris
T. helveticus
Motophythalamus
viridescens
Hyla cruel fer
Rana catesbeiana
R. sylvatica
R. temporaria
R. pipiens
Acrls gryllus
Bufo bufo
Common name
Yellow-spotted
salamander
Jefferson
salamander
Smooth newt
Palmate newt
Red-spotted newt
"Salamanders"
Spring peeper
Bullfrog
Wood frog
Common frog
Leopard frog
Cricket frog
Common toad
Observation Mechanism
Reproductive failure at Embryonic mortality
pH < 6.0
Egg nunber correlated ?
with pH
No effect of pH 5.0 ?
Not observed < pH 6.0 Cation concentration
Tolerant to pH 3.8 ?
Tolerant of pH 7.4-4.8 ?
Leave water at pH 4.0 Behavior change
Population density ?
correlated with pH
Mortality at pH 4.0 Embryonic mortality
Mortality below pH 5.0 ?
Mortality below 5.0 Embryonic mortality
Mortality below 5.0 ?
Mortality below 5.0 ?
Mortality below 4.7 Embryonic mortality
Reduction in sperm ?
mortality at pH < 6.5
Adults observed at pH ?
4.8
Mortality at pH 4.0 Embryonic mortality
Not observed < pH 4.2 ?
Evidence
Field obs.
Field obs.
Field obs.
Field correl .
Field obs.
Field obs.
Field pH manlp.
Field obs.
Lab study
Field obs.
Lab study
Field obs.
Field obs.
Lab study
Lab study
Field obs.
Lab study
Field obs.
References
Mushlnsky and Brodle 1975,
Pough and Wilson 1977
Clark and Euler 1980
Pough 1976
Cooke and Frazer 1976
Cooke and Frazer 1976
George et al . 1977,
pers. obs. (RS)
Hall and Likens 1980a,b
Clark and Euler 1980
Gosner and Black 1957
Clark and Euler 1980
Saber and Dunson 1978
Clark and Euler 1980
Hagstron 1977
Noble 1979
Schlicter 1981
pers. obs. (RS)
Gosner and Black 1957
Hagstrom 1977

-------
                                                      TABLE 5-15.   CONTINUED
en
 i
co
Taxa
B. calamita
BIRDS
Gavia immer

Mergus merganser
Megaceryle alcyon
Iridoprocne bicolor
Anas rubripes
Eucephala clangula
Tyranus tyranus
Passerines
MAMMALS
Procyon lotor
Capreolus capreolus
Al ces alces
Common name
Natterjack toad
Common loon
Common merganser
Kingfisher
Tree swallow
Black duck
Goldeneye duck
Eastern kingbird
Songbirds (4 sp)

Raccoon
Roe deer
Moose
Observation
Not observed < pH 5.0
Habitat restriction in
sensitive areas
Avoidance of acid lakes
Avoidance of acid lakes
Avoidance of acid lakes
Avoidance of acid lakes
Preference for acidic
lakes
Elevated (Hg) In eggs
Decreased egg weight
near acidic lakes
Breeding failure, thin,
porous eggs

5 x normal (Hg)
Cd accumulation
Cd accumulation
Mechanism
?
Land use changes,
fish losses?
Fish losses
Fish losses
Fish losses
Aquatic Insect losses
Abundance of preda-
tory Insect food
1 terns
From Hg In Insects
Aluminum toxicity?
Aluminum in Insect
prey

Bioaccumulation
Bioaccumulation
Bioaccumulation
Evidence
Field obs.
Field obs.
Field obs.
Field obs.
Field obs.
Field obs.
Field obs.
Lab analysis
Lab analysis
Lab analysis

Lab analysis
Lab analysis
Lab analysis
References
Beebee and Griffin 1977
Trivelpiece et al. 1979
Mclntyre 1979
DesGranges and Houde 1981
DesGranges and Houde 1981
DesGranges and Houde 1981
DesGranges and Houde 1981
Eriksson 1979
Eriksson et al. 1980a
Blancher 1982
Nyholm 1981, Nyholm and
Myhrberg 1977

Wren et al. 1980
Frank et al . 1981
Frank et al. 1981,
         Rangifer sp.
Caribou
Loss of winter browse
 over a long period
Sulfur sensitivity of  Lab study
  caribou lichen
  Mattson et al. 1981

Lechowicz 1982

-------
     0   Other grazing animals, including some domestic cattle,  may be
         subject to mineral  deficiencies, particularly  selenium,  if high
         SOX deposition continues for extended periods.  The
         seriousness of this impact is difficult to quantify and is
         highly speculative at this time.

     0   Mechanisms of impact include disrupted ionic balances in
         amphibians, metal  toxicity in higher trophic levels of
         wildlife, alterations in food chains, and nutrient
         deficiencies.

5.8  OBSERVED AND ANTICIPATED ECOSYSTEM EFFECTS (J. P.  Baker, F.  J.
     Rahel, and J. J. Magnuson)

     Acidification may produce changes in either ecosystem structure or
function.  Effects on structure involve changes in species composition
caused by species declines,  extinctions or replacements.   Effects on
ecosystem function refer to changes in such processes as  primary
production, energy transfer between trophic levels, detrital
decomposition and rates of nutrient cycling.   Most studies have
described the response of individual  taxa to the acidification process.
Thus most of our knowledge about the ecosystem-level  effects of
acidification concern changes in structure.  Little is  known about how
these structural changes influence ecosystem function.   The object of
this section is to note the ecosystem changes which have  been observed
in acidic habitats and to suggest potential ecosystem responses  that
need to be examined in future studies.

5.8.1  Ecosystem Structure

     Acidification produces changes in the basic structure of aquatic
ecosystems (Figure 5-17).  Certain taxa (e.g., fish and Daphnia)
disappear apparently as a direct result of acid toxicity"Direct
effects of acidity or aluminum are, however,  complicated  by interactions
among a complex web of consumers and their food resources (Section
5.10.2.3).  Important components of upper trophic levels-fish
populations decline or disappear.  As a result, large-bodied
acid-tolerant invertebrates  become top predators in the system
(5.3.1.5).  Shifts in the importance of invertebrate predators may alter
zooplankton community structure which, in turn, may alter the
phytoplankton community structure.  The reduction of grazers (snails,
amphipods, etc.) may allow periphyton to accumulate,  while the
inhibition of detritivores and decomposers apparently causes detritus to
accumulate.  Within benthic  and planktonic communities  the number of
species generally decreases.  The overall result is a general  decrease
in ecosystem complexity.

     Woodwell  (1970) considered simplification a system response  common
to all types of environmental pollution and also to natural  sources of
environmental  stress.  It is possible that simplification  increases
system instability (e.g., Woodwell 1970, Van Voris et al.  1980),
although the relationship between system complexity and system stability


                                  5-144

-------
                                   FISH
                               ZOOPLANKTON
                                           SMALL GRAZERS AND \
                                           DETRITIVORES       \


                                                  t    "\\
                                               MACROPHYTES AND   \\
                                               PERIPHYTON         \N
                                DECOMPOSERS

                                    I
                                 DETRITUS
                            [NON-ACIDIFIED LAKE]


                               INVERTEBRATE
                               PREDATORS
                                        V
                                              SMALL GRAZERS AND

                                              DETRITIVORES

                                                          \
                                                  MACROPHYTES AND

                                                  PERIPHYTON
Figure  5-17.
Trophic interactions in a neutral pH,  oligotrophic lake
compared to those  in an acidified lake.   Dotted  lines
indicate trophic  interactions  which may  be particularly
affected by acidification.   Note the  replacement of fish
by invertebrates  as the top-level predators.
                                     5-145

-------
is disputed (Allen and Starr 1982).  Marmorek (1983)  and Yan et al.
(1982) observed in field experiments that acidification indirectly
reduced the short-term stability and reslience of the plankton community
to nutrient additions.

     The physical structure of the aquatic system may also be slightly
altered with acidification.  The correlation between  increasing acidity
and increased water clarity has been well  established (Chapter E-4,
Section 4.6.3.4).  With an increase in light penetration,  some shift in
the thermal budget and patterns of thermal stratification  may occur  as
has been demonstrated for Lake 223 in the Experimental  Lakes Area of
Ontario (Schindler and Turner 1982).

5.8.2  Ecosystem Function

5.8.2.1  Nutrient Cycling--It has been suggested that nutrient cycling
and nutrient availability to primary producers are reduced in acidic
aquatic environments.  The rate of nutrient cycling is  thought to be
slowed primarily because of inhibition of bacterial decomposition and a
sealing-off of mineral sediments from the overlying water  column with
the accumulation of detritus and periphyton on the lake bottom (Section
5.3.2.1).   Grahn et al. (1974) speculated that acidification stimulated
lake oligotrophication as a result of these changes but definite
confirmation of this hypothesis is lacking.

     Nutrient availability could also be affected by  chemical  changes in
the water.  Of particular importance may be decreased phosphorus
availability because of aluminum-phosphorus interactions (Chapter E-4,
Section 4.6.3.5), decreased levels of dissolved inorganic  carbon due to
the decrease in pH (Section 5.5.2.3.2),  and also precipitation of
organics (Chapter E-4, Section 4.6.3.3)  and increased displacement of
these materials into benthic habitats.  Although all  of these postulated
chemical changes are theoretically plausible and potentially very
significant, effects on nutrient cycling in acidic waters  have not yet
been experimentally demonstrated.

5.8.2.2  Energy Cycling--Previous sections have discussed  four types of
possible reactions to acidification that are relevant to energy cycling
in aquatic systems:   1) a potential decrease in primary productivity, 2)
decreased  growth efficiencies, 3) decreased energy transfer between
trophic levels and 4) elimination of upper trophic levels.   The evidence
or lack of evidence for these hypotheses is discussed below.

     Biological  productivity in aquatic  ecosystems is supported by both
allochthonous organic carbon imported from sources external  to the
system plus autochtonous production of organic carbon by primary
producers  within the aquatic system.  As a result of  decreased nutrient
availability,  water column primary productivity in acidic  waters may be
altered.  Limited observations from field studies reviewed in Section
5.5.2.1.2  indicate,  however, that in most cases acidification has no
consistent adverse effect on primary productivity. Adverse effects  of
decreased  nutrient availability on water column primary productivity may
                                  5-146

-------
 be counterbalanced  by other changes (especially increased light
 penetration)  that stimulate primary production.  Although acidification
 does  not consistently decrease primary productivity, increased light
 penetration apparently does, to a certain extent, increase the
 importance of benthic primary producers relative to planktonic primary
 producers.  The effects of acidification on total primary production
 (including periphyton, macrophytes and phytoplankton) have not been
 studied.

      Energy transfers within aquatic systems can be examined both within
 a  given  trophic level and between trophic levels.  Growth efficiency
 usually  refers to within stratum transfer; the fraction of a given
 quantity  of energy  (food or light energy) consumed that is manifested as
 production (growth and reproduction).   Organisms that inhabit acidic
 waters may be inherently less efficient or may be less efficient because
 of  acid-induced stress, but examination of this phenomenon has been
 limited.  Fish have been observed in laboratory experiments to grow more
 slowly at lower pH levels (Section 5.6.2.8).  Primary producers in some
 acidic waters (Sections 5.5.2.1.2 and 5.3.2.2.3) have lower
 instantaneous rates of production per unit biomass.   Possible reasons
 for this lower production are numerous, however, and have not been
 clearly defined.  No studies of growth efficiencies  for zooplankton,
 benthos, or other aquatic organisms have been completed.  If growth
 efficiencies  are reduced in acidic environments, transfer of energy
 through the food chain would be reduced.

     Energy transfers between trophic  levels involve the percentage of
 available food actually used by consumers, or relative productivities in
 successive trophic levels.   In Section 5.5.3.3, it is postulated that
 the transfer  of energy between phytoplankton and zooplankton may be
 inhibited by  the inedible nature of many  of the phytoplankton species
 common in acidic lakes.   In stream systems,  a reduction in populations
 of  benthic invertebrate grazers apparently decreases conversion of
 primary production into secondary production (Section 5.3.2.5.4).
 Processing of detrital  particles may also be affected.   Again,  some
 evidence suggests there may be inhibition of energy  cycling  and energy
 transfer through the food chain.

     One of the best documented changes associated with acidification is
 the decline and loss of fish populations  which represent major
 components of upper trophic levels in  aquatic ecosystems.   Loss of fish
 populations results in a shortened aquatic food chain (Section 5.7).

 5.8.3  Summary

     Structural  changes  in  acidified aquatic ecosystems have been  well
 documented and include the  loss of fish populations,  reductions in the
number and diversity of  benthic and  planktonic invertebrates, and
 accumulations of periphyton and detritus.   How these structural  changes
affect ecosystem processes  such as primary production,  energy transfers
between trophic levels,  or  nutrient cycling  is largely  unknown.  To
date,  the limited evidence  available suggests that ecosystem functions


                                  5-147

-------
are relatively robust,  but additional  research is  required  before  final
conclusions can be reached.

5.9  MITIGATIVE OPTIONS RELATIVE TO BIOLOGICAL POPULATIONS  AT  RISK
     (C. T. Driscoll  and G.  C.  Schafran)

     The concept of surface water neutralization as  a  result of  base,
carbon, and phosphorus  additions is discussed in Chapter  E-4,  Section
4.4.  The biological  response to these additions and other  mitigative
options for fish populations at risk from acidification of  surface
waters follows.

5.9.1  Biological  Response to Neutralization

     In lakes where neutralization has resulted in large, rapid  pH
changes (e.g., Ca(OH)2  addition, see Chapter  E-4,  Section 4.7.1),
phytoplankton concentrations have been observed to decline  drastically
This phenomena may be either the result of stress  associated with  a
drastic change in pH  ("pH shock") or removal  of algal  biomass  with
metals through flocculation and precipitation processes (Scheider  and
Dillon 1976, Scheider et al. 1975).  Yan and  Dillon  (1981)  noted that a
small pH change, or a large pH change initiated gradually,  resulted  in
no change in biomass  of lake phytoplankton.

     After base addition, phytoplankton undergo a  taxonomic shift.
Certain species will  disappear while others appear.  Species dominance
has been observed to  shift,  and total  number  of species has been
observed to increase.  Species dominance/composition are  lake-specific,
so response of the phytoplankton population cannot be  generalized  for
all lakes.  Subsequent  to liming, Scheider et al.  (1975)  observed  a
shift in dominance to the genera Dinobryon and an  unidentified
chrysomonad.  The appearance of diatoms (Bacillariophycae--mostly
Navicula and Nitzschia) and blue-green algae  (Cyanophyta-Oscillatoria)
was also note
-------
have resulted In smaller pH changes have not affected  the  population
negatively (Scheider et al. 1975,  Dillon et al.  1979).   Swedish lakes
that have undergone a gradual  increase in pH through base  application
show a substantial  increase in zooplankton biomass, shifts in  species
composition,  and increases  in  species diversity  (Hultberg  and  Andersson
1982).

     Recovery of zooplankton populations is much slower  than that
observed for phytoplankton.  For two full  years  following  base addition,
zooplankton biomass was observed not to recover  to  pretreatment levels
(Van and Dillon 1981).  This relatively slow recovery  from base addition
stress may be due to slow life cycles and recolonization difficulties.

     The literature is not  consistent with respect  to  the  response of
benthic fauna to base addition.  In the first year  following large pH
increases due to base addition, Scheider et al.  (1975) observed numbers
of benthic organisms decrease  substantially.  Chironomids, which were
observed to be dominant prior  to neutralization  (Scheider  et al. 1975,
Van and Dillon 1981), contributed significantly  to  this  decline.  This
was attributed to an interruption of a life cycle in response  to the
sudden pH change.  However, this is not consistent  with  Swedish
observations.  Hultberg and Andersson (1982) observed  that the groups
Orthocladinae and Tanypodinae  increased, while no change was evident in
trichopteran  populations.  With benthic fauna constituting an  important
food source for fish, population perturbations resulting from
neutralization may affect fish positively or negatively.

     In some regions, a felt!ike structure of algal filaments, detritus,
and Sphagnum completely cover  lake sediments and deplete normal
populations of submerged vegetation like Isoetes and Lobelia (Grahn et
al. 1974, Hendrey and Vertucci 1980).  Hultberg  and Andersson  (1982)
indicate that liming appears to have a profound  effect on  Sphagnum.
After base addition, Sphagnum  was rapidly eliminated from  the  littoral
region where CaC03 was spread.  Populations were slowly  depleted (1 to
2 years) in the remainder of the treated lakes.   The few plants that
survived neutralization exhibited very slow growth  rates ( ~1  cm x
yr-1) as compared to acidic lake populations (8  to  10  cm x yr-1)
(Hultberg and Andersson 1982).  In lakes that were  allowed to  reacidify,
Sphagnum was observed to recolonize the benthic  region.

     Neutralization to improve the water quality of acidified  waters has
both a long- and short-term effect on fish.  Immediately following base
addition and subsequent pH  rise, aluminum hydrolysis generally occurs.
This perturbation, as previously described, may  be  detrimental to the'
existing fish population (Baker and Schofield 1980).   Mortality of  fish
may be lessened by incremental addition of base, resulting in  small pH
changes.  In some lakes this may not be deemed necessary as the fish
population may be negligible.

     The long-term consequence of lake neutralization, provided
reacidification is not allowed to occur, is a much  more  hospitable
environment for fish.  An immediate response (improvement) in


                                  5-149

-------
reproduction and survival  has been observed in one-year-old fry
(Hultberg and Andersson 1982).   An increase in recruitment and fish
survival tends to increase the biomass of the younger fish where
previously the population  had been dominated by older fish (Dickson
1978).  If pH is maintained, fish reproduction and survival will  show
marked improvement over acidified conditions and possibly restore the
population to pre-acidification levels.  Restocking of native species,
lost because of acidification, may be necessary in some waters.

5.9.2  Improving Fish Survival in Acidified Waters

      Three major approaches for improving fish survival  in acidified
waters deal directly with  the fish.  They are 1) screening existent  fish
strains to determine which strains exhibit high acid tolerance, 2)
selectively breeding a given strain for improved tolerance to low pH,
and 3) acclimating a group of fish to increase their resistance to
acidic water.

5.9.2.1  Genetic Screening—Several studies have shown differences in
acid tolerance between different strains within the same species
(Johnson 1975, brook trout; Gjedrem 1976, brown trout; Robinson et al.
1976, Swarts et al. 1978,  Edwards and Gjedrem 1979, Rahel and Magnuson
1980, yellow perch; and Schofield et al. 1981).

      Edwards and Gjedrem  (1979) determined that the method used  for
screening different strains was important in determining the hierarchy
of tolerance among strains.  They screened brown trout finger!ings (5.8
+ 0.8 g) in water synthetically acidified to pH values of 2.5, 3.0,  and
T.O and brown trout eggs and fry in naturally acidic water (pH 4.7)  and
in water adjusted from pH  4.7 to 5.2 with sodium hydroxide.  They found
a high correlation of ranking among strains tested at low pH values,
indicating that the pH level used within this range was unimportant.
However, when they compared ranking obtained from the finger!ings tested
at very low pH values and  those determined from the eggs and fry  tested
in the naturally acidic water, they found a low rank correlation  between
strains.  They concluded that the two different procedures were
apparently testing for different traits and thus could not be used
interchangeably.

     The results of Edwards and Gjedrem (1979) indicate that a
standardized screening procedure is very important in determining the
relative tolerance of strains within species.  Their results also
indicate that the life cycle stage screened is important in determining
relative strain tolerance.  Thus, it is important to develop a  screening
procedure consistent with the goals of the project.  Edwards and  Gjedrem
(1979) concluded that a screening program aimed at reestablishing viable
populations in acidified waters must select for strains with
acid-resistant egg and larval stages because the major cause of trout
population losses is thought to be poor recruitment caused by egg and
fry mortality (Beamish and Harvey 1972, Jensen and Snekvik 1972,
Leivestad et al. 1976, Schofield 1977).  However, if the goal of  a
screening program is to find a strain to be used in maintaining  stocked


                                  5-150

-------
populations, the screening procedure should target the life cycle
stages that will be stocked.

5.9.2.2  Selective Breeding—The logical  extension of a genetic
screening program is to select for acid tolerance within a few superior
strains and improve their acid tolerance through selective breeding.
Gjedrem (1976) and Edwards and Gjedrem (1979)  found high heritabilities
(ratio of genetic variance to total  variance)  for acid tolerance  in eggs
and alevins of brown trout.  They concluded that there was a good
possibility of producing acid-tolerant strains of brown trout through
selective breeding.

     Selective breeding tests with brook trout have produced mixed
results.  Swarts et al. (1978) performed a single selection with  NYSV
strain brook trout (selecting to 80 to 90 percent loss of equilibrium  at
pH 3.4 to 3.5) and found no increased tolerance in their offspring  in
field or laboratory tests.  Schofield et al.  (1981) selected yearling
(1977 year class) domestic strain brook trout to 50 percent, using
naturally acidified runoff water.  They then  challenged the offspring
(1979 year class) of the resistant and non-resistant cohorts as fry in
naturally acidified water.  The offspring of  the resistant cohort were
significantly more resistant (mean LTso 195.5 hr) than those of the
non-resistant cohort (LT$Q 72.0 hr;  P < 0.001).  However,  when an
identical  test was performed on the 1980 year class offspring of  the
1977 year class resistant and non-resistant cohorts, the offspring  of
the resistant cohort exhibited inferior performance to that of the
offspring of the non-resistant cohort (LTso values 76.6 and 77.1  hr vs
84.7 hr, respectively).  Included in the 1980 year class tests were
tests of hybrid crosses between resistant and non-resistant cohorts and
two wild strains from Canada (Assinica and Temiscamie).   In these tests
the resistant X Assinica and resistant X Temiscamie always performed
better than the non-resistant X Assinica and  non-resistant X Temiscamie.
From these results Schofield et al.  (1981) hypothesize that genetically
inherent physiological  acid tolerance may be  fixed within  the selected
cohorts.

     In a preliminary field trial, Schofield  et al. (1981)  separated
Assinica X domestic yearlings into resistant  and non-resistant cohorts
in March of 1979, stocked them in equal numbers in an acidified lake in
May, and sampled them in July.  They observed a 3:1 return of resistant
over non-resistant fish.  However, more extensive field trials performed
in 1980 produced a resistant/non-resistant ratio not significantly
different from the expected 1:1 ratio of the  no difference case.
Schofield et al. (1981) attributed the lack of an unbalanced ratio to
the relatively good water quality conditions  in the spring of 1980
caused by low snowfall  during the winter of 1980.  This study appears  to
give some evidence of improved acid tolerance of brook trout through
selective breeding, but it is far from conclusive.

     Hybrid vigor with  regards to acid tolerance has been  observed in
several studies.  Robinson et al. (1976)  found heterosis (hybrid  vigor)
in 66 percent of the strain crosses  tested.  Edwards and Gjedrem  (1979)
                                  5-151

-------
observed mean percent survival  in hybrid crosses of brown  trout  to  be
twice that of the parental  strains.   From this  they suggest  that the
most efficient way to produce acid-tolerant strains for  restocking
acidified waters would be to identify the best  strain  crosses  and then
maintain just a few pure bred strains in the hatchery.   These  strains
could be improved by selective  breeding  while hybrid fish  for  stocking
could be routinely produced by  crossing  a brood fish of  the  pure bred
lines.

5.9.2.3  Acclimation--A conceivable  method for  improving the success of
stocked populations in acidified waters  would be to. acclimate  the fish
to the acidic conditions before stocking.  The  question  of whether  fish
can acclimate to acidic conditions has been addressed  by numerous
authors, with mixed results.  Most of the studies in which fish  were
acclimated to sublethal pH values and then tested for  increased  survival
at lethal  pH values have produced negative results.  Lloyd and Jordan
(1964) acclimated rainbow trout to pH values of 6.55,  7.50,  and  8.40 and
found no difference in survivorship  when the fish were tested  at pH
values from 3.0 to 4.0.  Robinson et al. (1976)  held brook trout at pH
3.75 for one week and then tested them for survival  at pH  2.5  and 3.0.
They found that survival time was 20 to  25 percent less  in acclimated
fish than in fish not previously exposed to acidic water.  Falk  and
Dunson (1977) exposed brook trout to sublethal  pH values of  5.0  and 5.8
for two or 24 hours prior to testing for survival  at pH  3.15 or  3.5.
They found significant differences in survival  time between  acclimated
and non-acclimated fish in only three of nine tests.  Swarts et  al.
(1978) performed laboratory and field acclimation trials with  brook
trout.  In the laboratory they  acclimated the fish to  pH 4.25  for 10
days or 4.8 for 28 days and then tested  them for improved  survival  at pH
3.25 or 3.6 respectively.  They found no consistent differences  between
acclimated and non-acclimated fish in their laboratory trials.   In  three
field trials in which fish were held in  an acidified stream  (pH  4.8 to
5.8) and then tested in an acidic river  (pH 4.2),  the  acclimated fish
performed better than non-acclimated fish in only one  trial.

     In a study with embryos and alevins of Atlantic salmon  and  rainbow
trout which had been incubated at pH values ranging from 4.5 to  6.8 for
variable time periods, Daye (1980) could find no difference  in tolerance
between the different groups and thus concluded no acclimation had
occurred.   In a similar study,  performed by Trojnar (1977b), brook  trout
eggs were incubated at pH 4.6,  5.0,  5.6, and 8.0 and then  tested at
swim-up for survival at pH values from 4.0 to 7.86.  He  found  that  fish
incubated at pH 5.6 and below showed greatly increased survival  at  low
pH as compared with fish incubated at pH 8.0.  He attributed the
difference to acclimation.

     Physiological evidence for acclimation in  brown trout exposed  to
acidified water was provided by McWilliams (1980b),  who  suggested that
acclimation might occur through a progressive decrease in  the
diffusional  permeability of the gills to sodium.  However, actual
resistance to lowered pH levels, in  terms of increased survival ship, was
not determined in this study.
                                  5-152

-------
      In all of the aforementioned studies,  the acclimation  procedure
 consisted of holding the fish at a single  sublethal  pH  for  a  fixed  time
 period and then transferring them to the test pH  levels.  Guthrie  (1981)
 used a different method.  He hypothesized  that previous acclimation
 attempts had failed for three major reasons.   First,  if the acclimation
 pH was too high the fish might not need  to  adjust physiologically to
 maintain homeostasis.   The study by Lloyd  and Jordan (1964) might be an
 example of this.  Second,  if the acclimation  pH is too  low  then  it  might
 constitute a major stress in itself,  to  which the fish  are  unable to
 adjust.  The study by  Robinson et al.  (1976),  where  the fish  were
 acclimated to a pH of  3.75 before being  tested at a  lower pH,  is
 probably an example of this.   Third,  if  the test  pH  is  very low and the
 adaptive response of the fish is overwhelmed,  then no amount  of previous
 acclimation will improve survival.   This probably occurred  in  the
 studies where the test pH  was below 4.0  (Lloyd and Jordan 1964, Robinson
 et al. 1976, Swarts et al.  1978,  Falk  and Dunson  1977).

      To avoid these problems, Guthrie  (1981)  developed  a gradual
 acclimation procedure  in which the acidity  and aluminum concentration
 were increased from control  conditions to test conditions over a period
 of 4 to 5 days.   He used test pH values  of  5.0, 4.5, and 4.0 with
 nominal aluminum concentrations of 0.2 and  0.4 mg Al £-1.   In
 acclimation tests on brook  trout sac-fry and  swim-up fry, Guthrie (1981)
 found significantly improved  survival  at pH 5.0 and  4.5 at both aluminum
 levels, but no difference  in  survival  between  acclimated and  non-
 acclimated fish  at pH  4.0.   This lends credence to the  hypothesis that
 pH values below  4.0 are too low for testing for acclimation.   Guthrie
 also acclimated  brook  trout parr (55.7 ^6.8 mm)  to  naturally  acidic
 water (pH 4.9, 0.32 mg Al  £-1)  by gradually changing water  from
 non-acidified lake water (pH  6.5)  to acidic brook  water.  After 6 days
 in the acidic brook water,  80 percent  of the  acclimated fish  remained
 alive while only 40 percent of the  non-acclimated fish  (transferred into
 the acidic brook water at  the same  time  that  the  acclimation procedure
 was completed) were still  alive.   In experiments  with advanced fry  (28
 to 36 mm) and yearlings at pH 5.0 with 0.4  mg  Al  £-!, dramatic
 improvements in  the survival  of the acclimated fish  were also observed.
 However,  at pH 4.5 with the same  aluminum level,  acclimation did not
 improve survivorship in these life  history  stages.

      The studies performed  by Guthrie  (1981) clearly demonstrate the
 ability of brook trout to  resist increased  acidity and  aluminum levels,
 within specific  limits of  water quality  and developmental  sensitivities,
 as measured by improved survival  of fish in short term  gradual
 acclimation treatments.  These results indicate it may  be possible,
 through acclimation prior  to  stocking, to improve  initial  survival  in
 hatchery-reared  brook  trout destined for stocking  in waters of low  pH
 and high  Al  levels.

 5.9.2.4  Limitations of Techniquest to Improve Fish  Survival—In the
 future it appears that a combination of  these  three  techniques could be
 a  feasible strategy for maintaining a  sport fishery  in  waters where the
 extent of acidification is  such that a natural  fishery  is no longer
                                  5-153
409-262 0-83-17

-------
possible.  This could be accomplished by screening for the most acid
resistant strains of fish, selectively breeding those strains  and
acclimating them to the acid water before stocking.

     This strategy would probably be successful  in allowing the
maintenance of a sport fishery where none could exist otherwise, however
it would not be a solution.   It is doubtful  that these techniques  could
ever be used to re-establish a naturally reproducing  population where
one had been lost due to acidification.   Also since these techniques all
require a great deal of propagation work and clearly  defined genetic
strains, it would only be possible to use game fish.   The
reestablishment of non-game  fish in acidified waters  using these
techniques would not be feasible.

     When these techniques are used to re-establish sport fisheries in
acidified waters there is one foreseeable contraindication.  Toxic
metals such as mercury may be mobilized as a result of acidification.
This could result in a hazardous situation if stocked fish accumulated
these contaminants before they were caught.   Thus it  is important  that
fish stocked in acidified waters be closely monitored for toxic metals
contamination.

5.9.2.5  Summary--All three  techniques for producing  the fish  better
able to survive in acidified waters—genetic screening, selective
breeding, and acclimation--show promise as ameliorative strategies.
However, all are still in the early stages of development and  require
more laboratory and field testing before they will  be well  enough
defined to be useful as fish management tools.

5.10  CONCLUSIONS (J. J. Magnuson, F. J. Rahel,  J.  P. Baker, R. Singer,
      and J. H. Peverly)

     Although the literature regarding the response of aquatic biota to
acidification is sometimes conflicting,  some effects  have been well
documented.  These are summarized below (Section 5.10.1). Emphasis is
placed on those biological changes which are supported by a combination
of field observations, field experiments and laboratory experiments.
Together, these species declines, extinctions and replacements represent
major changes in the structure of acidified aquatic ecosystems. The
next section (5.10.2) focuses on the mechanisms by which acidification
affects aquatic ecosystems.   Although mechanisms by which acidification
may affect processes such as primary production, energy transfer between
trophic levels, and nutrient cycling have been hypothesized, few have
been critically evaluated using field and laboratory  experiments.   The
major conclusion is that many of these mechanisms are speculative  and
need to be examined in future research.   Section 5.10.3 describes
potential mitigative options from a biological perspective.  The final
section (5.10.4) presents an overview of biological changes expected if
current rates of acidic deposition continue in the northeastern United
States and southeastern Canada.
                                  5-154

-------
5.10.1  Effects of Acidification  on  Aquatic Organisms

    The effects of acidification  on  aquatic organisms  that are  supported
by numerous observations and experimental  studies  are  summarized in
Table 5-16 and in the following statements.

                                 Benthos

    0   The bottom community, which  provides  substrates  for many
        organisms and is the principal  site of  nutrient  recycling, is
        severely altered in clear waters  low  in pH,  as compared to
        otherwise similar,  but neutral  pH  waters.

    o   Bacterial metabolic rates are  decreased, between pH 6.0 and 4.0,
        and shredding invertebrate populations  are reduced in numbers,
        bringing about an increased  accumulation of  undecomposed organic
        materials.

    0   Most substrates are covered  with  an encrusting mat of algae and
        detritus in acidic  lakes  and streams  below pH  5.0.

    °   Many predatory insects (beetles,  true bugs,  dragonflies)
        increase in numbers below pH 6.0  in lakes  and  streams.  Their
        effect on the plankton and on  benthic detritivores is not known.

    o   Several preferred food sources for game fish (e.g., Gammarus
        snails, many mayflies and stoneflies) do not survive below pH
        5.0, but fisheries  impacts due to  food  shortages have not been
        observed.

                               Macrophytes

    0   Dominant macrophyte species  are the same in  both acidified (pH
        less than 5.6) and nonacidified (pH 5.6 to 7.5)  North American
        lakes.

    0   Shifts to Sphagnum-dominated macrophyte communities have been
        documented in six Swedish lakes acidified  for  at least  15 years.
        However, this does not seem  to be a general  property of
        acidified lakes as there is  currently no trend toward dominance
        of macrophyte communites by  Sphagnum  spp.  in 50  oligotrophic,
        soft-water lakes surveyed in North America.

    o   Standing crops of macrophytes vary widely  (5 to  500 g dry wt
        m-2) in soft-water, oligotrophic  lakes, and  acidification
        produces no consistent changes in  standing crop.  In Lobe!ia
        dortmanna, a common plant in soft-water, oligotrophic lakes,
        oxygen production was reduced 75  percent at  pH 4.0 vs pH 4.3 to
        5.5 in one flow-through laboratory experiment.
                                 5-155

-------
                     TABLE  5-16.   EFFECTS  OF  INCREASING  ACIDITY ON  AQUATIC ECOSYSTEMS.   "NUMEROUS"
            REFERS  TO  MANY OBSERVATIONS  OR  EXPERIMENTS, WHICH ARE DESCRIBED IN  THE SECTIONS  INDICATED
            Taxa or Process
                                                         Type of Evidence
 Field Observation   Field Experiment   Lab Experiment
                                                                                          Observed Effects
en
         Benthos

           Molluscs
           {most  species except
           fingernail clams, family
           sphaerlidae)
           Crayfish
           Amphi pods
           (Gammarus)
Mayfly larvae
(Ephemeroptera)
Numerous
(Section 5.2 and
5.3)
Aimer et al. 1978    Mills 1982
K. Okland 1980c
Sutcliffe and
Carrlck  1973
                                       Numerous (Section
                                       5.3)
                    Hall et al. 1980
                                                                  Mai ley 1980
                                                                  Costa 1967
                                                                  Borgstrom and
                                                                  Hendrey 1976

                                                                  Bell and
                                                                  Nebeker 1969,
                                                                  Bell 1971
Water striders  (Gerridae),     Numerous (Section
backswimmers  (Notonectidac),   5.3)
water boatmen (Corixidae),
beetles  (Dytiscidae,
Gyrinidae), dragonfl ies
(Odonata)

Benthos  community structure   Numerous (Section
                             5.2 and 5.3)
                                                           Hall  et al.  1980
           Benthlc algae
           (periphyton)
Numerous  (Section
5.3)
                                      Bell and
                                      Nebeker 1969,
                                      Bell 1971,
                                      Mai ley 1980
                                                Hall et al.  1980   Hendrey 1976
                                                Schindler 1980
The calcareous  shell of these animals
is soluble under acidic conditions
making this group highly sensitive  to
low pH.  Few species present below  pH
6.0 except for  several species of
fingernail  clams which may persist
down to pH  4.5- 5.0.

In soft water lakes, calcium uptake
and exoskeleton formation Inhibited in
the pH range 5.0-5.8. Reproduction
Impaired at pH  5.4.

Absent below pH 6.0, in the laboratory
avoids pH 6.2 and lower.
Most species  decline or are absent  1n
the pH range  4.5 to 5.5.
                                                                                              Tolerant of acidity.  Increase in
                                                                                              abundance in acidified lakes (below pH
                                                                                              5.0) after other invertebrate groups
                                                                                              and fish have been  eliminated.
With increasing acidity, species
richness  declines.  Entire groups  of
aquatic organisms are absent or poorly
represented below pH 5.0 (e.g., mol-
luscs, amphipobs, crayfish, mayflies).
Other taxa become dominant, particu-
larly after the loss of fishes (e.g.,
predacious beetles and true bugs).

Algal mass overgrow rooted plants
and cover bottom subtrates in
acidified lakes below pH 5.0

-------
                                                   TABLE 5-16.    CONTINUED
         Taxa or Process
                                                        Type of Evidence
                              Field Observation    Field Experiment   Lab Experiment
                                                                                                     Observed Effects
      Macrophytes

        Eriocaul on sp.
        Lobelia sp.
      Plankton
en
 i
en
Zooplankton community
structure
        Phytoplankton community
        structure
      Fishes

        Fatheal Minnow
        (Pimephales promelas)
                             Grahn 1977,  Best
                             and & Peverly 1981,
                             Miller et al. 1982
                                       Laake 1976         Rosette plant communities may
                                                         become overgrown by algal mats.
                                                         Tissue aluminum concentrations
                                                         increase as pH decreases
                                                         photosynthesis of rosette species
                                                         decreases by 75% as pH declines
                                                         from 5.5 to 4.0.
                                      Numerous (Section
                                      5.5)
Numerous (Section
5.2 and 5.5)
                                                                    Davis and
                                                                    Ozburn 1969
                                                                    Parent and
                                                                    Cheetham 1980
                             Numerous (Section
                             5.2 and 5.5.)
                    Van and Stokes
                    1978
                             Rahel  and Magnuson
                             1983
Mills 1982
                                                         Most  species are acid-sensitive
                                                         and absent below pH 7.0 to 5.5
                                     The number of species declines as
                                     acidity  increases.  Taxa
                                     characteristic of acid conditions
                                     include  certain genera of
                                     rotifers (Keratella. Kellicottia,
                                     Polyarthra); cladocerans
                                     TBosmina); and copepods
                                     (biaptonus) .

                                     The number of species declines as
                                     acidity  increases.  Dinoflagellates
                                     (Phylum  Pyrrophyta) frequently
                                     dominate acidified lakes (pH 4.0-5.0).
                                     Dinoflagellates are a less palatable
                                     food source for zooplankton compared
                                     to the phytopl ankton they frequently
                                     replace.
                                       Mount  1973        One of the most acid-sensitive
                                                        fish species. Reproductive failure
                                                        occurs near pH 6.0.   Generally
                                                        absent in waters below pH 6.5.

-------
                                                      TABLE  5-16.   CONTINUED
         Taxa or Process
                                                       Type  of Evidence
                               Field Observation   Field Experiment   Lab Experiment
                                                                Observed Effects
en
en
00
        Darters
        (Etheestoma exile, £.
        nigtum, Percina capnodes)
        and Mi nnows (several
         Notropis spp. Pimephales
         notatus)

         Smallmouth Bass
         (Micropterus dolomieui)
Lake Trout
(Salvelinus namayeusch)

White Sucker
(Catostomus commersoni)

Rainbow Trout
(Salmo gairdneri)

Atlantic Salmon
(Salmo salar)

Brown Trout
(Salmo trutta)

Brook Trout
(Salvelinus fontinalis)

Sunfishes
(Ambloplites rupestris,
Micropterus salmoides,
Lepomis spp.l
        Yellow Perch
        (Perca flavescens)
      Decomposition
                              Harvey  1980
                              Rahel and Magnuson
                              1983
Beamish 1976,
Harvey 1980,  Rahel
and Magnuson  1983

Beamish 1976,        Mills 1982
Beamish et al1.  1976
                                       Rahel and
                                       Magnuson 1983
Harvey 1980,  Rahel
and Magnuson  1983

Numerous (Section
5.6)

Numerous (Section
5.6)

Numerous (Section
5.6)

Numerous (Section
5.6)

Harvey 1980,
Rahel  and Magnuson
1983
                              Svardson 1976,
                              Keller et al. 1980,
                              Harvey 1980, Rahel
                              and Magnuson 1983

                              Hendrey 1976,
                              Leivestad et al.
                              1976
                                                          Mills 1982
                                                          Hall et al. 1980


                                                          Smith 1957
                                                                             Beamish 1972
                                                                             Trojnar 1977a
Numerous
(Section 5.6)

Numerous
(Section 5.6)

Numerous
(Section 5.6)

Numerous
(Section 5.6)
                                       Rahel  1983
                    Scheider et al.    Leivestad et
                    1976, Gahnstrom    al. 1976
                    et al. 1980, Hall
                    et al. 1980
                  Very acid-sensitive.  Generally
                  absent below pH  6.0 in both
                  naturally acidic and  anthro-
                  pogenically acidified waters.
                                                                                       Reproduction ceases and populations
                                                                                       become extinct below pH 5.2-5.5
                  Experiences reproductive failure  near
                  pH 5.0. Generally absent below pH 5.0
                  in both naturally acidic and
                  anthropogenically acidifed  waters.
Adversely affected by pHs below
5.0-5.5

Adversely affected by pHs below
5.0.

Lower pH limit between 4.5 to
5.0.

Lower pH limit between 4.2 to
5.0.

Lower pH limit near 4.5.
                  Lower pH limit 4.2  to  4.5.  May
                  become very  abundant after other
                  species have become extinct.
                  Bacterial  decomposition  is  signifi-
                  significantly  reduced  in the pH
                  range 4.0  to 5.0.   In  many
                  cases,  fungi replace bacteria as
                  the primary decomposers

-------
In the two published studies of metal  concentrations in
macrophytes from acidic lakes, tissue  concentrations of iron,
lead, copper and especially aluminum are higher,  while cadmium,
zinc and manganese are lower compared  to tissue concentrations
in plants from nonacidic lakes.

                        Plankton

Changes in species composition, standing crop,  and productivity
of the plankton community with acidification are complex and
probably result from not only lower pH levels and higher metal
concentrations, but also decreased fish predation, increased
water clarity, and perhaps decreased nutrient availability.

The structure of the plankton community in acidic lakes (pH 4.0
to 6.0) is markedly different from that in nonacidic lakes
within the same region.  With increasing acidity, the total
number of species decreases (by 30 to  70 percent) and biomass is
dominated by fewer species.

Comparisons between acidic and nonacidic lakes  within the same
region and experimental acidification  of a lake indicate no
consistent change in water column primary productivity with
increased acidity.

Data on zooplankton productivity are not available.   In three
studies, the biomass and/or numbers of zooplankton were lower in
more acidic lakes (pH 4.0 to 5.0).

                          Fish

The clearest evidence for impacts of acidification on aquatic
biota is adverse effects on fish.

Loss of fish populations associated with acidification of
surface waters has been documented in  the LaCloche Mountain
range of Ontario, Nova Scotia, and southern Norway.   Available
data for these regions include historic records of declining
fish populations coupled with historic records  of increasing
water acidity.   Additional  evidence for loss of fish populations
is available from the Adirondack region of New  York  State and
southern Sweden.

In the United States, only in the Adirondack region  have adverse
effects of acidification on fish populations been observed.  The
presence of fish in Adirondack lakes and streams  is  correlated
with pH level.   Particularly below pH  5.0, the  occurrence of
fish is reduced.   Loss of fish populations has  been  documented
for about 180 Adirondack lakes (out of a total  of approximately
2877), although historic records are not available at this time
to relate each loss specifically to acidification or acid
deposition.
                          5-159

-------
Fish kills have been observed during episodic acidification  of
surface waters in Norway and Ontario.   In  addition,  in
hatcheries receiving water directly from lakes or rivers,
unusually heavy mortalities of adult and young fish  have
occurred in the Adirondack region,  Nova Scotia,  and  Norway.
These mortalities are typically associated with  rapid decreases
in pH (generally to pH levels below 4.5 to 5.0)  during  snowmelt.

Many fish populations in acidic waters (pH 4.5 to 5.0)  lack
young fish, implying that failure to reproduce is a  common,
although not the only, cause for extinction of fish  populations
with acidification.  In Sweden, neutralization through  lake
liming resulted in the recurrence of young fish.

Field observations of growth of adult fish in acidic (pH  4.0 to
6.0) versus nonacidic waters, or through time with acidifica-
tion, typically indicate no change or increased growth  with
increased acidity.  In some cases,  increased growth  may be a
result of reduced competition for food as  fish populations
decline.

Experiments in the laboratory and the field have established a
direct cause-and-effect between acidification and adverse
effects on fish.  In the field, acid additions to Lake  223 in
the Experimental Lakes Area of Ontario produced pH declines  from
pH 6.5 to 5.9 in 1976 to pH 5.1 in 1981 and resulted in
reproductive failures and/or extinction of several fish
populations.  In laboratory bioassays, pH  and aluminum  levels
typical of acidified surface waters were toxic to fish.

                    Other Related Biota

Effects of acidification on amphibians, birds, and mammals are
still largely speculative.  Research is at an early  stage.
Decreased pH levels have been demonstrated in the laboratory to
decrease amphibian reproductive success, but the significance
and extent of breeding habitats acidified or sensitive  to
acidification have not yet been evaluated.

                    Ecosystem Effects

Changes in ecosystem structure have been well-documented  in
acidified aquatic habitats and include species declines,  local
extinctions and reduced species richness in many taxonomic
groups.  In some cases, acid-tolerant taxa which formerly were
rare, may become abundant.

The effects of acidification on ecosystem processes  such  as
primary production, energy transfer between trophic  levels and
nutrient cycling have not been well-studied and should  be
addressed in future research efforts.
                          5-160

-------
5.10.2  Processes and Mechanisms by Which Acidification Alters Aquatic
        Ecosystems

5.10.2.1  Direct Effects of Hydrogen Ions—Effects of low pH on aquatic
organisms are the best studied aspect of the acidification process.
Numerous laboratory bioassays have documented both the toxicity of
hydrogen ions to aquatic organisms and differences in sensitivity to
acid stress among taxonomic groups.  These studies provide insight into
physiological mechanisms of toxicity and offer guidelines for predicting
effects of various pH levels on aquatic biota. Mechanisms by which
various taxa are affected by low pH have been discussed elsewhere
(Section 5.3 through 5.6, Fromm 1980) and include disruptions in ion
transport, acid-base balance, osmoregulation and enzyme function.   Low
pH stress seldom exists alone in acidified waters and thus its effect on
aquatic organisms will be influenced by other stresses (Sections
5.10.3.2, 5.10.3.4, 5.10.3.7) and biological  interactions (Section
5.10.3.3).

5.10.2.2  Elevated Metal  Concentrations—The acidification process has
resulted in elevated concentrations of aluminum and other metals in  many
waters (Chapter E-4, Section 4-6).  Aluminum leached from the soil  in
response to acidic deposition has been implicated in fish kills in field
observations, field experiments, and laboratory studies (Section
5.6.4.2).  The interaction of acidity and aluminum is especially
important as fish may be killed by aluminum at a pH value not considered
harmful by itself.  The toxicity of aluminum is greatest in the pH range
4.5 to 5.5.

     In laboratory experiments, aluminum precipitates phosphorus from
water, with the greatest effect occurring in the pH range 5.0 to 6.0
(Aimer et al. 1978).  Phosphorus is the nutrient that typically limits
plant growth in oligotrophic lakes.  While increased aluminum due  to
acidification would be expected to reduce phosphorus concentrations  and
thereby reduce productivity, this process has not been confirmed by
in-lake studies.

     Aluminum concentrations are higher in macrophytes from acidified
lakes than in macrophytes from nonacidified lakes.   The biological
significance of these higher aluminum concentrations is not known.

     High mercury concentrations in fish are correlated with low pH
levels for lakes in Sweden,  Ontario and the Adirondack Mountains of  New
York (Section 5.6.2.5).  In  laboratory experiments,  bological  uptake of
most metals is enhanced at low pH but whether lake acidification will
significantly enhance bioaccumulation of mercury has not been
definitively demonstrated.   Furthermore,  there is considerable variation
in fish mercury concentrations between lakes and not all  acidified lakes
contain fish with elevated mercury concentrations.   Other factors, in
addition to pH,  which may contribute to between-lake variability in  fish
mercury concentrations include dissolved  organic carbon,  conductivity,
bioproductivity and watershed geology.
                                  5-161

-------
     Other metals which consistently  exhibit  increased concentrations In
acidic surface waters are manganese and  zinc  (Chapter E-4, Section
4.6.1).  Currently available toxicity data  indicate  that concentrations
of these metals in acidic surface  waters (unless  local sources of metal
emissions exist) are below toxic  levels.  However, a lack of  sufficient
bioassay data collected in soft,  acidic  waters  and the potential for
additive or synergistic effects with  other  toxic  components make this
statment tentative.

5.10.2.3  Altered Trophic-Level Interactions--The loss of fish from
acidified lakes has been documented  in Scandinavia,  Canada and the
United States (Section 5.6.2.1).   As  the top  predators in aquatic
habitats, fish are known to exert control over  trophic structure,
trophic dynamics and nutrient cycling in lakes  (Brooks and Dodson 1965,
Shapiro et al. 1975, Kitchell  et  al.  1979,  Clepper 1979, Zaret 1980).
For example, zooplanktivorous fish,  by influencing the species
composition and size distribution of  zoopl ankton, can alter the rate of
primary production in lakes (Shapiro  et  al. 1975).

     Changes in aquatic ecosystems following  the  loss of fish
populations are evident in nonacidified  lakes where  fish have been
intentionally removed (Stenson et al. 1978, Eriksson et al. 1980b,
Henrikson et al. 1980a,b).  Large invertebrate  predators (e.g.,
corixids, dytiscid beetles, Chaoburus) normally kept at low abundance by
fish predation become abunda"nT;Zoopl ankton  community composition
changes and dinoflagellates become dominant among the phytoplankton.
Many of these same changes have been  observed in  lakes which  have lost
their fish populations as a result of acidification. Thus biological
and limnological changes in a complex aquatic ecosystem undergoing
acidification may be difficult to ascribe directly to the  toxicity of
increased acidity or metal concentration.   Understanding the  role of
trophic-level interactions in producing  biological changes during
acidification will require holistic,  manipulative studies  of  consumer
regulation of ecosystem dynamics.

5.10.2.4  Altered Water Clarity—Water clarity  typically increases with
increased acidity (Section 5.5.2.3.2  and Chapter  E-4, Section 4.6.3.4).
This may be due to a reduction in algal  biomass in the water  column, the
precipitation of dissolved organics by aluminum or changes  in the
light-absorption capacity of aquatic  humic  materials.   Increased light
penetration would allow macrophyte and phytoplankton growth  at  greater
depths and would warm the water to a greater  depth.

5.10.2.5  Altered Decomposition of Organic  Matter--Decomposition of
organic material releases nutrients for reuse by  plants.   Reductions  in
decomposition rates have been reported in some  acidified lakes  as  a
result of decreased bacterial metabolic  rates and declines in
populations of  shredding invertebrates.   It has been suggested  that
decreases in  nutrient recycling as a  result of decreased decomposition
would lead to decreased productivity at all trophic  levels,  but this
hypothesis has  not been adequately tested nor have  consistent decreases
in productivity been observed.
                                  5-162

-------
5.10.2.6  Presence of Algal  Mats—Algal  mats  which cover  the lake bottom
down to the limit of light penetration are  characteristic of acidified
lakes.  While these mats would be expected  to interfere with water
column-sediment interactions important in the recycling of nutrients,
this hypothesis has not been experimentally tested.  The  degree to which
the physical alteration of the bottom substrate  affects benthic
invertebrates and fish is unknown.

5.10.2.7  Altered Nutrient Availability—Increased aluminum
concentrations could decrease the concentration  of phosphorus via
precipitation of aluminum-phosphorus complexes.   Reducing phosphorus
availability should decrease biological  production but this result needs
to be quantitatively evaluated.  Nitrogen added  via  acidic deposition is
used as a nutrient, but overall biological  effects on production would
be negligible since phosphorus is the limiting nutrient in most
oligotrophic waters.

5.10.2.8  Interaction of Stresses—Predicting the response of a
particular lake or stream to acidification  is difficult because
acidification results in many 1imnological  changes besides increased
acidity.  These changes interact with biotic  responses in complex and
often counterbalancing ways.  This  is illustrated by the  response of the
phytoplankton to acidification.  Phytoplankton biomass and productivity
have shown increases, decreases, or no change with respect to decreasing
pH (Section 5.8).  Certain types of algae (dinoflagellates) are
frequently dominant in acidic lakes, yet exceptions  are not uncommon.
Alga species that are rare one year may  dominate a lake the following
year (Van and Stokes 1978).   Variation in the response of plankton
communities to acidification may result  from  the interaction of many
factors.  Acidification eliminates  sensitive  algal species, may decrease
phosphorus and inorganic carbon concentrations,  and  may depress nutrient
cycling.  These changes would tend  to decrease phytoplankton biomass and
productivity.  Yet acidification may increase water  clarity, allowing
light to penetrate into the thermocline  and hypolimnion,  where nutrient
levels are generally higher.  This  would tend to increase productivity.
Zooplankton are similarly affected  by numerous factors besides pH,
including changes in their food supply and  the loss  of fish predators.

     The response of fish to acidification  is likewise complicated.
Aluminum and hydrogen ions interact to cause  fish mortalities.  Yet this
interaction may be most important during short time  periods (e.g.,
spring snowmelt) and may not be detected during  stream or lake surveys
done at other times of the year.  Laboratory  experiments  predict
decreased fish growth in acidified  waters (Section 5.6.4.1.3), yet
increased fish growth has been observed  in  the field.  The reason may be
that the increased metabolic demands at  low pH are outweighed by the
greater abundance of forage organisms available  to a continually
dwindling fish population.  Reproductive failures, not decreased growth,
the loss of food items, or adult mortality, appear responsible for most
fish extinctions.
                                  5-163

-------
     Contradictory responses should  not be  interpreted as evidence that
acidification has no effect, but rather as  an indication that poorly
understood interactions among stresses may  be involved.  The infrequency
of manipulative, whole-system experiments has contributed to this lack
of resolution.

5.10.3  Biological Mitigation

     Techniques for mitigating the effects  of acidification on aquatic
organisms include base additions to  neutralize the acidity (Section
5.9.1 and Chapter E-4, Section 4.7.1) and development of acid-tolerant
fish strains (Section 5.9.2).   Immediately  after base addition dramatic
reductions in phytoplankton, zooplankton and benthic fauna have been re-
served.  However, the long-term consequence of lake neutralization, pro-
vided that reacidification  is not allowed to occur, is repopulation by
aquatic organisms and an environment that is more hospitable for fish.

     Fish survival in acidic waters  may be  enhanced by genetic
screening, selective breeding and acclimation.  These techniques appear
to be a feasible strategy for maintaining a sport fishery in waters
acidified to the point where a natural fishery is no longer possible.
It is doubtful, however, that they could be used to reestablish
naturally reproducing fish  populations and  they do not address the
problem of restoring other  components of the biota to preacidified
conditions.  Because of the potential for increased metal concentrations
in fish from acidified waters (Section 5.6.2.5), fish stocked in such
waters should be monitored  for toxic metal  accumulation.

5.10.4  Summary

     Biological effects due to acidification become apparent as
alkalinities decline to 65-35  yeq jr1 and pH's decline to between
6.5 to 6.0 (Table 5-16). Since the  biological response to acidification
is a graded one, continuing pH declines below this range  will result in
escalating biological changes.   In Chapter  E-4 it was concluded that,
under current rates of acidic  deposition, a long-term pH of £ 4.9 can be
expected in low-alkalinity  lakes and streams in the northeastern United
States and southeastern Canada that  have pH levels in the mid 5's prior
to acidification (Section 4.4.4).  Episodic depressions down to pH 4.3
to 4.9 will occur during periods of  snowmelt and heavy rainfall and can
affect systems with a pH as high as  7.0 (Section 4.4.2).  These pH
levels, along with other changes associated with the acidification
process (e.g., increased aluminum clarity,  accumulation of detritus and
algal mats), will have significant harmful  effects on aquatic organisms.
In waters where pH values average 4.9 or lower, most fish species,
virtually all molluscs, and many groups of  benthic invertebrates will be
eliminated.  Increased aluminum concentrations may eliminate fish
species otherwise tolerant  of low pH.  The  plankton community will be
simplified and dominated by a few acid-tolerant taxa.  Benthic algal
mats will often cover the lake bottom, and  water clarity may increase.
These represent the best documented  effects of acidification.  Effects
on ecosystem processes remain largely unconfirmed and are an important
area for future research efforts.
                                 5-164

-------
 5.11   REFERENCES

 Adams, G. F. and C. H. Olver.  1977.  Yield properties  and structure  of
 boreal percid communities in Ontario.  J.  Fish.  Res.  Board Can.
 34:1613-1625.

 Allaway, W. H.  1970.  Sulphur-seleniun relationships in  soils and
 plants.  Sulphur Inst. J. 6(3):3-5.

 Allaway, W. H. and J. F. Hodgson.  1964.   Symposium on  nutrition, forage
 and pastures:  selenium in forages as related to the  geographic
 distribution of muscular dystrophy in livestock. J.  Animal  Sci.
 23:271-277.

 Allen, T. F. H. and T. B. Starr.   1982.  Hierarchy:   Perspectives for
 Ecological Complexity.  Univ. of  Chicago  Press,  Chicago,  IL. 310 pp.

 Aimer, B., W. Dickson, C. Ekstrom, and E.  Hornstrom.  1974.  Effects  of
 acidification on Swedish lakes.   Ambio 3:30-36.

 Aimer, B., W. Dickson, C. Ekstrom, and E.  Hornstrom.  1978.  Sulfur
 pollution and the aquatic ecosystem pp. 273-311. Ir±  Sulfur  in the
 Environment. Part II.  J. Nriagu, ed.  John Wiley and Sons,  New York.

 Anderson, R. S.  1974.  Crustacean plankton communities of 340 lakes  and
 ponds in and near the national  parks of the Canadian  Rocky Mountains.
 Journal of the Fishery Research Board of  Canada  31:855-869.

 Andersson, G., H. Berggren,  G.  Cronberg,  and C.  Gel in.  1978a.  Effects
 of planktivorous and benthivorous fish on  organisms and water chemistry
 in eutrophic lakes.   Hydrobiologia 59:9-15.

 Andersson, G., S. Fleischer,  and  W. Graneli.   1978b.  Influence of
 acidification on decompostion processes in lake  sediment.  Verh.
 Internat. Verein. Limnol. 20:802-807.

 Arnold, D., R. Light, and V.  Dymond.   1980.   Probable effects of acid
 precipitation on Pennsylvania waters.  U.S.  EPA  Corvallis  Environmental
 Laboratory, EPA-600/3-80-012.  (cited in  Schofield 1982).

 Baath, E., B. Lundgren,  and  B.  Soderstrom.   1979.  Effects of artificial
 acid rain on microbial activity and biomass.   Bull. Environ. Contam.
 Toxicol.  23:737-740.

 Baker, J. 1981.  Aluminum toxicity to fish as related to acid
 precipitation and Adirondack  surface  water quality.   Ph.D. thesis,
 Cornell University,  Ithaca,  NY.

Baker, J. p.   1982.   Effects  on fish  of metals associated  with
acidification.   Proc. Acid Rain/Fisheries  Symposium,  August 1981.
 Ithaca, NY.   In press.
                                 5-165

-------
Baker, J.  P.  and J.  J.  Magnuson.   1976.  Limnological responses of
Crystal Lake  (Vilas  County, Wisconsin) to intensive recreational use,
1924-1973. Transactions  of the Wisconsin Academy of Sciences, Arts and
Letters 64:47-61.

Baker, J.  P.  and C.  L.  Schofield.   1980.  Aluminum toxicity to fish as
related to acid precipitation  and  Adirondack surface water
quality.  J^n  Ecological  Impact Acid Precipitation, Norway, SNSF Project.
D. Drabljos and A. Tollan,  eds.  Proc.  Int. Conf.  pp. 292-293.

Baker, J.  and C. Schofield.  1982.  Aluminum toxicity to fish in acidic
waters.  Water, Air  and Soil Pollution 18:289-309.

Barr, J. F.  1973.  Feeding biology of the Common Loon (Gavia immer) in
oligotrophic  lakes of the Canadian Shield.  Ph.D. Thesis, Univ. of
Guelph.

Bayly, I.  A.   1964.   Chemical  and  biological studies on some acidic
lakes of East Australian  sandy coastal lowlands.  Australian Journal of
Marine and Freshwater Research 15:56-72.

Bayly, I.  A.  E. and  W.  D.  Williams.  1973.  Inland waters and their
ecology.  Longman Australia Pty Limited.  Victoria, Australia. 316 p.

Beamish, R. J.  1972.  Lethal  pH for  the white  sucker Catostomus
comrnersoni) (Lacepede).  Trans. Am. Fish. Soc.  101:355-358.

Beamish, R. J. 1974a.  Growth  and  survival of white suckers (Catostomus
commersoni) in an acidified lake.   J.  Fish.  Res. Board Can. 31:49-54.

Beamish, R. J. 1974b.  Loss of fish populations from unexploited remote
lakes in Ontario, Canada  as a  consequence of atmospheric fallout of
acid.  Water  Res. 8:85-95.

Beamish, R. J.  1976.  Acidification  of lakes in Canada by acid
precipitation and the resulting effects on fishes.  Water Air Soil
Pollut. 6:501-514.

Beamish, R. J. and H. H.  Harvey.   1972.  Acidification of the LaCloche
mountain lakes, Ontario,  and resulting fish mortalities.  J. Fish. Res.
Board Can. 29:1131-1143.

Beamish, R. J., W. L. Lockart, J.  C.  Van Loon,  and H. H. Harvey.  1975.
Long-term acidification of a lake  and resulting effects on fishes.
Ambio 4:98-102.

Beck, K. C.,  J. H. Reuter, and E.  M.  Perdue.  1974.  Organic and
inorganic geochemistry of some coastal plain rivers of the southeastern
United States.  Geochimica et  cosmochimica acta 38:341-364.

Beebee, T. J. C. and J. R. Griffin.  1977.  A preliminary investigation
into Natterjack toad (Bufo calamita)  breeding site characteristics in
Britain.  J.  Zool. Lori3on~181:341-350.
                                  5-166

-------
Bell, H. L.  1970.   Effects of pH  on  the life cycle of the midge
Tanytarsus dissimilis.   Can.  Entomol.  102:636-639.

Bell, H. L.  1971.   Effect  of low  pH  on the survival and emergence of
aquatic insects.   Water Res.  5:313-319.

Bell, H. L. and A.  V.  Nebeker.   1969.  Primary  studies on the tolerance
of aquatic insects  to  low pH.   J.  Kans. Entomol. Soc. 42(2):230-237.

Bell, G. R., G. E.  Hoskins, and J. W.  Bagshaw.  1969.  On the structure
and enzymatic degradation of the external membrane of the salmon egg.
Can. J. Zool. 47:146-148.

Benoit, D. and G.  Hoi combe.  1978.  Toxic effects of zinc on fathead
minnows (Pimephales promelas)  in soft water.  J. Fish. Biol. 13:701-708.

Berge, F.  1976.   Kiselalger og pH 5  noen elver og innsjoer i Agder
og innsjoer i Agder og Telemark.  En  sammenlikning mellom arene 1949 og
1975.  SNSF Project, Internal  Report  18/76, Osla-As, Norway.

Best, M. D. and J.  H.  Peverly.   1981.  Water and sediment chemistry and
elemental composition  of macrophytes  in thirteen Adirondack lakes
(Abstract), p. 348. Jji Proc.  Internet. Symp. on Acidic Precipitation
and Fishery Impacts in N.E. North  America, N.E. Div. Am. Fisheries Soc.
Cornell Univ., Ithaca,  NY,  Aug. 2-5,  1981.

Bick, H. and E. F.  Drews.   1973.  Selbstreinigung und Ciliatenbesiedlung
in saurem Milieu (Model 1 versuche).  Hydrobiologia 42:393-402.

Birge, E. A. and C. Juday.   1927.  The organic  content of the water of
small lakes.  Bulletin of the United  States Bureau of Fisheries
42:185-205.

Bishop, S.  1941.   The salamanders of New York.  New York State Museum,
Bulletin 324:1-365.

Bl ancher, P. J.  1982.   The effect of lake acidification on passerine
birds.  Wildlife Toxicology Division.  Canadian Wildlife Service.
Ottawa.  28 pp.

Blcornfield, J. S.  Quinn, R. Scrudato,  D. Long,  A. Richards, and F. Ryan.
1980.  Atmospheric  and watershed inputs of mercury to Cranberry Lake,
St. Lawrence County, New York,  pp. 175-210.  Jji Polluted Rain.  T.
Toribara, M. Miller, and P. Morrow, eds.  Plenum Press, New York, NY.

Bold, H. C.  1942.   The cultivation of algae.   Bot. Rev. 8:69-138.

Borgstrom, R. and G. R. Hendrey.  1976.  pH tolerance of the first
larval stages of Lepidurus  arcticus  (Pallas) and adult Gammarus
lacustris G. 0. Sars.   SNSF Project,  Internal Report 22/76, Oslo-fls,
Norway.37 pp.
                                  5-167

-------
Borgstrom, R.,  J.  Brittain,  and A. LiTlehammer.  1976.  Evertebrater og
surt vann.  Oversikt over  innsaml ings-lokaliteter.  IR 21/76, SNSF
Project, Norway.   33 pp.

Bricker, F. J.  and 0.  E. Gannon.  1976.  Limnological investigation of
Hoop Lake - a northern Michigan Bog.  The Michigan Academician Vol IX,
1:25-41.

Brinkhurst, R.  0.   1965.   The  biology of the Tubificidae with special
reference to pollution.  In  Proc. 3rd Seminar on Water Quality Criteria,
Cincinnati, Ohio.   pp. 57^56.

Brinkhurst, R.  0.   1974.   The  Benthos of Lakes.  St. Martin's Press, New
York.  190 pp.

Brock, T. D.  1973.  Lower pH  limit for the existence of blue-green
algae:  Evolutionary and ecological implication.  Science 179:480-483.

Brock, T. D.  1978.  Thermophilie Organisms and Life at High
Temperatures.  Springer-Verlag New York. 465 pp.

Brock, T. and M.  Brock.  1970.  The algae of Waimangu Cauldron (New
Zealand):  Distribution in relation of pH.  J. Phycol. 6:371-375.

Brooks, J. and S.  Dodson.  1965.   Predation, body size and composition of
plankton.  Science 150:28-35.

Brown, D.  1982.   The Effect of pH and calcium on fish and fisheries.
Water, Air, and Soil Pollution 18:343-351.

Brown, D. J. A.  1981.  The  effects of various cations on the survival
of brown trout.  Sal mo trutta, at low pH's. J. Fish Biol. 18:31-40.

Bua, B. and E.  Snekvik.  1972.  Hatching experiments with roe of
salmonid fish 1966-1971.   Effects of acidity and salt content of
hatchery water.  Vann 7:86-93  (cited in Muniz and Leivestad 1980a).

Budd, W. A. Johnson, J.  Huss,  and R. Turner.  1982.  Aluminum in
precipitation,  streams and shallow groundwater in the New Jersey Pine
Barrens.  Submitted for publication to Water Res. Res.

Burton, T. M.  1982.  The  effects of acidification on stream ecosystems,
Chapter 9.  Jjn Acid Precipitation: Effects  on Ecological Systems.  F. M.
D'ltri, ed.  Ann Arbor Science, Ann Arbor,  MI.  pp. 209-235.

Burton, T. M. and G. E.  Likens.   1975a.  Energy flow and nutrient
cycling  in salamander populations in the Hubbard Brook Experimental
Forest, New Hampshire.  Ecology 56:1068-1080.

Burton, T. M. and G. E. Likens.   1975b.  Salamander populations and
biomass  in the Hubbard Brook Experimental Forest, New Hampshire.   Copeia
1975(3):541-546.
                                  5-168

-------
Capelli, G. M.  1975.   Distribution,  life  history, and ecology of
crayfish in Northern Wisconsin,  with  emphasis on Orconectes propinquus
(Girard).  Ph.D.  thesis.  Zoology Department, University of Wisconsin,
Madison WI, 215 p.

Carrick, T. R.  1979.   The effect of  acid  water on the hatching of
salmonid eggs.  J.  Fish Biol.  14:165-172.

Carrick, T. R.  1981.   Oxygen  consumption  in the fry of brown trout
(Salmo trutta L.) related to pH  of the  water.  J. Fish Biol. 18:73-80.

Carter, J.  1971.  Distribution  and abundance of planktonic Crustacea in
ponds near Georgian Bay (Ontario, Canada)  in relation to hydrography and
water chemistry.   Arch. Hydrobid.  68:204-231.

Cassin, P.  1974.  Isolation,  growth, and  physiology of acidophilic
chlamydomonads. J.  Phycol.  10:439-447.

Cecil, S. G. and J. J.  Just.  1979.  Survival rate, population density
and development of a naturally occurring anuran larvae (Rana
catesbeiana).  Copeia 1979(3)-.447-453.

Chapman, P. M., L.  M.  Churchland, P.  A. Thomson, and E. Michnowsky.
1980.  Heavy metal  studies with  ol igochaetes, pp. 477-592.  In Aquatic
Oligochaete Biology.  R.  0. Brinkhurst  and D. G. Cook, eds. "Plenum
Press, New York,  NY.

Charles, D. F.  1982.   Limnological studies on 50 Adirondack Mountain
(NY) lakes.  Ph.D.  Dissertation, Indiana University, Bloomington.

Clark, K. L. and D. Euler.  1980.  Reptile and amphibian habitat and
human activity in central Ontario.  Technical Report for the Ministry of
Natural Resources Lakeshore Capacity  Study, Phase III.  40 p.

Clepper, H., Ed.   1979.  Predator-prey  systems for fisheries management.
Sport Fishing Inst., Washington, D.C. 504  pp.

Clymo, R. S.  1963.  Ion  exchange in  Spagnum and its relation to bog
ecology.  Ann. Bot. N.S.  27:309-324.

Clymo, R. S.  1967.  Control of  cation  concentrations and in particular
of pH in Sphagnum dominated communities, pp. 273-284.  In Chemical
Environment in the Aquatic Habitat H. L. Golterman and "RT S. Clymo, eds.
Amsterdam, N. V.  Noord-Hollandsche Uitgevers Maatschappij.

Coesel, P. F. M., R. Kwakkestein, and A. Verschoor.  1978.
01 igotrophication and eutrophication  tendencies in some Dutch moorland
pools as reflected in their desmid flora.  Hydrobiologia 61:21-31.

Cole, G. C.  1979.   Textbook of  limnology. C. V. Mosby Company, St.
Louis.  426 pp.
                                  5-169

-------
Collins, N.  C., A.  P.  Zimmerman, and R. Knoechel.  1981.  Comparisons
of benthic infauna  and epifaima in acidified and nonacidified Ontario
lakes, pp. 35-48.   In  Effects  of Acidic Precipitation on Benthos,
Proceedings of a Regional  Symposium  on Benthic  Biology.  R. Singer, ed.
North American Benthological Society,  Hamilton,  NY.

Colquhoun, J., J.  Symula,  M. Pfeiffer, and J. Feurer.  1981.
Preliminary report  of  stream sampling  for acidification studies 1980.
New York State Department  of Environmental Conservation, Technical
Report 81-2, Albany, NY.   116  pp.

Confer, J., T. Kaaret, and G.  Likens.  1983.  Zooplankton diversity and
biomass in recently acidified  lakes.   Can. J. Fish. Aquat. Sci.
40:36-42.

Conroy, N., K. Hawley, W.  Keller, and  C. LaFrance.  1976.  Influences of
the atmosphere on lakes in the Sudbury area.  J. Great Lakes Res.
2(Suppl. 1):146-165.

Cooke, A. S. and J. F. D.  Frazer.  1976.  Characteristics of newt
breeding sites.  J. Zool.  Lond. 178:223-236.

Costa, H. H.  1967.  Responses of Gammarus pulex (L.) to modified
environment II.  Reactions to  abnormal hydrogen  ion concentrations.
Crustaceana 13:1-10.

Cowling, E. B.  1980.   An  historical  resume  of  progress in scientific
and public understanding of acid precipitatation and  its biological
consequences.  SNSF Project, Norway  PR 18/80.   29  p.

Craig, G. R. and W. F. Baksi.   1977.   The effects  of  depressed pH on
flagfish reproduction, growth  and survival.  Water Res.
11:621-626.

Crisman, T. L., R.  L.  Schulze, P. L.  Brezonik,  and S. Bloom. 1980.  Acid
Precipitation:  The biotic response  in Florida  lakes,  pp. 296-297  In
Proceedings of the International Conference  on  the Ecological ImpacTTof
Acid Precipitation. D. Drab!0s and A.  Tollan, eds.  SNSF-project
report, Oslo, Norway.

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.

Dam, H. van, G. Suurmond,  and   C.  ter Braak. 1980.  Impact of acid
precipitation on diatoms and chemistry of Dutch moorland pools, pp.
298-299.  J^n Proc.  Int. Conf.  Ecol.  Impact Acid Precip.  D. Drabljbs
and A. Tollan, eds. SNSF Project,  Norway.

Dannevig, A.  1959.  Nedborens innflytelse pa vassdragenes  surhet og pa
fiskebestanden.  Jaeger og Fisker 3:116-118.
                                  5-170

-------
Davies, E. B. and J.  H.  Watkinson.   1966.  Uptake of native and applied
selenium by pasture species  II.   Effects of sulphate and of soil type on
uptake by clover.  N.Z.J.  Agr.  Res.  9:641-652.

Davis, P. and G.  Ozburn.   1969.   The pH tolerance of Daphnia pulex
(Leydig, emend.,  Richard).   Can.  J.  Zool.  47:1173-1175.

Davis, R. B. and  F. Berge.   1980.  Atmospheric deposition in Norway
during the last 300 years  as recorded in SNSF lake sediments.  II.
Diatom stratigraphy and  inferred  pH,  pp. 270-271.  In Proceedings of the
International Conference on  the Ecological Impact oT~Acid Precipitation,
March, 1980, D. Drabltfs  and  A.  Tollan, eds.  SNSF-project, Oslo,
Norway.

Davis, R., M. Smith,  J.  Bailey, and  S. Norton.  1978.  Acidification of
Maine (USA) lakes by acidic  precipitation.  Verh. int. Verein. Limnol.
20:532-537.

Davis, R. B., S.  A. Norton,  and C. T.  Hess.  1983.  Paleobiological
reconstruction of the effects of  atmospheric deposition of acids and
heavy metals on the chemistry and biology of lakes in New England and
Norway.  lr± Proc. Ill Intern. Symposium Paleol imnology.  J. Merilainen,
R. W. Battarbee,  and P.  Huttenen, eds. Devel. Hydrobiol.  W. Junk B.V.,
The Hague.

Daye, P. G. and E. T. Garside.  1975.  Lethal levels of pH for brook
trout, Salvelinus fontipalis (Mitchell).  Can. J. Zool. 53:639-641.

Daye, P. G. and E. T. Garside.  1975.  Lethal levels of pH for brook
trout, Salvelinus fontinail's (Mitchell).  Can. J. Zool. 53:639-641.

Daye, P. G. and E. T. Garside.  1976.  Histopathologic changes in
surficial tissues of brook trout, Salve!inus fontinalis (Mitchill),
exposed to acute  and chronic levels  of pH.Can. J. Zool. 54:2140-2155.

Daye, P. G. and E. T. Garside.  1977.  Lower lethal levels of pH for
embryos and alevins of Atlantic salmon, Sal mo salar L.  Can. J. Zool.
55:1504-1508.

Daye, P. G and E. T.  Garside.  1979.   Development and survival of
embryos and alevins of the Atlantic  salmon, Salmo salar L., ontinuously
exposed to acidic levels of  pH, from fertilization.  Can. Zool.
57:1713-1718.

Dean, H., J. Coloquhoun, and J. Symula.  1979.  Preliminary survey of
some Adirondack stream fish  populations in relation to the blackfly
control program and acid precipitation.  New York State Department of
Environmental Conservation,  unpublished report.  E.S.T. 1.6 Albany, NY.
15 pp.

Del Prete, A. and C. Schofield.   1981.  The utility of diatom analyses
of lake sediments for evaluatiing acid precipitation effects on dilute
lakes.  Arch. Hydrobiol.   9(3):332-340.
                                  5-171

-------
DesGranges, J.  L.  and B.  Houde.   1981.  Composition of bird communities
associated with lakes and other wetlands of varying acidity in the
province of Quebec:   Progress  report.  Unpubl. C.W.S. Manuscript.  37
pp.

Dickson, W.  1975.  The Acidification of Swedish Lakes. Rept. Inst.
Freshw. Res. Drottningholm  54:8-20.

Dickson, W.  1978.  Some  effects  of  the acidification of Swedish lakes.
Verh. Internat. Verein. Limnol. 20:851-856.

Dickson, W.  1980.  Properties of acidified waters, pp. 75-83.  Ij^
Ecological Impact of Acid Precipitation.  D. Drablos and A. Tollan, eds.
SNSF project, Oslo,  Norway.

Dillon, P., D.  Jeffries,  W.  Snyder,  R. Reid, N. Yan, D. Evans, J. Moss,
and W. Scheider.  1978.   Acid  precipitation in south-central Ontario:
recent observations.  J.  Fish. Res.  Bd. Can. 35:809-815.

Dillon, P., N.  Yan,  W. and  N.  Conroy.  1979.  Acid lakes in Ontario
Canada:  Characterization,  extent and responses to base and nutrient
additions.  Arch.  Hydrobiol. Beih. Ergeb. Limnol. 13:317-336.

Dively, J. L.,  J.  E. Midge,  W. H. Neff, and A. Anthony.  1977.  Blood
P02, PC02, and pH changes in brook trout  (Salvelinus fontinalis)
exposed to sublethal levels of acidity.   Comp. Bioche, Physlol.
57:347-351.

Dodson, S.  1974.   Zooplankton competition and predation:  An
experimental test of the  size-efficiency  hypothesis.  Ecology
55:605-613.

Donald, D. B.,  R.  S. Anderson  and D. W. Mayhood.  1980.  Correlations
between brook trout growth  and environmental variables for mountain
lakes in Alberta.   Transactions  of the American Fisheries Society
109:603-610.

Drablds, D. and I  Sevaldrud.   1980.  Lake acidification, fish damage
and utilization of outfields.  A compoarataive survey of six highland
areas in  Hedmark County, eastern Norway, pp. 354-355.  D. Drabltfs and
A. Tollan.  Prod, Int.  Conf.  Ecol.  Impact Acid Precip., Norway 1980,
SNSF project.

Drab!tfs, D., I. Sevaldrud,  and J. A. Timberlid.   1980.  Historical
land-use changes related  to fish status development  in different areas
in southern Norway, p. 367-369.   Eds. D.  Drabljis  and A. Toll an.
Procl,  Int. conf. ecol.  impact acid  precip., Norway  1980, SNSF project.

Driscoll, C. T.  1980.  Chemical  Characterization of Some Dilute
Acidified Lakes and Streams in the Adirondack Region of New York State,
Ph. D., Thesis, Cornell  University,  Ithaca, NY.
                                  5-172

-------
Driscoll, C. T., J.  P.  Baker, J. J. Bisogni, and C. L. Schofield.  1980.
Effect of aluminium  speciation  on  fish in dilute acidified waters.
Nature 284:161-164.

Dunson, W. A. and R. R. Martin.  1973.  Survival of brook trout in a
bog-derived acidity  gradient. Ecology 54:1370-1376.

Edmondson, W. T.  1944.  Ecological studies of sessile Rotatoria.
Ecolgical Monographs 14:31-66.

Edmondson, W.  1959.  Fresh-water  Biology.  2nd ed.  John Wiley and
Sons, New York, NY.

Edwards, D. and T. Gjedrem,  1979.  Genetic variation in survival of
brown trout eggs, fry,  and fingerlings in acidic water. SNSF-project
Norway FR 16-79. 28  pp.

Edwards, D. J. and S. Hjeldness.   1977.  Growth and survival of
salmonids in water of different pH.  SNSF-project, Norway, FR 10/77, 12
P-

Eriksson, M. 0. G.  1979.  Competition between freshwater fish and
goldeneyes, Bucephala clangula  (L.) for common prey.  Oecologia (Berl.)
41:99-107.

Eriksson, M. 0. G.,  L.  Henrikson,  and H. G. Oscarson.  1980a.
Fagelskydd.  Sjofagel och forsurning - nagra synpunkter.  Var Fagelvarld
39:163-166.

Eriksson, M. 0. G.,  L.  Henrikson,  B. I. Nilsson, G. Nyman, H. G.
Oscarson, and A. E.  Stenson.  1980b.  Predator-prey relations important
for the biotic changes in acidified lakes.  Ambio  9:248-249.

Evans, D.  1975.  Ionic exchange mechanisms in fish gills.  Comp.
Biochem. Physiol. 51A:491-495.

Falk, D. L. and W. A. Dunson.   1977.  The effects  of  season and acute
sublethal exposure on survival  times of brook trout at low pH.  Water
Res. 11:13-15.

Farmer,  G., T. Goff, D. Ashfield,  and H.  Samant.   1980.   Some effects  of
the acidification of Atlantic  salmon rivers in Nova Scotia.  Canadian
Technical Report of Fisheries  and   Aquatic  Sciences,  No.  972.

Fanner,  G. J., T. R. Goff, D.  Ashfield, and H. S.  Samat.  1981.   Some
effects  of the acidification of Atlantic  salmon  rivers in Nova  Scotia,
pp. 73-91.  Jji Acid Rain and the Atlantic Salmon.   IASF  Spec. Pub!. 10.

Fassett, N. C.  1930.  The plants  of  some northeastern Wisconsin  lakes.
Transactions of the Wisconsin  Academy of  Arts, Science and Letters
25:157-168.
                                  5-173

-------
Faucon, A. S. and W.  D.  Hummon.  1976.  Effects of mine acid on the
longevity and reproductive  rate  of the Gastrotricha Lepidodermellj
squammata (Dujardin).  Hydrobiologia 50:205-209.

Ferris, V. R., L. M.  Ferris,  and J. P. Tjepkema. 1976.  Genera of fresh
water nematodes (Nematoda)  of Eastern North America.  PB 224 - 365/413A
U.S. Environmental Protection Agency, Cincinnati, Ohio.

Fiance, S. B.  1978.   Effects of pH on the biology and distribution of
Epheitierella funeral is (Ephemeroptera).  Oikos 31:332-339.

Forman, R. T. (ed.).   1979.   Pine barrens:  Ecosystems and Landscape.
Academic Press, Inc,  New York.   601 p.

Frank, A., L. Petersson,  and  T.  Morner.   1981.  Lead and cadmium
contents in organs of moose,  roe deer, and hare.  Svensk
Veterinartidning 33(6):151-156.

Frey, D. G.  1951. The fishes of North Carolina's Bay lakes and their
intraspecific variation.  J.  of  the Elisha Mitchell Scientific Society
67(l):l-44.

Friberg, F., C. Otto, and B.  S.  Svensson.  1980.  Effects of acidifica-
tion on the dynamics  of allochthonous leaf material and benthic
invertebrate communities in running waters, pp. 304-305.  jji Proc. Int.
Conf. Ecol. Impact Acid Precip.   0. Drabltfs and A. Tollan, eds. SNSF
Project, Norway.

Fromm, P. 0.  1980.  A review of some physiological and toxicological
responses of freshwater fish  to  acid stress.  Environ. Biol. Fish.
5:79-93.

Fromm, P. 0.  1981.  Effect of acid stress on sodium transport by
isolated skins and on osmotic permeability of intact frogs.  Bull.
Environ. Contam. Toxicol. 27:160-166.

Fryer, G.  1980.  Acidity and species diversity in freshwater crustacean
faunas.  Freshwater Biology 10:41-45.

Gahnstrom, G., G. Andersson,  and S. Fleischer.  1980.  Decomposition and
exchange processes in acidified  lake sediment, pp. 306-307.  lr± Proc.
Int. Conf.  Inpact Acid Precip.  D. Drabltfs and A. Tollan, eds.  SNSF
Project, Norway.

Galloway, J. N. and E. B. Cowling.  1978.  The effects of precipitation
on aquatic and terrestrial  ecosystems - a proposed precipitation
chemistry network. J. Air  Poll. Cont. Assoc. 28(3) :229-235.

Galloway, J. N. and G. E. Likens. 1979.   Atmospheric enhancement of
metal deposition in Adirondack lake sediments.  Limnol.  Oceanogr.
24(3):427-433.
                                  5-174

-------
Garton, R. R. and R. C.  Ball.   1969.  Lime, fish and the soft-water bog
lake.  Michigan Academician  1:153-173.

Gates, F.  C.   1942.   The bogs  of  northern lower Michigan.  Ecological
Monographs 12:213-254.

Gaufin, A. R.  1973.  Water  quality  requirements of aquatic insects.
EPA-660/3-73—004, U.S.  Environmental Protection Agency, Washington,
D.C.  89pp.

Geering, H. R., E. E. Gary,  L.  H.  P. Jones, and W. H. Allaway.  1968.
Solubility and redox criteria  for the possible forms of selenium in
soils.  Proc. Soil Sci.  Amer.  32:35-40.

George, C. J., C. W. Boylen, and  R.  B. Sheldon.  1977.  The presence of
the red-spotted newt, Notophthalmus  viridescens Rafinesque (Amphibia,
Urodela, Salamandridae), in  waters exceeding 12 meters in Lake George,
New York.   J. Herpetol.  11(1):87-90.

Gilani, S. H. and M. Chatzinoff.   1981.  Aluminum poisoning and chick
embryogenesis.  Env. Res.  24:1-5.

Gissel-Nielsen, G.  1973.  Uptake and distribution of added selenite and
selenate by barley and red clover as influenced by sulphur.  J. Sci. Fd.
Agric. 24:649-655.

Gissel-Nielsen, G.  1975.  Selenium  concentration in Danish forage
crops.  Acta Agricul. Scand. 25:216-220.

Gjedrem, T.  1976.  Genetic  variation in tolerance of brown trout to
acid water.  SNSF-project, Norway, FR5/76.  11 pp.

Godfrey, W. E.  1966. The Birds  of  Canada.  Nat. Museums of Canada,
Bulletin 203.  428 pp.

Goldman, J.  1973.  Carbon dioxide and pH:  Effect on species succession
by algae.   Science 182:306-307.

Goodnight, C. J.   1973.   The use  of  aquatic macroinvertebrates as
indicators of stream pollution.   Trans. Am. Microsc. Soc. 92:1-12.

Gosner, K. L. and I. H.  Black.  1957.  The effects of acidity on the
development and hatching of  New Jersey frogs.  Ecology 38:256-262.

Grahn, 0.   1976.   Macrophyte succession in Swedish lakes caused by depo-
sition of airborne acid  substances,  pp. 519-530.  In Preceedings of the
International Symposium  on acid precipitation and "tFfe Forest Ecosystem
(1st), Columbus,  Ohio.   A. S.  Dochinger and T. A. Seliga, eds.
USDA Forest Service General  Technical Report NE-23, Upper Darby, PA.
                                  5-175

-------
Grahn, 0.  1977.  Macrophyte succession  in  Swedish  lakes caused by
deposition of airborne acid substances.   Water,  Air, and Soil Pollut.
7:295-305.

Grahn, 0.  1980.  Fish kills in two  moderately  acid lakes  due to  high
aluminum concentration, pp. 310-311.  Jhi Ecological  Impact of Acid
Precipitation.  D. DrabUs and A.  Tollan, eds.   SNSF-project, Oslo,
No rway.

Grahn, 0., H. Hultberg, and L. Landner.   1974.   01 igotrophication - A
self-accelerating process subjected  to  excessive supply of acid
substances.  Ambio 3:93-94.

Grande, M., I. P. Muniz, and S. Andersen.   1978. The  relative tolerance
of some salmonids to acid waters.  Verh.  Internat.  Verein. Limnol.
20:2076-2084.

Greeley, J. R. and S.  C.  Bishop.  1932.  Fishes of the area, pp. 54-76.
jji A Biological Survey of the Oswegatchee and Black  River  Systems.
Supplement to 21st Annual Report,  State  of  New  York  Conservation  Dept.,
J. B. Lyon Co., Albany, NY.

Green, A. and R. Root.  1933.  The Equilibrium  between hemoglobin and
oxygen in the blood of certain fishes.   Biol. Bull.  64:383-404.

Green, R. H. and G. L. Vascotto.  1978.   A  method for  the  analysis of
environmental factors controlling  spatial patterns  in  species
composition.  Water Research 12:583-590.

Griffiths, D.  1973.  The structure  of  an acid  moorland pond community.
Journal of Animal Ecology 42:263-283.

Guthrie, C. A.  1981.   Acclimation of brook trout (Salyelinus
fontinalis) to acidified waters.  M.S.,  Thesis,  Cornel 1 University,
Ithaca, NY.

Hagen, A. and A. Langeland.  1973.  Polluted snow in southern Norway
and the effect of the meltwater on freshwater and aquatic  organisms.
Environ. Pollut. 5:45-57.

Hagstrom, T.  1977.  The extinction  of  frogs in a lake acidified  by
atmospheric pollution.  Sveriges Natur.  11:367-369.

Haines, T. A.  1981 a.   Effects of  acid  rain on  Atlantic salmon rivers
and restoration efforts in the United States, pp. 57-63.   Jji Acid Rain
and the Atlantic Salmon.  IASF Spec. Publ.  10.

Haines, T. A.  1981b.   Acidic precipitation and its consequences  for
aquatic ecosystems:  A Review. Trans. Amer. Fish. So.  110:669-707.

Hakanson, L.  1980.  The Quantitative  impact of pH, bioproduction and
Hg-contamination on the Hg-content of fish  (pike).   Environ. Pollut.
61:285-304.
                                  5-176

-------
Hall, D., W. Cooper,  and E.  Werner.   1970.  An  experimental approach to
the production dynamics and  structure of  freshwater animal communities.
Limnol. Oceanogr. 15:839-928.

Hall, R. J. and G. E. Likens.  1980a.  Ecological effects  of experi-
mental   acidification on a stream  ecosystem, pp. 375-376.  In Ecol.
Impact Acid Precip.  D. Drabljos and A.  Toll an,  eds.  SNSF Project,
Sanderfjord, Norway.

Hall, R. J. and G. E. Likens.  1980b.  Ecological effects  of whole-stream
acidification, pp. 443-451.   In Atmospheric Sul fur Deposition,
Environmental Impact and Health" Effects.   D. S.  Shriner,  C. R. Richmond,
and S. E. Lindberg, eds.  Ann  Arbor Science Publishers,  Inc., Ann  Arbor,
MI.

Hall, R. J., G. G. Likens, S.  B. Fiance,  and G.  R. Hendrey.  1980.
Experimental acidification of a stream  in the Hubbard Brook Experimental
Forest, New Hampshire.  Ecology 61(4):976-989.

Hamilton, A. L.  1971.  Zoobenthos of fifteen lakes in the experimental
lakes area, northwestern Ontario.  Journal  of the Fisheries Research
Board of Canada 28:257-263.

Hanson, W. C., F. W.  Whicker,  and  J.  F. Lipscomb.  1975.  Lichen forage
ingestion rates of free-roaming caribou estiamted with fallout
cesium-137, pp. 71-79.  ^n Proc. First  Internat. Reindeer and Caribou
Symp. Luick, J. R., P. C. Lent, D. R. Klein, and R. G. White, eds.
Biological Papers of the Univ. Alaska,  Special  report 1,  Fairbanks.

Hargreaves, J. W., E. J. H.  Lloyd, and  B. A. Whitton.  1975.  Chemistry
and vegetation of highly acidic streams.   Freshwat. Biol. 5:563-576.

Harp, G. L. and R. S. Campbell. 1967.  The distributions of Tendipes
plumosus (Linne) in mineral  acid water.  Limnology and Oceanography
12:260-263.

Harr, J. R.  1978.  Biological effects  of selenium.  Chapt. 18 In
Toxicity of Heavy Metals in the Environment.  Part  I.  F. W. OeTime, ed.
M. Dekker, Inc., New York.

Harriman, R. and B. Morrison.   1980.   Ecology of acid streams draining
forested and non-forsted catchments  in  Scotland, p. 312-313.  D.
Drablj6s and A To!Ian, eds.  Proc., Int. Conf. Ecol.  Impact Acid
Precip., Norway 1980, SNSF project.

Harriman, R. and B. Morrison.   1982.   Ecology of streams draining
forested and non-forested catchments  in an area of  central Scotland
subject to acid precipitation.  Hydrobiol. 88:251-263.

Hart, C. W. Jr., and S. L. Fuller. 1974.  Pollution Ecology of
Freshwater Invertebrates.  Academic Press, New  York, NY.
                                  5-177

-------
Harvey, H. H.  1975.   Fish populations  in  a  large group of acid-stressed
lakes.  Verh. Internat.  Verein.  Limnol.  19:2406-2417.

Harvey, H. H.  1979.   The  acid  deposition  problem and emerging research
needs in the toxicology  of fishes.   Proc.  Fifth Ann. Aquatic Toxicity
Workship, Hamilton, Ontario,  Nov. 7-9,  1978.   Fish, Mar. Serv. Tech.
Rep. 862:115-128.

Harvey, H. H.  1980.   Widespread and diverse changes in the biota of
North American lakes  and rivers coincident with acidification, pp.
93-98.  D. Drabljos and A.  Tollan, eds.   Proc., Int. Conf. Ecol. Impact
Acid Precip., Norway  1980, SNSF project.

Harvey, H. and C. Lee.  1980.  Fishes of the LaCloche Mountain lakes of
Ontario, 1965-1980.  Report to  the  Ontario Ministry of Natural Resources
(cited in NRCC 1981).

Harvey, H. and C. Lee.  1982.  Historical  fisheries changes related to
surface water pH changes in Canada,  pp.  45-56. In Acid Rain/Fisheries,
Amer. Fisheries Soc., Bethesda, MD.

Harvey, H. P. Dillon, G. Fraser, K.  Somers,  P. Fraser, and C. Lee.
1982.  Elevated metals and enhanced metal  uptake  in fishes in
acid-stressed waters.  Abstract, 185th National Meeting, American
Chemical Society 22(1):438-441

Hastings, R. W.  1979.  Fish of the Pine Barrens.  In Pine Barrens:
Ecosystem and Landscape.  R.  T. T.  Forman, ed. Acacfemic Press, New
York, pp. 489-504.

Havas, M.  1981.  Physiological response of  aquatic animals to low pH.
pp. 49-66 Ir± Effects of Acidic  Precipitation on Benthos.  Proc. of a
Regional Symposium on Benthic Biology.  R. Singer, ed.  North American
Benthological Society, Hamilton, NY.

Havas, M. and T. C. Hutchinson.  1982.  Aquatic invertebrates from the
Smoking Hills, N.W.T.:  effect of pH and metals on mortality.  Can. J.
Fish.  Aquatic Sci. 39:890-903.

Heinselman, M. L.  1970.  Landscape evolution peat!and types and the
environment  in the Lake Agasslz Peatlands  Nature  Area, Minnesota.
Ecological Monographs 40:235-261.

Hendrey, G.  1976.  Effects of low pH on the growth of periphytic algae
in  artificial stream channels.   IR 2576, SNSF Project, Oslo, Norway.

Hendrey, G.  1980.  Effects of acidity on  primary productivity in lakes:
phytoplankton, pp. 357-371.  In Atmospheric  Sulfur Deposition-
Environmental Impact and HealTfi" Effects.  D. Shriner,  C.  Richmond, and
S.  Lindberg, eds.  Ann Arbor Science, Ann  Arbor,  MI.
                                  5-178

-------
Hendrey, 6. R. and F. W. Barvenik.   1978.   Impacts  of  acid  precipitation
on decomposition and plant communities in  lakes,  pp. 92-103.  Jji
Scientific Papers from the Public Meeting  on Acid Precipitation, Science
and Technology Staff, New York State Assembly.  H.  H.  Izard and J. S.
Jacobson, eds.  Center for Environmental Research,  Cornell  University,
Ithaca, NY.

Hendrey, G. R. and F. A. Vertucci.   1980.   Benthic  Plant  Communities in
Acidic Lake Colden New York:   Sphagnum and the  Algal Mat.   In Proceeding
of the International Conference on  Ecological Impact of Aci'lT"
Precipitation.  D. Drabljos and A. Tollan,  eds.  SNSF project, pp.
314-315.

Hendrey, G. R. and R. F. Wright.  1976.  Acid precipitation in Norway:
Effects on aquatic fauna.  J.  Great Lakes  Res.  2  (suppl.  1)-.192-207.

Hendrey, G. R., J. N. Galloway, and C. L.  Schofield. 1980a.  Temporal
and spatial trends in the chemistry of acidified  lakes under ice
cover, pp. 266-267.  ^ni Proc.  Int.  Conf. Impact Acid Precip., Norway.
D. Drabltfs and A. Tollan, eds.  SNSF Project, Norway.

Hendrey, G. R., N. D. Yan, and K. J. Baumgartner.   1980b.   Response of
freshwater plants and invertebrates to acidification.  Jji Proc.
Internat. Symp. for Inland Waters and Lake Restoration, Portland, Maine,
Sept. 8-12, 1980.  EPA/DECD.

Hendrey, G. R., K. Baalsrud,  T. S.  Traaen,  M. Laake, and G. Raddum.
1976.  Acid precipitation:  Some hydrobiological  changes.   Ambio 5:
224-227.

Henriksen, A. and R. Wright.   1978.  Concentrations of heavy metals in
small Norwegian lakes.  Water Res.  12:101-112.

Henrikson, L. and H. G. Oscarson.  1978.   Fish  predation limiting
abundance and distribution of Glaenocorisa p. propinqua.  Oikos
31:102-105.                   	 	

Henriksen, L., H. Nyman, H. Oscarson, and  J.  Stenson.  1980a.  Trophic
changes, without changes in the external nutrient loading.
Hydrobiologia 68:257-263.

Henrikson, L., H. Nyman, G. Oscarson, and  J. A. E.  Stenson.  1980b.
Does the change of predator system  contribute to  the biotic development
in acidified lakes? pp. 316-317. Jji Proc.  Int. Conf.  Impact Acid
Precip. D. Drablj6s and A. Toll an, eds. SNSF Project,  Norway.

Hidiroglou, M., R. B. Carson,  and G. A. Brossard.   1965.  Influence of
selenium on the selenium contents of hair  and on  the incidence of
nutritional muscular disease in beef cattle.  Can.  J.  Anim. Sci.
45:197-202.
                                  5-179

-------
Hobaek, A. and G. Raddum.   1980.   Zooplankton  communities in acidified
lakes in southern Norway.   Internal  Report IR  75/80, SNSF-Project, Oslo,
Norway.

Hoglund, L. B.  1961.  The reactions of fish in  concentration gradients.
Chapter 3, Reaction to pH  and  C02  Report No. 43; pp. 95-147.  Rep.
Inst. Freshw. Res., Drottningholm.

Hopkins, E. F. and F. B. Wann.   1926.   Relation  of hydrogen-ion
concentration to growth of Chi ore!la and to the  availability of iron.
Bot. Gaz. 81:353-376.

Howmiller, R.  1977.  On the abundance of Tubificidae  (Annelida:
Oligochaeta) in the profundal  benthos  of some  Wisconsin lakes.  Am.
Midi. Nat. 97:211-216.

Huckabee, J. W., C. P. Goodyear, and R.  D.  Jones.  1975.  Acid rock in
the Great Smokies:   Unanticipated  impact on aquatic biota of road
construction in regions of sulfide mineralization.  Trans. Am. Fish.
Soc. 104(4):677-684.

Hulsman, P. and P.  Powles.  1981.  In  situ  effects of  low pH on
incubation and early life  stages of  walleye, Stizostedion vitreum
yitreum.  Unpublished report to  the  Ontario Ministry of Natural
"Resources, Toronto, Ontario (cited in  U.S./Canada MOI  1982).

Hultberg, H.  1977.  Thermally  stratified acid water in late winter - a
key factor inducing self oligotrophication  processes which increase
acidification.  Water, Air, and  Soil  Pollut. 7:279-294.

Hultberg, H. and I. Andersson.   1982.   Liming  of acidified lakes and
streams - aspects on induced physical-chemical and biological changes.
Water, Air, and Soil Pollut. 18:311-331.

Hultberg, H. and 0. Grahn.  1976.  Effects of  acid precipitation on
macrophytes in oligotrophic Swedish  lakes.  In Atmospheric Contribution
to the Chemistry of Lake Waters.   Proc.  FirsT~Specialty Symp., Internat.
Assoc. Great Lakes  Research, Sept. 28-Oct.  1,  1975.  J. Great Lakes Res.
2(Suppl. 1):208-217.

Hummon, M. R. and W. D. Hummon.  1979.   Reduction in fitness of the
gastrotrich Lepidodermella squammata by  dilute acid mine water and
amelioration or tne errect oy  caroonates.   Int.  J. Invert. Repro.
1:297-306.

Hutchinson, G. E.  1957.  A treatise on  limnology.  I.  Geography,
Physics and Chemistry.  J. Wiley & Sons,  New York.  1015 p.

Hutchinson,- G. E.  1967.  A treatise on  limnology.  II.  Introduction in
lake biology and the 1imnoplankton.  J.  Wiley  &  Sons,  New York.  1115 p.
                                  5-180

-------
Hutchinson, G. E.  1975.   A treatise on  limnology.   III.  Limnological
Botany.  John Wiley & Sons, New York.  660  p.

Hutchinson, T. C., W. Gizyn,  M. Havas  and V. Zobens.  1978.  Effect of
long-term ligmite burns on artic ecosystems at  the Smoking Hills N.W.T.
In Trace Substances in Environmental  Health-XII.  D. D. Hemphill, ed.
Wiversity of Missouri, Columbia.

Jacobsen, 0. J.  1977.  Brown trout (Salmo  trutta L.) growth at reduced
pH.  Aquaculture 11:81-84.

Janicki, A. and J. DeCosta.  1979.   A  Multivariate analysis of the
crustacean plankton community of an acid reservoir.  Arch. Hydrobio.
85:465-481.

Janzen, D. H.  1974.   Tropical  blackwater rivers, animals and mast
fruiting by the Dipterocarpacea.  Biotropica 6:69-103.
Jensen, K. W. and E.  Snekvik. 1972.  Low pH levels wipe out salmon and
trout populations in southernmost Norway.   Ambio 1:223-225.

Jensen, K. W. and E.  Snekvik.  1972.   Low pH levels  wipe out salmon and
trout populations in  southernmost Norway.   Ambio 1:223-225.

Jernelov, A.  1980.  The effects of acidity on  the uptake of mercury in
fish, pp. 211-217.  In Polluted Rain.  T. Y. Toribara, M. W. Miller, and
P. E. Morrow, ed.  Plenum Press, NY,  London.

Jewell, M. E.  1939.   An ecological study of the freshwater sponges of
Wisconsin, II.  The influence of calcium.   Ecology 20:11-28.

Jewell, M. E. and H.  W. Brown.   1924.  The  fishes of an acid lake.
Trans, of the Amer. Micro. Soc. 43:77-84.

Jewell, M. E. and H.  W. Brown.   1929.  Studies  on northern Michigan bog
lakes.  Ecology 10:427-475.

Johansson, N. andJ.  E. Kihlstrom.   1975.   Pikes (Esox lucius L.) shown
to be affected by low pH values during first weeks after hatching.
Environ. Res. 9:12-17.

Johansson, N. and G.  Milbrink.   1976.  Some effects  of acidified water
on the early development of roach (Rtitilus  rutilus L.) and perch (Perca
fluviatilis L.).  Water Resour. Bull.  12:39^4~B"T

Johansson, N., J. E.  Kihlstrom, and A. Wahlberg.  1973.  Low pH values
shown to affect developing fish eggs (Brachydam'o rerio Ham. - Buch.)
Ambio 2:42-43.

Johansson, N., P. Runn, and G.  Milbrink. 1977. Early development of
three salmonid species in acidified water.  Zoon 5:127-132.
                                  5-181

-------
Johnson, C. M.  1975.   Selenium  in  soils and plants.  Contrasts in
conditions providing safe  but  adequate amounts of selenium in the food
chain, pp. 165-180.   In Trace  Elements in Soil-Plant-Animal Systems.  D.
J. D. Nicholas and A.T. Egan, eds.  Academic Press.  New York.

Johnson, D. W.  1975.   Spawning  Behavior and Strain Tolerance of Brook
Trout (Salvelinus fontinalis)  in Acidified Water.  M.S. thesis, Cornell
University, Ithaca,  NY.

Johnson, D. W.  1981.   The natural  acidity of some unpolluted waters in
southeastern Alaska  and Potential  impacts of acid rain.  Water, Air and
Soil Pollution 16:243-252.

Johnson, D. W. and D.  A. Webster.   1977.  Avoidance of low pH in
selection of spawning sites by brook trout (Salvelinus fontinalis).  J.
Fish. Res. Board Can.  34:2215-2218.

Johnson, M. G., M. Michalski,  and A. Christie.  1970.  Effects of acid
mine wastes on phytoplankton communities of two northern Ontario lakes.
J. Fish. Res. Board Can. 27:425-444.

Jones, J. R. E.  1948.  A  further study of the reactions of fish to
toxic solutions.  J. Exp.  Biol.  25:22-34.

Jop, K.  1979.  Quality evaluation of  roe-deer antlers from an
industrial region in southern Poland.  Acta Theriol. 24:23-34.

Jordan, M. K. and G. E. Likens.   1975.  An organic carbon budget for an
oligotrophic  lake in New Hampshire, U.S.A.  Verh. Internat. Verein.
Limnol. 19:994-1003.

Jozuka, K. and H. Adachi.   1979.  Environmental physiology on the pH
tolerance of  the teleost.  2.  Blood properties of medaka, Oryzias
latipes, exposed to low pH environment.  Annot. Zool. Jap. bZ:iu/-113.

Kalman, S.  1959.  Sodium  and water exchange  in the  trout egg.  J.  Cell.
Comp. Physiol. 54:155-162.

Keller, W., J. Gunn, and N. Conroy.  1980.  Acidification  impacts on
lakes in the  Sudbury, Ontario, Canada  area, pp. 228-229.  D. Drablrfs
and  A.  Tollan, eds.  Proc, Int.  Conf.  Ecol.  Impact  Acid Precip., Norway
1980, SNSF project.

King, G. M.,  T. Berman, and W. J. Wiebe.   1981.   Methane  formation  in
the  acidic peats of Okefenokee Swamp,  Georgia.  American  Midland
Naturalist 105:386-389.

Kirk, P. W.,  Jr.  (ed.).   1979.  The Great  Dismal  Swamp.   Old Dominion
University Research Foundation,  Inc.,  427  p.
                                  5-182

-------
Kitchell, J. F., R. V.  O'Neill,  D.  Webb, G. W. Gallepp, S. M. Bartell,
J. F. Koonce, and B. S.  Ausmus.   1979.  Consumer regulation of nutrient
cycling.  Boscience 29:28-34.

Kortmann, R.  1980.  Benthic and atmospheric contributions to the
nutrient budgets of a soft-water lake.  Limnol. Oceanogr. 25:229-239.

Kramer, D. L., C. C. Lindsey,  G. E.  E. Moodie, and E. 0. Stevens.  1978.
The fishes and the aquatic environment of  the  central Amazon basin, with
particular reference to  respiratory  patterns.  Can. J. Zool. 56:717-729.

Kubota, J., W. H. Allaway, D.  L. Carter, E. E. Cary, and V. A. Lazar.
1967.  Selenium in crops in the  U.S.  in relation to selenium responsive
diseases of animals.  J. Agric.  Food  Chem. 15(3):448-453.

Kwain, W.  1975.  Effects of temperature on development and survival of
rainbow trout, Sal mo gairdneri,  in  acid waters.  J. Fish. Res. Board
Can. 32:493-497"!

Kwiatkowski, R. and J.  Roff.   1976.   Effects of acidity on the
phytopi ankton and primary productivity of  selected northern Ontario
lakes.  Can. J. Bot.  54:2546-2561.

Laake, M.  1976.  Effekter av  lav pH  pa produksjon, nedbrytning og
stoffkretslop i littoralsonen.   SNSF  project,  Internal Report
29/76, Oslo--As, Norway.

Laerm, J., B. J. Freeman, L.  J.  Vitt, J. P. Meyers, and L. Logan.  1980.
Vertebrates of the Okefenoke Swamp.   Brimleyana 4:47-73.

Lazarek, S.  1980.   Cyanophycean mat  communities in acidified lakes.
Naturwissenschaften 67:97-98.

Lazarek,  S.  1982.  Structure and  productivity of epiphytic algal
communities on Lobelia  dartmanna L.  in acidified and limed lakes.
Water, Air & Soil Pollut. 18:333-342.

Lechowicz, M. J.  1982.   The effects  of simulated acid precipitation on
photosynthesis in the caribou  lichen  Cladina stellaris (Opiz) Brodo.
Water Air Soil Pollut.  18:421-430.

Lee, R. M. and S. D. Gerking.  1981.  Survival and reproductive
performance of the desert pupfish,  Cyprinodon  N. nevadensis in acid
waters.  J. Fish Biol.  17:507-515.

Leivestad, H. and I. P.  Muniz.   1976.  Fish kill at low pH in a
Norwegian river.  Nature 259:391-392.

Leivestad, H., G. Hendrey, I.  P. Muniz, and E. Snekvik.  1976.  Effects
of acid precipitation on freshwater organixms, pp. 87-111.  J[n Impact of
acid precipitation on forest and freshwater ecosystems in Norway, F. H.
Braekke, ed. SNSF Project, Norway FR  6/76.
                                 5-183

-------
Leivestad, H., I.  P.  Muniz,  and B. 0. Rosseland.  1980.  Acid stress in
trout from a dilute mountain stream, pp. 318-319.   D. Drabltfs and A.
To!Ian, eds.  Proc.,  Int.  conf. ecol. impact acid precip., Norway 1980,
SNSF project.

Levesque, M.  1974.  Some  aspects of selenium relationships in Eastern
Canadian soild and plants.   Can. J. Soil Sci. 54:205-214.

Lewis, M.  1976.  Effects  of low concentrations of manganese sulfate on
eggs and fry of rainbow trout.  Prog. Fish. Cult. 38:63-65.

Lind, 0. and R. Campbell.  1970.  Community metabolism in acid and
alkaline strip-mine lakes.   Trans. Amer. Fish. Soc.  99:577-582.

Lloyd, R. and D. H. M. Jordan.  1964.   Some factors affecting the
resistance of rainbow trout  (Salmo guirdneri, Richardson) to acid
waters.  Int. J. Air Wat.  Poll. 8: 393-403.

Lockhart, W. L. and A. Lutz.  1977.  Preliminary biochemical
observations of fishes inhabiting an acidified lake in Ontario, Canada.
Int. J. Water Air Soil Pollut.  7:317-332.

Loefer, J. B.  1973.   Carbon dioxide and pH: Effect on species
succession of algae.   Science 182:306-307.

Lowe, R. L.  1974.  Environmental requirements and pollution tolerance
of freshwater diatoms.  EPA-670/4-74-005,  U. S. Environment! Protection
Agency, Cincinnati, Ohio.

Lowndes, A.  1952.  Hydrogen ion concentration and the distribution of
freshwater Entomostraca.  Ann Mag. Nat. Hist.  5:58-65.

Lynch, M.  1979.  Predation, competition,  and zooplankton community
structure:  An experimental  study.  Limnol. Oceanogr.  24:253-272.

Lynch, M.  1980.  The evolution of cladoceran life histories.  Quat.
Rev. Biol.  55:23-42.

Lynch, M. andJ. Shapiro.   1981.  Predation, enrichment,  and
phytoplankton community structure.  Limnol. Oceanogr.  26:86-102.

Maetz, J.  1973.  Ma+/MH4+,  Na+/H+ exchanges and NH^ movement across
the  gill of Carassius auratus.   J.  Exp. Biol. 58:255-275.

Makarewicz, J. and G. Likens.  1979.   Structure  and  function of the
zooplankton community of Mirror Lake,  New  Hampshire. Ecol.  Monogr.
49:109-127.

Malley, D. F.  1980.  Decreased survival  and calcium uptake by the
crafish Orconectes virilis in low  pH.   Can. J.  Fish. Aquatic Sci.
37:364-372.
                                  5-184

-------
Malley, D., D.  Findlay,  and P.  Chang.   1982.   Ecological effects of acid
precipitation on zooplankton,  pp.  297-327.   I_n Acid Precipitation
Effects on Ecological  Systems.   F. D'Intri,  ed.  Ann Arbor Science, Ann
Arbor, MI.

Marmorek, D.  1983.   Changes in the temporal behavior and size structure
of plankton systems  in acid lakes.  Am.  Chem.  Soc. In press.

Marshall, H. G.  1979.  Lake Drummond:   with a discussion regarding its
plankton composition.   In The Great Dismal Swamp.  P. W. Kirk, Jr., ed.
University Press of  Virginia,  Charlottesville, 427 p.

Mashiko, K., K. Jozuka and K.  Asakura.   1973.  Different type of
chloride cells in the  gills of Tribolodon hakonensis from Lake
Osoresan-ko. Ann. Rep. Noto. Marine Lab 13:33-37.  In Exceptional
Tolerance to Low pH  of Some Tropical  Blackwater  FisTiT  W. A. Dunson, F.
Swarts, and M. Silvestri, eds.   J  of Exp. Zoology 201:157-162.

Matthias, U.  1982.   Der Einfluss  der Wasserstoffionekonzentration auf
die Zusammensetzung  von  Bergbachbiozonosen,  dargestellt an einigen
Mittelbebirgsbachen  des Kaufunger  Waldes (Nordhessen/Sudniedersachsen).
PhD thesis.  Universitat des Landes Hessen.  Kassel, West Germany.  135
pp.

Mattson, P., L. Albanus, and A. Frank.   1981.  Cadmium and some other
elements in liver and  kidney from  moose (Alces alces).  Var  Foda 8-9:
335-345.

McCauley, E. and F.  Briand.  1979.  Zooplankton  grazing and
phytoplankton species  richness:  Field tests of  the predation
hypothesis.  Limnol.  Oceanogr. 24:243-252.

McDonald, D. and C.  Wood.  1982.  Branchial  and  renal  fluxes in the
rainbow trout, Salmo gairdneri. at low environmental pH.  J. Exp. Biol.
93:101-118.

McDonald, D. G., H.  Hobe, and C. M. Wood.   1980. The  influence of
calcium on the physiological responses of the  rainbow  trout, Sal mo
gairdneri, to low environmental pH.  J. Exp. Biol. 88:109-131.

Mclntyre, J. W.  1979.  Status of common loons in New  York  from an
historical perspective,  pp. 117-121.  In The Common Loon.  S. A.
Sutcliffe, ed.  Proc.  Second N. A. ConTT Common  Loon Research and
Management.  National  Audobon Society, New  York. 162 pp.

McLachlan, A. J. and S.  M. McLachlan.  1975.  The physical environment
and bottom fauna of a bog lake.  Archives fur  Hydrobiologie  76:198-217.

McWilliams, P.  1982.   The Effects of calcium  on sodium fluxes  in the
brown trout, Salmo trutta, in neutral and acid water.  J.  Exp. Biol
96:439-442.
                                  5-185
409-262 0-83-18

-------
McWilliams, P.  G.   1980a.   Effects  of  pH  on  sodium uptake on Norwegian
brown trout (Salmo trutta).  J.  Exp.  Biol. 88:259-267.

McWilliams, P.  G., 1980b.   Acclimation to an acid medium in the brown
trout (Salmo trutta).   J.  Exp.  Biol. 88:269-280.

McWilliams, P.  G.  and W.  T.  W.  Potts.   1978.  The effects of pH and
calcium concentrations on  gill  potentials in the brown trout, Salmo
trutta.  J. Comp.  Physiol.   126:277-286.

McWilliams, P.  G., D.  J.  A.  Brown,  G.  D.  Howells, and W. T. W. Potts.
1980.  Physiology  of fish  in acid waters, pp. 282-283.  D. Drabltfs and
A. Tollan, eds.  Proc., Int. conf.  ecol.  impact acid precip., Norway
1980, SMSF project.

Menendez, R.  1976.  Chronic effects of reduced pH on brook trout
(Salvelinus fontinalis).   J. Fish.  Res. Board Can. 33:118-123.

Merritt, R. W.  and K.  W.  Cummins.   1978.  An Introduction to the Aquatic
Insects of North America.   Kendall/Hunt Publishing Co., Dubuque, Iowa.

Milbrink, G. and N. Johansson.   1975.   Some  effects of acidification on
roe of roach, Rutilus rutilus L., and  perch, Perca fluiatilis L. - with
special reference  to the  Avaa Lake  system in eastern Sweden. Rep. Inst.
Freshw. Res. Orottningholm 54:52-62.

Miller, G. E.,  I.  Wile, and G.  G. Hitchin.   1982.  Patterns of
accumulation of selected  metals in  members of the soft-water macrophyte
flora of central Ontario lakes.  Aquat. Bot. 15:53-64.

Mills, K. H.  1982.  Fish population responses during the experimental
acidificatoin of a small  lake.   Presented at American Chemical Society
Meeting, Las Vegas, NV.  March.

Mitchell, M. J., D. H. Landers, and D. F. Brodowski. 1981a.  Sulfur
constituents of sediments and their relationship to lake acidification.
Water, Air, Soil Pollut.  16:351-359.

Mitchell, M. J., G. B. Lawrence, D. H. Landers, and K. A. Stucker.
1981b.  Role of benthic invertebrates  Hexagenia, in affecting sulfur and
nitrogen dynamics of lake sediments, pp.  67-76.  ^n Effects of Acid
precipitation on Benthos.   R. Singer,  ed. Proceedings of a Regional
Symposium on Biology, North American Benthological Society, Hamilton,
NY.

Moeller, R. E.   1975.  Hydrophyte biomass and community structure in a
small, oligotrophic New Hampshire Lake.  Verh.  Internet. Verein. Limnol.
19:1004-1012.

Moeller, R. E.   1978.  Seasonal changes in biomass, tissue chemistry,
and net production of the evergreen hydrophyte, Lobelia dortmanna.  Can.
J. Bot. 56:1425-1433.
                                  5-186

-------
Moore, P. D. and D.  J.  Bellamy.   1974.   Peatlands.  Elek Science,
London, 221 p.

Mortimer, C. H.   1941.   The  exchange of  dissolved substances between mud
and water in lakes (Parts I  and  II).  J.  Ecol. 29:280-329.

Mortimer, C. H.   1942.   The  exchange of  dissolved substances between
mud and water in lakes  (Parts  III  and V).  J. Ecol. 30:147-201.

Moss, B.  1973.   The influence of environmental  factors on the
distribution of freshwater algae:   An experimental  study.  II. The role
of pH and the carbon dioxide-bicarbonate system.  J. Ecol. 61:157-177.

Mount, D. I. 1973.  Chronic  effect of low pH on  rathead minnow survival,
growth and reproduction.  Water  Res. 7:987-993.

Muller, P.  1980.  Effects of  artificial  acidification on the growth of
periphyton.  Can J.  Fish. Aquat. Sci.  37:355-363.

Muniz, I. P.  1981.   Acidification and the Norwegian salmon, pp. 65-72.
j_n Acid Rain and the Atlantic  Salmon.  IASF Spec. Publ. 10.

Muniz, I. P. and H.  Leivestad.  1980a.   Acidification  effects on
freshwater fish, pp. 84-92.  D.  Drabltfs  and A. Tollan, eds.  Proc.,
Int.  Conf. Ecol. Impact Acid Precip., Norway 1980,  SNSF project.

Muniz, I. P. and H.  Leivestad.  1980b.   Toxic  effects  of  aluminum on the
brown trout, Salmo trutta L, pp. 320-321. D.  Drabltfs  and A.  Tollan,
eds.  Proc., Int. Conf. Ecol.  Impact Acid Precip.,  Norway 1980, SNSF
project.

Mushinsky, G. R. and E. D. Brodie, Jr.   1975.  Selection  of  substrate pH
by salananders.  Am. Midi. Nat.  93(2):440-443.

Muth, 0. M. and W. H. Allaway.  1963.  The relationship of White Muscle
Disease to the distribution of naturally occurring  selenium.  J. Am.
Vet.  Med.  Assoc. 142:1379-1384.

Muth, 0. H., J. E. Oldfield, L.  F. Remmert,  and  J.  R.  Schubert.  1958.
Effects of selenium and vitamin E on white muscle  disease.   Science
128:1090.

National Research Council Canada.  1981.  Acidification  in the  Canadian
aquatic environment:  Scientific criteria for  assessing the  effects of
acidic  depostion on aquatic ecosystems.   NRCC  18475,  Environmental
Secretariat, Ottawa, Canada.

Needham, J. G. and J. T. Lloyd.  1916.   The Life of Inland Waters.
Comstock Publishing Company,  Ithaca, NY.  438  pp.

Nelson, J.  1982.  Physiological observations  on developing  rainbow
trout,  Salmo gairdneri  (Richardson), exposed to  low pH and varied
calcium  ion concentrations. J. Fish. Biol. 20:359-372.
                                  5-187

-------
Neville, C. M.  1979a.   Sublethal  effects  of environmental acidification
on rainbow trout (Salmo gairdneri).  J.  Fish. Res.  Board Can.
36:84-87.

Neville, C. M.  1979b.   Ventilatory  response of rainbow trout (Sal mo
gairdneri) to increased H+  ion  concentration in blood and water.
uomp. Biochem. Physiol. 63A:373-376.

New York State Department of Environmental Conservation.  1982.
Clarification of issues raised  concerning  the effects of acid
precipitation on fish  resources in the  Adirondack  Region of New York
State.  Testimony presented before the  Senate Committee on Environment
and Public Works, May  27, 1982.

Newman, J. R.  1979.   Effects of industrial air pollution on wildlife.
Biol. Conserv. 15:81-190.

Nilssen, J. P.  1980.   Acidification of a  small watershed in southern
Norway and some characteristics of acidic  aquatic  environments.   Int.
Revue ges. Hydrobiol.   65:177-207.

Nilssen, N. and B. Pejler.   1973.   On the  relation between fish fauna
and zooplankton composition in  north Swedish lakes.  Rep. Inst.
Freshwater Res. Drottningholm,  Sweden.   53:51-74.

Nilsson, N. A.  1972.   Effects  of introductions of salmonids into barren
lakes.  J. Fish. Res.  Board Can.  29:693-697.

Nilsson, S. G. and I.  N. Nilsson.   1978.  Breeding bird community
densities and species  richness  in  lakes.  Oikos 31:214-221.

Noble, K. M.  1979.  The effect of p'ri stress on developing Ran a pi pi ens
and the effect of calcium ions  in relieving pH stress.  Honours B.Sc.
thesis, Univ. of Toronto.  9 pp.

Norton, S. A. and C.  T. Hess.  1980.    Atmospheric deposition in  Norway
during the last 300 years as recorded in SNSF lake sediments I.
Sediment dating and chemical  stratigraphy,  pp. 268-269.  Ir± Ecological
Impact of Acid Precipitation.  D.  Drablos  and A. Tollan, eds.  SNSF
rreject, Sandefjord,  Norway.

Norton, S. A., R. B.  Davis, and D. F. Brakke.  1981.  Responses of
northern New  England lakes to  atmospheric inputs  of acids and heavy
metals.  Land and Water Resources Center,  Project  A-048-ME.  Univ. of
Maine at Ocone, 82 pp.

Nyholm, N. E. I.  1981.  Evidence of involvement of aluminum in
causation of defective formation of eggshells and  of impaired breeding
in wild passerine birds.  Env.  Res.  26:363—371.

Nyholm, N. E. I. and H. E.  Myhrberg.  1977.  Severe eggshell defects and
impaired reproductive  capacity  in small  passerines in Swedish Lapland.
Oikos 29:336-341.
                                  5-188

-------
O'Brien, J.  1979.  The Predator-prey  interaction of planktivorous fish
and zooplankton.   Amer. Sci.   67:572-581.

0kland, J.  1969a. Om forsuring  av  vassdrag og betydningen av
surhetsgraden (pH) for fishkens  naeringsdyr i ferskvann. Fauna (Oslo)
22:104-174.

0kland, J.  1969b.  Distribution and ecology of  the fresh-water snails
(Gastropoda) of Norway.  Malacologia 9(1):143-151.

0kland, J.  1976.  Uttbredelsen  av  noen  ferskvannsmuslinger  i Norge,
og littom European Invertebrate  Survey.   Fauna (Oslo)  29:29-40.

0kland, J.  1979a.  Kalkinnhold, surhetsgrad (pH) og snegler i norske
innsjoer.  Fauna (Oslo) 32-96-111.

0kland, J.  1979b. Environment and  snails (Gastropoda):  Studies of
1000 lakes in Norway, pp. 322-323.   In Ecological Impact of  Acid
Precipitation, D. Drabljbs and A. TolTan,  eds.  SNSF Project,
Sanderfjord, Norway.

0kland, J.  1980.  Environment and  snails (Gastropoda), pp.  322-323.
lr± Proc.  Int. Conf. Ecol. Impact Acid  Precipitation.   D. Drablj6s and
ATlollan, eds.  SNSF Project, Norway.

0kland, J. and K. A. 0kland.   1978. Use of fresh-water littoral
fauna for environmental monitoring: Aspects related to studies of 1000
lakes in Norway.  In Preceedings from  a  conference arranged  by the
Swedish Society OiTos and the Nordic Council for Ecology on  "The Use of
Ecological Variables in Environmental  Monitoring",  Swedish  University
of Agricultural Sciences, Uppsala,  Sweden  (SNSF Contribution FA 37/78).

0kland, J. and K. A. 0kland.   1980.   pH level and food organisms
for Studies of 1000 lakes in  Norway, pp.  326-327.  Jji  Ecological
Impacts   of Acid Precipitation,   D. Drablj6s and  A. Tollan, eds.  SNSF
Project,  Sanderfjord, Norway.

0kland, K. A.  1971. On the ecology of Sphaeriidae in  a high mountain
area in south Norway.  Norway.  J.  Zool.   19:133-143.

0kland, K. A.  1979a.  Localities with Asellus aquaticus  (L) and
Gammarus  lacustric G. 0. Sars  in Norway, and  a  revised system of
faunistic  regions.  Jji^ SNSF Project,  Technical Note 49/79, Oslow-As,
No rway.

0kland, K. A.  1979b.  A project for studying  ecological  requirements
and for   mapping  distribution patterns of small  bivalvis   (Sphaeriidae)
and some  Crustacea (Asellus,  Gammarus) in fresh  water  in  Norway.  C. R.
Soc. Biogeog. (Mem. 3E Ser.,  No. 1) 1979:37-43.
                                  5-189

-------
0kland, K.  A.  1979c.  Sphaeriidae  of  Norway:  A project for studying
ecological  requirements  and  geographical  distribution.  Malacologia
18:223-226.

0kland, K.  A. 1980a.   Ecology  and  distribution of Asellus aquaticus
(L.) in Norway, including  relation to acidification in lakes.  Ir±
SNSF Project, Internal Report  52/80,  Oslo-As, Norway.

0kland, K.  A.  1980b.  Mussels and crustaceans:  Studies of 1000 lakes
Norway, pp. 324-325.  Jji Ecological Impact  of Acid Precipitation,  D.
Drabljos and A. Toll an, eds.  SNSF  Project,  Sanderfjord, Norway.

0kland, K.  A. 1980c.   Okologi  og utbredelse til Gammarus lacustric G.
0. Sars i Norge, med  vekt  pa forsuringsproblemer.  SMSF Project,
Internal Report 67/80, Oslo-As, Norway.

0kland, K.  A. and J.  G.  J. Kuiper. 1980.   Distribution of small
mussels (Sphaeriidae)  in Norway, with notes on their ecology.  Haliotis
10(2):109.

Orser, P. N. and D. K. Shure.   1972.  Effects of urbanization on the
salamander, Desmognathus fuscus fuscus.   Ecology 53:1148-1154.

Overrein, L., H. Seip.,  and  A. Toll an.   1980.  Acid precipitation -
effects on forest and fish.  Final Report of the SNSF-project 19/2-1980.

Packer, R.  K.  1979.   Acid-base balance  and gas exchange in brook trout
(Salvelj.nus fontfnalis)  exposed to acidic environments.  J. Exp.
Biol. 79:127-134.

Packer, R.  K. and W.  A.  Dunson.  1970.   Effects of low environmental pH
in blood pH and sodium balance of  brook  trout.  J.  Exp. Zoo!.
174:65-72.

Packer, R.  K. and W.  A.  Dunson.  1972.   Anoxia and sodium loss
associated  with the death  of brook trout at low pH.  Comp. Biochem.
Physiol. 41A:17-26.

Parent, S.  and R. Cheetham.  1980. Effects of acid precipitation on
Daphnia magna.  Bull.  Environ. Contam. Toxicol. 25:298-304.

Parsons, J. D.  1968.  The effects of acid  strip-mine effluents on the
ecology of a stream.   Arch.  Hydrobiol. 65:25-50.

Parsons, J. D.  1977.   Effects of  acid mine wastes on aquatic
ecosystems.  Water, Air, and Soil  Pollution 7:333-354.

Patalas, K.  1971. Crustacean plankton  communities in forty-five lakes
in the Experimental Lakes  Area, northwestern Ontario.  J. Fish. Res.
Board Can.  28:231-244.
                                  5-190

-------
Patrick, R., V. P. Binetti,  and S.  G.  Halterman.   1981.  Acid lakes from
natural and anthropogenic causes.   Science 211:446-448.

Patrick, R., B. Matson, and  L.  Anderson.  1979.  Streams and lakes in
the Pine Barrens.  In Pine Barrens:   Ecosystem  and Landscape.  R. T. T.
Forman, ed.  Academic Press, New York.  601  p.

Pennak, R. W.  1953.  Freshwater Invertebrates  of  the United States.
The Ronald Press Company, New York.   769  p.

Pennak, R. W.  1978.  Fresh-water Invertebrates of the United States.
John Wiley and Sons, New York,  NY.

Peterson, R. H., P. G. Daye, and J.  L.  Metcalfe.   1980a.  Inhibition of
Atlantic salmon hatching at  low pH.   Can. J.  Fish. Aquat. Sci.
37:770-774.

Peterson, R. H., P. G. Daye, and J.  L.  Metcalfe.   1980b.  The effects of
low pH on hatching of Atlantic  salmon eggs,  pp. 328.  D. Drablj6s and
A. Tollan, eds.  Proc. Int.  Conf.  Ecol.  Impact  Acid Precip., Norway
1980, SNSF project.

Pfeiffer, M. and P. Festa.  1980.   Acidity status  of lakes  in the
Adirondack region of New York in relation to  fish  resources.  FW-P168
(10/80), New York State Department of Environmental Conservation,
Albany, NY.

Plonka, A. C. and W. H. Neff.  1969.   Mucopolysaccharide histochemistry
of gill epithelial secretions in brook  trout exposed to acid pH.  Proc.
Pa. Acad. Sci. 43:53-55.

Porter, K.  1977.  The plant-animal  interface in freshwater ecosystems.
Amer. Sci. 65:159-170.

Potts, W. and G. Fryer.  1979.   The effects  of  pH  and salt  content on
sodium balance in Daphnia magna and Acantholeberis curyirostris
(Crustacea:  Cladocera).  J. Comp.  Physiol.  129:289-294.

Pough, F. H.  1976.  Acid precipitation and  embryonic mortality of
spotted salamanders, Ambystoma  maculatum.  Science 192:68-70.

Pough, F. H. and R. E. Wilson.   1977.   Acid  precipitation and
reproductive success of Ambystoma  salamanders.  Water Air Soil Pollut.
7:307-316.

Pratt, J. M. and R. J. Hall.  1981.   Acute effects of stream
acidification on the diversity  of macroinvertebrate drift,  pp. 77-96.
In Effects of Acid  Precipitation  on Benthos.  R.  Singer, ed.
Proceedings of a Regional Syumposium on  Bethic Biology, North American
Benthological Society, Hamilton, NY.
                                  5-191

-------
Prossner,  C. and F.  Brown.   1961.  Comparative Animal Physiology.  W. B.
Saunders Co., Philadelphia.

Raddum, G. G.  1976.   Preliminaere data oom bunnfunaen i j6vre botnat
jj6nn.  jji SNSF Project,  Technical Note 27/76, Oslow-As, Norway.

Raddum, G. G.  1978.   Invertebrates:  Quality and quantity of fish food,
pp. 17-24.  ^n Limnological  Aspects  of Acid Precipitation, Proceedings
of the International  Workshop  held at the  Sagamore Lake Conference
Center, September 25-28, 1978.   G. Hendrey, ed.  BNL 51074, Brookhaven
National Laboratory,  Upton,  NY.

Raddum, G. G.  1979.   Virkninger av  lav pH pa insektlarver.  SNSF
Project, Internatal  Report 45/79, Oslow-As, Norway.

Raddum, G. G.  1980.   Comparison of  benthic invertebrates in lakes with
different acidity, pp. 330-331.   In  Ecological Impact of Acid
Precipitation.  D. Drabltfs and A.Tollan,  eds.  SNSF Project,
Sanderfjord, Norway.

Raddum, G. G. and 0.  E.  Saether. 1981.  Chironomid communities  in
Norwegian lakes with different degrees of  acidification.  Verh.  Int.
Verein. Limnol.  21:367-373.

Raddum, G. G. and A. L.  Steigen. 1981.  Reduced survival and calorific
content of stoneflies  and caddis flies in acid water.  Jh^ Effect of
Acidic Precipitation on Benthos.   R. Singer, ed.  Proc. of a Regional
Symposium on  Bethnic Biology, North American Benthological Society,
Hamilton, NY.  pp. 97-101.

Raddum, G. G., J. Gastry, B. 0.  Rosseland, and I Sevaldrud.  1979.
Vannteger i. Sor-Norge og deres  betydning  some fiskefode i vann  med  ulik
pH.  SNSF Project, Internal  Report 50/79,  Oslo-As, Norway.

Raddum, G.,  A. Hobaek, E. Lomsland,  and T. Johnsen.   1980.
Phytoplankton and zooplankton  in acidified lakes in  southern Norway, pp.
332-333.  Jji Proceedings of  the  International Conference on the
Ecological  Impact of Acid Precipitation, March, 1980, D. Drab!0s and
A. Toll an, eds.  SNSF-project, Oslo, Norway.

Rahel, F. J.  1982.  Fish assemblages  in Wisconsin bog  lakes.  Doctoral
thesis.  Zoology Department, University of Wisconsin-Madison.  197 p.

Rahel, F. J.  1983.  Population  differences  in acid  tolerance between
yellow perch, Perca flayescens,  from naturally acidic and alkaline
lakes.  Can. J. zool. bl:l47-ib2.

Rahel, F. J. and J. J. Magnuson.  1980.   Fish  in Naturally Acidic Lakes
of Northern  Wisconsin, U.S.A.  Jji Proceedings  of the International Conf.
Ecol.  Impacts Acid Precipitation.  D.  Drabl^s  and A.  Tolan eds., SNSF
project,  pp. 334-335, Sandefjord, Norway,  March 11-14,  1980.
                                  5-192

-------
Rahel, F. J.  and J.  J.  Magnuson.   1983.  Low pH and the absence of fish
species in naturally acidic  Wisconsin lakes:  inferences for cultural
acidification.   Can. J.  Fish.  Aquat. Sci. 40:3-9.

Redfield, G.   1980.   The effect of zooplankton on phytopl ankton
productivity  in the  epilimnion of  a subalpine lake.  Hydrobiologia
70:217-224.

Regier, H. A. and V. C.  Applegate.   1972.   Historical review of the
management approach  to  exploitation and  introduction in SCOL lakes.  J.
Fish. Res. Board Can. 29:683-692.

Reinecke, K.   1979.   Feeding ecology and development of juvenile black
ducks in Maine.  Auk 96:737-745.

Retzsch, W.,  A. Everett, P.  Duhaime, and R. Nothwanger.  1982.
Alternative Explanations for Aquatic Ecosystems Effects Attributed to
Acidic Deposition.  Utility  Air Regulatory  Group, Everett and
Associates, Rockville,  MD.

Roback, S. S.  1974.  Pollution Ecology  of  Freshwater Invertebrates, pp.
313-376.  In  C. W. Hart, Jr. and S. L. H. Fuller, eds.  Academic Press,
New York.

Roberts, D. A., C. W. Boylen,  and  R. Singer.  1982.  Acid precipitation
and the  Adirondacks:   Acid effects on  the biota indigenous to Lake
George,  pp.  106-113.  ^n The Lake George Ecosystem.  II.  Proc.,
Second Lake  George   Research Symposium, Lake George, NY.

Robinson, G.  D., W.  A.  Dunson, J.  E. Wright, and G. E. Mamolito.  1976.
Differences in low pH tolerance amoung strains of brook  trout
(Salveilnus fontinalis)  J.  Fish Biol. 8:5-17.

Roff, J. and  R. Kwiatkowski.  1977. Zooplankton and zoobenthos
communities of selected northern Ontario lakes of different acidities.
Can J. Zool.   55:899-911.

Rosseland, B. 0., I. Sevaldurd, D. Svalastog, and I. P.  Muniz.  1980.
Studies on freshwater fish populations-effects of acidification on
reproduction, population structure, growth  and food selection,  pp.
336-337.  D.  Drablj6s and A.  Tollan, eds. Proc., Int. Conf. Ecol.
Impact Acid Precip., Norway 1980,  SNSF project.

Ruby, S. M.,  J. Aczel,  and G. R. Craig.  1977.   The effects of depressed
pH on oogenesis in  flagfish Jordanella floridae.  Water  Res. 2:757-762.

Ruby, S. M., J. Aczel,  and G. R. Craig.  1978.   The effects of depressed
pH on spermatogenesis in flagfish  Jordanella  floridae.   Water Res.
12:621-626.
                                  5-193

-------
Runn, P., N.  Johansson,  and  G. Mil brink.  1977.  Some effects on low pH
on the hatchability of eggs  of perch, Perca fluviatilis L.  Zoon
5:115-125.

Ryan, P. and H. Harvey.   1977.   Growth of rock bass, Ambioplites
rupestris, in relation to the morphoedaphic index as an indicator of an
envimoment stress.  J.  Fish. Res.  Board Can. 34:2079-2088.

Ryan, P. M. and H.  H. Harvey.  1980.  Growth responses of yellow perch,
Perca flavescens (Mitchill), to  lake acidification in the La Cloche
Mountain Lakes of Ontario.  Env.  Biol. Fish.  5:97-108.

Ryan, P. and H. Harvey.   1981.   Factors accounting for variation in the
growth of rock bass (Amblopli tes rupestri s) and yellow perch (Perca
flavescens) in the acidifying LaCloche Mountain lakes of Ontario,
Canada.  Verh.  Internat. Verein.  Limncl. 21:1231-1237.

Saber, P. A.  and W. A. Dunson.   1978.  Toxicity of bog water to
embryonic and larval anurian amphibians.  J. Exp. Zoo!. 204:33-42.

Sand-Jensen,  K. and M. Sondergaard.  1979.  Distribution and
quantitative development of  aquatic macrophytes in relation to sediment
characteristics in oligotrophic  Lake Kalgaard, Denmark.  Freshwater
Biol. 9:1-11.

Sawicka-Kapusta, K.  1978.  Estimation of the cntents of heavy metals in
antlers of roe deer from silesion woods.  Arch. Ochrony Srodowiska
1:107-121.

Sawicka-Kapusta, K.  1979.  Roe  deer antlers as bioindicators of
environmental polution in southern Poland.  Environ. Pollut. 13:283-293.

Scheider, W. and P. J. Dillon.   1976.  Neutralization and fertilization
of acidified lakes near  Sudbury,  Ontario.  Wat. Pollut. Res. Can. 11:
93-100.

Scheider, W., J. Adamski, and M.  Paylor.  1975.  Reclamation of
Acidified Lakes near Sudbury, Ontario, By Neutralization and
Fertilization.  Ontario  Ministry  of the Environment, Rexdale, Ontario,
Canada.

Scheider, W.  A., J. Jones, and B.  Cave.  1976.  A preliminary report on
the neutralization of Nelson Lake near Sudbury, Ontario.  Ontario
Ministry off the Environment, Rexdale, Ontario, Canada.

Schindler, D.  1975.  Whole-lake eutrophication experiments with
phosphorus, nitrogen and carbon.   Verh. Internat. Verein. Limnol.
19:3221-3231.

Schindler, D. W.  1980.   Experimental acidification of a whole lake:  A
test of the oligotrophication hypothesis, pp. 370-374.  In Proc. Int.
Conf. Impact Acid Precip., SNSF  Project, Norway.  D. DraoTtfs and A.
Toll an, eds.  SNSF Project,  Norway.
                                  5-194

-------
Schindler, D.  W. and S.  K.  Holmgren.   1971.   Primary production and
phytoplankton  In the Experimental  Lakes Area, northwestern Ontario, and
other low-carbonate waters,  and a  liquid  scintillation method for
determining 14C-activity in  photosynthesis.   J. Fish. Res. Board Can.
28:189-201.

Schindler, D.  and M. Turner.   1982.   Biological, chemical and physical
responses of lakes to experimental acidification.  Water, Air, and Soil
Pollution 18:259-271.

Schindler, D.  W., F. A.  J.  Armstrong,  S.  K.  Holmgren, and G. J.
Brunskill, 1971.  Eutrophication of  lake  227, experimental lakes area,
northwestern Ontario, by addition  of phosphorus and nitrate.  J. Fish.
Res. Bd. Canada 28:1763-1782.

Schindler, D.  W., R. H.  Hesslein,  R.  Wagenann, and W. S. Broecker.
1980a.  Effects of acidification on  mobilization of heavy metals and
radionuclides  from the sediments of  a  freshwater lake.  Can. J. Fish.
Aquat. Sci. 37:373-377.

Schindler, D., R. Wagemann,  R. Cook,  T. Ruszczynski, and J. Prokopowich.
1980b.  Experimental acidification of Lake 223, Experimental Lakes Area:
background data and the first three  years of acidification.  Can. J.
Fish. Aquat. Sci.  37:342-354.

Schlichter, L. C.  1981.  Low pH affects  the fertilization and
development of Rana pi pi ens eggs.  Can. J. Zool. 59:1693-1699.

Schofield, C.  L.  1965.   Water quality in relation to survival of brook
trout, Salvelilnus fontinalis (Mitchill). Trans. Am. Fish. Soc.
94:227-735^

Schofield, C.  L.  1976a.  Acid precipitation:  Effects on fish.  Ambio
5:228-230.

Schofield, C.  L.  1976b.  Lake acidification in the Adirondack Mountains
of New York:  causes and consequences.  Proc. First Int. Symp. Acid
Precipitation and the Forest Ecosystem, May  12-15, 1975.  Columbus,
Ohio.  USDA Forest Serv. Gen.  Tech.  Rep.  NE-23. p. 477.

Schofield, C.  L.  1976c.  Dynamics and management of Adirondack fish
populations.  Project Report April 1,  1975 - March 31, 1976, No.
F-28-R-4, Dept. of Environmental Conservation, Albany, N.Y.

Schofield, C.  L.  1977.   Acid snow-melt effects on water quality and
fish survival  in the Adirondack Mountians of New York State.  U.S. NTIS
Report PB-277801.

Schofield, C. L.  1978.   Toxicity  of metals, pp. 31-34.  Jji Limnological
Aspects of Acid Precipitation.  G. Hendrey,  ed.  Brookhaven National
Laboratory, Upton, NY.
                                  5-195

-------
Schofield, C. L.  1981.   Acid rain and the Adirondack  trout, pp. 93-96.
Jji Acid Rain and the Atlantic Salmon.   IASF Spec.  Publ.  10.

Schofield, C. L.  1982.   Historical  fisheries  changes  as related to
surface water pH changes in the United States.   Acid Rain/Fisheries
Symposium, August 1980,  Ithaca, NY,  in press.

Schofield, C. L. and J.  R. Trojnar.   1980.  Aluminum toxicity to brook
trout (Salvelinus fontinal1s) in acidified waters,  pp. 341-362.  In
Polluted Rain.  T. Y. Toribara, M. W.  Miller,  and  P. E.  Morrow, e"cfs.
Plenum Press, NY, London.

Schofield, C. L., D. A.  Webster, C.  A. Guthrie,  and W. A.  Flick, 1981.
Management of Acidified  Waters.  Final Report  C-164480,  New  York State
Department of Environmental Conservation.

Schwarz, K. and C. M. Foltz.  1957.   Selenium  as an integral part  of
Factor 3 against dietary necrotic liver degeneration.  J.  Am.  Chem.
Soc. 79:3292-3293.

Schweibert, E. G., Jr.  1974.  Matching the Hatch:  A  Practical Guide to
Imitation of Insects Found on Eastern and Western  Trout  Waters.
Macmillan, New York.  221 pp.

Scotter, G. W.  1963.  Growth rates of Cladonia  alpestris, £. mitis, and
C. rangiferina in the Taltson River region,"N.'W.T  K. F.  Can. J.  Bot.
Tl:1199-1202.

Seddon, B.  1972.  Aquatic macrophytes as limnological  indicators.
Freshwater Biology 2:107-130.

Sevaldrud, I.,  I. P. Muniz, and S. Kalvenes.  1980.  Loss  of fish
populations in  southern Norway.  Dynamics and  magnitude  of the problem,
pp. 350-351.  D. Drabltfs and A. Tollan, eds.  Proc. Int. Conf. Ecol.
Impact Acid Precip., Norway 1980, SNSF Project.

Shannon, E. E.  and P. L. Brezonik.  1972.  Limnological  characteristics
of North and Central Florida lakes.   Limnol. and Oceanogr. 17:97-110.

Shapiro, .]., V. Laroarra, and M. Lynch.  1975.   Proceedings of  a
symposium on water quality management through  biological control,  pp.
85-96.  P. L. Brezonik and J. L. Fox, eds.  Univ.  of  Florida,
Gainesville.

Shaw, G. G.  1981.   The effect of SO? on selenium  concentrations  in
serviceberry.  Proc. 38th NE Fish and Wildlife Conf.,  Virginia Beach,
April 19-22.  8 pp.

Shaw, J.  1960.   The absorption of sodium  ions by  the crayfish Astacus
pallipes Lereboullet.  III.  The effect of other cations in the  external
solution.  J. Expt.  Bol.  37:342-354.
                                  5-196

-------
Singer, R. (ed.).  1981a.   Effects  of  Acidic Precipitation on Benthos,
Proceedings of a Regional  Symposium on Bnthic  Biology, North American
Benthological  Society,  Hamilton,  NY.   156  pp.

Singer, R.  1981b.   Effects of pH on the morphology and ultrastructure
of the shell  of the freshwater clam Anodonta grandis  (Unionidae).  ^ji
Effects of Acidic Precipitation on  Benthos, Proceeding of a Regional
Symposium on Benthic Biology,  North American Benthological Society,
Hamilton, NY.   pp.  103-111.

Singer, R., D. A. Roberts,  C.  W.  Boylen.   1983.  The macrophytic
vegetation of an acidic   Adirondack (New  York, USA) lake:  A new depth
record for aquatic macrophytes.  Aquat. Bot.   In press.

Sinley, J., J. Goettl,  and P.  Davies.   1974.   The effects of zinc on
rainbow trout (Salmo gairdneri) in  hard and soft water.  Bull. Environ.
Contamin. Toxicol.  12:193-201.

Smith, E. and H. Swingle.   1939.  The  relationship between plankton
production and fish production in ponds.   Trans. Amer. Fish. Soc.
68:309-315.

Smith, M. W.  1961.  A  limnlogical  reconnaissance of a Nova Scotian
brown-water lake.  J. Fish. Res.  Board Can. 18:463-478.

Smith, R. F.   1957.  Lakes and ponds.   Fishery Survey Report Number 3.
New Jersey Department of Conservation  and  Economic Development, Division
of Fish and Game, Trenton,  New Jersey, 198 p.

Sondergaard,  M.  1979.   Light  and dark respiration and the effect of the
lacunal system on refixation of CO? in submerged aquatic plants.  Aq.
Bot. 6:269-283.

Sondergaard,  M. and K.  Sand-Jensen. 1979.  Carbon uptake by leaves and
roots of Littorella uniflora (L.) Aschers.  Aq.  Bot. 6:1-12.

Sorokin, C.  1962.   Carbon dioxide  and bicarbonate in cell division.
Arch. Fur Mikrobiol. 44:219-227.

Sprules, W.  1975a.  Midsummer crustacean  zooplankton comunities in
acid-stressed lakes. J. Fish. Res. Board  Can. 32:389-395.

Sprules, W.  1975b.  Factors affecting the structure of limnetic
crustacean zooplankton  communities  in  central  Ontario lakes.  Verh.
Internat. Verein. Limnol.  19:635-643.

Spry, D. C., C. M.  Wood, and P. V.  Hodson.  1981.  The effects of
environmental  acid on freshwater  fish  with particular reference to the
softwater lakes in Ontario and the  modifying effects of heavy metals.  A
literature review.   Can. Tech. Rep. Fish.  Aquat.  Sci. 999.  50 pp.
                                  5-197

-------
Stanley, R. A.  1974.   Toxicity  of  heavy metals and salts to Eurasian
watermilfoil  (Myriophyllum spicatum L.).  Arch. Environ. Contam.
Toxicol. 2:331.

Steigen, A. L. and G.  6.  Raddum.  1981.  Effects of acidified water on
behavior and energy content in the  water louse Asellus aquaticus^ (L.).
lr± Effects of Acidic Precipitation  on  Benthos, Proc. of a Regional
Symposium on Benthic Biology, North American  Benthological Society,
Hamilton, NY.  pp. 113-118.

Stenson, J., T.  Bohlin, L. Henrickson,  B. I.  Nilsson, H. G. Nyman, H. G.
Oscarson, and P. Larsson   1978.   Effecs of  fish removal from a small
lake.  Verh. Internat. Verein. Limnol.  20:794-801.

Stickney, F.  1922.  The relation of the nyumphs of a dragonfly
(Libellula pulchella Drury) to acid and temperature.  Ecology
3{3):437-462.

Stokes, P.  1980.  Phytopl ankton of acidic  lakes in Killarney Ontario:
Community structure related to water chemistry.  Jjl Abstracts for the
International Conference on the  Ecological  Impact  of Acid Precipitation,
March, 1980.  SNSF-project, Oslo, Norway.

Stokes, P. M.  1981.  Benthic algal communities in acid lakes, pp.
119-138.  In Effects of Acidic Precipitation  on Benthos.  R. Singer, ed.
North American Benthological Society,  Hamilton, New York.

Stoneburner, D. L. and L. A. Smock. 1980.  Plankton communities of an
acidic, polymictic, brown-water  lake.   Hydrobiologia 69:131-137.

Strijbosch, H.  1979.  Habitat  selection  of amphibians  during their
aquatic phase.  Oikos 33:363-372.

Suns, K., C. Curry, and D. Russell.  1980.  The effects of water quality
and morphometric parameters on mercury uptake by yearling yellow perch.
OME Tech. Rept. LTS 80-1  (cited  in  U.S./Canada MO I 1982).

Sutcliffe, D. W. and T. R. Carrick.  1973.  Studies on mountain streams
in the English Lake District.   Freshwater Biology  3:437-462.

Svardson, G.  1976.   Interspecific  population dominance in  fish
communities of Scandinavian lakes.   Report  of the  Institute of
Freshwater Research at Drottningholm 55:144-171.

Swarts, F. A., W. A. Dunson, and J. E. Wright.   1978.   Genetic  and
environmental factors involved  in increased resistance  of brook trout  to
sulfuric  acid solutions and mine acid polluted waters.   Trans.  Am.  Fish.
Soc. 107:651-677.

Taylor, M., A. Demayo, and K. Taylor..  1982.   Effects  of  zinc on humans,
laboratory and farm animals, terrestrial  plants,  and  freshwater aquatic
life.  CRC Critical Reviews in  Environ. Contr.  12:113-181.
                                  5-198

-------
Tessenow, U. and Y. Baynes.   1978.   Experimental  effects  of  Isoetes
lacustris L. on the distribution  of  EH,  pH,  Fe  and  Mn  in  lake
sediment.  Verh. Internat.  Verein. Limnol. 20:2358-2362.

Thompson, D. C. and K.  H.  McCourt.   1981.  Seasonal  diets of the
Porcupine Caribou Herd.  Am.  Midi. Nat.  105(1):70-76.

Tomlinson, G. H.  1978.  Acidic precipitation and mercury in Canadian
lakes and fish, pp. 104-118.   lr± Scientific  Papers  from the  Public
Meeting on Acid Precipitation.   H. H.  Izard  and J.  J.  Jacobson, eds.
Science and Technology  Staff, New York State Assembly  Center for
Environmental Research, Cornell  University,  Ithaca,  NY.

Torke, B. G.  1979.  Crustacean zooplankton  data  for 190  selected
Wisconsin inland lakes.  Wisconsin Department of  Natural  Resources
Research Report 101. Madison,  WI, 70  p.
Traaen, T. S.
"litterbags".
Norway.
1976.   Nedbrytning av  organisk  materiale.  Forso/k med
SNSF Project,  Technical  Note 19/76, Osla-AOs,
Traaen, T. S.  1977.   Nedbrytning  av  organisk  materiale.  Forso/k med
"litterbags", resul tater etter 2 ars  feltforso/k.   SNSF Project,
Technical Report, Oslo-fts,  Norway.
Traaen, T. S.  1978.  Bakterieplankton  i.  innsjtfer.
Technical Note 41/78,  Oslo-A°s,  Norway.
                                     SNSF Project,
Trivelpiece, W., S. Brown,  A.  Hicks,  R.  Fekete,  and N. J. Volkman.
1979.  An analysis of the distribution and  reproductive  success of the
common loon in the Adirondack  Park,  New  York,  pages 45-55.   In The
Common Loon.  S. A. Sutcliffe, ed.,   Proc.  Second  N. A.  Conf7~on Common
Loon Research Management.  National  Audobon Society, New York.  162 pp.

Trojnar, J. R.  1977a.  Egg and larval  survival  of white suckers
(Catostomus commersoni)  at low pH.   J. Fish. Res.  Bd. Can. 34:262-266.
Trojnar, J. R.  1977b.   Egg
(Salyelinus fontinalis)  fry
574-579.
             hatchability  and  tolerance of brook trout
             at low pH.  J.  Fish.  Res. Bd. Can. 34:
Troutman, D. E. and N.  E.  Peters.   1982.   Deposition and  transport of
heavy metals in three lake basins  affected by  acid  precipitation  in
Adirondack Mountains, New  York,  pp.  33-61.  Jji Energy and Environmental
Chemistry, II;  Acid Rain.   L.  H. Keith,  ed.  Ann  Arbor  Sci., Ann  Arbor,
MI.

U.S./Canada Memorandum  of  Intent on  Transboundary Air Pollution.  1982.
Impact Assessment Group 1, Phase III Report April 1982.   Vol.  II, 519
pp.  Final Report January  1983.
                                  5-199

-------
Ueno, M.  1958.  The disharmonious  lakes  of Japan.  Verh. Internal. Ver.
Limnol. 13:217-226.

Ultsch, G. R.   1978.  Oxygen  consumption  as a  function of pH in three
species of freshwater fishes.   Copeia  No. 2 1978:272-279.

Ultsch, G. R.  and G. Gros.   1979.   Mucus  as a  diffusion barrier to
oxygen:  possible role in 02  uptake at low pH  in carp (Cyprinus
carpio) gills.  Comp. Biochem.  Physiol.   62A:685-689.

Ultsch, G., M. Ott,  and N.  Heisler. 1981.  Acid-base and electrolyte
status in carp (Cyprinus carpio) exposed  to low environmental pH.  J.
Exp. Bioi. 93:65^8Tn

Uutala, A. J.   1981.  Composition and  secondary production of the
chironomid  (Diptera) communities in two  lakes in the Adirondack Mtn.
Region, New York.  Pp. 139-153  j_n Effects of Acidic Precipitation on
Benthos, Proc. of Regional  Symposium on Benthic Biology.  R. Singer, ed.
North American Benthological  Society,  Hamilton, NY.

Van Everdingen, R. 0.  1970.  The Paint Pots,  Kootenay National Park,
British Columbia-acid spring  water  with extreme heavy-metal content.
Canadian Journal of Earth Science 7:831-852.

Van Voris, P., R. O'Neill,  W. Emmanuel, and H. Shugart, Jr.  1980.
Functional complexity and ecosystem stability. Ecology 61:1352-1360.

Vangenechten,  J. D.  and 0.  L. J. Vanderborght. 1980.   Effect of pH on
sodium and chloride balance in  an inhabitant of acid freshwaters:  The
waterbug Corixa punctata (Illig.) (Insecta, Hemiptera), pp. 342-343.
In Ecological  Impact of Acid Precipitation.  D. Drablj6s and A. Toll an,
eTTs.  SNSF Project,  Sandefjord, Norway.

Vangenechten,  J. H.  D., S.  Van  Puymbroeck, and 0. L. J. Vanderborght.
1979.  Effect of pH  on the  uptake of sodium in the waterbugs Corixa
dentipes (Thorns.) and Corixa  punctta (Illig.)  (Hemiptera,
Heteroptera).   Comp. Biochem. Physiol.  64A:509-521.

Vitt, D. H.  and N. G. Slack.  1975. An analysis of the vegetation of
Sphagnum-dominated kettle-hole  bogs in  relation to environmental
gradients.  Can. J.  Bot. 53:332-359.

Von Ende, C. N.  1979.  Fish predation, interspecific predation and the
distribution of two  Chaoborus species.  Ecology 60:119-128.

Waring, G. A.   1965.  Thermal springs  of  the United States and other
countries of the world.  A  summary. U.S. Geol. Sur. Prof. Paper No.
492, 383 p.

Warner, R. W.   1971.  Distribution  of  biota in a stream polluted by acid
mine-drainage.  OhioJ. Sci.  71:202-215.
                                  5-200

-------
Watt, W. D.  1981.  Present and potential  effects  of  acid  precipitation
on the Atlantic salmon in eastern Canada,  pp.  39-45.  ^ji Acid Rain and
the Atlantic Salmon.  IASF Spec.  Publ.  10.

Watt, W., C. Scott, and W. White.  1983.   Evidence of acidification of
some Nova Scotian rivers and its  impact on  Atlantic salmon, Salmo salar.
Can. J. Fish. Aquat. Sci. 4 (In press).

Welch, P. S.  1952.  Limnology.  McGraw-Hill  Book  Company, New York.
538 p.

Wenblad, A. and A. Johansson.   1980.   Aluminum i foer surade
vaestsvenska sjoear.  Vatten 36:154-157 cited in Aquat. Sci. Fish.
Abstracts 11(1) :100.

Wetzel, R. G.  1975.  Limnology.   W. B. Saunders Company, Philadelphia,
PA, 743 p.

Wiederholm, T. and L.  Eriksson.  1977.  Benthos of an acid lake.  Oikos
29:261-267.

Wiener, J. G. and J. P. Giesy,  Jr.  1979.   Concentrations of Cd, Cu, Mn,
Pb, and Zn in fishes in a highly  organic  softwater pond.  J. Fish. Res.
Board Can. 36:270-279.

Wiener, J. G. and W. R. Hanneman.  1982.  Growth and  condition of
bluegills in Wisconsin lakes:   Effects  of population  density and lake
pH.  Trans. Amer. Fish. Soc. 111:761-767.

Wile, I.  1981.  Macrophytes, pp. 216-223.  ^n Acidification in the
Canadian Aquatics Environment.  National  Res.  Council, Canada, No.
18475.

Wiurn-Andersen, S. and J.  M.  Andersen.   1972.   Carbon  dioxide content of
the interstitial  water in the sediment  of Grane Langso, a Danish Lobelia
lake.  Limnol. and Ocean. 17:943-947.

Winter, K. A. and U. C. Gupta.  1979.   Selenium content of forages grown
in Nova Scotia, New Brunswick and Newfoundland.  Can. J.  Animal Sci.
59:107-111.

Woelkerling, W. J. and S. B. Gough.  1976.  Wisconsin Desmids.  Ill
Desmid community composition and  distribution  in relation to lake type
and water chemistry.  Hydrobiologia 51:3-32.

Wood, C. and D. McDonald.  1982.   Physiological mechanisms of acid
toxicity to fish, pp.  197-226.  J^n  Acid Rain  Fisheries, Amer. Fish.
Soc., Bethesda, MD.

Woodwell, G.  1970.   Effects of pollution on the structure and
physiology of ecosystems.  Science  168:429-433.
                                 5-201

-------
Wren, C., H.  MacCrimmon,  R.  Frank,  and P.  Suda.  1980.  Total and
methyl mercury levels in wild mammals  from  the Precarnbrian Shield area of
south-central Ontario, Canada.   Bull. Environ. Contam. Toxicol.
25:100-105.

Wright, R. F. and E. Snekvik.   1978.  Acid precipitation: chemistry and
fish populations in 700 lakes  in southernmost Norway.  Verh. Intern.
Verein. Limnol. 20:765-775.

Wright, R., N. Conroy, W. Dickson,  R. Harriman, A. Henriksen, and C.
Schofield.  1980.  Acidified lake districts of the world:  A comparison
of water chemistry of lakes  in  southern  Norway, southern Sweden,
southwestern Scotland,  the Adirondack Mountains of New York, and
southeastern Ontario, pp. 377-379.   In Ecological Impact of Acid
Precipitation, D. Drablj6s and  A. TolTan, eds.  SNSF project, Oslo,
No rway.

Wright, R. F., T. Dale, E. T.  Gjessing,  G.  R. Hendrey, A. Henriksen,
M. Johannessen, and I.  P. Muniz.  1975.  Impact of acid precipitation on
freshwater ecosystems in  Norway.  SNSF Project, Research Report FR
3/75, Oslo-AOs, Norway.

Wright, R. F., T. Dale, E. T.  Gjessing,  G.  R. Hendrey, A. Henriksen, M.
Johannessen, and I. P. Muniz.   1976.  Impact of acid precipitation on
freshwater ecosystems in  Norway.  Water, Air, Soil Pollut. 6:483-499.

Yamagami, K.  1973.  Some enzymological  properties of a hatching enzyme
(chorionase) isolated from the fresh  water teleost, Oryzias laptipes.
Comp. Biochem Physiol. 468:603-616.

Yamagami, K.  1981.  Mechanisms of hatching in fish:  Secretion of
hatching enzyme and enzymatic  choriolysis.   Ameri. Zool. 21:459-471.

Yan, N.  1979.  Phytoplankton  community  of an acidified, heavy metal
contaminated lake near Sudbury, Ontario:   1973-1977.  Water, Air, Soil
Pollut. 11:43-55.

Yan, N. D. and P. J. Dillon.  1981.  Studies of Lakes and Watersheds
near Sudbury, Ontario.  Draft,  Ontario Ministry of the Environment
Technical Report, Water  Resources Branch,  Rexdale, Ontario, Canada.

Yan, N. and P. Stokes.  1978.   Phytoplankton of an acidic lake, and its
responses to experimental alterations of pH. Environ. Conserv.
5:93-100.

Yan, N. and R. Strus.  1980.  Crustacean zooplankton communities of
acidic, metal-contaminated lakes near Sudbury, Ontario. Can. 0. Fish.
Aquat. Sci. 37:2282-2293.

Yan, N., C. Lafrance, and G. Hitchin.  1982. Planktonic fluctuations in
a fertilized, acidic lake.  lr± Proceedings of the Acid Rain? Fisheries
Symposium, Cornell University,  Ithaca, New York,  August 3-6, 1981.
                                  5-202

-------
Zaret, T. M.  1980.   Predation  and  freshwater communities.  Yale
University Press,  New Haven,  CT.  187 pp.

Zaret, T. M. and W.  C.  Kerfoot.   1975.  Fish predation on Bosmina
longirostris; body size selection versus visibility selection.  Ecology
56:232-237.

Zotin, A.  1965.  The Uptake  and  movement of water in embryos, pp.
365-385.  Ir± The State and Movement of Water in Living Organisms.
Cambridge University Press, Cambridge, England.
                                 5-203

-------
            THE ACIDIC DEPOSITION  PHENOMENON AND ITS EFFECTS
                     E-6.   INDIRECT EFFECTS ON HEALTH

6.1  INTRODUCTION (T. W.  Clarkson)

     Indirect health effects  that  may be causally related to acidic
deposition have not been  demonstrated in human populations.  This lack
of documented effects may  mean  that no such effects exist in individuals
or populations.  On the other hand, interest in the pheomenon of acidic
deposition is recent and  few  investigations, if any, have been made into
the possibility of indirect health effects.  In principle, acidic
deposition may influence human  exposure to toxic chemicals via two main
pathways:   the accumulation of  chemicals in food chains leading to man
and the contamination of  drinking  water.  The format of this chapter is
organized according to these  exposure pathways, i.e., Food Chain
Dynamics (Section 6.2) and Ground, Surface and Cistern Waters (Section
6.3).

     The substances requiring special attention are methyl mercury, due
to its accumulation in aquatic  food chains, and lead, due to the
potential  for contaminating drinking water.  Aluminum is a special case
where its  presence at elevated  concentrations in water used in dialysis
therepy may be a cause of brain damage.  Other elements and chemicals
will only  be briefly mentioned  as  information is limited.  These include
arsenic, asbestos, cadmium, copper, nickel, and the nitrosamines.
Furthermore, reference will be  made to other metals and elements that
may interact with mercury, lead, and aluminum to modify human exposure
and toxicity.

6.2  FOOD  CHAIN DYNAMICS  (T.  W. Clarkson)

6.2.1  Introduction

     Human exposure could  result from bioaccumulation processes.
Aquatic organisms, particularly predatory fish at the top of the food
chain, may concentrate certain  toxic elements, leading to substantial
human exposure as in the  case of mercury.  Accumulation may occur in
wildlife in contact with  the  contaminated water or consuming aquatic
organisms.  Water used for irrigation could lead to contamination of
edible vegetation.  Concentrations of toxic elements in meat, eggs, and
diary products could be produced by contamination of livestock.  This
could occur from drinking  water or from contamination of livestock food.

     Each  of these potential  bioaccumulation pathways to humans should
be considered in light of possible health hazards.  Data, however, are
very limited with regard  to measurement of the toxic elements and to the
kinetics of transfer and  uptake in bioaccumulation processes.  This
discussion will, therefore, be  limited to only a few toxic elements and
the major pathways of exposure.
                                  6-1

-------
6.2.2  Availability and Bioaccumulation  of Toxic Metals

     Mercury and its compounds have  been extensively  studied  in terms  of
availability and bioaccumulation.  The  impetus  for this work  came from a
discovery made in the late I9601s  (see  below) that inorganic  mercury may
be methylated in the aquatic  environment to the highly neurotoxic
species - methyl mercury - and thereby  accumulate in  aquatic  food chains
leading to man.   Mercury is the most dramatic example of a change in
speciation produced in the environment  that ultimately leads  to
increased levels in human food. Alkylation of  certain other  toxic
metals may also occur in the  environment (Wood  1974). Organic forms of
arsenic are known to accumulate in shellfish but organic arsenic is much
less toxic to man and animals than the  inorganic species.  Cadmium
accumulates in plants and certain marine Crustacea, although  the role  of
aquatic acidification in these accumulation processes is not  well
documented.  In short, this section  will  deal primarily with  our
knowledge concerning the bioaccumulation of methyl mercury in aquatic
food chains and the possible  role of acidification.   Other metals and
elements will be discussed briefly as a  group.

6.2.2.1  Speciation (Mercury)—The different chemical and physical  forms
of mercury each have their own distinctive biological activity (for a
detailed review, see Carty and Malone 1979). Each differs from the
others in the extent of bioaccumulation  in food chains and in toxicity
to humal life.  The speciation of mercury in natural  bodies of water is,
therefore, an important consideration in assessing potential  hazard to
man.

     Mercury exists in a variety of  physical and chemical forms.  The
inorganic forms have three oxidation states: Hg° or  "metallic" mercury
is in the zero oxidation state.  It  is  a liquid metal ("quicksilver")
and possesses a high vapor pressure. The vapor is a  monatomic gas, is
highly diffusible, and possesses a low  solubility in  water.   It is
commonly referred to as "mercury vapor"  despite the fact that certain
other forms of mercury (e.g., dimethyl  mercury) also  readily  vaporize.
If Hg° is produced in aquatic bodies of  water,  it will readily diffuse
into the atmosphere.

     Mercury vapor in the presence of water and oxygen is readily
oxidized to the first oxidation state Hg22+, called mercurous
mercury and to the second oxidation  state, Hg2+, known as mercuric
mercury.  Indeed, the interconversion of these  three  oxidations states
via the disproportionation reaction

          Hg22+ Z Hg2+ + Hg°

is an important reaction in the environmental  transport  of mercury  (Wood
1974).  The direction of the  reaction is affected not only by the
relative concentrations of the three species of mercury  but by the
ambient redox potential and by certain  microorganisms capable of
reducing Hg2+ to Hg° (Wood 1974).
                                  6-2

-------
     Most mercurous salts of mercury possess a low solubility in  water.
 Furthermore, the mercurous action disproportionates to Hg°  and Hg2+  in
 the  presence of protein and other substances containing ligands having  a
 high affinity for Hg2+.  Thus, inorganic mercury in the environment
 tends either to be present as Hg° (usually as the vapor)  or Hg2+.

     The mercuric cations are capable of forming a wide variety of
 chelates and complexes with electron donating groups (ligands).  For
 example, four complexes are formed with chloride anions--HgCl + ,
 HgCl2, HgCl3~, and HgCl42~.  The mercuric cation possesses
 such high affinities for many organic ligands expected to be present in
 sediments, water, and aquatic biota that it is unlikely that the  free
 cations, Hg2+, will ever be detected in measurable quantities. Its
 highest affinity is for sulfur anions S2~, S-H and the sulfhydryl
 anion in proteins and ami no acids--R-S~ where the affinity  constants
 are usually in the range of 10 to 20.  It is not surprising,  therefore,
 that the naturally occurring ore of mercury, cinnabar,  is the sulfide
 complex HgS.  The reaction of Hg2+ with sulfide ions is important  in
 the geochemical cycles of mercury (see below).  Mercuric  sulfide  is
 highly insoluble in water, (solubility product 10-53 M),  so reaction
 of mercury with sulfides in water and sediments leads to  immobilization
 of the metal.  However, in the presence of well-oxygenated  water  (Jensen
 and Jernelov 1972) and also in the presence of aerobes,  HgS can be
 oxidized to the much more  soluble sulfite and sulfate salts,  thus
 leading to remobilization  of mercury (see below).

     Mercuric mercury can  form a wider variety of organometallic
 compounds in which the mercuric atom is linked covalently with at  least
 one carbon atom.  These organometallic compounds are usually  referred to
 as "organic mercury."  Examples of two organic forms of mercury are
 depicted in Table 6-1.  Phenyl mercury has long been used as a fungicide
 in the paint industry and  as a slimicide in the paper pulp  industry.
 The latter use led to contamination of many bodies of fresh water  in
 Europe and North America,  and its use has now been banned.   Phenyl
 mercury may be broken down rapidly to inorganic mercury (Hg2+)  by
 microorganisms present in  the aquatic environment and by  enzymes in
 mammalian tissues.  It has a low toxicity to man.

     Methyl  mercury possesses unique environmental  and toxicological
 properties that make it the most dangerous mercury compound to human
 health and one of the most hazardous chemicals found in the natural
 environment.  Methyl  mercury is known to be produced by methylation  of
 inorganic (Hg2+) mercury by methanogenic bacteria present in  sediments
 in natural  bodies of water (for review, see Wood 1974).   It is avidly
 accumulated by fish and attains the highest concentration in  species of
 predatory fish.  Like Hg2+, it has a high affinity for  organic
 ligands, prticularly the sulfhydryl  anion in proteins.  It  appears to
have a low toxicity to fish and other aquatic species but is  highly
 toxic to the human central  nervous system (see Section  6.2.4.2).

     Dimethyl  mercury (CH3)2Hg is also produced by  methanogenic
bacteria.   Like mercury vapor, it possesses a low solubility  in water


                                  6-3

-------
         TABLE 6-1.   SOURCES  OF MERCURY IN THE ENVIRONMENT 1971
                         (WHO 1976, NRIAGU 1979)
          Source                                 Amount
                                            Metric tons yr-1
Natural
   degassing of earth's  crust                        -30,000

Anthropogenic
   worldwide mining                                  10,000
   combustion of coal                                  3,000
   combustion of oil                                 400-1500
   smelting of metal  sulfide ores                      1,500
   steel  cement phosphates                               500
                                  6-4

-------
and has a high vapor pressure.  Thus, dimethyl mercury tends to escape
from the aquatic system into the atmosphere, where it may be broken down
by sunlight to Hg°  and methyl  free radicals.

6.2.2.2  Concentrations and Speciations in Water (Mercury)--The early
findings of Stock and Cucuel (1934) that rain water contains mercury
between 50 to 500 ng Hg &"1 is generally supported by more recent
findings.  Brune (1969)  reported average values of approximately 300 ng
Hg £~1  in Sweden,  and Eriksson (1967) also in Sweden found most
samples of rain water in the range of 0 to 200 ng £"1.

     Values for snow depend greatly on the collection conditions and how
long the snow has laid on the  ground.  Straby (1968) found values of 80
ng g"1 in fresh snow, but values as high as 400 to 500 ng Hg g
were found in snow  samples that had partly melted and evaporated over
the winter.  Analysis of the samples deposited in Greenland prior to the
1900s yielded values of 60 ng  g'1 (Weiss et al. 1971).

     Bodies of fresh water for which there is no known source of
contamination generally yield  values less than 200 ng jT1 .  Most
values fall in the  range of 10 to 40 ng£-1  and drinking water
usually has values  less than 30 ngji'1 (WHO 1976).

     Few reports exist on the  speciation of mercury in water, probably
because of analytical difficulties.  A recent review by McLean et al.
(1980) found that methyl mercury accounted for a small fraction of the
total  ~ of the order of 1 percent.  However, a more recent report by
Kudo et al. (1982)  found that  methyl mercury accounted for about 30
percent of total mercury in samples taken from Canadian and Japanese
rivers.  Mercuric mercury (Hg2+) accounted for about 50 percent.

     Two important  conclusions may be drawn from these data.  First,
that precipitation  is an important source of mercury to fresh water (see
next section), and  second, that mercury in drinking water offers no
health threat.  Concentrations on the order of a few hundred nanograrns
per liter would result in a negligible intake of mercury on the assumed
intake of two liters per day  (U.S. EPA 1980a).  This intake, less than 2
yg day"1, is well below the advised maximum safe intake of 30 vg
Hg day-1 (WHO 1972b); thus, additional mobilization of mercury into
water by acidic deposition should not pose a health threat in terms of
contaminated drinking water.

6.2.2.3  Flow of Mercury in the Environment--This topic has been the
subject of a number of reviews (WHO 1976, NAS 1978, U.S. EPA 1980a) and
will be briefly summarized here.  The subject is one of intensive
research, particularly by the  Coal-Health-Environment Project (KHM 1981)
in Sweden.  This topic's development  is hampered by the need for more
sensitive and more  specific methods for measuring the various physical
and chanical species of mercury believed to be present at extremely low
concentrations in the atmosphere and  in bodies of natural water.
                                  6-5

-------
6.2.2.3.1  Global  cycles.   The global  cycles  of mercury have recently
been reviewed by Nriagu (1979)  and  by  the  National Academy of Sciences
(1978).  The global  cycle  of mercury  involves degassing of the element
from the earth's crust and evaporation from natural bodies of water,
atmospheric transport believed to be  mainly in the form of mercury
vapor, and deposition of mercury  back  onto land and water.  Mercury
ultimately finds its way to sediments  in water, particularly to oceanic
sediments where the  carry-over is very slow.   The ocean and oceanic
sediments are believed to  be the  ultimate  destination of mercury in the
global cycle.

      Andren and Nriagu (1979)  have indicated that mercury's residence
time in the atmosphere may vary from  approximately 6 to 90 days.
Residence times of mercury in soils are on the order of 1000 years,
oceans on the order  of 2000 years,  and sediments on the order of
millions of years.

     Estimates of the quantities  of mercury entering the atmosphere from
degassing of the surface of the planet vary widely, but a commonly
quoted figure is 30,000 tons yr-1 (Table 6-1). Estimates of the
proportion of the mercury  in the  atmosphere due to anthropogenic sources
vary greatly; figures from 10 percent to 80 percent of atmospheric
mercury have been credited to man.  Estimates of the yearly amount of
mercury finding its  way to the ocean  indicate that atmospheric
deposition accounts  for the major amount,  approximately 11.000 tons
yr-1, with land runoff accounting for about 5,000 tons yr~*.

     The measurement of mercury in  extremely  low environmental
concentrations is frequently close  to the  limit of detection of many
current methods.  With this caveat, it would  appear that the vastly
predominant reservoir for  mercury is  the ocean water, containing on the
order of 40 million  tons (Table 6-2).   In  contrast, the atmosphere and
fresh water contain  much less.  As  one might  expect, therefore, the
impact of man-made release of mercury  is much greater on these smaller
reservoirs, especially those to which man-made release is direct.  Thus,
the impact on levels of atmospheric mercury and mercury in fresh waters
is appreciable, whereas it is estimated that  oceanic concentrations have
not appreciably changed in recent history. For example, it is estimated
that the mercury content of lakes and rivers  may be increased by a
factor of 2 to 4 due to man-made  release  (Nriagu 1979).

6.2.2.3.2  Biogeochemical  cycles  of mercury.   This overall global cycle
of mercury results from extremely complex  physical, chemical, and
biochemical processes occurring in  the main reservoirs and interfaces
between these reservoirs.   Most of these  processes are poorly
understood; nevertheless,  certain very import fundamental discoveries
have been made in recent years and are summarized below.

     The most important single discovery  in  understanding the chemical
and biochemical cycles of  mercury in  the  environment was made by Swedish
investigators in the 1960s (for a review  see  NAS  1978, Nriagu 1979).   An
intensive investigation into the  source of the methyl mercury compound


                                  6-6

-------
TABLE 6-2.  THE AMOUNT OF MERCURY  IN SOME GLOBAL  RESERVOIRS  (NAS  1978)
     Reservoir                          Mercury  Content
                                         (metric tons)
     Atmosphere                                850
     Fresh water                              2,000
     Freshwater biota                          400
     Ocean water                        41,000,000
     Oceanic biota                          200,000
                                 6-7

-------
in freshwater fish revealed that microbial  activity  in  aquatic  sediments
can result in the methylation of inorganic  mercury  (Jensen and  Jernelov
1967).  The most probable mechanism  involves  the  non-enzymatic
methylation of mercuric mercury ions  by  methyl-carboning compounds
(Vitamin 612) that are produced as a  result of  bacterial synthesis.
However, other pathways both enzymatic and  nonenzymatic may  play a role
(Beijer and Jernelov 1979).

     The methylation of ionic mercury in the  environment appears to
occur under a variety of conditions:   in both aerobic and anaerobic
waters; in the presence of various types of microbial populations, both
anaerobes and aerobes; and in different  types of  freshwater  bodies such
as both eutrophic and oligotrophic lakes.

     The methylation of mercury can  result  in a formation of either
monomethyl  or dimethyl mercury compounds (Figure  6-1).  The  monmethyl
mercury compound is avidly accumulated by fish  and  shellfish, whereas
the dimethyl  compound, having a low  solubility  and high volatility,
tends to vaporize from the water phase to the atmosphere where  it may be
subjected to photolytic decomposition (Figure 6-1).

     However, these reactions are not understood  in  detail and  there
does not appear to be general agreement  in  the  literature as to those
conditions that favor the formation  of monomethyl or the dimethyl form;
neither is there complete agreement  as to the extent that the dimethyl
species actually vaporizes from the  water phase into the atmosphere.

     Methyl mercury compounds are subject to  decomposition in the water
phase probably by the action of a variety of  microorganisms.  These
demethylation microbes appear to be  widespread  in the environment,
occuring in water sediments and soils and in  the  gastrointestinal tract
of mammals, including humans.  This  biogeochemical cycle involving
bacterial methylation and demethyl ation  is  part of a more general cycle
of mercury that describes global  transport  of mercury.  Professor
Brosset and colleagues (KHM 1981) have described  a large-scale  cycle
that has the following aspects.

     1)  Mercury is introduced to the atmosphere  from the ground and
         water surfaces.   It occurs  primarily in  the form of mercury
         vapor (Hg°).

     2)  The total concentration of  mercury diminishes  while the
         proportion of water soluble mercury  increases  as a  function of
         height over the ground.   The origin  of the  soluble  mercury is
         not yet completely understood.

     3)  Water soluble mercury is deposited in  wet  and  dry forms in the
         water phase of terrestrial  and  aquatic systems and  probably in
         other phases if the mercury  compounds  are  soluble in those
         phases.
                                  6-8

-------



m


CH4 +
\
(CH,
J
FISH
i
1 +
^ CH~Hu — "^ (tH0
BACTERIA 3 3 BACTERIA ;
^^s^ '
°+ to.
Hn -f Ha

4H6- ;
5)2Hg
i
SHELLFISH
CH3SHgCH3
1
t
pHg CrUS-HgCHo
"+ + He
BACTERIA
\o
1 AIR
WATER

SEDIMENT

3
Figure 6-1.  The mercury cycle, demonstrating chemical  transformation
              by chemical  and biological  processes and  the accumulation
              of monomethyl  mercury by fish.  Adapted from NAS  (1978).
                                 6-9

-------
     4)  The deposited forms  of water soluble mercury, once in the water
         or terrestrial  phase,  partly undergo reduction to Hg°, and are
         partly absorbed temporarily  or permanently on sediments.

     5)  The rates of deposition into and removal  from the water phases
         determine the steady state levels of each mercury species in
         water.

     6)  The concentration of each mercury species in the water phase
         determines the concentration on the sediment in contact with
         the water phase.

     7)  The reduction product Hg° returns (i.e.,  is re-emitted) to the
         atmosphere.

     Neither the detailed  chemical mechanisms nor the kinetics of these
processes are understood at this time;  for example, the extent to which
mercury may be deposited and  remitted from water or land surfaces to the
atmosphere is still not understood in quantitative terms.  Nevertheless,
the general picture that emerges is one in which long distance transport
of mercury in the vapor phase is possible, its  deposition in water and
remission probably occurs  extensively,  and the  chemical conversion of
mercury from the elemental to the ionic and to  the organic forms is much
more extensive than was originally believed.  Therefore, methyl mercury
may occur not only as a result of microbial action in aquatic sediments
as indicated in Figure 6-1 but may have a more  general source, including
the atmosphere.

6.2.3  Accumulation in Fish (T. W. Clarkson and J. P. Baker)

     Once methyl mercury enters the water phase as a soluble compound,
it is rapidly accumulated  by  most aquatic biota and attains highest
concentrations in the tissues of large carnivorous fish.  Indeed, it is
generally believed that the major amount of methyl mercury compounds in
bodies of water are contained in the  biomass of the system. The
bioconcentration factors,  that is, the ratio of the concentration of
methyl  mercury in fish tissue to concentrations in fresh water can be
extremely large, usually of the order of 10,000 to 100,000 (U.S. EPA
1980a).

     In principle, fish can accumulate methyl mercury both directly from
water and from the food supply.  Hultberg and Hasselrot (1981) have
reviewed available data and suggested that pike obtain virtually all
their methyl mercury from  their food  supply.  Methyl mercury
concentrations correlate well between different trophic levels of fish
and other aquatic organisms,  implying the importance of the food chain.
In a survey of several lakes, levels  of methyl  mercury in pike were
closely correlated (r = 0.92) with methyl  mercury  concentration in
plankton (Hultberg and Hasselrot 1981).  Thus factors that affect
bioaccumulation of methyl  mercury at  this early stage of the food chain
should also affect methyl  mercury levels at the highest level  (e.g., in
predatory fish).
                                  6-10

-------
     The concentration of methyl  mercury  in fish  tissue  is  of  special
interest in terms of human exposure.   Bioaccumulation  of methyl mercury
in fish is the main if not the sole source of human  exposure,  barring
episodes of accidental discharge or misuse of man-made methyl  mercury
compounds.  Thus, factors that affect concentrations of  methyl mercury
in edible fish tissue are of considerable  importance  in assessing
potential human health risks from this form of mercury.

6.2.3.1  Factors Affecting Mercury Concentrations in Fish--In  general,
for any body of water one might expect to see an  eventual steady-state
distribution of methyl mercury—a balance of synthetic and  degradation
reactions.  Concentrations of methyl  mercury in sediment, water, and
biomass at steady-state are influenced by a wide  variety of experimental
conditions, perhaps only a few of which have so far  been identified.  No
detailed review will be given in this chapter,  but the reader  is
referred to other references that give a  more specific treatment of this
topic (Nriagu 1979).

     Theoretical considerations,  experimental  data,  and  observations in
field studies have indicated or suggested that methyl  mercury
concentrations in fish are affected by:   (1)  the  species of fish, (2)
the age of the fish, (3) concentrations of mercury in  surface  sediments
and/or in water, (4) the biomass or biomass production index,  (5)
salinity, (6) concentrations of dissolved organics,  (7)  the microbial
activity associated with sediments, (8) the degrees  of oxygenation of
water and redox potential, (9) the pH and/or alkalinity  of  water
(Hultberg and Hasselrot 1981, Jensen and  Jernelov 1972,  Fagerstrom and
Jernelov 1971, Jernelov 1980).  This list is not  exhaustive and, indeed,
recent evidence suggests that other as yet unknown factors  are involved
(for discussion see Hultberg and Hasselrot 1981).  In  view  of  the
current interest in the relationship between the  use of  fossil fuels,
particularly coal, and possible acidification of  large bodies  of fresh
water, the influence of aquatic pH on levels of methyl mercury in fish
will  be given special  attention here.

     An indirect result of acidification  of surface  waters  may be
increased accumulation of mercury (and perhaps other metals) in fish.
Evidence for this relationship derives from correlations between metal
concentrations in fish and lake and stream pH levels,  and evalutions of
metal chemistry and availability in oligotrophic,  acidic waters.

     Elevated levels of mercury in fish from acidic  waters  have been
measured in Sweden, Norway, Ontario,  and  the Adirondack  region of New
York (Hultberg and Hasselrot 1981, Overrein et al. 1980,  Suns  et al.
1980, Jernelov 1980, Schofield 1978).  In each case, although  fish
mercury content was statistically correlated with  pH level, the data
points still exhibit significant scatter.  At any  particular pH level,
for a given age and species of fish,  the  range observed  between lakes in
values of mg Hg kg~l flesh was considerable,  even to the extent that
not all lakes with low pH exhibited elevated mercury concentrations in
fish and some lakes without low pH had fish with  high  mercury  content.
Obviously, other factors in addition to pH control the accumulation of


                                  6-11

-------
mercury in fish as noted  above.  Waters of low productivity
(oligotrophic lakes)  and  low alkalinity tend to be more sensitive to
mercury contamination and mercury accumulation in fish.  Because these
conditions are also strongly associated with low pH levels, the effect
of pH on mercury bioaccumulation may be somewhat confounded.  The
correlation between pH and fish mercury content may in part be a result
of the observation that low pH waters tend to be oligotrophic, soft
waters with low alkalinities.  On the other hand, the association
between low alkalinity and elevated mercury accumulation and low pH
waters have low alkalinities.  Results from these correlations must be
interpreted carefully.

     The most extensive studies on factors controlling mercury levels in
fish have been carried out in Sweden.  In the 1960's pike and other
edible fish were found to have unacceptably high levels of mercury
(greater than 1 yg Hg g'M.  For some lakes, local industrial
"mercury emitters" with direct outlets to the lakes were identified as
the cause.  Many lakes, however, had inexplicably high mercury levels in
fish.  This led to extensive studies in Sweden on the dynamics of
mercury chemistry and uptake by fish and the role of acidity in these
processes.

     Data collected by Jernelov et al. (1975), Grahn et al. (1976),
Landner and Larsson (1972), and Hultberg and Jernelov (1976), as
reported by Jernelov  (1980), all indicated an overall strong correlation
between mercury levels in fish and pH values of lakes.  Jernelov (1980)
concluded that in Swedish lakes in general, extremely few lakes with pH
values below 5.0 have pike (weighing 1 kg) with mercury concentrations
of less than 1 mg kg"1.  At a pH value of 6.0, the normal level for
the same pike would be approximately 0.6 mg kg'1-

     Hultberg and Hasselrot (1981) reviewed ten years of Swedish work on
factors affecting mercury in fish.  In a study involving over 152
Swedish lakes mercury level in pike muscle was inversely correlated with
water pH (Figure 6-2). Water samples collected during the fall overturn
were analyzed for pH, humic material (water color at an adjusted pH) and
specific conductivity (salt content).  Multiple linear regression
analysis (Table 6-3)  suggested that a one unit decrease in pH would
elevate mercury in the muscle tissue of pike (weighing 1 kg) by 0.14
ppm.  The influence of pH on fish mercury content was generally greater
than that associated  with humic content or conductivity.

     Hakanson (1980), also using the Swedish data base, developed (based
on a combination of statistics and deductive reasoning) a quantitative
model expressing mercury  content in a 1-kg pike as a function of pH, the
mercury content in the top one cm of lake sediments, and a bioproduction
index.  The model was validated using an independent data set from 107
Swedish lakes.  The correlation coefficient between observed and
predicted mercury content was 0.79.
                                 6-12

-------
CO
                 cu

                 ro  o
                    cr —i
                       3
                       ro
                    ro  QJ
                    3  -s
                                   o
    -j. 
 m 3 eu
 c  ro 01

 rt- CU  I
 cr o to
 ro  o -Q
 -s  o c:
«2  -S Co
    Q. -S
 co  ->• ro
 3  3 01
^ <•+ ro
Co O tQ
01    -s
to cu ro
ro    01
—i 3" O1

o -a o
<-+• ro 3

^-~~ cr cu
h-> O 3
^O —' Co
co -•• —•
I—' 3 <<
	    to
.  -+, _i.
      to
               -br+
            —' c:  -"•
            —' 3  o
            CU O  3
         -S 3 ch 01
         o a. -••
         ro  i  o  -1-
            CO 3  3
         O 3
         -h eu o  3
s CO
ro 3
-s o-
o
c o

3 ~*
•  co

   ro
—i a.
3- ro
ro 3
   •
01

^ —i
Cu 3"
-•• ro
ia 01
3- ro
r+
               — I (D
               3- -S
               ro o
                  c
               to -s
                               C
                               Q. O
                               ^< o

                               so
                               a> ro
                               to 3
                                  <-*•
                               o -s
                               o a>
   _       j> Q. -••
   3    -"• 7T C. O
   o co 3 ro o 3

   -•• Q.       ro
   o    s 3" a. ->•
   3 rt CU cu    3
   .  3" 01 < -i.
      ro    ro 3 -a
         a_       -j.
   o o -5 3 —i TT
   co c co o cu ro
   r»- -s S    A-
   cu < 3 TC ro 3
      ro    3 01 c
      o ^T ^3    (/)
         ^ s -•• o
            33—"
                  ro
                                            METHYL MERCURY  CONCENTRATION  IN  PIKE  MUSCLE

                                                                  Hg  g-1  wet wt.)
                                   -P>O
                                                    on
                                                    •
                                                    o
                                                    en
 73    in

T3    00
                                    PO
                                                    CT>
                                                    •

                                                    CTl
                                                         o
                                                         •
                                                         en
                                                                                                                    ro
                                                                                                                    •
                                                                                                                    o

-------
   TABLE 6-3.   THE RESULTS OF  A  STATISTICAL ANALYSIS  INDICATING THE
    CONTRIBUTION OF pH,  HUMIC  CONTENT AND  SPECIFIC CONDUCTIVITY
  TO METHYL MERCURY CONCENTRATIONS  IN THE  MUSCLE TISSUE OF 1 KG PIKE
              (ADAPTED FROM HULTBERG AND HUSSELROT 1981)
Change in water pH
one pH unit
two pH units
three pH units
Change in mercury concentration
          mg Hg kg-1

              0.14
              0.28
              0.42
Color increase
 10 mg Pt
 50 mg Pt £
100 mg Pt rj
           -1
              0.015
              0.075
              0.150
Change in specific conductivity
 5 tnS nrl
10 mS nr1
20 mS m-1
              0.075
              0.150
              0.300
                                 6-14

-------
     Hakanson's formula was as follows:


                     4.8 x log (1  +  H950)
             F(Hg)  = 	200
                      (pH-2)  x log(BPI)


     where

      F(Hg) = the concentration of methyl  mercury  in  a  1  kg  pike  in
              pg g~l wet weight,

       Hgso = tne weighted mean mercury content of surface sediments,
              0 to 1 cm, in ng Hg  g"1 ds  (ds  =  dry substance),

         pH = the mean pH of the water system,  i.e.,  the  mean of
              at least five measurements  of which  two should be
              obtained at different  seasons,  and

        BPI = Bioproduction Index  -  for details, see  Hakanson (1980).

     Calculations based on the Hakanson formula yield results similar  to
those from Hultberg and Hasselrot  (1981).   For  example, if it is  assumed
that a 1 kg pike at pH 6.0 contains  0.75  ppm  Hg (e.g.,  Figure 6-2),  then
a pH change from 6.0 to 5.0 would  increase fish mercury concentration  by
approximately 0.13 ppm.  Overlap in  the data  bases used by both Hakanson
and Hultberg-Hasselrot may have occurred,  however, accounting in  part
for this close agreement.

     If the Hakanson formula is valid, then a question  might be raised
on the appropriateness of linear regression analyses  relating pH  to
mercury concentration (e.g., in Figure 6-2 and  the multilinear analysis
used for Table 6-3).  The Hakanson formula has  the general form of a
rectangular hyperbole:

          FHg =
                pH - 3

where Hg   and BPI are constant.
Regression analysis of the data in Figure 6-3 according to a  hyperbolic
equation yielded a value of the correlation coefficient (r2 = 0.81)
appreciably higher than that obtained by linear regression analysis
(r2 = 0.3).  Thus, for the Swedish study, .change in pH accounted for
about 80 percent of the total  variance in methyl  mercury concentrations
in 1 kg pike.  The hyperbolic aspects will become more pronounced at
lower pH values and will be discussed later with regard to apparent
scatter of points around linear regression lines.

     Additional and as yet unknown factors seem to be operative in
determining mercury concentrations in fish.  For example, Hultberg and
                                  6-15

-------
              200
          ~  140-
          cc
          r>
          <_>
          a:
              40-
                     4.5
                                           PH
                          1. DUCK LAKE
                          2. LITTLE CLEAR LAKE
                          3. HARP LAKE
                          4. BIGWIND LAKE
                          5. NELSON LAKE
                          6. CHUB LAKE
                          7. CROSSON LAKE
                          8. DICKIE LAKE
                          9. LEONARD LAKE
10.
11.
12.
13.
14.
15.
16.
17.
HENEY LAKE
CRANBERRY LAKE
HEALEY LAKE
CLEAR LAKE
FAWN LAKE
BRANDY LAKE
MCKAY LAKE
LEECH LAKE
18. MOOT LAKE
Figure 6-3.   Mercury concentrations  in yearling yellow  perch  and
               epilimnetic  pH in  lakes  in the  Muskoka-Haliburton  area
               of  Ontario  (Suns et al.  1980, U.S./Canada  1983).
                                        6-16

-------
Hasselrot (1981) noted that lakes In more northern regions of Sweden
tend to have higher concentrations of mercury  in pike.  Possible
explanations include 1)  the impact of snow on  water quality during the
spring melt, 2)  loss of  sensitive prey species (in this case roach,
Rutihis rutilis) adversely  affected during acid episodes during spring
melt and a shift to predation  on  higher trophic levels (in this case
perch, Percas gluvicotilis)  that  contain greater amounts of mercury, 3)
the importance of snow itself  as  a source of mercury including methyl
mercury (Brouzes et al.  (1977)  and 4)  lower water temperature and
salinity generally associated  with northern latitudes.

     In Norway,  concentrations  of mercury in muscle of trout, perch,
char, and pike were studied by  Muniz,  Rosseland, and Paus (Overrein et
al 1980).  Again, fish populations in  acidic waters generally had higher
levels of mercury than did  reference populations from areas without
acidified lakes.

     Studies in  Canada (Suns et al.  1980)  have also found a
statistically significant (r =  0.65, p < 0.05)  inverse correlation
between water acidity and mercury levels in fish, for yearling perch in
14 pre-cambrian  lakes in Ontario  (Figure 6-3).* For lakes with similar
pH, mercury levels were  higher  in fish from lakes with a higher drainage
area/lake volume ratio.

     Suns et al. (1980)  failed  to see  a relationship between mercury in
fish and water alkalinity,  whereas Scheider et al. (1979) reported that
for walleye (Stigostedion yitreum)  of  equal length, caught in Ontario
lakes with alkaline water (<_ 15 mg CaCOs &~1)  had significantly
higher mercury levels than  walleye caught in lakes with high alkalinity
(> 15 mg CaC03 £  )•  Comparisons based on fish length may,
however, be somewhat misleading.   If fish from waters with lower
alkalinity grow  slower (possibly  as  a  result of lower primary
prooductivity or lower temperatures),  than the higher mercury content at
a given length may actually only  reflect the older age of the fish.

     Statistical evaluations of mercury in fish and water acidity have
not been published for fresh water fish caught in the United States.  A
graph of mercury levels  in  brook  trout muscle  as a function of fish
length for Adirondack lakes indicated  that fish from acid drainage lakes
(pH < 5.0) in general had higher  mercury levels (for a given length)
than fish from limed, seepage,  or bog  lakes (Schofield 1978).  However,
high mercury level in fish  were also found in  some lakes without low pH,
indicating that the unusual  mercury  bioaccumulation may be, in part or
in total, independent of pH level.   From 1969  to 1972, over 3500 fish
from New York State lakes and  streams  were collected and analyzed for
mercury content.  Two Adirondack  Mountain reservoirs, Cranberry Lake and
Stillwater Reservoir, yielded  fish with particularly high mercury
levels, despite  the undeveloped nature of their watersheds (Blcornfield
et al. 1980) (Figure 6-4).   The highest mercury levels in the New York
State survey were recorded  for  Onondaga Lake fish and caused, for the
most part, by mercury contributions from a chlor-alkali plant during the
1960s.  Other areas, such as. Lake Ontario, Lake Champlain, and the St.
                                  6-17

-------
      250


     2.25


     2.00


  ^ 1.75
  e-H
  O>
  o> 1.50


  R 1.25
  g 1.00
     0.75

     0.50

     0.25

        0
           =20
     LEGEND
.+	   ADIRONDACK LAKES
-o—   ST.  LAWERENCE  RIVER
-*	   ONONDAGA LAKE  (No Fish>35 cm)
-•	   LAKE ONTARIO
-6	   LAKE CHAMPLAIN
-A	   LAKE ERIE - NIAGRA RIVER
        CRANBERRY LAKE and
        STILLWATER RESERVOIR
     20-25
25-30
30-35
35-40
                                  LENGTH  OF  FISH  (cm)
Figure 6-4.   Average mercury concentration vs length class in New York
             State smallmouth bass.  Adapted from Bloomfield et al.  (1980)
                                   6-18

-------
Lawrence River also either have or have had in the past sources  of
direct contamination, yet mercury levels found in  fish  from  these areas
were lower than in fish from Stillwater Reservoir  and Cranberry  Lake.
Other Adirondack lake samples also tended to have  lower fish mercury
levels.  The source of mercury to Cranberry Lake and Stillwater
Reservoir remains unresolved.  Stillwater Reservoir is  an  acidic (pH  <
5.0) clear water lake.  Cranberry Lake, on the other hand, had a mean pH
of 6.90 in the fall of 1978 (Bloomfield et al. 1980).   Cranberry Lake
waters are, however, relatively highly colored with high concentrations
of humic organics.

     In summary, field studies in Sweden, Norway,  and Canada have
identified several factors that correlate (positively or negatively)
with mercury levels in fish.  This includes fish species and age (length
and weight are frequently used instead of age), mercury levels in
surface sediments, the biomass or bioproductivity  of the lake, the
salinity (specific conductivity) and pH.  Other factors may  also be
operative, such as morphometric parameters (drainage area/lake volume
ratios) and geographic (northern latitude).  However, in virtually all
such studies published to date, elevated mercury levels in fish  muscle
(most notably the pike and the perch)  have been statistically associated
with higher levels of acidity.

     However, a number of factors influencing mercury levels in  fish  may
also change in parallel with acidity.   Thus, a true cause-effect
relationship between acidity and elevated mercury  in fish  has not been
established by the available data.  Absolute proof may  be  unattainable
in field studies, given the large number of variables and  the
probability that, in any given field study, not all of  these will be
controlled or even measured.

     To resolve whether correlations observed between lake pH level and
mercury content in fish actually reflect a cause-and-effect  relationship
and whether acidification will enhance bioaccumulation  of  mercury, the
effects of pH and acidity on mercury chemistry, mobilization, and uptake
must be understood.  Field and laboratory research on mercury cycles
have resulted in several proposed mechanisms (Jernelov  1980, Wood 1980,
Haines 1981):

     1)  Acidic precipitation may scavenge mercury from the  atmosphere
         more effectively than nonacidic precipitation.

     2)  The rate of methylation of inorganic mercury by microorganisms
         is pH-dependent, the maximum occurring at pH 6.0--methylation
         is higher from pH 5.0 to 7.0 than above 7.0.  Thus, at  lower pH
         more methyl mercury would be present and, because methyl
         mercury is the form most rapidly taken up by fish,
         bioaccumulation presumably would be enhanced.
                                  6-19

-------
     3)  Low pH levels favor the formation of monomethyl  mercury rather
         than dimethyl mercury.  Dimethyl  mercury is unstable and
         volatile and thus more quickly lost from the aquatic system
         (Figure 6-1).

     4)  Under aerobic conditions, inorganic mercury is more soluble at
         reduced pH and thus more available for methylation reactions.
         Retention of mercury in the water column is enhanced with
         increased acidity (Jackson et al. 1980), thus increasing the
         exposure of fish to mercury.

     5)  Since the biomass of fish is often lower in acidic lakes, the
         available mercury is concentrated in a smaller biomass,
         resulting in higher body burdens  per fish.   Also,  if growth
         rate is reduced, fish in an acidic lake would be older than
         fish of an equivalent size in a nonacidic lake and would have
         been accumulating mercury longer.

     Laboratory experiments will be useful, if not essential, in order
to unravel  mechanisms associating pH change with mercury accumulation in
fish.  Laboratory experiments have shown that, for a given  amount of
total mercury in an aquatic ecosystem, higher levels of mercury were
found in fish at low pH values than at high pH values (for  review, see
Jernelov 1980).

     Miller and Akagi (1979) presented experimental  evidence that low pH
levels mobilize methyl mercury absorbed on sediments.  Natural  water
from the Ottawa River was incubated with various types of sediment
materials for periods of approximately three weeks.   Irrespective of the
type of sediment, a reduction in water pH  shifted, by a factor of 2  for
each unit change in pH, the distribution of methyl mercury  from the
sediment to the water phase (Figure 6-5).   Miller and Akagi  (1979)
asserted that the effect of pH on the equilibrium of methyl  mercury
between water and sediment, may be the principal  factor responsible  for
higher levels of mercury in fish in low pH aquatic environments.

     That acidification of surface waters  will significantly enhance
bioaccumulation of mercury has not been definitively demonstrated.   The
chemistry and environmental sampling of mercury are extremely complex.
More research is needed to identify all  factors that affect mercury
accumulation in fish and the relative importance of each.   The
significance of a one unit pH decrease (or a decline in alkalinity by
100 yeq rl) relative to the effects of the large number of other
factors that influence bioaccumulation has not been  quantified.   This
need is especially urgent in the United States, where few data are
available at this time.

     Other  metals in addition to mercury occur at elevated
concentrations in acidic waters and potentially may  accumulate in fish
and other biota.  Data on these accumulations are, however,  very
limited.   Dickson (1980)  reported that concentrations of  cadmium ta  pike
increased with increased acidity.   Harvey  et al.  (1982) determined

       *
                                  6-20

-------
     150
%£>  100
  OJ
El
1     50
                                         LEGEND

                                     H SAND

                                     E3 SAND CHIP  SEDIMENT

                                     D WOOD CHIPS
                                           MONOMETHYL MERCURY
                                              •I
                                  6

                                 pH
Figure 6-5.
            The partition coefficient of methyl mercury between water
            and three different types of sediments.  The units of the
            ordinate have been multiplied by  a  factor of 10,000.  The
            data are taken from Miller and Akagi  (1979).
                                 6-21

-------
manganese concentrations in  the  vertebrae of white suckers from six
lakes in sourthern Ontario.   Fish from the most acidic lake, George Lake
(pH 4.65) had particularly high  manganese content.  The remaining five
lakes had pH levels from 5.02  to 6.59, and fish manganese level appeared
relatively independent of pH.  George Lake also had aqueous manganese
concentrations that were 50  percent greater than in any of the other
lakes.  The Ontario Ministry of  Environment (U.S./Canada 1983) analyzed
yearling yellow perch for body burdens of lead, cadmium, aluminum, and
manganese in 14 Ontario lakes (Figure 6-6). Lead (p < 0.01) and cadmium
(p < 0.05) were significantly correlated with lake pH level.  No data
are available to evaluate the environmental significance of these
accumulations.  No correlations  between lake acidity and body levels of
aluminum or manganese were evident.  Aluminum has, however, been
observed to accumulate on gills  of fish during fish kills in Plastic
Lake, Ontario, and in two lakes  in Sweden.  Grahn (1980) measured 40 to
47 yg Al g"1 wet weight of tissue on gills from dead ciscoe from
lakes Ransjon and Amten, Sweden, but only 6 yg Al g"1 for fish from
reference lakes without fish kills.  Aluminum concentrations on fish
gills from dead and moribund pumpkinseed and sunfish from Plastic Lake
ranged from 83 to 250 mg g'1 dry weight (Harvey et al. 1982).

6.2.3.2  Historical and Geographic Trends in Mercury Levels in
Freshwater Fish—Presently it is difficult to assess quantitatively the
contribution of acidic deposition to elevations of mercury
concentrations in freshwater fish.  The problem in part is a lack of
data showing temporal and regional changes in mercury as related to
water pH and in part due to  the  operation of other processes affecting
mercury levels in fish.

     81oomfield et al. (1980)  have reviewed the results of an extensive
mercury screening involving  some 3500 freshwater fish collected in New
York State from 1960 to 1972.  Less than 10 percent of the fish had
mercury levels in excess of  the  current federal guideline of 1.0 ppm.  A
sizeable portion of the high mercury fish came from Onondaga Lake—
known to be polluted by a local  industrial source of mercury.  Predatory
species of fish such as walleye, pike, and smallmouth bass had levels
sometimes exceeding 1 ppm in certain Adirondack Lakes remote from known
sources of mercury.  Bloomfield  et al. (1980) quote unpublished work
indicating that concentrations in smallmouth bass were still high in
1975, and Armstrong and Sloan (1980) reported elevated mercury levels in
predatory fish species collected in certain Adirondak Lakes (Cranberry,
Great Sacandaga, Raquette) in 1978.  In contrast, fish from rivers and
lakes previously contaminated with mercury now show declining fish
levels (Armstrong and Sloan  1980).  For example, following cessation of
mercury discharge, levels of mercury in smallmouth bass in Lake Onondaga
declined by 55 percent over  the  period 1972 to 1978.  The Ontario
Ministry of Environment (1977) has reported substantial declines in
mercury in fish caught in Lake St. Clair following curtailment of
industrial discharge of mercury.
                                  6-22

-------
               520
     500



     400
7  300
 o>
 en
 c
 .0
 O.
    200
     100
                          J	L
    250



    200
7   150
 CO
 CO
 E
 T3
 O
100



 50



  0
 co
 CD
      25
      20
      15
 S    10
       0
       4.5   5.5    6.5    7.5   8.5

                    pH
     25
     20
  ^  15
  co


  S 10
      0
      4.5   5.5   6.5    7.5   8.5

                   PH
Figure 6-6.  Metal concentrations in yearling yellow perch  and
             epilimnetic pH in 1981 in lakes in the Muskoka-Haliburton
             area of Ontario (U.S./Canada 1983).
                                   6-23

-------
     Based on very limited data  in  the United States, a general picture
emerges of declining mercury  levels in freshwater fish caught in areas
where direct discharge of mercury has been  curtailed but of continued
high levels of mercury in certain lakes  remote from industrial activity.
Reasons for these high mercury levels are being investigated  (Section
6.2.2.3).  Wet deposition of  mercury from the atmosphere has been shown
to occur in several Adirondack Lakes.  These lakes, in general, are
characterized by low pH and low  alkalinity.  The role of long distance
transport of mercury and lake acidification merits careful
investigation.

6.2.4  Dynamics and Toxicity  in  Humans (Mercury)

6.2.4.1  Dynamics in Man (Mercury)--The  U.S. EPA (1980a) has reviewed
information on uptake, distribution, and excretion of methyl mercury In
man.  Meth/l  mercury is almost completely absorbed from the diet (90 to
100 percent,  i.e., between 90 to 100 percent of the amount ingested is
absorbed).  After absorption  in  the gastrointestinal tract, methyl
mercury passes into the bloodstream and  is  distributed to all organs in
the body.  Approximately 5 percent  of the absorbed dose goes to the
blood compartment and 10 percent to the  brain—the target organ for
toxic effects.

     After the initial distribution is completed, usually a matter of a
few days in man, the brain to blood concentration ratio is roughly
constant, having value between 5:1  to 10:1.  Methyl mercury is
accumulated in growing hair.   At the time of formation of the head hair,
the ratio of the concentration of mercury in hair to the simultaneous
concentration in blood is roughly constant  and has an average value of
about 250:1.   Once incorporated  into the hair, the mercury concentration
remains constant.  Because human head hair  grows about 1 centimeter per
month, analyzing centimeter segments of  hair can recapitulate average
monthly concentrations of methyl mercury in blood.  Measurements of
mercury in samples of blood or hair are  now routinely used to assess the
body burden of methyl  mercury in humans  and as an indicator of brain
concentrations.

     Methyl mercury is excreted  from the body mainly in feces.  Before
excretion in the feces, methyl mercury is converted into inorganic
mercury.  The site of this conversion is not known, but microflora in
the lower gut are known to possess  this  capability.  The rate of
elimination from the body is  directly proportional to the body burden.
It is well described by a single exponential function characterized by a
half-time of about 70 days.  An  important conclusion from this kinetic
information is that it will take about one  year for humans to attain a
state of balance, i.e., to attain maximum steady body burden of methyl
mercury for any given daily intake  in the diet.  After cessation of
given exposure, it will take  one year for the body burden to fall to
pre-exposure levels.  Thus, dietary intake  of methyl mercury  from fish
should be evaluated over a matter of months.  Intake on any one single
day does not normally make an important  contribution to the overall body
burden.
                                  6-24

-------
     Considerable individual  differences exist in biological half-times
in man although the average  value  is 70 days with a range of 30 to 180
days.  The distribution is bimodal  with 90 percent of the values
distributed about an average  value  of about 65 days and 10 percent
distributed about an average  value  of 120 days.  The reasons for this
wide range of biological  half-times are not known, except that lactating
women have a short half-time  averaging about 40 days.

     Methyl mercury readily  crosses the placental barrier and enters the
fetus.  It distributes to all  tissues in the fetus, including the fetal
brain, which is the principal  target for prenatal toxicity of methyl
mercury.  Levels of methyl mercury  in cord blood are usually higher than
the maternal blood concentrations.

     Methyl mercury is secreted in  milk.  Thus body burdens of methyl
mercury acquired by the infant before birth may be maintained by breast
feeding if the nursing mother continues to be exposed to methyl mercury.

     The rate of elimination of methyl mercury from the human fetus and
suckling infant is not known.   Experiments on animals indicate that
elimination in suckling animals is much slower than in adults.  The
adult rate of excretion appears to  commence at the end of the suckling
period.

     In brief, methyl  mercury accumulates in the human body over a
period of about one year.  Blood and hair analyses may be used as
indicators of human absorption of mercury.   In assessing hazard to human
health, chronic exposure over weeks or months is important.

6.2.4.2  Toxicity in Man—Methyl mercury damages primarily the human
central nervous system.  When ingested  in sufficient amounts, methyl
mercury destroys neuronal cells in certain areas of the brain, the
cerebellum and the visual cortex,  resulting  in permanent loss of
function.  Symptoms of damage include loss of sensation, constriction of
the visual fields, and impairment of  hearing.  Coordination functions of
the brain are also damaged,  leading to  ataxia and  dysarthria.  Severest
damage causes mental incapacitation,  coma, and death.  The mildest and
earliest effect in adults is usually  a  complaint of paresthesia, an
unusual sensation in the extremities  and around  the mouth.  In the
Japanese population poisoned by methyl  mercury from contaminated fish,
paresthesia was usually permanent.   In  the  Iraqi population, paresthesia
was frequently reported to be transient.  This population had consumed
homemade bread from wheat contaminated  with  a methyl mercury fungicide.

     The effects on the fetal brain differ  qualitatively from  those  seen
in adults.  Methyl mercury interferes with  the normal growing  processes
of the brain.  It inhibits migration  of neuronal cells  to their final
destination, thus affecting the brain's architecture.   This damage
manifests  itself as diminished head size  (microcephaly)  and gross
neurological manifestations such as cerebral palsy.  The mildest effects
are delayed achievement of developmental milestones  in  children and  the
presence of abnormal reflexes and mild  seizures.


                                  6-25

-------
     Brain concentrations  associated with the onset of human methyl
mercury poisoning are  in the range of 1 to 5 yg Hg g"1 wet tissue.
Blood concentrations for the onset of the mildest effects have been
established to be between  200 and 500 ng Hg m -1 whole blood.
Corresponding hair concentrations would be 50 to 125 yg Hg g-1 hair
(Table 6-4).  The chronic  daily intake of methyl mercury that would lead
to a miximum blood level of 200 ng m -1 has been established to be
300 yg Hg.  However, in the mother during pregnancy, the blood level
associated with the earliest damage to the fetus has not yet been
determined.

     The conclusions reported in Table 6-4 were based on observations of
affected populations in outbreaks of poisoning in Niigata, Japan and in
Iraq (the 1971-72 outbreak).  In effect, the numbers in Table 6-5 refer
to the lowest effect levels observed in an outbreak of poisoning from
methyl  mercury contaminated fish in Niigata, Japan (Swedish Expert Group
1971) and lowest effect levels estimated from an affected population in
the Iraqi outbreak of  1971-72 (Bakir et al. 1973).  With such low
observed effect levels on  humans, it is usual to apply a safety factor
of ten (WHO 1972a) to  arrive at an acceptable "safe" body burden or
"allowable daily intake."

     A direct estimation of absolute risks associated with a given
long-term daily intake of  methyl mercury was reported by Nordberg and
Strangert (1976, 1978).  In their approach they combined the data from
dose-response relationship published in the Iraqi outbreak (Bakir et al.
1973) with the range of biological half-times, also obtained in the
Iraqi outbreak (Shahristani and Shihab 1974) to calculate the
relationship depicted  in Figure 6-7.

     Their calculations indicated that an intake of 50 yg day1 in
an adult gives a risk  of about 0.3 percent of the symptom of
paresthesia, whereas an intake of 300 yg dayl would give a risk of
about 8 percent of symptoms of paresthesia.  As pointed out by Nordberg
and Strangert (1976),  the  background frequency of these non-specific
symptoms such as paresthesia plays a key role in determining the
accuracy of the estimates  of response of low frequencies.  They
estimated from the same Iraq data the background frequency of
paresthesia of 6.3 percent.  However, there is considerable uncertainty
in determining the precise value of the background frequency, and this
uncertainty becomes the dominant cause of error at low rates of
response.

     Since the studies on  the Iraqi outbreak, a major epidemic!ogical
study has been carried out in Northwestern Quebec on Cree Indians
exposed to methyl mercury  in freshwater fish (Methyl "Mercury Study Group
1980).  The authors claim  to find an association in men over age 30 and
women over age 40 of a set of neurological abnormalities and the
estimated exposure to  methyl mercury.  However, it should be pointed out
that this association  has  been seen by only four of seven observers who
                                 6-26

-------
  TABLE 6-4.   THE CONCENTRATIONS  OF  TOTAL MERCURY  IN INDICATOR MEDIA AND
        METHYL MERCURY ASSOCIATED WITH  THE  EARLIEST EFFECTS  IN THE
             MOST SENSITIVE  GROUP IN THE ADULT POPULATION3
                         (ADAPTED FROM  WHO  1976)
Concentrations in indicator media

    Blood                  Hair         Equivalent long-term daily intake
(ng mH)                (yg g-1)              (yg kg-1 body weight)
200 to 500                50 to 125                     3 to 7
aThe risk of the earliest effects  can  be expected to be between 3 to 8
 percent, i.e.,  between  3 to  8  percent of a  population having blood
 levels in the range 200 to 500 mg ml-1, or  hair levels between 50
 to 125 yg g-1 would be  expected to  be affected (for further
 details, see text).
                                  6-27

-------
       I/)
       z;
       o
       o.
       I/)
       LU
       a:
       LU
       a.
       X
                    0.1
                  0.4    1.0    2


               DAILY INTAKE (mg)
Figure 6-7.
The calculated relationship between frequency of

paresthesia in adults and long-term average daily intake of

methyl mercury.  The calculations were performed by

Nordberg and Strangert (1978).  The broken line is the

estimated background frequency of paresthesia in the

population.  Data are taken from publications on the
Iraqi outbreak of methyl  mercury poisoning  (Bakir

et al. 1973; Shahristani  & Shihab 1974).
                                  6-28

-------
reviewed video taped recordings  of  the neurological screening tests.
The severity of these neurological  abnormalities was assessed by
neurologists as mild or questionable.  It was not possible to estimate
any threshold body burden or hair levels because this population had
been exposed possibly for most of their lives; therefore peak values in
previous years are unknown.   However, observations on this population
over several years indicate  that maximum blood concentrations are below
600 ppb and most below 200 ppb (Wheat!ey 1979).  A WHO expert group
(1980), on examining the reports from these studies, raised the
possibility that this might  be the  first example of an endemic disease
due to exposure to methyl  mercury in freshwater fish.  However, another
epidemiological and clinical  study  of the same population of Cree
Indians failed to find any effects  associated with methyl mercury
(Kaufman, personal communication to EPA).

     The safety factor of ten applied to the lowest effect levels in
Table 6-4 was intended to take into account inter alia the greater
sensitivity of the fetus.  Since the WHO evaluation of 1976, data have
been published relating methyl mercury levels in the mother during
pregnancy to effects such as psychomotor retardation in the offspring
(Marsh et al. 1980).  These  data were the basis of a recent risk
estimate (Berlin 1982) relating  concentrations of mercury in maternal
hair to risk of mental retardation  in prenatally exposed infants (Figure
6-8).  Berlin calculated a background frequency in the Iraqi children of
approximately 4 percent as compared to a background frequency of mental
retardation in Sweden of 2 percent. He also noted that in the case of
adults that the error in determining background frequency is probably
the major source of error when researchers look at low rates of
responses.  Berlin calculated that  there was a risk of doubling the
background frequency of mental retardation at methyl mercury levels in
the mother on the order of 20 ppm in hair and a risk of a 50 percent
increase in background frequency at hair concentrations of about 10 ppm.

     The McGill Group (Methyl Mercury Study Group 1980) in their study
of Cree Indians exposed to methyl mercury in fish, found an association
"... between findings on examination of tone and reflexes in Cree boys
and the concentration of methyl  mercury in the mothers' hair during
pregnancy.  This association was shown at levels of methyl mercury
exposure which are very low  in relationship to those previously reported
to be associated with effects of methyl mercury in utero....  These
findings were isolated and the variation from normal was mild."  The
highest range of maternal  hair concentration was 13 to 23.9 g g~l.

     These hair levels overlap the  range estimated by Berlin associated
with the earliest detectable effects in Iraq.  However, the association
noted in the McGill study may have  been due to chance as their
observations on tone and reflexes were part of a number of observations,
the rest of which did not correlate with mercury levels.

     These observations on human infant-mother pairs agree with animal
data indicating the greater  sensitivity of prenatal life to methyl
mercury (for review, see Clarkson 1983).  However, the risk


                                 6-29

-------
       Q-
       O
       Q.
            20-
                           300
                                                         500
       LLJ
       O
            30
             20
             10
             0
                       10
                           30
50
Figure 6-8.
            MERCURY  IN HAIR  (ppm)

A dose-response relationship between the frequency of mental
retardation in a population of children prenatally exposed
to methyl  mercury and the maximum hair concentrations of the
mothers during pregnancy.  The maximum hair concentrations
in the mothers during pregnancy was used as a measure of the
prenatal dose.  The curves are drawn according to logit
analysis,  assuming the presence of a background frequency.
Figure 6A gives the complete dose-response curve and the
logit equation.  Figure 6B gives the low frequency end of
the dose-response relationship, indicating the presence of a
background frequency, i.e., the vertical intercept at zero
mercury concentration in the mothers'  hair.  The analysis
was carried out by Berlin (1982) on data from the Iraqi
outbreak (Marsh et al.  1980).
                                    6-30

-------
estimation described in Figure 6-7 should only  be  regarded  as
approximate, as they are based on  small  numbers.   We greatly need to
obtain more precise estimates of human  health risks associated with
prenatal  exposure to methyl  mercury.

6.2.4.3  Human Exposure from Fish  and Potential  for Health  Risks--
Dietary intake accounts for the greatest fraction  of total  mercury
intake by man (Table 6-5).   Methyl  mercury intake  is exclusively from
the diet and almost entirely from  fish  and fish products.   The evidence
comes from dietary studies   showing close correlation of blood levels
with fish consumption (Swedish Expert Group 1971)  and from  large-scale
analyses of food items in several  countries, indicating that significant
concentrations of methyl  mercury are found only in fish and fish
products (U.S. EPA 1980a).

     Based on data from the National Marine Fisheries, Cordle et al.
(1979) have reported a ranking of  species of fish  according to annual
consumption in the United States (Table 6-6).   The table clearly
demonstrates that oceanic fish, especially tuna, account for the major
amount consumed.  However,  when consumption is  expressed according to
the consumer use, a different picture emerges.   On this basis,
freshwater fishes dominate  the rankings, with northern pike consumed at
17.4 g day"1, followed by freshwater trout at 12.3 g day"1, bass
(freshwater) and catfish at 12.1 g day"1.  The  highest user
consumptions of seafood are crabs  and lobster at 10.6 g day"1, with
tuna down to 6.1 g day"1.

     The highest average mercury concentrations are also found in
freshwater fish - pike at 0.61 yg  Hg g"1 and trout at 0.42  yg Hg
g"1.  Thus a pike consumer would have a daily average intake of methyl
mercury of 10.4 yg exclusively from pike, and a trout consumer would
have had an average intake  of 5.2  yg Hg.  These average values are
well below the recommended maximum safe intake  of  30 yg day"1.

     The National Marine Fisheries developed an extensive data bank on
fish consumption by individuals according to fish  species (U.S.
Department of Commerce 1978).  These data were  based on a Diary Panel
Survey of approximately 25,000 individuals chosen  to be representative
of the U.S. population.  These data, along with additional  information
on mercery concentration of edible tissues of various fish  species,
allowed a calculation of the number of  individuals who would be expected
to exceed the maximum safe  daily intake of 30 yg.  It was calculated
that 47 individuals would exceed this limit by  a small margin from
consumption of fish and that 23 of these were consumers mainly of
freshwater fish.  According to calculations by  Nordberg andfctrangert
(Figure 6-7) the risk at this level of  intake will be small—on the
order of 0.3 percent.

     The risk of prenatal poisoning cannot be estimated with any
precision, given the small  number  of cases used in Figure 6-8.  The
daily intake of about 30 yg Hg roughly  corresponds to a hair
concentration of 6 to 10 ppm.  The dose-response data in Figure 6-8


                                  6-31

-------
  TABLE 6-5.  ESTIMATES OF AVERAGE AND MAXIMUM DAILY INTAKES OF
   MERCURY BY THE "70 kg MAN" IN THE UNITED STATES POPULATION
               (ADAPTED FROM U.S. EPA 1980a)



Media              Mercury intake vg day-1 70 kg-1     Predominate
                             (average)                    form


Air                             0.3                        Hg°

Water                           0.1                        Hg2+

Food                            3.0                        CH3Hg+
                                  6-32

-------
    TABLE 6-6.   ESTIMATED FISH AND SHELLFISH CONSUMPTION IN THE  UNITED
       STATES RANKED  ACCORDING TO ANNUAL CONSUMPTION FOR THE PERIOD
          SEPTEMBER 1973 TO AUGUST 1974 (ADAPTED FROM  U.S. EPA
                       1980a AND CORDLE  ET AL. 1979)
Amount
Rank 1()6 lb yr-1
Total
Tuna (mainly
Canned)
Unclassified
(mainly
breaded,
Including fish
sticks)
Shrimp
Ocean Perchd
Fl ounder
Clams
Crabs/lobsters
Salmon
Oysters/scallops
Troutf
Codd
Bassf
Catfish^
Haddockd.
Pollock*1
Herring/smelt
Sardines
Pikef
Halibutd
Snapper
Whiting
All other
classified


1




2
3
4
5
6
7
8
9
9
11
12
12
12
15

16
17
18
18
20


2957

634




542
301
149
144
113
110
101
88
88
78
73
73
73
60

54
35
32
32
25

152
Percent of
total by
weight
100.0

21.4




18.4
10.2
5.0
4.9
3.8
3.7
3.4
3.0
3.0
2.7
2.5
2.5
2.5
2.0

1.8
1.2
1.1
1.1
0.9

5.1
Number of
actual users
(millions)
197.0

130.0




68.0
45.0
19.0
31.0
18.0
13.0
19.0
14.0
9.0
12.0
7.6
7.5
11.0
11.0

10.0
2.5
5.0
4.3
3.2


Mean Amount
per user,
(g day1)
18.7

6.1




10.0
8.3
9.7
8.€
7.6
10.6
6.7
7.8
12.3
8.1
12.0
12.1
8.6
6.8

6.7
17.4
8.0
9.3
9.7


Average cone.
of mercury
vg Hg g-1*

0.14b
0.27
0.35



c
0.05
0.13
0.10
0.05
0.07-0.14^
0.08
0.03
0.42
0.14
c
0.15
0.11
0.14

0.02
0.61
0.19-0.53
0.45-35g
c

c
aU.S. Chamber of Commerce (1978).
^Average values for skipjack, yellow fin, and white tuna, respectively.
cData not available.
dflainly imports.
CKing crab - all others, respectively.
fpresh Water.
9Red Snapper - other.
                                        6-33

-------
would indicate that the background frequency  of mental  retardation would
be increased by less than 50 percent.

     Estimates of increased rates  risks  due to acid  precipitation would
depend upon a number of assumptions,  including whether  increases in
freshwater acidity would elevate levels  of methyl mercury  in freshwater
fish and by how much, the effect of acidity on the supply  of freshwater
fish, as well as actions taken  by  local  state and Federal  agencies to
limit fishing and sales of fish if methyl mercury levels  increase.
Nevertheless, information on methyl  mercury is now reaching the point
where rough estimates can be made  of health risks in this  country for
consumption of methyl mercury from freshwater fish,  and information may
be forthcoming on the impact of acidity  on methyl mercury  levels in
fish.  At least the direction of future  research is  now more clear—to
obtain more quantitative information on  human dose-response relationship
and to further test hypotheses  on  cause-effect relationship between pH
and methyl  mercury levels in freshwater  fish.

6.3  GROUND SURFACE AND CISTERN WATERS AS AFFECTED BY ACIDIC DEPOSITION
     (W. E. Sharpe and T. W. Clarkson)

     For reasons given in Section  6.1, this section  will deal only with
those metals whose concentrations  and/or speciation  in  drinking water
may be affected by acidic deposition.  As discussed  in  the previous
section, mercury concentrations, including any potential changes due to
pH, should not offer any conceivable threat to human health.  Lead is
the one metal of greatest concern  and will be given  special attention in
this section.  Other metals such as aluminum, cadmium,  and copper, will
be discussed briefly.

6.3.1  Water Supplies

     An understanding of the modes of hydrologic interactions between
acid deposition and various types  of water supplies  is  essential to
assessing the potential indirect health  effects  to users of drinkinq
water obtained from such systems.   In addition,  the  physical facilities
used to store, treat, and distribute water are of primary  importance, as
are the chemical methods used to treat water  prior to use.  Principal
water sources in continental North America are usually  either surface or
groundwater, with other sources such as  direct use of precipitation of
much lesser importance.  Health risk is  directly related  to the source
of drinking water.

     Health risk in drinking water supplies is also  closely related to
the management of the drinking water supply.  Risks  are generally
greater the smaller the water supply,  with small privately owned water
systems serving a single dwelling  at greatest risk.   These systems
typically do not routinely monitor water quality nor do they provide
even rudimentary water treatment.   Data  on the impacts  of  atmospheric
deposition on drinking water quality are extremely scarce; however, by
using available information on the impacts to surface water aquatic
ecosystems, we may assess impacts.


                                  6-34

-------
6.3.1.1  Direct Use of Precipitation (Cisterns)--The  direct  use of
precipitation by collection in artificial  catchments  is one  of the
oldest forms of water supply, having been used widely by  ancient
civilizations.  The Romans used lead-lined cisterns for the  storage of
water, and it has been reported that plumbism  (chronic lead  poisoning)
was a major reason for the fall of the Roman Empire (Gil fill an 1965,
Nriagu 1983).

     Direct use of precipitation has been practiced in North America
from very early times and is still  common where there are no other water
supply alternatives.  Island communities in the equatorial regions of
the world still rely heavily on rainwater cisterns to supply their
freshwater needs, .and this method of water supply  is  being seriously
considered as appropriate technology for the developing counties of the
world.

     Roof catchments consist of an impervious  surface, usually a house
or auxiliary building roof, connected by means of conventional roof
gutters and downspouts to a below ground concrete or  cinder  block
cistern.  Water is pumped from cistern storage to points  of  use within
the house.  Since, in most systems, precipitation is  used directly with
no treatment, the quality of precipitation and the amount of dry
deposition on the catchment between precipitation events  are of
paramount importance to the quality of drinking water at  the user's tap.
The major impacts are twofold.  First, direct  deposition  of  atmospheric
pollutants such as lead and copper may occur and, second,  the acid
components of atmospheric deposition may cause increased  corrosion of
metallic plumbing system components.

     In a study of 40 roof-catchment cistern systems  in western
Pennsylvania, Young and Sharpe (1983a) report  that lead in atmospheric
deposition accumulates in the sediments that collect  at the  bottoms of
cisterns and that this particulate lead could  appear  in the  drinking
water of cistern users when conditions allowing the suspension of this
material in cistern water are present.  They did not  report  on the
frequency of such conditions, but they did point out  that in the systems
they studied there were no safeguards to prevent the  ingestion of lead
contaminated cistern sediments.  However,  cistern systems  with gross
particulate filters for incoming catchment runoff had much lower lead
concentrations in sediments.

     Young and Sharpe (1983a) also report accumulations of cadmium in
cistern sediments, although such accumulations were less  frequent than
lead.  The cadmium concentrations in atmospheric deposition  in the Young
and Sharpe study are generally very low, indicating that  some other
source such as corrosion of galvanized gutters and downspouts might be
present.

     Young and Sharpe (1983a) found that precipitation was highly
corrosive as measured by the Langelier Saturation Index (LSI) and that
cistern water, although still corrosive in all  but a  few  systems, was
less corrosive than bulk precipitation.   The decreased corrosion


                                  6-35

-------
potential of cistern water  was  attributed to dissolution of the calcium
carbonate building materials  in the cistern, a fact confirmed by the
much higher LSI's of cisterns with impermeable vinyl liners.

     Young and Sharpe measured  the concentrations of copper and lead in
tapwater that had stood  in  the  plumbing system overnight.  In nine of
the 40 systems studied (22  percent) average lead concentrations exceeded
drinking water limits (U.S. EPA 1979b), copper exceeded drinking water
standards (U.S. EPA 1979b)  in 11 of the 40 systems.  All of the systems
(100 percent) having all  copper plumbing showed an increase in copper
concentration in standing tapwater as compared to cistern water,
indicating that corrosion was taking place.

     Young and Sharpe (1983b) attempted to establish a relationship
between the corrosivity  of  precipitation and the concentration of copper
in the tapwater of cistern  systems.  Their basic premise was that
although atmospheric deposition is more so, and this increased
corrosivity has caused an increase in the concentration of copper in the
tapwater of the cistern  systems studied.  Table 6-7 contains a
comparison of pH, corrosivity (as indicated by the Ryznar Index) and
corresponding tapwater copper concentrations.  The potential impact of
increasing the corrosivity  of atmospheric deposition on tapwater copper
concentrations is obvious.  If  the relationship proposed by Young and
Sharpe is valid, increasingly polluted atmospheric deposition would
appear to be responsible for  tapwater copper concentrations in excess of
drinking water standards.  Unpolluted atmospheric deposition (pH 5.6)
would appear to result in tapwater copper concentrations within drinking
water standards for roof catchment cisterns.

     The actual number of roof-catchment cisterns in those areas of the
country impacted by polluted  atmospheric deposition is unknown.  Kincaid
(1979) estimated that there were 67,000 cistern systems in the state of
Ohio, although there is  good  reason to believe that this figure may be
too high.  Determination of the population at risk is difficult, but
these data indicate that it is  likely to be substantial.

     Cistern systems can be modified to minimize the risk (Young and
Sharpe 1982).  However,  these modifications are likely to be expensive
with minimum estimated costs  of $500 to $1000 per household for water
treatment equipment and  the necessary changes to plumbing systems
(Sharpe 1980).

     Young and Sharpe (1983a) conclude that "The Presence of lead and
copper in the tapwater of cistern water supplies in western Pennsylvania
was sufficient to constitute  a  hazard to users of such systems.  Users
involved in the study were  advised to discontinue use of cistern water
for drinking purposes until such time as proper safeguards were employed
to reduce the hazards implicit  from this study."

6.3.1.2  Surface Mater Supplies--Very little work has been done on the
specific effects of atmospheric deposition on surface water supplies,
although quite a bit can be inferred from the surface water quality work
                                  6-36

-------
      TABLE 6-7.  COMPARISON OF BEST AND WORST  CASE AND MEAN RYZNAR
       STABILITY INDEX (RI)  OF  BULK  PRECIPITATION COLLECTED WEEKLY
      DURING 1979, 1980 AND  1981 IN  CLARION  AND INDIANA COUNTIES, PA
                      (FROM  YOUNG AMD SHARPE 1983b)
Parameter
PH
Alkalinity
(mg r1 CaCOs)
Best
case
5.29
2.22
Meana
3.87
0
Worst
case
3.40
0
Specific conductance
( mhos cnrl)
Calciun (mg £-1)
RI @ 20 C
Predicted Cu
cone, (yg jr1)
127
12.8
14.73

922
70
1.14
18.76

2724
220
0.23
19.58

3299
aBased on 138 samples collected over three years.
                                  6-37

-------
done to determine impacts on  aquatic  biota.   In most regions where
atmospheric deposition is of  concern  the same types of surface water are
used for both water supply and fish propagation; consequently, the water
quality changes reported for  one are  applicable to the other.  The chief
area of concern is for surface water  supplies providing drinking water
for humans.

     Two main drinking water  impacts  exist.   The quality of the source
water may be impaired and/or  increases  in  the corrosivity of the water
could lead to the same types  of tapwater quality problems evident with
cistern water supplies.  As reported  elsewhere in this document (Chapter
E-5), aluminum concentrations may  be  increased in surface waters.  In a
1981 study of the surface water quality of a  stream (Card Machine Run)
feeding a small water supply  reservoir, DeWalle et al. (1982) reported
that aluminum concentrations  in the stream directly above the water
supply intake increased from  .05 mg jr1 to 0.70 mg £-1 in
response to a February rain and snowmelt event on the watershed.  These
data are illustrated in Figure 6-9.   High  concentrations of aluminum
have been reported elsewhere  by Cronan  ahd Schofield (1979) and Herrman
and Baron (1980).  The health significance of aluminum concentrations of
this magnitude are addressed  elsewhere  in  this chapter.  Other metals
not as readily leached from acidified soils are not likely to increase
as dramatically as aluminum.

     Increasing corrosivity is probably the most significant potential
impact of atmospheric deposition on surface water supplies.  The
corrosivity of the dilute water often used for surface water supplies in
the northeastern United States is  mostly controlled by H+ concen-
tration.  As the H+ concentration  increases so does the corrosivity of
the water.

     Corrosivity in surface water  supplies has been widely reported, and
its impacts are well documented.   Where lead  water distribution pipes
are in use, clinical lead poisoning of  children has been reported as a
consequence of corrosive drinking  water conveyance.  A notable example
of such a problem is Boston,  Massachusetts.   Less well known is the case
of Mahanoy City, PA (Kuntz 1983).  A  case  of  copper toxicity from a
corroded water fountain has also been reported by Semple et al. (1960).
Where pipes are of other metals such  as copper, iron, or galvanized
steel the respective corrosion products of copper, lead, iron, zinc, and
cadmium can be problems.

     Because these corrosion  problems can  lead to elevated
concentrations of toxic metals in  drinking water, the U.S. EPA (1979a)
has recommended that all drinking  water supplies be noncorrosive and
that a minimum pH of 6.5 be maintained.  The  results of a review of the
drinking water standard for corrosivity, which was completed by U.S. EPA
in 1980, have not as yet been released.  Numerous studies of surface
water chemistry have shown dramatic increases in the H+ concentration
of surface waters in response to acidification by atmospheric
deposition.  In dilute surface waters such increases are almost certain
to produce corresponding increase  in  the corrosivity of that water.  If


                                  6-38

-------
CO
                   I
                   o?

                   o>
                      1.5
                      1.0
                      0.5
                                               CARD MACHINE  RUN
LEGEND
   ALUMINUM
                                 	  DISCHARGE
                           . j.^f^'L.j-nf. mmmpmm m^m in uu^-__-j_^_-J—A'
10      12        14        16        18

                                FEBRUARY
                                                                      20
           5 ^

           4 5
             i—i

           3 ^

           2 g
                                                                                                   1  o
                                                                                                     »—i
                                                                                                     o
                                         22
24
26
   Figure 6-9.  Aluminum concentration and discharge for Card Machine  Run.   Adapted from DeWalle  et al
                (1982).

-------
the pH and computed Ryznar Stability  Index  (RI) for the data of DeWalle
et al. (1982) are plotted for a  rain  and  snowmelt event on Card Machine
Run In February 1981 (Figure 6-10)  a  strong relationship between the two
is identified.  Linear regression  techniqes were used to quantify the
relationship between pH and RI for this runoff event, and a correlation
coefficient of r  = -1.00 was obtained.   This indicates that large
changes in the pH of dilute surface waters, weakly buffered by CaC03,
are almost certain to produce correspondingly large increases in the
corrosivity of such waters.

     If RI values are plotted with streamflow (discharge) for the same
event on Card Machine Run (Figure  6-11),  it is obvious that as
streamflow increases as a result of acid  snowmelt and rainfall runoff,
the corrosivity as indicated by  the Ryznar  Index also increases
dramatically.  Regression analysis again  yields a very good correlation
(r^ = .80) for these two variables.

     Although the data presented are  limited, there would appear to be
strong indications that the corrosviity of  raw water entering surface
water supplies located in headwater areas of the Laurel Hill is
increased substantially as a result of acid snowmelt and rainfall
runoff.  If the model for the relationship  of pH and RI holds true for
all dilute surface waters, then  increased corrosivity is likely anywhere
that the pH of such waters changes dramatically subsequent to acid
runoff events.  Where surface water storage facilities are small,
necessitating the direct use of  raw water during stormflow periods, and
where corrosion control is not practiced  in the water system,
populations served are at increased risk  of being exposed to higher
concentrations of corrosion products  such as Cu, Pb, Cd, and Zn.

6.3.1.3  Groundwater Suppl 1_es—Acidification of groundwater as a
consequence of atmospheric (leposition has been reported in Sweden by
Hultberg and Wenblad (1980).  Such changes  have not as yet been well
documented in North America.  Funs (1981) reports that atmospheric
deposition in sensitive regions  of New York State has decreased the pH
and increased the Al  concentration of shallow groundwater and indicates
that pH of groundwater is significantly correlated with depth, with
deeper groundwater sources having  higher  pH.  Fuhs also reports on the
concentrations of Pb and Cu  in private individual water supplies
obtaining water from shallow circulation  springs and shallow wells.
Fuhs indicates that the Al concentrations measured in these types of
water sources would make such water unsuitable for hemodialysis units.
Although Fuhs demonstrates that  standing  tapwater derived from shallow
groundwater systems in atmospheric  deposition sensitive areas of New
York contains high concentrations  of  Cu and Pb, he does not make a clear
case linking these results to the  acidity of atmospheric deposition.  As
Fuhs correctly states, shallow groundwater  in these areas would be
corrosive even without acid  deposition; consequently, the degree to
which atmospheric deposition makes these  waters more corrosive and the
concomitant increases in tapwater  metals  concentrations must be
determined.  Neither has yet been  demonstrated conclusively.
                                  6-40

-------
0)
6.0
5.9
5.8
5.7
5.6
5.5
5.4
5.3
5.2
5.1
5.0
4.9
4.8
4.7
4.6
              4.5
                                                                          LEGEND
                                                                       •  pH
                                                                          RYZNAR INDEX
                 17     18    19    20    21     22     23     24
                                                   FEBRUARY
                                                   25
26
27
28    29
                          18.9
                          18.8
                          18.7
                          18.6
                          18.5
                          18 A
                          18.3
                          18.2
                          18.1
                          18.0
                          17.9
                          17.8
                          17.7
                          17.6
                                                                                                        X
                                                                                                        LU
                                                                                                        Q
                                                                                                        CtL
                                                                                                        o:
   Figure 6-10.   pH  and  Ryznar  Index  for  Card  Machine Run.

-------
                                  Zfr-9
                                       RYZNAR INDEX
N
3
CD
Q.
tn
x
O)
Q.
Q.
l/>
O
QJ
-s
05
O
-s
o
OJ
-s
a.
CD
                             DISCHARGE (£  s~l ha~l)

-------
     Unpublished data collected by Sharpe and DeWalle indicate a
probable link between acid  recharge water and the decreasing pH and
alkalinity of a deep circulation  spring on Pennsylvania's Laurel Hill.
The data were collected during an acid snowmelt and rainfall runoff
event in March of 1982 and  are depicted in Figure 6-12.  Unfortunately,
flow data for the spring are not  available; consequently, flow data for
Wildcat Run,  a stream whose watershed makes up a significant part of the
spring's recharge area, are used  for comparison.  Wildcat Run, at the
point of flow measurement,  is only several feet from the spring
discharge and groundwater is an important component of its total flow.
Thus, the run's temporal response to acid runoff recharge is likely to
be quite similar to that of the spring.  The pH and alkalinity of the
spring water appear to drop in concert with the increased streamflow in
Wildcat Run,  with the most  dramatic change occurring in alkalinity.

     As discussed in an earlier section of this chapter there is a
strong correlation between  pH change and corrosivity for dilute waters;
therefore, it could be reasonably assumed that the corrosivity of the
water in this spring increased during the acid recharge event.

     The lack of data is greatest with respect to groundwater impacts
from atmospheric deposition.  Much additional work is indicated, but
preliminary information seems to  indicate that adverse impacts to
drinking water quality are  possible in water supplies using shallow
groundwater in areas edaphically  and geologically sensitive to
atmospheric deposition.

6.3.2  Lead

6.3.2.1  Concentrations in  Noncontaminated Waters—The U.S. national
interim primary drinking water standard for lead is 50 yg£-!.
The United States Environmental Protection Agency (U.S. EPA 1979a)
summarized data in two surveys on lead in drinking water.  The median
lead concentration in municipal drinking water supplies is about 10 yg
s."1.  In certain areas, such as Metropolitan Boston, it may contain
lead in excess of the 50 yg rl standard.  This is believed to be
due to very soft water (low pH) and the presence of lead piping in the
domestic water distribution system (The Nutrition Foundation Expert
Advisory Committee 1982).  Lead piping is no longer used for new potable
water systems in the United State (U.S. EPA 1979a).

     A recent national survey of  Canadian drinking water supplies
involving 71 municipalities representing 55 percent of the population,
indicated a medial level of lead  equal to or less than 1 yg £~
and values ranged from < 1  yg  £-1 to 7 yg £~1.

     Most natural ground waters have concentrations ranging from 1 to 10
yg r1.

6.3.2.2  Factors Affecting  Lead Concentrations in Water, Including
Effects of pH--In areas where"the home water supply is stored in lead
lined tanks and where it is conveyed to the household taps by lead
                                  6-43

-------
     8.
     8.0
     7.5
en
i
     7.0
     6.5
    6.0
           21
           18
           15
          ro
          o
          o
          (O
       T    t
            11
12
                                                 LEGEND

                                               ALKALINITY (Spring)

                                               pH (Spring)

                                               DISCHARGE (Wildcat Run)
13
14
  15


MARCH
                                                                        16
                                                            17
                                                            18
                                                                                                                     (O

                                                                           0.091




                                                                           0.084




                                                                           0.077



                                                                           0.070



                                                                           0.063



                                                                           0.056 ~




                                                                           0.049 ^


                                                                                L±J

                                                                           0.042 g

                                                                                3C


                                                                           0.035 S
                                                                                Q


                                                                          0.028



                                                                          0.021



                                                                          0.014



                                                                          0.007
                                                            19
         Figure 6-12.  Alkalinity, pH, and discharge for Wildcat Run.

-------
pipes, the concentration may  reach  several hundred micrograms per liter
and even exceed 1000 yg r1  (Beattie et al. 1972).  The
concentration of lead in water conveyed through lead pipes is affected
by several factors.   The longer the water  is held in the pipes, the
higher the lead concentrations (Wong and Berrang 1976).  The so-called
"first flush" sample generally has  lead concentrations about three times
higher than free-running tap  water  (Nutrition  Foundation Expert Advisory
Committee 1982).  The lower  the pH  of  the  water and the lower the
concentration of dissolved salts, the  greater  the solubility of lead in
water.

     Leaching of lead from plastic  pipes has also been reported (Heusgem
and DeGraeve 1973).   The source of  lead was probably lead stearate,
which is used as a stabilizer in the manufacture of polyvinyl plastics.

6.3.2.3  Speciation  of Lead  in Natural Water--Lead does not present the
wide range of chemical and physical forms  that mercury does.  Metallic
lead and its inorganic compounds possess a negligible vapor pressure at
room temperatures, so volatile forms of lead are not important in the
geochemical cycle.  The organometallic forms of lead, such as the
tetra-alkyl leads, although  synthesized for use as antiknock compounds
in gasoline, do not occur naturally as in  the  case of methyl mercury
compounds.  The inorganic salts of  lead are numerous.  The solubility of
these compounds differs greatly.

     The soluble salts will  dissociate in  water to liberate the reactive
lead cation Pb^+, which will  form complexes and chelates with a
variety of organic ligands present  in  water and sediments.  Sibley and
Morgan (1977) have described different forms of lead in freshwater:
complexed ions, lead absorbed to precipitate,  solid precipitate, and
free lead ions.  Lead present as the complexed ion is by far the most
predominant species.

     No studies have reported on the effect of acidic deposition on the
speciation of Pb in natural  bodies  of  water.   Lead has been reported to
bind to a wide range of organic fractions  in river water (Ramanoorthy
and Kusher 1975).  As pointed out in Chapter E-4 of this document,
decreasing water pH will reduce the fraction of heavy metals bound to
organic components and increase the concentration of free inorganic
metal species.  This should increase lead  levels in aquatic biota,
possibly affecting human dietary intake.

6.3.2.4  Dynamics and Toxicity of Lead in  Humans—Excellent reviews of
this topic have been published in recent years (WHO 1977, U.S. EPA
1980b, Nutritional Foundation Expert Advisory  Committee 1982).

6.3.2.4.1  Dynamics of lead in humans.  The  uptake, distribution, and
excretion of lead have recently been  reviewed  in detail  (U.S.  EPA
1980b).  Approximately 8 percent of dietary  lead  is absorbed  in  the
gastrointestinal tract in adults.  Children  absorb about 50 percent of
the  ingested lead.  Lead in water and other beverages may be  absorbed
with greater efficiency than lead presented  in food.


                                  6-45
409-262 0-83-20

-------
     Lead is distributed to all  tissues  in  the body and to all
compartments within cells.   Most of the  lead  in blood is associated with
the red blood cells.  The skeleton is  the main site of lead storage,
with about 95 percent of the total  lead  in  the body in the skeleton of
adults.  Lead readily crosses the placenta.   It also crosses the
blood-brain barrier but more readily in  children than in adults.

     Lead is excreted in urine and feces, with the human urinary route
probably being more important.  The half-time of lead retention in soft
tissues is about six weeks  following exposure of a few months.  The
half-time may be longer following years  of  occupational exposures to
lead.   Lead is accumulated  in the skeleton  throughout most of the human
life-span, and the half-time in  skeletal tissue is very long.

     Lead concentration in  whole blood is the most commonly used
indicator for assessing the burden of  lead  in soft tissues.  The
relative contributions of airborne lead, lead in food, and other sources
of lead are often assessed  in terms of their contributions to the
blood-lead concentration.

     A positive correlation exists between  the concentration of lead in
domestic water supply and the concentration of lead in blood.  The
United States Environmental Protection Agency, based on a study by Moore
et al. (1977), has estimated blood concentrations associated with levels
of lead in free-running tap water (Table 6-8).

     If the relationship is valid,  the impact of lead concentrations in
running tapwater is greatest in  the lower range of lead in water.
According to Table 6-8, the median lead  level in U.S. drinking water (10
yg JT1) would contribute approximately 3.4  yg dl-1.  Assuming
the median blood level  in the absence  of the water contribution to be 11
yg dl'1, the U.S. water supply contributing about 30 percent
additional blood lead and lead present in tapwater at the current
interim primary drinking water standard  would contribute about 10 yg
dl-1 to blood lead concentration,  i.e.,  about equal to the lead
contribution from all  other sources.   However, blood levels in the
United States are affected  by a  number of factors such as age, sex, and
urban  versus non-urban  locations.   Urinary excretion of lead may be used
on a group basis to indicate the soft  tissue burden.  Lead in hair,
unlike the case of methyl mercury,  is  not a useful  indicator because it
represents external contamination of the hair sample.

6.3.2.4.2  Toxic effects of lead on humans.  Lead damages a variety of
human  organs and tissues.Damage to the human hemopoietic system is
usually the first observable effect of lead (Figure 6-13).  The
inhibition of enzymes  involved in  synthesizing hemoglobin results in the
accumulation of precursor substances:  s-aminolevulinic acid (s-ALA)
in plasma and urine, and free erythrocyte protoporphyrin (FEP) on the
red blood cells.  Measuring FEP  has become  a routine method for checking
the earliest effects of lead.
                                 6-46

-------
      TABLE 6-8.  THE ESTIMATED RELATIONSHIP  BETWEEN LEAD
         CONCENTRATIONS IN RUNNING TAP  MATER  AND HUMAN
     BLOOD LEAD LEVELS (MOORE ET AL.  1977  IN  U.S.  EPA 1980b)
Lead in running            Total  lead             Lead  in blood
    tap water            in blood (PbB)            due to water
    (ng A'1)               (yg dl-1)                 (yg dl'1)
       0                    11                          0

       1                    14.4                       3.4

       5                    16.7                       5.8

      10                    18.4                       7.4

      25                    21.0                      10.0

      50                    23.6                      12.6

     100                    26.8                      15.8
                                  6-47

-------
ENZYMIC STEPS
 INHIBITED
 BY LEAD
              NORMAL PATHWAYS
     METABOLITES AND
ABNORMAL PRODUCTS ACCUMULATED
  IN HUMAN  LEAD POISONING
                   PROPHYRIN FORMATION
                        IRON UTILIZATION

1

2Pb 	
3
4
"7PK ..... .

MITOCHON

CYTOPLASM
~
o
3;
o

Pb 	
rKREBS CYCLE 	 j Fe TRAf
__.. 	 . J (C,EPUM
SUCCINYL CoA + GLYCINE ETICU,
ALAS 1
1 Fe
d-AMlNOLEVULINIC ACID (ALA) 	 •
1 ALAD
PORPHOBILINOGCN 	
IURO I SYN
URO II COSYN
UROPORPHURINOGEN III 	
• UROGENASE
rflPDfiPfl[?nwYPTNn£FN T T T

COPROGENASE
PROTOPURRHYRIN S
HEMESYNTHETASE
«FERRIN
INTO
.OCYTES
t
+++



Pb
1 Pb
nrMr


^ ' 	 	 . 	 __ 	 Pb 	 	
HEMOGLOBIN
                                                                 Serum Fe
                                                                 may be increased
                                                                 ALA in urine (ALAU)
                                                                 and serum increased
                                                                 = urine
                                                               •»-  COPRO in rbc urine (CPU)
                                                                 Zn Protoporphyrin
                                                                 (ZnP) in RBC
                                                                 Ferritin, Fe micelles
                                                                 in rbc
                                                                 Damaged Mitochondria and
                                                                 immature rbc fragments
                                                                 (basophilic stippled cells)

                                                                 Globin
    Figure 6-13.
The initial  and  final steps associated with disturbances
in the biosynthesis of hemoglobin  due to lead are mediated
by intramitochondrial enzymes and  the intermediate  steps by
cytoplasmic  enzymes.   The enzymes  most sensitive to lead
(steps 2 and 7)  are the SH-dependent enzymes, 6-amino-
levulinate dehydrase (ALAD) and  heme synthetase.  Accumulation
of the substrates of these enzymes (ALA and FEP) is charac-
teristic of  human lead poisoning as  is increased urinary
coproporhyrin excretion.  Although zinc protoporphyrin  (ZnP)
accumulates  in erythrocytes in lead  poisoning (and  iron
deficiency), it  is usually measured  as "free" erythrocyte
protoporphyrin (FEP).  Lead reduces  the bioavailability of
iron for heme formation.  A compensatory increase in the
activity of  the  first enzyme  in  the  pathway, 6-amino-
levulinic  acid synthetase (ALAS),  occurs in response to
reduced heme formation.  Other compensatory responses
include erythroid hyperplasia  ,  reticulocytosis and micro-
cytosis.   Non-random shortening  of erythrocyte life span
has been demonstrated in lead workers.  Amicrocytic,
hypochromic  anemia results including some morphological
features noted above.  Adapted from Chisolm (1978).
                                        6-48

-------
     During recent years, measurement of FEP  has come  into wide use as
the most practical screening tool  in both epidemiologic studies and in
monitoring populations at high risk for lead  toxicity.  Figure 6-14
shows the curvilinear relationship between FEP and lead concentration in
blood.  The curvilinear shape is typical  of the relationship between
blood lead and other intermediate metabolites of porphyrin synthesis,
such as 
-------
              FEP « 0.043 x (blood lead) +. 0.45(blood lead) - 2.14
              r - 0.79
              n * 1056
          CO
          10
          o
          o
          o
          o
          O)
          3.
1200
1080
 960
 840
 720
 600
 480
 360
 240
 120
   0
                 0   15  30  45  60   75  90  105 120  135  150
                         BLOOD LEAD (yg 100 ml'1)
Figure 6-14.  Free erythrocyte protoporphyrin (FEP) vs blood level
              Shoshone County, Idaho, August 1974.  Adapted from
              Landrigan et al. (1976).
                                 6-50

-------
        TABLE 6-9.  NO DETECTED EFFECT LEVELS  IN RELATION TO PbB
                         (ADAPTED  FROM WHO  1977)
No-detected effect
level (yg 100 ml-1)
      Effect
 Population
 < 10
 20-25
 20-30
 25-35
 30-40
 40
 40
 40
 40-50
 50
 50-60
 60-70
 60-70
 > 80
Erythrocyte ALAD inhibition
FEP
FEP
FEP
Erythrocyte ATPase inhibition
ALA excretion in urine
CP excretion in urine
Anemi aa
Peripheral neuropathy
Anemi aa
Minimal brain dysfunction
Minimal brain dysfunction
Encephalopathy
Encephalopathy
Adults, children
Children
Adults, female
Adults, male
General
Adults, children
Adults
Children
Adults
Adults
Children
Adults
Children
Adults
aThe term anemia here is used to  denote  earliest  statistically
 demonstrable decrease in blood hemoglobin.   In adult workers a decrease
 in blood hemoglobin within the normal range  has  been reported during
 the first 100 days of employment.  Other  studies of workers indicate
 that frank anemia is not statistically  demonstrable until PbB > 100
 yg, as cited elsewhere in the WHO  report.  An increased frequency of
 early anemia has been reported at  PbB > 40 yg of groups of children
 in whom concurrent iron deficiency anemia was not ruled out but is
 highly likely.
                                  6-51

-------
                                       tQ
                                       CT>
                                        I
                                       cn
                                                                  PREVALENCE  OF ABNORMAL VALUES  (%.POPULATION)
 I
on
ro
-s oo :> —• -h -H
fD CO CT fD  -5  3"
00 ..   3  <  fD  fD
      O  0)  fD
                         fD
                         n
                                 00 fD
                                       fD
 < CTl
 fD CO <
 •  (Q C
      fD
   I—" 00
 3=- O
 0- O X
 01    fD
T3 3 -S
 r+ —i ft)
 fD   I
 Q- I—O.
      fD
 -h -h -h
 -S O -••
 O -S 3
 3    fD
   O Q.
 70 3-
 O -•• OJ
 fD —' l/>
 —' a.
 (/> -s ~n
   re m
 fD 3 -a
                                 3 ct- rt-
                                    3" -••
                                 3& -S O
                                 »  O 3
                                    O 00
                                 O << 3"
                                 3" rh -"•
                                 -i. fD -O
                                       CT
                                       fD
Q.-0
-S -S
fD O r+
3 rH «
>•• O fD
   •a fD
00 O 3
»  -s
   •a n-
o> 3- 3-
O.^ fD
C -S
_i -i.-a
c+ 3 fD

-h Q> O
fD 3 fD
30-3
                          i O
                           3 c
                           fD fD
                              < —i O> 01
                              O> fD < CQ
                              —i 00 fD fD
                        ^1 (a
                        OO 3
            01 O
      00 o> CQ -h
         3 fD
         Q^    Qj
            cr cr
         o —i 3
      fD "  O O
      x    o -s
      O O) Q- 3
      fD Q.    O>
      00
      00
                                   i fD — '
      O    Q-
      -h 3

      00 —'
      ro ro
      »  00
                                       CQ
                                       3-
                                                            o  r
                                                                                ro
                                                                                o
                                                                                   cn
                                                                                   o
                                                                                        oo
                                                                                        o
                                                                                o
                                                                                o
00
I—
o
O
a
                                                     (Q
                                                     o
                                                     o
                                                            ro
                                                            o
                                                            GO
                                                            o

-------
ami no aciduria, glycoseria, and hyperphosphoturia,  usually  do not occur
until blood levels exceed 70 yg dl'1.   Chronic  lead nephropathy is
not usually recognized in humans until  it has reached  an  irreversible
stage. The disease is characterized by  the slow development of
contracted kidneys with pronounced arterioscelerotic changes, fibrosis,
glomerular atrophy and hyaline degeneration of  blood levels.  These
changes portend progressive disease sometimes resulting in  acute renal
failure.  The duration of excessive exposure to lead is believed to play
an important role in the development of the disease.   Although
information on blood levels is inadequate, it is unlikely that the
general child and adult populations, even in the upper 2  to 5 percentile
of the "normal" U.S. range are sufficient to produce chronic renal
effects.

     Studies in the 19th and early 20th centuries  indicated that
occupational  exposures to lead (presumably higher  than current
exposures) caused increased frequency of abortions and stillbirths
(Oliver 1911).  Indeed, following the publication  of Oliver's findings,
women have largely been excluded from occupational  exposures to lead
until very recently.

     Lancranjan et al. (1975)  have reported reduction  in  sperm counts
and abnormal  sperm morphology in occupationally eposed men.  The
functional significance on fertility is not known.

     Prenatal  exposure to lead may be associated with  mental retardation
in children (Moore 1980).  The human data are consistent  with
experimental  findings on animals that modestly  elevated blood levels,
- 40 yg dl~l,  during prenatal  and early postnatal  life may  be
associated with subtle and long lasting adverse consequences to the
offspring.

     Lead has been shown to be a carcinogen in  animal  tests, but
epidemiologial studies have failed to reveal an association between lead
exposure and human cancer.  Measurement of precursor metabolite of heme
synthesis such as FEP or 6-ALA provide  the earliest warning of the
effects of lead.  If protected against, effect  on  heme synthesis will
protect against the more serious clinical  effects  of lead,  such as
anemia and encephalopathy.

6.3.2.4.3  Intake of lead in water and  potential for human  health
effects.  Mahaffey (1977) estimated that the daily intake of drinking
water ranged from 300 ml for children to as much as 2000  ml for adults.
An expert group of the National Academy of Sciences (WAS  1980) stated a
value of 1630 ml day1 for water intake of adults  (not including
amounts used to prepare foods and beverages) and a range  of 100 ml to
3000 ml for children.

     A study in Canada by Armstrong and McCullough quoted by the
Nutrition Foundation's Expert Advisory  Committee (1982) indicated that
the total daily intake including water  used as  a food  ingredient was 760
ml averaged for 0 to 6 years, and 1140  ml  for the  6- to 18-year-old


                                  6-53

-------
group.  The highest average  daily  intake was 1570 ml for the 55 and
older age group.   However, up  to 3000 ml total water per day was
consumed by sorne  children in the 0- to 6-year-old age group and up to
4300 ml  total  water was  consumed by certain individuals in each of the
remaining age groups.

     Using the MAS reported  range  of 100 to 3000 ml for children and a
U.S. median level  of 10  yg £-1, the range of intake for children
would be 1 to 30  yg Pb and for adults 16 yg, assuming a water intake
of 1600 ml day-1  (Table  6-10).  If average lead concentrations
attained the interim drinking  water standard of 50 yg £-!, these
intake values would be five  times  greater.

     The review of the human toxicity of lead in Section 6.3.2.4.2
identified children as the most susceptible group in the general
population.  Blood lead  levels in  children in the United States cover a
broad range of values  (Mahaffey et al. 1982a).  A criterion of 30 yg
Pb 100 ml"1 whole  blood  has  been used in estimating the prevalence of
elevated blood lead (Center  for Disease Control 1978).  If this
concentration of blood lead  is accompanied by an erythrocyte proto-
porphyrin concentration  of 50  to 250 yg 100 ml"1 of whole blood, the
child is thought to have undue lad absorption.  Community based lead
poisoning prevention programs  report that approximately 75 percent of
children with blood lead levels of > 30 yg 100 ml"1 also have
erythrocyte protoporphyrin values  oT >_ 50 yg 100 ml"1 (Mahaffey et
al. 1982a).  The review  of human toxicity data in Section 6.3.2.4.2 also
indicate that blood lead levels in children >_ 30 yg 100 ml"1 indi-
cates a risk of biochemical, if not neuropsychological, dysfunctions.

     A survey of blood lead  levels in children in the years 1976 to 1980
in the United States indicated that substantial numbers of children have
blood leads >_ 30 yg dl'1 (Table 6-11).  The prevalence of elevated
blood lead values  is highest in children of low income families
(approximately 11  percent of children in families having an income less
than $6000) and in children  living in large cities (7.2 percent of
children living in cities of population more than one million).
However, elevated  blood  lead is widely distributed in the general
population, including children in  families earning more than $15,000
annual income (1.2 percent)  and in children living in rural areas (2.1
percent).

     Section 6.3.1 reviewed  available data to indicate that reduced pH
increases the corrosivity of water and can mobilize metals such as lead,
resulting in increased concentrations in drinking water.  Lead piping in
home plumbing is  rare and no longer used in this country except in
certain parts of  New England.  However, lead can be mobilized from other
types of piping where it is  used as a solder (copper piping) or in
stabilizers (certain types of  plastic pipes).  Homes using
roof-catchment cisterns  for  collecting drinking water seem especially
vulnerable to corrosive  rain water.  Young and Sharpe (see Section
6.3.1.1) noted that 22 percent of  such systems yielded lead
                                  6-54

-------
       TABLE 6-10.   DAILY INTAKE  OF  LEAD FROM DRINKING WATER

Age Group             Daily  Water Intake3       Daily Lead Intakeb
                               ml                       yg Pb

Children                  100  -  3000                    1 - 30
Adults                       1630                         16

aNational  Academy of Science (1980).
bAssumes U.S. median concentration of lead  in drinking water to be 10
 yg Pb A-l.
                                  6-55

-------
TABLE  6-11.  BLOOD LEAD LEVELS IN  CHILDREN  6 MONTHS THROUGH 5 YEARS
   BY ANNUAL FAMILY INCOME AND DEGREE  OF  URBANIZATION OF PLACE OF
        RESIDENCE IN THE UNITED STATES FROM  1976 TO 1980a
Demographic variable




Estimated
population
(thousands)



No. of
persons
exami ned



Bloodb lead
yg 100 ml"1



Revalence of
blood lead
levels < 30
yg 100 ml"1
% persons
examined
Annual Family Incomec

< $6000                    2465       448      20  +_ 0.6      10.9 +_ 1.4
$1000 - 14,999             7534      1083      16+0.5      4.2+0.7
> 15,000                   6428       774      14  + 0.4      1.2 +~ 0.4
Degree of Urbanization
urban
urban
Rural
>
<

106
106

persons
persons

4344
6891
5627
544
944
884
18
16
14
+ 0
+ 0
TO
.5
.7
.6
7.2
3.5
2.1
+
T
-
0.7
0.6
0.9
aAdapted from of Mahaffey et al.  (1982a).

bMean+_ S.E.M.

CA11 values shown for this variable reflect the exclusion (from
 analysis and tests of significance) of children in households  that
 declined to reported their income.
                                  6-56

-------
concentrations in tap water (having  stood  overnight) in excess of the
drinking water standard of 50  yg  Pb  fc'1.

     From the point of view of human health risks, any increases of lead
concentrations in drinking water  should be viewed as an additional
burden of lead.  This is especially  important with children where
substantial numbers already have  elevated  blood levels.  Drinking water
at the median concentration of 10 yg Pb &"1 already makes an appreciable
contribution to blood lead levels (approximately 30 percent add on to
other sources of lead; see Section 6.3.2.4.1).  Thus the drinking water
standard of 50 yg Pb fc-1 will  not provide  sufficient protection to those
children already having high blood lead from other sources of exposure.

     Unfortunately quantitative data are lacking on the contribution of
acidic deposition to lead in drinking water.  Roof-catchment cistern
systems believed to be widely  used in rural areas of Ohio and Western
Pennsylvania appear to be a probable target for the effect of acidic
deposition.  Thus, it is of great importance to ascertain the extent of
usage of these systems in those areas of the USA subject to acidic
deposition and to check the extent to which charged corrosivity of this
water affects lead levels in tap  water.

6.3.3  Aluminum

     Inorganic aluminum is toxic  to  fish and may be the main cause of
fish kills due to acidification of natural bodies of water.  Acidic
deposition dissolves aluminum  in  clay materials in soils and sediments,
thereby increasing concentrations of the A13+ ions and inorganic salts
of aluminum (for details, see  Chapter E-2).  Fish mortality appears to
be due to damage to the gills  of  the fish.  The toxic properties of
aluminum are self-limiting with regard to  bioaccumulation; when the
aluminum levels in water reach toxic levels, the ensuing mortality of
fish stops further accumulation in aquatic food chains.  The behavior of
aluminum is thus in sharp contrast to methyl mercury, which is of lower
toxicity to fish and is avidly accumulated.

     Aluminum in drinking water,  unlike lead, is not directly toxic to
humans.  However, a special  circumstance may lead to human toxicity--
that is the use of aluminum containing water in hemodialysis procedures.
This is believed to lead to direct entry of aluminum into the blood
stream and eventually damage to the  central nervous system.

6.3.3.1  Concentrations in Uncontaminated  Water--Burrows (1977) has
reviewed the literature on concentrations  of aluminum in natural bodies
of water.  He draws attention  to  two factors that are important in
assessing published values. First,  many publications do not clearly
distinguish between dissolved  and suspended aluminum in water.  He notes
that many investigators now use a 0.45 ym  millipore filter to
distinguish between dissolved  and parti oil ate aluminum.  The second
factor is that procedures for  trace  analysis of aluminum have only
recently become available and  most of the  literature data have been
                                  6-57

-------
collected without using  these  techniques.  Burrows states that as a
general  rule all  aluminum  values reported before 1940 should be regarded
with skepticism.   Unfortunately, very few analyses have been reported
for the most recent times  (from 1970).  The Maumee River Basin (Ohio)
was reported to have a mean  value of 0.01 mg r~l for the period
1971-73.  A phosphate limestone lake in Florida had a mean value of 0.05
mg 5,"1 at a water pH 7.0 to  9.6.  Tributaries to Lake Michigan had
mean values of 0.353 in  1972 but pH was not specified.  The above values
have been taken from Burrows (1977).

6.3.3.2  Factors  Affecting Aluminum Concentrations in Water—Burrows
(1977) notes a number of factors that influence aluminum concentrations
in bodies of natural water:

     1)   Acidic waters consistently contain much more soluble aluminum
         than neutral or alkaline waters.  Schofield and Trojner (1980)
         report that in  a  brook in the Adirondack Wilderness region of
         New York State, aluminum concentrations rose from about 0.2 mg
         r1 at pH 5.5 to  6.5  to 0.8 to 1.0 mg £-1 as the pH
         fell to  less than 5.0 during the spring snow melt.

     2)   Highly saline waters  contain higher aluminum concentrations
         than fresh waters.

     3)   Hot waters (e.g., hot water springs) tend to have higher levels
         of aluminum than  cold water.

     4)   Moving waters tend  to give higher aluminum analysis than
         quiescent waters.  This effect  is probably due to mobilization
         of suspended material.

6.3.3.3  Speciation of Aluminum in Water—The species of aluminum in
bodies of natural water  have been discussed in Chapter E-4.  Most of the
dissolved aluminum is present  as complexes with organic ligands.  The
inorganic fractions consist  of A13+ and aluminum fluoride, hydroxide,
and sulfate complexes.   The  fluoride complex is probably the predominant
inorganic species, according to thermodynamic calculations (Driscoll et
al. 1980).

     The inorganic monomeric species are more toxic to fish than are the
organic complexes of aluminum. Of the inorganic species, the fluoride
complex is probably the  least  toxic because addition of fluoride ion
reduces the toxicity of  aluminum.  Lowering the pH in natural bodies of
water increases the labile (inorganic) monomeric aluminum and thereby
increases toxicity to fish.  Driscoll et al. (1980) found that seasonal
variations in organically  chelated-aluminum were not affected by
seasonal variations in  pH  in lakes in the Adirondack region of New York
State.  The organic aluminum correlated with total carbon measurements
in water.

6.3.3.4  Dynamics and Toxicity in Humans--This topic has been the
subject of a number of  reviews (Norseth  1979).
                                  6-58

-------
6.3.3.4.1  Dynamics  of aluminum  in humans.  Data on absorption,
distribution, and excretion of aluminum compounds in man have been
reviewed recently {Norseth 1979).  Aluminum is absorbed in the
gastrointestinal  tract.   The  fraction of dietary intake absorbed into
the blood stream  is  believed  to  be small, but precise figures are not
available.  When  aluminum was given as the hydroxide salt to uremic
patients, approximately  15 percent of the dose was absorbed, with
considerable differences between individuals (Clarkson et al. 1972).
Unfortunately, information is not available on the absorption of other
forms of aluminum or in  people with normal kidney function.  Aluminum is
distributed to all tissues in the body and has been reported in fetal
tissues.  When aluminum  in food  was given to rats, increased levels were
reported in blood, brain, liver, and testes (Ondreicka et al. 1966).

     Little information  on the relative importance of urine versus fecal
pathways of excretion is available.  Renal clearance of aluminum may be
as high as 10 percent of the  glomerular filtration rate (creatinine
clearance) as indicated  in patients with compromised renal function.
These data would  suggest a high  urinary rate of excretion in normal
subjects and a correspondingly short biological half-time (on the order
of days or hours).  Animal experiments indicate that biliary excretion
of aluminum contributes  to fecal excretion of the metal.

     Aluminum is  found in both cow and human milk.  Normal levels of
aluminum in human blood and other biological fluids exhibited a very
wide range of values relative to the different laboratories making the
analyses.  Apparently considerable problems remain, particularly those
related to change contamination  by the ubiquitous metal, in determining
reliable values for the low levels in human plasma.

6.3.3.4.2  Toxic  effects of aluminum in man.  Toxic effects in terms of
fibrosis of lung  tissue  have  been reported in workers inhaling aluminum
or its compounds.  The situation with regard to toxic effects in humans
due to oral intake of aluminum is equivocal.  An early claim (Crapper et
al. 1973) that Alzheimer's Disease—a chronic degenerative disease of
the central nervous  system leading to presenile dementia—was associated
with accumulation of aluminum in the brain has not been substantiated by
later studies (Markesbery et  al. 1981).  However, a chronic neurological
disease "Dialysis Dementia,"  that develops in a number of patients
receiving dialysis therapy may be associated with elevated aluminum
intake (Alfrey et al. 1976, McDermott et al. 1978).  Intake of aluminum
may be directly from the water used in the dialysis fluid or from the
aluminum hydroxide compounds  given orally to remove phosphate from the
uremic patients.   Aluminum has been shown to be harmful to the central
nervous system in animals when directly administered in brain tissue
(Kopeloff et al.  1942) and to damage neuroblastoma cells in culture
(Miller and Levine 1974).

6.3.3.5  Human Health Risks from Aluminum in Water—Acute or chronic
disease in man has not been related to normal dietary intake of aluminum
from food or drinking water.  However, a potential risk may exist under
the special circumstances of  patients with compromised kidney function


                                 6-59

-------
who undergo regular therapeutic dialysis.  Driscoll et al. (1980)  have
reported levels  of aluminum in natural bodies of freshwater in the
Adirondack Region of New  York State to attain values as high as 800 yg
Al &"1 under the influence of acidic deposition.  A concentration  of
50 yg A"! of aluminum in  dialysis water is claimed to be dangerous
(Registration Committee,  European Dialysis and Transplant Association
1980).

     Of the various species of aluminum known to exist in bodies of
natural water, only data  on aluminum hydroxide are available.   This is
absorbed across  the human gastrointestinal tract.  In areas of the
country where drinking  water is fluoridated or where elevated fluoride
concentrations occur naturally, it is likely that aluminum flouride
complexes will be present in tap water in substantial amounts.
Unfortunately, we know  nothing of the gastrointestinal absorption  or
about its potential  toxicity in humans.

6.4  OTHER METALS

     A number of other  metals such as cadmium, copper, manganese,  and
zinc have been mentioned  with regard  to the possibility of indirect
heath effects.  In general, evidence  to justify a detailed report for
each metal is lacking.   However, it should be noted  that this chapter
has not considered at least one potential pathway of human intake of
environmental chemicals,  i.e., the food supply other than fish and fish
products.  Cadmium is known to be accumulated by plants, including
cereals, and the possible effects of  acidic deposition have not been
considered chiefly because of a lack  of studies.

6.5  CONCLUSIONS

     Chapter E-6 examines the direct  health effects and in doing so
mainly discusses lead and mercury availabilities as  affected by acidic
deposition.  The following statements summarize the content of this
chapter.

     0   No adverse human health effects  have been documented as being a
         consequence of metal mobilization by acidic deposition.  On the
         other hand, interest in the  phenomena of acidic  deposition  is
         recent and few investigations, if any, have been made into the
         possibility of indirect effects  on human health  (Section
         6.2.1).

     o   The substances requiring special attention  are methyl mercury,
         due  to  its accumulation  in  aquatic food chains,  and  lead due  to
         the  potential  for contaminating  drinking  water  (Section
         6.2.1).

      0    In virtually all studies  published to  date  elevated  methyl
         mercury levels in fish  muscle  (most notably the  pike and perch)
          have been statistically  associated with higher levels of
         acidity in water.   However,  a  number of factors  influencing


                                  6-60

-------
         mercury levels in fish may also change in parallel  to  acidity
         {Section 6.2.3).

     o    More research is needed to identify all  the factors that affect
         mercury accumulations in fish and the relative importance  of
         each.  This need is especially urgent in the United States
         where  few  data are available at this time (Section  6.2.3).

     8    The contamination of freshwater fish by direct discharge of
         mercury has been curtailed in recent years.  The role  of
         long-distance transport of mercury merits careful  investigation
         as an  explanation for high mercury levels in lakes  remote  from
         mercury-related industries (Section 6.2.2).

     0    Potential  impacts in acidic deposition of methyl  mercury
         concentrations in freshwater are of interest for several
         reasons (Section 6.2).

         a)  Fish and fish products are the major if not only
            sources of methyl mercury for humans.

         b)  Consumers of freshwater fish have a greater possibility of
            exceeding a allowable daily intakes of methyl  mercury  than
            do consumers of other forms of fish.

         c)  Pike and trout, freshwater fish among the most likely
             species to be affected by acidic deposition, have  the
            highest user consumption figures and the highest average
            methyl mercury levels.

     0    Prenatal life is a more sensitive stage of the life cycle  for
         methyl mercury.  More information is needed on fish consumation
         patterns of women of child-bearing age in order to quantita-
         tively assess the potential impact on human health of  elevated
         methyl mercury levels in freshwater fish (Section 6.2.4).

     Data on  the  impacts of acidic deposition on drinking water quality
are scarce.  However, by using available information, tentative
assessments of  impacts on ground and surface water systems were made.

     0    The lack of data is greatest with respect to groundwater.
         Preliminary information seems to indicate that adverse impacts
         to drinking water quality are possible in water supplies using
         shallow groundwater in areas edaphically and geologically
         sensitive  to acidic deposition (Section 6.3.1.3).

     0    Increasing corrosivity is probably the most significant
         potential  impact of acidic deposition on surface water
         supplies.  Populations are at increased risk of being exposed
         to higher  concentrations of corrosive toxicants, such as lead
         and  possibly cadmium, where surface water storage facilities
                                  6-61

-------
         are  small, necessitating the direct use of  raw  water during
         storm  flow periods and where corrosive control  is not practiced
         in the water system (Section 6.3.1.2).

     «   People receiving drinking water from roof catchment cistern
         systems  should be considered at potential  risk  of increased
         intake of lead in areas of acidic deposition  and especially if
         cisterns are used that have no particulate  filters {Section
         6.3.2).

     o   From the point of view of human health risks, any increases in
         lead concentrations in drinking water should  be viewed  as an
         additional burden of lead.  This is especially  important in
         children where substantial numbers already  have elevated blood
         lead levels (Section 6.3.2.4).

     °   Acute or chronic diseases in humans have not  been related to
         normal dietary intake of aluminum from food or  drinking water.
         However, a potential threat exists for patients undergoing
         hemodialysis if aluminum concentrations in  the  water used in
         this treatment exceed 50 yg of aluminum per liter  (Section
         6.3.3).

     Generally, the indirect effects on human health attributable to
acidic deposition require further study.  Data are very  limited  with
regard to measurement of the toxic elements and their  speciation and to
the kinetics  of transfer and uptake by accumulation  processes.   Studying
less toxic essential metals may be important in that elevated
concentrations  of some or all of them might affect the food chain
dynamics or the toxicity of lead or mercury.
                                  6-62

-------
6.6  REFERENCES

Alfrey, A. C., G. R. LeGendre, and W. D.  Kaehny.   1976.   The dialysis
encephalopathy syndrome:  Possible aluminum intoxication.   N.  Engl. J.
Med. 294:184-188.

Andren, A. W. and J. 0. Nriagu.  1979.  The global  cycle  of mercury, pp.
1-22.  Ir± The Biogeochemistry of Mercury in the Environment.   J. 0.
Nriagu, ed.  Elsevier/North Holland Biomedical  Press,  Amsterdam.

Armstrong,  R. W.  and R. J.  Sloan.   1980.  Trends in levels of  several
known chemical  contaminants in fish  from  New York State Waters.  DEC
Publication,  Technical  Report 80-2 Bureau of Environmental Protection
Division  of Fish  and Wildlife, Department of Environmental
Contamination,  New York State, Albany, NY.

Bakir,  F.,  S. F.  Damluji,  L.  Amin-Zaki,,  M. Murtadha, A. Khalidi, N. Y.
Al-Rawi,  S. Tikriti, H. I.  Dhahir,  T.  W.  Clarkson, J. C. Smith, and R.
A. Doherty.  1973.  Methylmercury  poisoning in  Iraq.  Science
181:230-241.
                        *
Beattie,  A. D., M. R. Moore,  W. T. Devenay, A.  R. Miller, and A.
Goldberg.  1972.   Environmental lead pollution  in an urban soft-water
area.  Brit.  Med. J. 2:4901-4903.

Beijer, J.  and A. Jernelov.  1979.  Methylation of mercury in  aquatic
environments, pp. 203-210.   In The Biogeochemistry of Mercury  in the
Environment,  J. 0. Nriagu,  e'37  Elsevier/North  Holland Biomedical Press,
Amsterdam.

Berlin, M.   1982.  Health  effects  due to  methyl mercury.  Seminar given
at the University of Rochester School  of  Medicine, Rochester,  NY,
December 1982.

Blcornfield, J.  A., S. 0. Quinn, R. J.  Scrudata, D. Long, A. Richards,
and F.  Ryan.   1980.   Atmospheric and watershed  impact of mercury to
Cranberry Lake, St.  Lawrence County, New  York,  pp. 175-210.  In Polluted
Rain.  T. Y.  Toribara,  M.  W.  Miller and  P.  E. Morrow, eds.  Plenum
Press,  New York.

Boulton,  P. and L. J. Hetling.  1972.  A  statistical analysis  of the
mercury content of freshwater first in New  York State.  Research and
Development Unit.  NYDEC,  Tech. Paper 19.  p. 37.

Brosset,  C.  1973.  Air-borne acid.   Ambio. 2:2-9.

Brouzes,  R. J.  P., R. A. N. McLean,  and G.  H. Tomlinson.  1977.  The
link between pH of natural  waters  and the mercury content of fish.
Domtar Research Center Report.  Senneville, Quebec, Canada,  p. 37.

Brune,  D.  1969.   Aspects  of low-temperature irradiation in neutron
activation analysis.  Anal. Chim.  Acta 44:15-20.


                                  6-63

-------
Burrows, W. D.   1977.   CRC  Critical Reviews in Environmental Control,
pp. 167-216. ^n Aquatic  Aluminum:  Chemistry, Toxicology and
Environmental Prevalence.

Carty, A. J. and S.  F.  Malone.   1979.  The Chemistry of Mercury in
Biological Systems,  In  The  Biogeochemistry of Mercury in the
Environment. J. 0.  Wiagu,  ed.   Elsevier/North Holland, Amsterdam,
Chapter 17.

Center for Disease Control.   1978.  Preventing lead poisoning in young
children.  J.  Pediatr.  93:709-720.

Chi sol m, J. J., Jr.   1978.   Heme metabolite blood and urine in relation
to lead toxicity and their  determination.  Clinical Chem. 20:225-265.

Clarkson, E. M., V.  A.  Luch,  W.  V.  Hynson, R. R. Bailey, J. B. Eastwood,
J. S. Woodlead, V. R. Clements,  J.  L. H. O'Riodan, and H. E. DeWardener.
1972.  The effect of aluminum hydroxide on calcium, phosphorus and
aluminum balances.  Clin. Sci. 43:519-531.

Clarkson, T. W.  1975.   Exposure to methylmercury in Grassy Narrows and
White Dog Reserves,  An  Interim Report, Medical Services Branch,
Department of health and Welfare, Ottawa, Canada.

Clarkson, T. W.  1983.   Methylmercury toxicity to the mature and
developing nervous system,  pp. 183-197.  In Biological Aspects of Metals
and Metal-Related Diseases,  B. Sarkar, ed7~ Raven Press, New York.

Cordle, F. P.  Corneliusjen,  C. Jelinck, B. Hacklay, R. Lehman, J.
Mclaughlin, R.  Rhodes,  and  R. Shapiro.  1979.  Human exposure to
polychlorinated biphenyls and polybrominated biphenyls.  Environ. Health
Perspectives 24:157.

Crapper, D. R., S. S. Krishnan,  and A. J. Dalton.  1973.  Aluminum
distribution in Alzheimer's disease and experimental neurofibrillary
degeneration.   Science  180:511-513.

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.

DeWalle, D. R., W. E. Sharpe, R. S. Dinicola, R. T. Leibfried, W. G.
Kimmel, and L.  S. Sherwin.   1982.   Causes of acidification of four
streams on laurel hill  in southwestern Pennsylvania.  Institute for
Research on Land and Water  Resources, The Pennsylvania State University,
University Park, PA. 62 pp.

Dickson, W.  1980.  Properties of acidified waters,  pp. 75-83.  _Iji
Ecological Impact of Acid Precipitation.  D. Drablos and A. Tollan, eds.
SNSF Project,  Oslo,  Norway.
                                  6-64

-------
Driscoll, C. T., J. P.  Baker,  J.  J.  Bisogni,  and C. L. Schofield.  1980.
Effect of aluminum speciation  on  fish  in  dilute acidified waters.
Nature 284:161-164.

Eriksson, E.  1967.  Mercury in Nature.   Oikos, Suppl. 9:13.

Fagerstrom, T. and A. Jernelov.  1971.  Formation of  methy!mercury from
pure mercuric sulphide  in aerobic organic sediment.   Water Research,
Pergamon Press, 5:121-122.

Fuhs, G. W.  1981.  Acid precipitation effects on drinking water
quality.  Paper presented at the  109th Annual Meeting of the American
Public Health Association, Los Angeles, CA.

Gilfillan, S. C.  1965.  Lead  poisoning in the fall of Rome.  J.
Occupational Medicine 7:63.

Grahn, 0.  1980.  Fish  kills in two  moderately acid lakes due to high
aluminum concentration,  pp. 310-311.  j£ Ecological  Impact of Acid
Precipitation. D. Drablos and  A.  Tollan,  eds.  SNSF Project, Oslo,
Norway.

Grahn, 0., H. Hultberg, and A. Jernelov.   1976.  IVL  B 291 Report from
Swedish Institute for Water and Air  Pollution Research.

Haines, T. W.  1981. Acidic precipitation and its consequences for
aquatic ecosystems:  A  Review. Trans. Amer.  Fish. So. 110:669-707.

Hakanson, L.  1980.  The quantitative  impact  of pH bioproduction and Hg
contamination on the mercury content of fish  (pike).  Environ. Pollut.
Bl:   285-304.

Harada, M., T. Fujimo,  T. Akagi,  and S. Nishigaki.  1976.
Epidemic!ogical and clinical Study and historical background of mercury
pollution on Indian Reserve in North Western  Ontario, Canada,  pp.
169-184.  Bull. Inst. Consit.  Med.,  Kumamoto  Univ., Kumamoto Japan. 26.

Harvey, H., P. Dillon,  G. Graser, K. Somers,  P. Fraser, and C. Lee.
1982.  Elevated metals  and enhanced  metal  uptake in fishes in
acid-stressed waters.  Abstract,  185th National Meeting, American
Chemical Society 22, 1:438-441.

Herrman, R. and J. Baron.  1980.   Aluminum mobilization in acid stream
environments. Great Smoky Mountains  National  Park, USA.  Ecological
impact of acid precipitation.   D. Drablos and A. Tollan eds.  Oslo,
No rway.

Heusgem, C. and J. DeGraeve.  1973.  Importance de 1'apport ailmentaire
en plomb Test de la Belgique, p. 85.  In Proc. Int.  Symp. Environ.
Health Aspects of Lead.  Amsterdam,  2-6~Dctober 1972.  Comm. Eur. Commn.
Luxembourg.
                                  6-65

-------
Hultberg H. and A.  Jernelov.   1976.   IVL B Pub!. 316, Swedish Institute
for Water and Air Pollution Research.

Hultberg, H. and B. Hasselrot.  1981.  Quicksilver in the Ecosystem.
pp.  33-51.  In Rapport 192 Ingenjorsvetenskapsak Kademiem.  KHM
Project, SwedTsh State  Power  Board, Vallingby Sweden.

Hultberg, H. and A. Wenblad.   1980.   Acid groundwater in southwestern
Sweden.  Proc. Int. Conf.  Ecol. Impact Acid Precip.  Norway 1980. SNSF
Project, pp. 220-221.

Jackson, T., R. Kipphut,  R. Hesslein, and D. Schindler.  1980.
Experimental study  of  trace metal  chemistry in soft-water lakes at
different pH levels.  Canadian J.  Fish. Aq. Sci. 37:387-402.

Jensen, S. and A. Jernelov.   1967.  Biosynthesis of Methyl Mercury 1.
Nordforsk Bioacid Inf.  10,4,  Swedish.

Jensen, S. and A. Jernelov.   1972.  Behavior of Mercury in the
environment.  In Mercury Contamination in Man and His Environment, World
Health Organization, Technical Reports Series No. 137, International
Atomic Energy Agency,  Vienna.

Jernelov, A.  1980. The effects of acidity on the uptake of mercury in
fish, pp. 211-222.   In  Polluted Rain.  T. Y. Toribara, M. W. Miller and
P. E. Morrow, eds.   PTenum Press,  New York.

Jernelov, A., Linse, and Hansson.  1975.  IVL B publ. 245, Swedish
Institute for Water and Air Pollution Research.

Johnston, S. E., S. A.  Norton, C.  T.  Hess, and R. B. Davis.  1981.
Chronology of atmospheric deposition  of acids and metals in New England
based on the records of lake  sediments.  Presented before Div. of
Environ. Chem. Amer. Chem. Assoc.  Atlanta, Georgia October.

KHM.  1981,  Kvicksilver i atmosfor och vatten.  pp.  279-300.  In Coal:
Health Environment Project Report.  The Swedish State Power BoaroV
Vallingby, Sweden.

Kincaid, T. C.  1979.   Alternative individual water  sources.  In Quality
Water for Home and Farm:  Proc. Third Domestic Water Qual. Sympos., pp.
73-78.  American Society of Agric. Engineers,  St. Joseph, Mich.

King, T. L. and P.  B.  Bedient. 1982.  Effect of acid rain upon cistern
water quality, pp.   244-248.   ^Proceedings of the  International
Conference on Rain Water Cistern  Systems held  at the Water Resources
Research Center, University  of Hawaii at Manuoa, Honolulu, Hawaii.

Kopeloff, L. M., S. E. Barrera, and N.  Kopeloff.  1942.   Recurrent
convulsive seizures in animals produced by  immunologic and chemical
means.  Am. J. Psychiatr. 98:881-902.
                                  6- 66

-------
Kudo, A., H. Nagase,  and Y.  Ose.   1982.  Proportional of Methyl mercury
to the total amount of mercury  in  river waters in Canada and Japan Water
Res.  16:1011-1015.

Kuntz, G.  1983.  Personal  communication,  Project Engineer, Garrett
Fleming Water Resources Engineers,  Harrisburg, PA.

Lancranjan, J., H. I. Popescu,  0.  Gavanescu,  I. Klepsch, and M.
Serbanerscu.  1975.  Reproductive  ability  of  workmen occupationally
exposed to lead.  Arch. Environ. Health 30:396-401.

Landrigan, P. J.,  E.  L. Baker,  R.  G.  Seltmann, D. H. Cox, K. V. Bdem, W.
A. Orenstein, J. A. Mather,  J.  A.  Yankel,  and I. H. Van Linder.  1976.
Increased lead absorption with  anemia and  slow nerve conduction in
children near a lead  smelter.   J.  Pediatrics  89:904-910.

Landner and Larsson.   1972.  IVL B  115, Swedish Institute for Water and
Air Pollution Research.

Mahaffey, K. R.  1977.  Quantities  of lead producing health effects in
humans:  Sources and  bioavailability.  Environ. Health Perspect.
19:285.

Mahaffey, K. R., J. L. Annest,  J.  Roberts, and R. S. Murphy.  1982a.
National Estimates of Blood  Lead Levels:   United States, 1976-1980, New
Eng. J. Med. 307:573-579.

Mahaffey, K. R., J. F. Rosen, R. W.  Chesney,  T. J. Peeler, C. M. Smith,
H. F. DeLuca.  1982b.  Association between blood lead concentration and
plasma 1,25-dihydroxycholecalciferol  levels in children.  Am. J. Clin.
Nutr. 1982; 35:1327-31.

Markesbery, W. R., W. D. Ehmann, T.  I. M.  Hossain, M. Alauddin, and D.
T. Goodin.  1981.   Instrumental Neutron Activation Analysis of Brain
Aluminum in Alzheimer's Disease and Aging.  Ann. Neurol. 10:511-516.

Marsh, D. 0., G. J. Myers, T. W. Clarkson, L. Amin-Zaki, S. Tikriti, and
M. A. Majeed.  1980.   Fetal  methylmercury  poisoning:  clinical and
toxicological data on 29 cases.  Ann. Neurol. 7:348-355.

McCabe, L. J.  1970.   Metal  levels  found in distribution samples.  AWWA
Seminar on Corrosion  by Soft Water,  Washington, D.C.

McDermott, J. R.,  A.  I. Smith,  M.  K.  Ward, I. S. Parkinson, and D. N. S.
Kerr.  1978.  Brain aluminum concentration in dialysis encephalopathy.
Lancet 1:901-904.

McKim, J. M.  G. F. Olson, G. W. Holcome, and E. P. Hunt.  1976.  Long
term effects of methylmercury chloride on  three generations of brook
trout (Salvelinus  fontinalis).  Toxicity, accumulation, distribution and
elimination.  jT~Fis. Res. Board,  Can. 33:2726-30.
                                 6-67

-------
McLean, R. A. N. M. 0. Farkas,  and D.  M.  Findlay.   1980.  Determination
of mercury in natural  waters.   Sampling and  Analysis methods, pp.
151-174.  In Polluted Rain.   T. Y.  Toribara,  M.  W.  Miller and P. E.
Morrow, ed?7  Plenum Press,  New York.

Methyl Mercury Study Group.   1980.   Methyl Mercury  Study.  McGill Univ.,
Montreal, Canada.

Miller, C. A. and E. M. Levine.  1974.  Effects  of  aluminum  salts on
cultured neuroblastoma cells.   J. Neurochem.  22:751-758.

Miller, D. R. and H. Akagi.   1979.   pH affects mercury distribution, not
methylation.  Ecotoxicol. Envir. Safety 3:36-38.

Moore, M. R.  1980.  Prenatal  exposure to lead and  mental retardation.
In Low Level Lead Exposure:   The Clinical  Implications of Current
"Research.  L. Needleman, ed.   Raven Press, New York.

Moore, M. R., P. A. Meredith,  B. C. Campbell, and A. Goldberg.  1977.
Contribution of lead in drinking water to blood-lead.  Lancet 2:717-718.

NAS.  1978.  An assessment of mercury  in  the  environment.  National
Academy of Sciences, Washington D.C.

NAS.  1980.  Lead in the Human  Environment.   National Academy of
Sciences, Washington,  D.C.

Needleman, H. L. C. Gunnoe,  A.  Leviton, R. Reed, H. Peresie, C. Maher,
and P. Barrett.  1979.  Deficits in psychologic  and classroom
performance of children with elevated  dentine lead  levels.   N. Eng. J.
Med. 300:688-95.

Nordberg, G. F. and P. Strangert.  1976.   Estimations of a dose-response
curve for long-term exposure  to methyl mercury compounds in  human beings
taking into account variability of critical  organ concentrations and
biological half times.  A preliminary  communication, p. 273.  In Effects
and Dose-Response Relationships of Toxic  Metals.  G. F. Nordbe"rg, ed.
Elsevier/North Holland, Amsterdam.

Nordberg, G. F. and P. Strangert.  1978.   Fundamental aspects of
dose-response relationships  and their  extrapolation for non-carcinogenic
effects of metals.  Environ. Health Perspect. 22:97-108.

Norseth, T.  1979.  Aluminum,  pp. 275-281.  ^Handbook on the
Toxicology of Metals.   L. Friberg,  C.  F.  Nordberg,  and V. B. Vouk, eds.
Elsevier, New York.  Chapter 15.

Nriagu, J. 0.   1979.   Production and  uses of mercury, pp. 23-40.  Ir±
The Biogeochemistry of Mercury  in the  Environment.  J. 0. Nriagu, ed.
Elsevier/North-Holland Biomedical Press,  Amsterdam.
                                  6-68

-------
 Nriagu, J. 0.  1983.  Saturnine Gout Among Roman  Aristocrats.  New Eng.
 J. Med. 308, 11:660-663.

 Nutrition Foundation's Expert Advisory Committee.   1982.   Assessment of
 the Safety of Lead and Lead Salts In Food.  The Nutrition  Foundation,
 Inc., 489 Fifth Avenue, New York 10017.

 Oliver, T.  1911.  A lecture of lead poisoning and  the  race.   Brit. Med.
 J. 1:1096-1098.

 Olson, G. F., D. L. Mount, V. M. Snarski,  and T.  W. Thorslund.   1975.
 Mercury in fathead minnows, Pimephales promelas Rafinesque, chronically
 exposed to methyl mercury in water.  Bull.  Environ.  Contam. Toxicol.
 14:129.

 Ondreicka, R., E. Ginter, and J. Kortus.   1966.  Chronic toxicity of
 aluminum in rats and mice and its effects  on phosphorus metabolism.  Br.
 J. Ind. Med. 23:305-312.

 Ontario Ministry of the Environment.  1977.  The  decline in mercury
 concentration in fish from Lake St. Clair, 1970-1976.   Report  No.
 AQS77-3, Laboratory Services Branch, Ministry of  the Environment,
 Ontario, Canada.

 Ottawa River Project.  1976.  Distribution and transport of pollutants
 in flowing water ecosystems.  University of Ottawa, National Research
 Council of Canada, Ontario, KIA OR6 Report #3.

 Overrein, L., H. Seip, and A. Tollan.  1980.  Acid  precipitation -
 effects on forest and fish.  Final report  of the  SNSF-project  1972-1980.

 Piomelli S.  1980.  The effect of low-level lead  exposure  on heme
 metabolism, pp. 67-74.  In Low Level Lead  Exposure:  The Clinical
 Implications of Current "Research.  H. D. Needlemen, ed. Raven Press,
 New York.

 Ramanoorthy, S. and D. J. Kusher.  1975.   Binding of metal ions  in river
 water, pp. 19-21.  Jji International Conference on Heavy Metals in the
 Environment.  Abstracts.  Toronto, Canada.  October 27-31.

 Registration Committee, European Dialysis  and Transplant Association.
 1980.  Dialysis Dementia in Europe.  Lancet 3:190-192.

 Reinert, R. E., L. J. Stone, and W. A. Williford.  1974.   Effect of
 temperature on accumulation of methyl mercury chloride and  p,p-DDT on
 rainbow trout (Salmo gairdneri).  J. Fis.  Res. Board,  Can. 31:649-655.

 Roels, H. A., J. P. Buchet, A. Bernard, J. Hubermont,  R. R. Lawerys, and
 P. Masson.  1978.   Investigations of factors influencing exposure and
 response to lead, mercury and cadmium in man and  animals.  Environ.
 Health Perspect. 25:91-96.
                                   6-69
409-262 0-83-21

-------
Rosen, J. F., R. W.  Chesney,  A.  Hamstra, H. F. DeLuca, and K. R.
Mahaffey.  1980.  Reduction in 1,25 dihydroxyvitamin 0 in children with
increased lead absorption.  N. Eng. J. Med. 302:1128-31.

Scheider, W. A., D.  S.  Jeffries,  and P. J. Dillon.  1979.  Effects of
acidic precipitation on precambrian freshwaters in southern Ontario.  J.
Great Lakes Res. 1979,  Internat.  Assoc. Great Lakes Res. 5, 1:45-51.

Schofield, C. L.  1978.  Toxicity of metals,  pp.  31-34.  In
Limnological Aspects of Acid  Precipitation.  G. Hendrey, edT~ Brookhaven
National Laboratory, Upton, NY.

Schofield, C. C. and J. R.  Trojner.  1980.  Aluminum Toxicity to Brook
Trout Salvelinus Fontinalis in Acidified Waters,  pp. 341-362.  Jji
Polluted Rain, T. Toribara, M. W. Miller and P. E. Morrow eds.  Plenum
Press, NY.

Semple, A. B., W. H. Parry, and  D. E.  Phillips.  1960.  Acute copper
poisoning.  Lancet 2:7152,  700.

Shahristani-al, H. and K. M.  Shihab.   1974.  Variation of biological
half-life of methyl mercury in man. Arch.  Environ. Health 27:342-.

Sharpe, W. E.  1980.  Statement  of Dr. William  E. Sharpe, Assistance
Professor of Forest Resources Extension  and Water Resources Specialist,
The Pennsylvania State University to the Select Committee on Small
Business and the Committee on Environment  and Public  Works, United  State
Senate, September 23, 1980.   Economic  Impact of Acid  Rain, U.S.
Government Printing Office,  Washington.

Sharpe, W. E. and E. S. Young.   1982.   Occurrence of  selected heavy
metals in rural roof-catchment cistern systems, pp.   249-256.  ^n_
Proceedings of the International Conference on  Rain Water Cistern
Systems, Water Resources Research Center,  University  of  Hawaii at Mauoa,
Honolulu, Hawaii.

Sherbin,  I. G.  1979.  Mercury in the  Canadian  Environment.  Economic
and Technical Review Report E.P.S. 3-EC-79-6 Environment Canada,
Ottawa.

Sibley, T.  H. and J. J. Morgan.   1977.  Metal  speciation in  aquatic
systems and equilibria, p. 319.   ^n_ First  International  Conference  on
Heavy  Metals  in the Environmental, Analytical  Papers.   Vol.  1.   T.  C.
Hutchinson, ed.  Toronto.

Stock, A. and F. Cucuel.  1934.   Propagation  of mercury.  Die
Naturwissenschafter 22:390-393.

Straby,  A.   1968.  Analysis of snow and water,  In Development of
Analytical  Methods for Mercury and Studies of iT? Dissemination  from
Industrial  Sources.  T. Westermark and K.  Ljunggeren, eds.   Swedish
Technical Research Council Stockholm.
                                  6-70

-------
Suns, K.,C. Curry, and D.  Russell.   1980.  The effects of water quality
and morphometric parameters on  mercury  uptake by yearling yellow perch.
Tech. Rep. ITS 80-1, Ontario Ministery  of  Environmental, Rexdale,
Ontario.

Swedish Expert Group.  1971.  Methylmercury  in fish  - a
toxicological-epidemiological  evaluation of  risks.   Report  from an
expert group.  Nordisk. Hyg. Tidskript. Suppl. 4.

Thornton, J. D., S. J. Eisenreich,  J. W. Hunger, and E. Gorham.  1981.
Trace metals and strong acid compositio. of  rain and snow in northern
Minnesota, pp.  261-284.  Jji Impact of  Pollutants to Natural Waters, S.
J. Eisenreich, ed. Science Publ.,  Inc., Ann  Arbor, MI.

U.S./Canada Memorandum of  Intent on Transboundary Air Pollution.  1983.
Working Group I.  Final Report. February  1983.

U.S. Department of Commerce.  1978.  Report  on the chance of U.S.
seafood consumers exceeding the current acceptable daily intake for
mercury and recommended regulatory  controls.  National Marine Fisheries
Service, Seafood Quality and Inspection Division, U.S. Department of
Commerce, Washington, D.C.

U.S. Environmental Protection Agency.   1980a.  Ambient water quality
criteria for mercury.  EPA 440/5-80-058.   Office of  Water Regulation and
Standards, Criteria and Standards  Division,  Washington, D.C. EPA

U.S. Environmental Protection Agency.   1980b.  Ambient Water Quality
Criteria for Lead.  Office of Water Regulation and Standards.  Criteria
of Standards Division, Washington,  D.C.

U.S. Environmental Protection Agency.   1979a.  The Health and
Environmental Impacts of Lead and  an Assessment of a needs  for
limitations.  Batelle Columbus Labs. OH.   U.S. Department of Commerce,
National Technical Information Service, PH-296-903.

U.S. Environmental Protection Agency.   1979b.  National Secondary
Drinking Water Regulations.  Federal Register 44:140, July  19, 1979.

Weiss, H. V., M. Koide and E. D. Goldberg.  1971.  Mercury  in a
Greenland Ice Sheet:  Evidence of  recent input by man.  Science 174:692.

Wheatley, B.  1979.  Methyl mercury in  Canada.  Medical Services Branch,
Ministry of National Health and Welfare, Ottawa, Canada.

WHO.  1972a.  Evaluation of certain food additives and the  contaminants
mercury, lead and cadmium.  World  Health Organization Technical Report
Series No. 505, Geneva.

WHO.  1972b.  The Hazards  of Persistent Substances in Drinking Water.
World Health Organization, Copenhagen.
                                  6-71

-------
WHO.  1976.  Environmental  Health Criterial  Mercury.   World health
Organization, Geneva.

WHO.  1977.  Environmental  Health Criteria 3.   Lead.   World Health
Organization, Geneva.

WHO.  1980.   Health Risk Evaluation  for methyl mercury - An Interim
Review.  EHE/EHC/80. 22.   World Health  Organization, Geneva.

Wildt, K., R. Eliasson, and M.  Berlin.   1983.   Effects of occupational
exposure to lead on sperm and semen.  _In_ Reproductive  and Developmental
Toxicity of Metals.  T. W.  Clarkson,  G.  Nordberg, and  P. R. Sager, eds.,
Plenum Press, NY.  In press.

Wong, C. S. and P.  Berrang.  1976.  Contamination of tap water by lead
pipe and solder.  Bull. Environ.  Contam. Toxicol. 15:530.

Wood, J. M.  1974.   Biological  Cycles for Toxic  Elements in the
Environment.  Science 183:1049-1052.

Wood, J. M.  1980.   The role of pH  and  oxidation-reduction potentials in
the mobilization of heavy metals.  Jji Polluted  Rain, pp.  223-232.  T.
Y. Toribara, M. W.  Miller and P.  E. Morrow,  eds.  Plenum Press, New
York.

Young, E. S. and W. E. Sharpe.   1982.  Rainwater cisterns design
construction and management.  Cooperative Extension Service, The
Pennsylvania State  University,  University Park,  PA.  15 pp.

Young, E. S. and W. E. Sharpe.   1983a.   Atmospheric deposition and
roof-catchment cistern water quality  submitted  for publication to
Journal of Environmental  Quality.

Young, E. S., Jr. and W.  E. Sharpe.  1983b.  Corrosivity of acid
precipitation and its relationship  to tapwater  copper  concentrations.
Submitted for publication to Science.

Yule, W., R. R. Lansdown, I. B. Millar,  M. A. Urbanowicz.  1981.  The
relationship between blood lead concentration,  intelligence and
attainment in a shoo! population:  a  pilot study.  Dev. Med. Child
Neurol. 23:567-76.
                                  6-72

-------
            THE ACIDIC DEPOSITION PHENOMENON  AND  ITS EFFECTS
                       E-7.  EFFECTS ON MATERIALS

                      (J. E. Yocom and N.  S.  Baer)

7.1  INTRODUCTION

     In the popular press many articles ascribe damaging  effects to
acidic deposition (LaBastille 1981).  Damage  to non-living materials and
structures is invariably listed as one of  the important effects of this
phenomenon.  Furthermore, damage to irreplaceable historic buildings and
monuments, works of art, and other cultural properties is emphasized as
one of the most important aspects of such  damage.   If one narrowly
considers the "acid rain syndrome" as precipitation that  has been
rendered more acidic as a result of long-range transport  of acid rain
precursors, this mechanism alone probably  accounts  for only a small
fraction of the total  damage to materials  attributable to the effects of
anthropogenic air pollutants.

     In general, the distinction between the  effects on materials of
near or intermediate sources from distant  sources is difficult if not
impossible to make.1  If the discussion is broadened to "acidic
deposition," which includes all  of the mechanisms by which acidic
pollutants (gases and solid and liquid particulate  matter) may contact
and damage surfaces, one is able to  point to a considerable body of
experimental evidence for damage to materials by  acidic deposition.  For
most cases, in urban areas where most materials are located, the
atmospheric load from local sources tends  to  dominate over the smaller
amounts of pollutants arriving from remote upwind sources (U.S.-Canada
1982).  This broad definition is used for  this chapter.

     Damage to materials from acidic deposition takes a variety of forms
including the corrosion of metals, erosion and discoloration of paints,
decay of building stone, and the weakening and fading of  textiles.  All
of these effects occur to a significant degree as a result of natural
environmental conditions, even in unpolluted  atmospheres.  Moisture,
atmospheric oxygen,  carbon dioxide,  sunlight,  temperature fluctuations,
and l'™ action of microorganisms all contribute to  the deterioration  of
materials.  Quantifying the specific contributions  of anthropogenic air
pollutants to such damage is a formidable  task.   Furthermore,
distinguishing the relative amount of damage  caused by specific
    Chapter A-9 of this document the  following definitions for the
 scales of acidic deposition  transport are given:  short range
 (< 100 km), intermediate range  (100  to 500 km), and long range
 (> 500 km).
                                  7-1

-------
pollutant transformation and contact processes  (for example, acid
precipitation)  becomes even  more  elusive.

     This chapter deals with the  effects on materials of anthropogenic
acidic air pollutants.  Later in  this chapter several typical broad
mechanisms for acidic deposition  are discussed.  They include adsorption
and absorption of acidic primary  pollutant gases such as S02 and N02
on moist surfaces and their  conversion to  strong acids and the  processes
in which precipitation is acidified by condensation around acidic
particles or washout of acidic primary gases.   While this chapter's
scope is extremely broad in  concept, the literature describing  research
on any one specific contact-and-effect scenario may be limited  or even
non-existent.

     There is a significant body  of literature  describing the effects of
primary air pollutants on materials as determined  by both laboratory and
field experiments.  This literature has been  summarized in detail by the
U.S. Environmental Protection Agency in its Criteria Documents
supporting the establishment of air quality  standards,  for example, the
document on sulfur oxides and particulate  matter (U.S. EPA 1981).
Several other reviews have also been published  (Yocom and Grappone 1976,
Yocom and Upham 1977, Yocom and Stankunas  1980).   A recent review in
draft form by Haagenrud et al. (1982) deals  primarily with effects of
sulfur compounds.  The draft U.S.-Canada Transboundary Report contains a
review of the literature on the effects of acidic  deposition on
materials (U.S.-Canada 1982).

     Among the documented effects of air  pollution on materials are many
that may broadly be described as associated with acidic deposition.
Table 7-1 summarizes the potential  damaging effects of air pollutants
and other environmental conditions on several classes of materials.  One
should note that sulfur oxides, other acidic  gases, and particulate
matter figure prominently among the important,  potentially damaging
pollutants, and note that moisture (as atmospheric humidity and surface
wetness) is an extremely important factor.

7.1.1  Long Range vs Local Air Pollution

     Acidic pollutants whether they are present as primary  pollutant
gases (e.g., S02 and NOX)» as fully oxidized  acids or  salts  (e.g.,
sul fates and nitrates) or in the form of acidified precipitation may,
have arrived at a material surface from local  pollutant sources or may
have been transported many miles from distant sources.  Table 7-2
summarizes the characteristics of long-range  and local  air  pollutants
and their effects.  As the table shows, several mechanisms may  be
described as acidic deposition.  The separation of long-range and local
characteristics is  somewhat artificial since phenomena  associated with
long-range transport may be generated by  local  sources  under  the
appropriate conditions.  For example, acidic precipitation may  be
produced close to sources of primary pollutants under  the  proper
meteorological conditions.  The distinction between  different acidic
                                  7-2

-------
                                       TABLE  7-1.   AIR  POLLUTION DAMAGE TO MATERIALS
                        Type of       Principal air       Other
         Materials       impact        pollutants      environmental
                                                      factors
Methods of measurement
Mitigation measures
i
CjO
Metal s


Building
Stone






Ceramics
and Glass

Paints and
Organic
Coatings



Paper




Photo-
graphic
Materials
Textiles




Textile
Dyes
Leather



Rubber


Corrosion,
tarnishing

Surface erosion,
soiling, black
crust formation





Surface erosion,
surface crust
formation

Surface erosion
discoloration,
soiling



Embrittlement,
discoloration



[•11 c rob! emi she s


Reduced tensile
strength,
soiling


Fading, color
change
Weakening,
powdered surface


Cracking


Sulfur oxides
and other acid
gases

Sulfur oxides
and other acid
gases





Acid gases,
especially
fluoride-
containing
Sulfur oxides,
hydrogen
sulfide



Sulfur oxides




Sulfur oxides


Sulfur and
nitrogen
oxides


Nitrogen
oxides
Sulfur oxides






Moisture, air,
salt, parti cul ate
matter

Mechanical ero-
sion, parti cul ate
matter, moisture,
temperature
fluctuations,
salt, vibration,
COg, micro-
organisms
Mol sture


Moisture, sun-
light, ozone,
parti cul ate
matter, mechan
ical erosion,
microorganisms
Moisture, phys-
ical wear,
acidic materi-
al s introduced
in manufacture
Parti cul ate
matter,
moisture
Parti cul ate
matter,
moisture,
light, physical
wear, washing
Ozone, light,
temperature
Physical wear,
residual acids
introduced in
manufacture
Ozone, sun-
light, physical
wear
Weight loss after removal of
corrosion products, reduced
physical strength, change In
surface characteristics
Weight loss of sample, surface
reflectivity, measurement of
dimensional changes, chemical
analysis




Loss in surface reflectivity
and light transmission, change
in thickness, chemical
analysis
Weight loss of exposed painted
panels, surface reflectivity,
thickness loss



Decreased folding endurance,
pH change, molecular weight
measurement, tensile strength


Visual and microscopic
examination

Reduced tensile strength,
chemical analysis (e.g.,
molecular weight) surface
reflectivity

Reflectance and color value
measurements
Loss in tensile strength,
chemical analysis


Loss 1n elasticity and
strength, measurement of crack
frequency and depth
Surface platinq or coating,
replacement with corrosion-
resistant material, removal to
controlled environment.
Cleaning, impregnation with
resins, removal to controlled
environment.





Protective coatings,
replacement with more
resistant material, removal to
controlled atmosphere.
Repainting, replacement with
more resistant material




Synthetic coatings, storing
controlled atmosphere
deacidi fication, encapsula-
tion, impregnation with
organic polymers.
Removal to controlled
atmosphere

Replacement, use of substi-
tute materials, impregnation
with polymers


Replacements, use of
suhsti tute material s, removal
to controlled environment.
Removal to a controlled
atmosphere, consolidated with
polymers, or replacement

Add antioxidants to
formulation, replace with more
resistant materials

-------
 TABLE 7-2.   CHARACTERISTICS  OF  LONG-RANGE AND LOCAL AIR  POLLUTION
 Pollutant
 or Effect
     Long-range
      Local
Pollutant Concen-   Low  concentrations and uniform
tration Patterns    distribution.
Sulfur Oxides
Nitrogen Oxides
Participate Matter
(includes
aerosols)
Ozone and Other
Oxidants
Dry Acidic
Deposition
Acidic
Precipitation
Acidic Fog
(includes 1 iquid
aerosol s)
S02 tends to be oxidized  to
particulate sulfates.
Significant conversion  to
particulate nitrates.
Only the smallest primary
particle sizes persist. Large
component of material  converted
from gases and vapors  to
particulate form such  as
sul fates.

Ozone and other oxidants are
produced from hydrocarbons and
NOX over moderate to long-range
transport in presence  of
sun!ight.

Dry deposition of acidic
particles (for example, sul fates)
Is possible.
Acidic rain nechanisns may be
predominantly  through droplet
condensation around acidic
particles.

Acidic fog  may be  formed by drop
condensation around small acidic
particles or other acidic
condensation nuclei.
High to moderate
concentrations and strong
gradients in time and space.

Exist primarily as S02;
however, under light  winds and
stable atmospheric conditions
conversion to particulate sul fate
can occur.

Exist primarily as MO and M02,
but under low wind speed, stable
conditions and sunlight, conversion
to organic or inorganic nitrates in
particulate form is possible.

Exists in wide range  of sizes which
may be bimodal.  Particles are
capable of producing  surface
soiling and participates in the
formation of corrosion layers
(e.g., black crust on stone).

The formation of ozone and other
oxidants is likely only under low
winds and sunlight if precursors
are present.
Dry deposition  of  acidic particles
is possible,  especially under
stable conditions, often enhanced
by moist surfaces.

Acidic rain formation may be
predominantly through rain washout
of acidic particles and pollutant
gases.

Same as for long-range.
                                         7-4

-------
deposition scenarios is especially  important when  the cost of damage
related to such deposition is  considered  and when  control strategies to
ameliorate such damaging effects are being  developed.  The transport,
deposition, damage,  and cost scenarios  of greatest economic importance
must be defined before the effectiveness  of any  control  strategy can be
estimated.

7.1.2  Inaccurate Claims of "Acid Rain" Damage to  Materials

     The popular literature contains frequent references to "acid rain"
damage to cultural property.  In most cases no attempt is made to
distinguish between  local  pollution sources and  long-range transport.
In some cases the damage is caused  by factors entirely independent of
acidic deposition.

     Perhaps the most egregious  example is  the damage to the granite
Egyptian obelisk, "Cleopatra's Meedle," located  since 1881 in Central
Park in New York City.  In one account, it  was stated that, "The city's
atmosphere has done  more damage  than 3  1/2  millenia in the desert, and
in another dozen years the hieroglyphs  will  probably disappear (New York
Times 1978a). "  A careful  study of the monument's complex history makes
it clear that the damage can be  attributed  to advanced salt decay, high
humidity of the New  York climate, and unfortunate  attempts at
preservation (New York Times 1978b, Winkler 1980).

7.1.3  Complex Mechanisms of Exposure and Deposition

     The work done to date to  measure damage to  materials from acidic
deposition has not considered  to any significant degree  the specific
mechanisms of exposure, deposition, and subsequent damage.  As will be
discussed in the next section, most of  the  studies that  have used
laboratory chamber exposure or field exposure in the ambient atmosphere
are unable to isolate specific deposition mechanisms from the many
interrelated chemical  and physical  processes involved.   The following
list presents a series of simplified mechanisms  that the authors believe
occur in one form or another.  These mechanisms  are based upon the
presence of acidic gases such  as $03 and  N02, their transformation
products, and moisture in some form.

     1.  Dry Gas, Dry Surface:   An  acid gas is adsorbed  on a relatively
         dry material  surface  (for  example,  building stone) and
         exposure to moisture  forms acids that attack the material.

     2.  Dry Gas, Wet Surface:   An  acid is  absorbed in moisture
         (condensed  dew or collected precipitation) already on surfaces
         and results in acid attack.

     3.  Large, Dry  Particle,  Dry Surface:   Large  particles containing
         acid components fall  on the material's  surface  and lead to
         damage directly.   An  example would be acid-containing soot from
         an oil-fired boiler.
                                 7-5

-------
     4.  Small Particle,  Dry or Wet Surface:   A  small  particle
         containing acidic  compounds such  as  sulfuric  or nitric acid
         salts capable of reacting with  moisture to  form acids  settles
         on or impacts on a dry or wet  surface and subsequently leads
         to acid attack.

     5.  Acid Precipitation:  Rain or snow containing  acidic components
         falls on the material  surface and leads to  damage directly.

     The above group of simplified mechanisms is not intended to be
exhaustive or completely  rigorous.  They are  illustrative of the wide
spectrum of processes that  operate to produce acidic deposition and each
of the listed mechanisms  may have  one or more variations.  For example,
in Mechanism 1 (Dry Gas,  Dry Surface) it is likely,  in the case of S02
contact, that some surface  oxidation may take place  within a relatively
dry adsorbed layer or that  S02 may react directly with a reactive
surface to produce a sulfite salt.  Nevertheless, as will become
apparent later in this chapter, acidic deposition and  subsequent damage
accelerates in the presence of moisture.

     The end result of each of these mechanisms  is acidic deposition
capable of damaging materials.   Yet certain of these mechanisms are
undoubtedly more important  than others  in  causing economically
significant damage.  In large population centers where levels of
primary, gaseous pollutants, and total material  inventories are high,
Mechanisms 1, 2, and 3 may  be more important  than 4  and 5.  In rural
areas where the inventory of exposed materials is likely to be different
than in urban areas, and  the pollutant mix may include a higher portion
of secondary, particulate pollutants, Mechanisms 4 and 5 may dominate.

     These factors and others such as the  distinction  between wet and
dry deposition mechanisms are important  because  of the link between
pollutant levels and meteprological  factors.   For example, if a local
source has an elevated emission point, the kind  of surface inversion
associated with radiational cooling and  dew formation  may also act to
keep the pollution from reaching ground  level.   Thus,  Scenario 2 may not
be especially important,  even though all the  critical  components (active
pollutant, susceptible material, wet surface)  are all  present on an
annual average basis.  Conversely, materials  on  elevated terrain may be
subject to pollutant plume  impact  only rarely, but when they are
affected, the conditions  (such as wetness) may be such that the maximum
degree of damage occurs.

     Note in Table 7-2 that Mechanisms 4 and  5 (small  particle, dry or
wet surface; and acid precipitation) may occur both  locally and after
long-range transport.  Stable atmospheric  conditions and low wind speeds
may provide the time necessary for atmospheric transformations  to create
effects on a local scale  that would otherwise be associated with
long-range transport.
                                  7-6

-------
7.1.4  Laboratory vs Field Studies

     The effects of acidic deposition on  materials  have been  studied
under both laboratory and field conditions.   In laboratory  studies, the
conditions of exposure can be controlled,  and the specific  effects of a
single pollutant or environmental  parameter can be  isolated.  However,
to produce measurable material  damage in  a reasonable  time  period, the
material is often exposed continuously to severe environmental
conditions (e.g., extremely high pollutant concentrations and/or high
humidity) completely unrepresentative of  field conditions.  Furthermore,
the exposure conditions are programmed through predetermined  cycles that
may only remotely resemble the complex interactions of temperature,
humidity, surface wetness, sunlight,  pollutant concentration, and other
environmental  factors occurring in the ambient atmosphere.  In this
context, laboratory experiments have  thus far been  unable to  represent a
true picture of the effects of pollutants under conditions  of long-range
transport, where such transformation  would have had ample opportunity to
take place.

     Field studies normally consist of exposing samples of  materials to
ambient atmospheres representing various  combinations  of  pollutant
concentrations and other environmental  factors.  By comparing damage
level (e.g., loss of surface material)  with  pollutant  concentration and
other environmental factors (e.g., humidity,  "time-of-wetness", or pH of
rainwater) statistical models may be  developed for  the damage.  The
principle difficulties with this approach are

         Materials exposed may not represent  materials in actual use.

     0   In normal use materials are  found in combination.  Field
         studies may not include interactions of other materials in
         contact with test materials.

     0   Damage is a complex function of  many environmental conditions,
         and the effect of one condition  is  difficult  to  isolate.

         Measured variables may be interrelated (e.g., pH of  rain may be
         dependent upon S02 level).


7.1.5  Measurement of Materials Damage

     Material  damage is usually measured  by  noting  quantitative changes
in some physical or chemical feature  of the  material  (e.g., weight or
thickness of a sample, surface color  or reflectivity,  chemical analysis,
and identification of corrosion products).

7.1.5.1  Metals—Corrosion of metals  may  be  measured  by weight changes
resulting from the accumulated corrosion  products before  and  after a
predetermined exposure period.   However,  for long exposures,  corrosion
products tend to spall or wear off.  Thus corrosion products  formed
during the exposure period are usually removed chemically to  determine


                                  7-7

-------
damage by weight of metal lost.  Another method applicable  to metals  is
measurement of changes in sample thickness,  which  in  some  cases may  be
obtained from the electrical  resistance.   Mechanical  tests involving
bending are frequently used to test for stress corrosion.

     Physical  methods such as scanning  electron microscopy, x-ray
diffraction, and x-ray fluorescence can be used to characterize the
physical  and chemical nature  of corrosion  products.

7.1.5.2   Coatings—Paint and other coatings damaged  by environmental
exposure usually erode,  so measurement  is  most conveniently done by
measuring weight loss of painted panels.   Surface  darkening by deposits
of particulate matter or reactions between pigments and air pollutants
are usually measured photometrically.

7.1.5.3  Masonry—Samples of  building materials such  as stone, mortar,
and concrete can be weighed before and  after exposure to determine
erosion rates.  Caution  must  be  exercised  in interpreting  such data
because conversion to new phases may involve weight gain without obvious
change in physical  appearance.  Discoloration of such samples from
exposure to dark particulate  matter can be measured photometrically.  A
series of photographs of buildings taken over sufficient time periods
may provide a qualitative assessment in the  form of soiling and/or loss
of surface detail.

7.1.5.4  Paper and Leather—The  embrittlement of paper is  accelerated by
exposure to acidic deposition.  Excess  acidity can be observed by
combination surface electrode pH measurements.  Resulting  damage may be
determined by measuring  folding  resistance.   Weakening of  leather caused
by acidic deposition can be quantified  by  means of tensile  strength
tests.

7.1.5.5  Textiles and Textile Dyes—Certain  textile materials are
weakened by acidic deposition.Such damage  is best determined by
measuring loss in tensile strength.  Cotton  is also weakened by
biological  processes (e.g., mildew), and methods have been developed to
differentiate between acidic  deposition and  these  biological mechanisms
by determining the relative molecular weight of the exposed material.
Damage from acidic chemical attack causes  depolymerization and reduction
in average molecular weight,  while biological  attack  causes essentially
no reduction in average  molecular weight.

     Textile dyes are affected by N02.   Changes in color values from
such damage are measured by specially designed colorimeters or
spectrophotometers capable of detecting small  changes in color within
narrow ranges of the visible  spectrum.

7.2  MECHANISMS OF ACIDIC DEPOSITION

     Acidic deposition damages a wide range  of materials.  This section
will cover some of the principal  damage mechanisms for selected classes
of materials that are widely  used and economically important.
                                 7-8

-------
7.2.1  Metals

     The atmospheric corrosion of metals  is  generally an electrochemical
process governed by diffusion of moisture, oxygen, and acidic  pollutants
(e.g., S02) to the surface.   The EPA  Draft Criteria Document  for
Sulfur Oxides and Particulate Matter  (U.S. EPA  1981)  provides  a review
of the primary mechanisms governing the corrosion of metals in the
presence of S02 and moisture.  This review is based on the research of
many workers, and it deals primarily  with the effects of S02 and
moisture on metals and other materials.   However, most of the  scenarios
discussed fall within the general definition of acidic deposition.

     The rusting of metals is an oxidation  process that is accelerated
by the presence of acidic pollutants.   Barton (1976) has proposed the
following set of reactions involving  the  oxidation of $03 to  sulfate
on iron surfaces:

     S02 + 02 + 2e~ + S042-                                      [7-1]
or
     4 HS03" + 3 02 + 4e- •*  4 S042' + 2 H20.                    [7-2]

     The electrons are provide by the oxidation of the metal  (M):

     M -> Mn+ + ne-                                              [7-3]

     Barton (1976) noted that rusting of  iron occurs  first at  isolated
sites and then spreads across the entire  surface.  This phenomenon  is
not well understood but may  relate to a variety of factors including
differential deposition rates of S02  or acidic  partial!ate matter,  the
influence of rust deposits on subsequent  corrosion, and variations  in
"time-of-wetness" in relation to electrolyte concentrations at various
points on the surface.  Rice et al. (1982) believe that moisture forms
in "clusters" on metal surfaces even  in indoor  environments and at  the
site of these clusters, corrosion is  initiated. While rust deposits
increase the absorption of S02, a thin layer of iron  oxide on  steel
will provide some degree of protection from  subsequent atmospheric
corrosion.  In fact, special steel alloys whose iron  oxide layers
provide considerable protection against further corrosion have been
developed for bold, unprotected exposures.   The corrosion products  on
several non-ferrous metals (zinc, copper, and especially aluminum)  tend
to suppress the absorption of S02.

     Moisture is always required for  metal  corrosion, each metal tending
to have a critical humidity  above which corrosion tends to accelerate.
Depending on the specific metal, these critical humidities are in the
range of 60 to 80 percent RH.  The relative  length of time a  metal
surface is wet {"time-of-wetness") is the single most important variable
affecting the acceleration of corrosion by  acidic deposition.  Some
workers (U.S. EPA 1981) have found that hygroscopic corrosion products
(e.g., iron sulfate) cause metal surfaces to remain wet at lower RH than
if these products were not present.
                                  7-9

-------
7.2.2  Stone

     Stones composed almost entirely  of calcium  carbonate  (limestone,
marble, travertine, etc.)  or stones whose cementing  material  is calcium
carbonate are particularly vulnerable to damage  from acidic  deposition.

     The attack of sulfur dioxide on  such carbonate  stones has been
studied for over a century.  Yet, no  quantitative  relationship has been
developed between ambient S02 levels  and resulting materials damage.

     The general decay mechanism includes aerodynamic factors
controlling delivery of S02 to the stone surface,  oxidation  of S02
to sulfate and the subsequent reaction with the  carbonate  surface,
mechanical  stress by which reaction products destroy the  stone
structure,  and removal of the stone and its alteration products by
rainfall and other weathering phenomena (Livingston  and Baer 1983).

     Although the primary air pollutants causing damage to stone  are
sulfur compounds, a comprehensive decay mechanism  must include the roles
of nitrogen compounds, carbon dioxide, and water.  For the carbonate
stone/sulfur compound system three general modes of  attack pertain:

     Gaseous SO?

          S02 + CaCOa + CaSOa + C02      (Step 1)                  [7-4]

          CaS03 + 1/202 ->CaS04          (Step 2)                  [7-5]

     Wet Deposition

                                        C02                       [7-6]
     Dry Deposition is exemplified by the reaction between sul fates in
particul ate matter and calcium carbonate either in the form of sulfuric
acid as in wet deposition, or in the form of ammonium sul fates (Stevens
et al 1980).
                    + CaCOs -> CaSCty + (NH4)2C03                    [7-7]

          NH4HS04 + CaCOs -> CaSCH + NfyHCOa                        [7-8]
     The anhydrous CaS04 is hydrated to form gypsum, which is highly
 susceptible to surface erosion.

     Humidity plays a key role in all aspects of the interactions of
 S0v with carbonaceous stone.  In autoradiographic experiments using
 sulfur-35, Spedding (1969b) showed surface saturation of oolitic
 limestone samples by SO? at 81 percent RH occurring in less than ten
 minutes.  However, at the same concentrations but at 11 percent RH only
 a few distinct sites showed reaction after 20 minutes exposure, with
 approximately 25 percent of the total S02 uptake measured for the high
 humidity case.  Tombach (1982) has summarized the factors contributing
 to stone decay as shown in Table 7-3.
                                  7-10

-------
               TABLE  7-3.    CLASSIFICATION  OF  MECHANISMS CONTRIBUTING TO  STONE  DECAY3
                                         (ADAPTED FROM TOMBACH  1982)
Mechani sm
                                                                                  Temper-
                                                          Rainfall  Fog  Humidity   ature
                                                                                              Solar
                                                                                            Insolation   Wind
  Gaseous
Pollutants  Aerosol
External  Abrasion
  Erosion by wind-borne particles
  Erosion by rainfall
  Erosion by surface Ice

Volume Change of Stone
  Differential expansion of mineral  grains
  Differential bulk expansion due to uneven heating
  Differential bulk expansion due to uneven moisture
    content
  Differential expansion of differing materials at
    joints

Volume Change of Material In Capillaries and Interstices
  Freezing of water
  Expansion of water when heated by sun
  Trapping of water under pressure when surface freezes
  Swelling of water-imbibing minerals by osmotic pressure
  Hydration of efflorescences, internal impurities, and
    stone constituents
Crystallization of salts
Oxidation of materials  into more voluminous forms

Dissolution of Stone or Change of Chemical Form
  Dissolution in rainwater
  Dissolution by acids  formed on stone  by atmospheric
    gases or  particles  and water
  Reaction of stone  with SOg to  form water-soluble
    material
  Reaction of stone  with acidic clay aerosol particles

Biological Activity
  Chemical  attack  by  chelating,  nitrifying,  sulfur-
    reducing or sulfur-oxidizing bacteria
  Erosion by  symbiotic  assemblages  and higher  plants
    that penetrate stone or  produce damaging excretions
       circles denote principal atmospheric  factors; open circles denote  secondary factors.

-------
7.2.3  Glass

     Glass weathering is the  process  of  removing alkali cations (e.g.,
Na+ and K+) from glass by reaction with  water or sulfur dioxide.
The reaction with water involves  the  exchange of sodium ions by hydrogen
ions with the rate of reaction  limited by the diffusion of sodium ions
to the surface.   The reaction with sulfur dioxide in the range 20 to 100
C in gas saturated with SOg involves  the same process at approximately
the same rate as with water alone (Douglas  and Isard 1949).

7.2.4  Concrete

     Cement, concrete, and steel  reinforced concrete structures are all
subject to complex actions reducing their durability.  The alkaline
nature of cement has led to general neglect of the effects of acid
deposition and acidified water  runoff on concrete/cement durability
although it is recognized that  any reaction reducing matrix alkalinity
will be harmful.  The role of chloride ion  as a major contributor to
corrosion of reinforced concrete  is well established (Browne 1981,
Volkwein and Springenschmid 1981).  The  alkalis in the hardened cement
passivate the reinforcing steel while penetrating chlorides depassivate
the iron.  Other factors in corrosion of the steel include the
development of electrolytic corrosion cells and the penetration of
atmospheric 02 through the concrete to the  steel.  The reaction of
SOo and 50^2- yith cement involves the formation of calcium
sulfate and calcium sulfate aluminum  hydrate (ettringite).

7.2.5  Organic Materials

     Most organic materials exposed boldly  to the atmosphere are quite
resistant to the effects of acidic deposition.  Deterioration of such
materials is determined primarily by  the effects of atmospheric oxygen,
ultraviolet (UV) light, and atmospheric  oxidants such as ozone.  Painted
surfaces are the most widely  employed organic surfaces exposed to the
atmosphere and to some degree are susceptible to acidic deposition.
However, these effects are relatively minor when compared with natural
environmental factors such as sun and precipitation (Haynie et al 1976).

     Paint damage from acidic deposition is strongly related to the
paint formulation.  Such factors  as the  ratio of pigment and extenders
to film forming  ingredients determine the hardness, flexibility, and
permeability of  the surface.  It  has  been shown that the presence of
extremely high concentrations of  $03, a  reducing gas, can interfere
with the drying  process, which  is an  oxidation-polymerization reaction
(Hoibrow 1962).   However, it  is doubtful that S02 concentrations at
the present time in any area  of the United  States would be high enough
to cause this potential problem.

     The most realistic mechanism for damage to paint by acidic
deposition is reaction between  acidic materials and pigments (e.g., ZnO)
                                  7-12

-------
and extenders such as CaCCh.  The long-term effect is  the  loss of
paint surface through erosion.

     The degradation of paper and textiles is dominated by three
factors:  light, humidity, and acidity.   Paper and other cellulosic
materials (e.g., cotton, linen, and rayon)  are highly  susceptible to
acid hydrolysis at the glucosidic linkage in the  cellulose chain.  Among
proteinaceous textile materials silk is  most susceptible to damage by
light.  In bright light silks may lose 60 percent of their strength in
as little as 8 weeks of exposure (Leene  et al.  1975).

7.2.6  Deposition Velocities

     Chemical reaction between exposed surfaces and air pollutants leads
to removal of the pollutant from the atmosphere.   Deposition  rates are
quantified using the expression:

     Flux = Vg C ,                                             [7-9]

which relates the flux of a pollutant gas to a  surface to  the
atmospheric concentration C above the surface.  The deposition velocity,
Vg, depends on the specific gas/surface  combination.   Other factors
influencing Vg are humidity, surface roughness, air velocity, and
turbulence.  The determination of Vg is  usually made by measuring the
change in concentration above the surface or measuring the rate of
deposition at the surface.  Judeikis (1979)  has compiled deposition
velocities for various materials in contact with  sulfur dioxide and
ozone.  Table 7-4 presents the deposition velocities for sulfur dioxide.
(More extensive discussion of deposition processes can be  found in
Chapter A-7.)

7.3  DAMAGE TO MATERIALS BY ACIDIC DEPOSITION

     A wide range of sensitive materials can be damaged by acidic
deposition.  However, this chapter will  deal  only  with those  choices
that are judged to be economically and culturally  most important.  These
material classes are:

     °   metals
     o   masonry

     °   paint and other coatings
     o   cultural  property (historically and culturally valuable
         structures and objects)

     *   other materials (paper,  photographic materials, textiles, and
         leather)

7.3.1  Metals

     The position of metals in the electromotive  series determines their
relative reactivity.   However, the solubility of  the particular metal
                                  7-13

-------
TABLE 7-4.  MEASURED DEPOSITION VELOCITIES FOR SOo ON VARIOUS SURFACES
                       (COMPILED BY JUDEIKIS 1979)
   Surface^                                       Vg (m m1n~l)b
Cement (5)
Limestone (6)
Copper
Leather (18)
Steel
Fabric (2)
Wood (7)
Aluminum (2)
Gloss Paint
Asphalt
Carpeting (3)
Wallpaper (17)
Solid Floor Materials (25)
0.6
> 0.021
> 0.001
> 0.1
> TJ.001
"0.010
0.016
0.001
0.001
0.024
0.005
0.002
0.0003
- 1.6
- 0.63
- 0.26
- 0.2
- 0.13
- 0.033
- 0.031
- 0.029
- 0.025

- 0.014
- 0.010
- 0.003
^Number in parentheses  indicates the number of different
 materials examined if  greater than one.

bAs defined by  Equation 7-9.
                                 7-14

-------
salt and the stability of the metal  oxide coatings  that  tend  to  form  in
the atmosphere determine metals'  abilities to  corrode as a  result of
acidic deposition.  For example,  aluminum is high in the series, but
aluminum oxide coatings that form in the  atmosphere resist  corrosion
even in the presence of significant amounts of acidic deposition.
However, even aluminum may be pitted in atmospheres containing sea  salt
or large, acidic particles.

     Thermodynamic considerations governing electrochemical corrosion
are conveniently examined with the help of Pourbaix potential-pH
diagrams.  By plotting electrical potential  against solution  pH, regions
of stability for various chemical species can  be  indicated.   In
simplified form, when reactions to form  soluble species  occur, one  has
"corrosion;" when the free metal  is stable the region is designated
"immune" to corrosion; and when a chemically stable oxide or  salt film
forms on the surface, leaving the metal resistant to subsequent  attack,
the region is one of "passivation" or mitigation  of corrosion.   Pourbaix
(1966) has developed diagrams that show areas  of  stability, corrosion,
and passivity for various combinations of electrode potential and pH,
several of which are presented as Figure  7-1.

     In using these diagrams to determine the  effect of  lowered  pH  on
corrosion one must determine the potential attained by the  metal in the
natural environment.  Moreover, reduced pH tends  to increase  the
solubility of corrosion products.  While  the corrosion products  in
unpolluted atmospheres may be relatively  insoluble, in polluted
atmospheres quite different corrosion products may  form  which may be
considerably more soluble.  This potentially synergistic problem is
sometimes overlooked in traditional  writings on corrosion.  The Pourbaix
diagram can give much insight into this process.  However,  caution  must
be exercised in interpreting these diagrams because kinetic factors with
non-equilibrium behavior may govern corrosion.

7.3.1.1  Ferrous Metals--Corrosion of iron and steel in  polluted
atmospheres has received a great deal of  attention  over  the years.
Steel, unless it is an alloy designed for unprotected exposure is
usually coated by painting or plating (e.g., zinc)  when  used  in  outdoor
exposures.  Nevertheless, data on iron and steel  corrosion  provide
valuable information on the relative importance of  acidic deposition
components and the mechanisms causing damage.   The  Pourbaix diagram for
the iron system is presented as Figure 7-2.  It illustrates the
relationships among normal corrosion products  and the equilibrium pH  and
potential conditions for their stability.

     Some of the earliest work on the nature of iron corrosion in
atmospheres containing acid gases and moisture was  that  of  Vernon
(1935). He showed that in the presence of S02  and moisture, iron
corrosion proceeds from randomly distributed centers which  he associated
with the deposition of particulate matter.

     As pointed out in Section 7.2.1, many theories have been advanced
on the principle chemical reactions that  describe iron and  steel
                                  7-15

-------
                   14
0     7     14
                                                     0    7     14
            GOLD
   SILVER
 COPPER
            LEAD
    IRON
                                                     0     7    K
            ZINC
   CHROMIUM
ALUMINUM
                                       LEGEND:
                                                     STABLE (IMMUNE)


                                                     CORROSION
                                                     PASSIVATION
Figure 7-1.   Pourbaix  diagrams  for various metals.  The ordinate is in
             volts (electron  potential  standard  hydrogen electrode) and the
             abscissa  is  in units of  pH.  The upper thin diagonal line is
             the 0? evolution line while  the lower ine is that for H2
             evolution.   Adapted from Pourbaix (1966).
                                   7-16

-------
         LU UJ

         o t/i
         CL.

         UJ 
         O >
         O
         o; >
         H-*—•
         O
         LU
Figure 7-2.  Pourbaix diagram for the system  Fe, Fe2+f Fe3+,
             Fe3d4, and F6203-  The thin diagonal lines indicate
             regions of water stability.  Compare with  Figure 7-1  for
             designated regions of "corrosion," "immunity,  and
             "passivation."  The reactions considered are:
             1) Fe = Fe2+ + 2e-
             2) 3Fe + 4H20 =
             3) 3Fe2+ + 4H20 =
             4) 2Fe2+ + 3H20 =
             r- \ i- O±   _ Ox
             5) Fe^+ = Fe6  +
             6) 2Fe3+ + 3H20 =
             7) 2Fe304 + H20 = 3Fe203
8H+ + 8e-
+ 8H+ + 2e~
+ 6H+ + 2e-


+ 6H+
   2H+ + 2e-
                                   7-17

-------
corrosion In the presence of SO? and moisture.   Various chemical
routes are possible and it is likely that several  of them  operate  in
actual atmospheric exposures.

     According to Nriagu (1978), once corrosion has been initiated, the
progress of the reaction is controlled largely  by  sulfate  ions  produced
from the oxidation of absorbed or adsorbed S02. However,  the actual
mechanism of S02 oxidation on the surface is poorly understood.  The
work of Johnson et al. (1977) appears to show that sulfur  or sulfates
are only a minor constituent of the corrosion products  of  steel.   Mild
steel samples were exposed to two urban areas near Manchester.   One area
was heavily polluted, and the other was lightly polluted.   Scanning
electron microscopy, energy dispersive x-ray analysis,  and x-ray
diffraction analysis of corrosion products showed  them  to  be
predominantly Y -Pe2Q3'^2Q» a -Fe203'H20 and a -FeOOH.   Some  minor
amounts of sulfur were found in a few of the samples.   While not
discussed in the article, the possibility exists that any  sulfates
formed were soluble and washed away.   The relative amount  of corrosion
produced was strongly dependent on whether the  sample was  initially wet
at the beginning of the exposure.

     An iron oxide corrosion layer tends to  reduce the  rate of  further
corrosion of iron and steel.  Nriagu  (1978)  and Sydberger  (1976) showed
that steel samples exposed initially  to low  concentrations of sulfur
oxides were more resistant to further corrosive attack  than samples
exposed continuously to high concentrations.  This suggests that the
composition of the initial layer is critical  in determining the  nature
and extent of subsequent corrosion.

7.3.1.1.1  Laboratory studies.  Exposing iron and  steel  samples  to S02
and moisture under controlled laboratory conditions has two prinicipal
advantages:

     1.   The pollutant concentrations and other influencing factors can
         be independently controlled  in a factorial  experiment and
         permit the quantification of each factor's impact.

     2.   Exposure conditions can be made more severe than  in nature to
         accelerate the corrosion effect,  thereby  reducing the duration
         of the experiment.

     While many of the early experiments showed clearly that corrosion
rates correlate with both S02 and humidity,  exposure consisted of
S02 concentrations many times higher  than those found in the ambient
atmosphere,  or what are referred to  as "reflux" conditions, where  water
and excess S02 were continuously flushing the surface of the samples.

     The set of laboratory experiments most  clearly  approximating  field
conditions was conducted by  Haynie et al.  (1976).   Various materials
were exposed to controlled pollutant  concentrations and moisture
conditions at levels encompassing those found in ambient urban
atmospheres.   Sunlight and the formation of  dew were also  simulated.
                                  7-18

-------
Steel corrosion was determined in  terms of weight loss of the steel
panels by chemically removing  the  corrosion  products, and the results
showed a strong, statistically significant relationship between steel
corrosion and S02 concentration, together with high humidity.

7.3.1.1.2  Field studies.   A  problem  inherent with field studies is that
iron and steel corrosion occurs even  in unpolluted atmospheres, and the
impact of specific acidic  deposition  scenarios is difficult to isolate
completely.  Therefore,  the effects of acidic deposition can only be
inferred by statistical  treatment  of  the data.

     Upham (1967) exposed  mild steel  samples in a number of sites in and
around St. Louis and Chicago.   He  showed that corrosion correlated well
with sulfur oxide levels and  increased with length of exposure.
Starting in 1963, Haynie and  Upham carried out a five-year progam in
which three different types of steel  were exposed in eight major
metropolitan areas in the United States.  Multiple regression analyses
showed significant correlations between average S02 concentrations and
corrosion for all three  types of steel.  No attempt was made to relate
damage to the joint occurrance of  S02 and moisture (relative humidity
or time-of-wetness).

     In 1964, Haynie and Upham (1971) exposed steel samples for 1 and 2
years at 57 stations of  the National  Air Sampling Network.  Pollutants
of interest were S02, total suspended particulate matter, and the
sulfate and nitrate content of the particulate matter.  An empirical
function was developed relating sulfate in particulate matter and
humidity to corrosion.   However, the  authors believed that S02 rather
than sulfate was the causative agent  in producing corrosion, and the
relationship was transformed  into  one based on S02 from a linear
regression between sulfate and S02.   The corrosion or damage function
is:

     cor = 325  /T etO.00275  S02-(163.2/RH)]                     [7-10]

where

     cor = depth of corrosion, ym,
       t = time, years,
     S02 = S02 concentration,  yg nr3,
      RH = average annual  relative humidity, percent.

     Figure 7-3, based on  the above damage functions, shows the
relationship between pseudocorrosion  rate (cor /t~l), relative
humidity, and S02 concentrations.  This graph shows that the corrosion
rate is much more sensitive to humidity than to S02, especially at
levels of S02 normally experienced in urban areas.

     For example, referring to Figure 7-3, if one were comparing
relative corrosion at 55 percent relative humidity in two areas with
average levels of 100 and  150  yg nr3, a very significant difference
in relative air pollutant  levels,  the difference in relative corrosion


                                  7-19

-------
         oc

         z
         o
         o
         cei
         C£.
         O
         o
         o
         o
         Z3
         LU
         I/O
         D_
100


 90


 80


 70


 60


 50


 40


 30


 20


 10

  0
                         100
                      200
 55% RH
300
400
                    AVERAGE S02 CONCENTRATION, jjg m-3
Figure 7-3.  Steel  corrosion behavior as a function of average sulfur
             dioxide concentration and average relative humidity.  Adapted
             from Haynie and Upham (1974).
                                    7-20

-------
would be approximately three pseudocorrosion units.   On the other hand,
if one were comparing relative corrosion at a constant $03  level  of
100 yg m~3 between two areas with a moderate difference in  average
relative humidity (55 and 65 percent), the difference in relative
corrosion rate would be approximately 15 pseudocorrosion units.

     This damage function shows that the sensitivity of corrosion to
humidity is far greater than that to S02, especially at levels  of
S02 normally experienced in urban areas.

     A number of other damage functions relating steel  corrosion  to
S02 and humidity (or time-or-wetness) have been developed by several
other workers which have been summarized by EPA (1981)  and  Haagenrud  et
al. (1982).  It should be noted that nearly all  metal  corrosion damage
functions have been developed by regression analysis and do not include
terms for precipitation.

     A recent study of material  damage in the St.  Louis area in 1974-75
by Mansfeld (1980)  included the use of special  atmospheric  corrosion
monitors which measure the length of time that a corrosion  panel  was  wet
enough for electrochemical corrosion to take place (time-of-wetness).
His sample exposure array included weathering steel, galvanized steel,
house paint, and Georgia marble.  Concentrations of $62 measured  in
this study were an  order of magnitude lower than those  measured in
Upham1s earlier study (Upham 1967).  Mansfeld was  unable to show  any
significant correlation between corrositivity and  pollutant levels.

     Some of the experiments of Vernon (1935)  showed that moist air
polluted with S02 and particles of charcoal produced corrosion much
more rapidly than air containing S02 and moisture  alone.  He reasoned
that the effect of the particles was primarily  physical in  that they
increase the S02 concentration.   Sanyal  and Singhania  (1956)  stated
that particulate matter had a "profound" effect on corrosion rates.
They believed that  the influence of particulate matter  on corrosion was
related to its electrolytic, hygroscopic and/or acidic  properties,  and
its ability to absorb corrosive pollutant gases.  While these laboratory
studies appear to show a strong influence of particulate matter with
corrosion, field studies have not confirmed this effect.

     Haynie (1983)  has attempted to address the effects of  small
particles on materials.  Lacking a significant body  of experimental
data,  he has approached the question theoretically,  using data  on
deposition velocities.  He considered four species of small  particles:
carbon, sulfuric acid, ammonium sulfates, and ammonium  nitrate.   He
concludes that data from one study (Harker et al.  1980) confirmed the
chemical  models for damage, and based on calculated  pollutant fluxes,
S02-induced damage  will tend to dominate over H2S04  effects in
most urban areas.

     Measurement of the effects of pollutants associated with long-range
transport (e.g., acid precipitation) as compared with locally generated
                                  7-21

-------
pollutants (e.g.,  primary  pollutant  gases)  is just getting under way in
the United States.   The Scandanavians  have  been addressing this question
for some years.   In summarizing  several year's work in Norway, Haagenrud
(1978) states that  monthly corrosion rates  for carbon steel are strongly
influenced by long-range transport of  both  acid precipitation and
However, episodes of precipitation of  < pH  4.0 occur so seldom that
these episodes do  not strongly  influence  long-term corrosion rates.
Similarly, episodes of high S02  concentration also affected monthly
corrosion rates, but had little  effect on long-term values because they
occurred so seldom.

7.3.1.2  Nonferrous Metals--The  corrosion rates of commercially
important nonferrous metals in  polluted atmospheres are generally less
than those for steel  but cover a wide  range.  Figure 7-4, from the work
of Sydberger and Vannenberg (1972),  shows adsorption of S02 with time
at 90 percent relative humidity  for  iron  and three nonferrous metals.
Copper and aluminum have relatively  low adsorption capacities for $03,
confirming the lower sensitivity of  these metals to attack by S02 in
the presence of moisture.

     These tests were carried out by exposing polished metal surfaces to
the test conditions over very short  exposure periods.  While the results
appear to confirm the relative  sensitivity  of these metals to acidic
deposition and attack, the exposure  conditions bear little relationship
to real life conditions.  Rice  et al.  (1982) point out that a pure metal
surface rarely presents itself  to the  atmosphere for more than a few
microseconds.  Water is rapidly  absorbed  in the surface films and may
exist as moisture clusters as pointed  out in Section 7.2.1.  Further-
more, corrosion products and salts  from  surface contamination (e.g.,
chlorides) greatly  influence corrosion rates, principally through
lowering of the critical humidity—the point where corrosion rates begin
to accelerate.

     Only limited evidence links NOX with damage to non-ferrous
metals, though a number of corrosion problems with telephone equipment
have been traced to NOX and high nitrate  concentrations in airborne
dust.  In a laboratory study of nickel-brass wire  springs, stress
corrosion cracking was observed  when surface concentrations of nitrate
reached 2.1 mg cm'2 and RH was  about 50  percent.  To  avoid the
nickel-brass corrosion problem,  zinc has  been eliminated  from the alloy,
and the cooling systems for existing equipment  have been  modified to
keep the RH below 50 percent in  NOX  impacted areas (Harrison 1975).

     Such damage to components  in communications switch gear is an
insidious problem because a simple  malfunction  can  put a  large  system
out of service.

7.3.1.2.1  Aluminum.  Aluminum  is quite  resistant  to  S02-related
acidic deposition.   However, the presence of particulate  matter may
produce a pitted or mottled surface  in the  presence of $03 and
moisture.  In view of the reductions in  emissions  of  S02  and
particulate matter—especially  larger  particles or agglomerates that
                                  7-22

-------
  CM
         3 -
   u

   CD
   CVI
   o
   CO
   o;
   o
   to
   Q
         1  —
                           I     1     I      I
2 —
                                                              10
                              EXPOSURE TIME, hr
Figure 7-4.  Adsorption of sulfur dioxide on polished metal  surfaces is
             shown at 90 percent relative humidity.   Adapted from Sydberger
             and Vannenberg (1972).
                                   7-23

-------
could act as centers for corrosion initiation--S02  related  acidic
deposition and surface corrosion of aluminum does not  appear  to be  a
significant problem (Fink et al. 1971).

7.3.1.2.2  Copper.  Copper and copper alloys in  most atmospheres develop
thin, stable surface films, which inhibit further corrosion.   Initial
atmospheric corrosion is a brown tarnish of mostly  copper oxides and
sulfides that can thicken to a black film.   Then in a  few years, the
familiar green patina forms.  Analysis of this film indicates it to be
either basic copper sulfate or,  in marine atmospheres, basic  copper
chloride.  However, in coastal urban areas,  the  sulfate may still
predominate (e.g., the Statue of Liberty)  because of the continuous
availability of S0£ over many years.   Nevertheless, both the  sul fate
and chloride-based patinas are generally resistant  to  further attack
(Yocom and Upham 1977).

7.3.1.2.3  Zinc.  Zinc is used primarily for galvanizing steel to make
it resistant to corrosion in the atmosphere  and  as  an  alloying metal
with copper to produce brass.  Zinc as a coating on steel is  anodic with
respect to steel such that when  zinc  and steel are  in  contact with
eletrolyte, the current flow protects the steel  from corrosion at the
expense of some oxidation of zinc.

     Because of its economic importance, the behavior  of zinc in the
presence of acidic deposition has been studied intensively  by a number
of workers.  Guttman (1968) carried out long-term measurements of
atmospheric corrosion of zinc from which he  developed  an empirical
damage function for zinc corrosion in relation to SOg  concentrations
and time-of-wetness.  Time-of-wetness was measured  by  means of a dew
detector.  S0j> was measured by lead peroxide sulfation candles and
conductions trie S02 measurements.   His empirical damage function is

     Y = 0.005461(A)0.8152x (B + 0.02889),                       [7_H]

where

     Y = corrosion loss, mg for a 3 x 5  in.  panel,
     A = time of wetness, hr,
     B = atmospheric S02 content during the  periods that the  panels
         were wet, ppm.

     Haynie and Upham (1970) carried  out an  extensive  zinc  corrosion
study in eight cities where zinc panels were exposed, while concurrently
collecting data on SO?,  temperature,  and humidity.  They developed  the
following empirical damage function relating zinc corrosion to S02
levels and relative humidity:

     y = 0.001028 (RH -  48.8) S02,                             [7-12]
                                  7-24

-------
where

      y = corrosion rate,  m yr~l,
      RH = average annual  relative  humidity.
     S02 = average S02  concentration,  yg m~3.

     Note that in Equation 7-11 moisture is in terms of time of wetness
while in Equation 7-12  annual  average  relative humidity is used. Time of
wetness is a far more relevant indication of  surface moisture than
average relative humidity  when corrosion and  other forms of
moisture-enhanced material damage are  being considered.  For example, if
Equation 7-12 is applied in an area that has  annual average relative
humidity significantly  less than 48.8  percent, no corrosion is implied.
Yet in such areas, surfaces become  wet with dew or seasons of high
humidity occur and corrosion proceeds  even when annual average relative
humidity is below the critical  value obtained by regression analysis.

     The damage coefficients for these two functions plus functions
developed from other studies were compiled by EPA (1981).  These
coefficients are compared  in Table  7-5.

7.3.2  Masonry

     The term "masonry" is applied  to  a large number of building and
decorative materials exhibiting a broad range of surface activities to
physical and chemical  stresses imposed by the environment.  The
importance of acidic deposition to  this class of materials may be
related to the effect produced directly on a  single material (e.g.,
limestone or marble) or direct or indirect damage to composite masonry
systems.  An example of direct damage  to composite systems involves the
rusting of steel reinforcing bars in concrete, which expand and crack
the concrete.  Indirect damage includes damage to brick-mortar systems
in which the relatively reactive mortar is damaged directly by acidic
materials and rainfall  and then, the salts released by these reactions
diffuse into the brick, causing stress and subsequent spalling.

7.3.2.1   Stone--The accelerated decay of stone buildings and monuments
in highly industrialized areas has  been documented by comparison of
current condition with  historic photographs and plaster casts.
Photographs taken in 1908  and 1969  of  a sandstone sculpture carved in
1702 in Westphalia, West Germany demonstrate  a dramatic loss of material
in the past 60 years with  virtual obliteration of the object (Winkler
1982).  Similarly, comparison of a  plaster cast made in 1802 with a
photograph taken in 1938 demonstrates  substantial deterioration of a
sculpture from the west frieze of the  Parthenon (Plenderleith and Werner
1971).  A detailed account of the restorations on the Acropolis and
measurements of the thickness of gypsum layers formed on its exposed
marble surfaces is presented by Skoulikidis (1982).  The deteriorating
condition  of the Caryatids of the  Erechthion led to their replacement
with replicas and their removal to  the controlled environment of the
Acropolis Museum (Yocom 1979).
                                  7-25

-------
      TABLE 7-5.  EXPERIMENTAL REGRESSION COEFFICIENTS WITH ESTIMATED
          STANDARD DEVIATIONS FOR SMALL ZINC AND GALVANIZED STEEL
               SPECIMENS OBTAINED FROM SIX EXPOSURE SITES
           Study
Time of wetness
 coefficient     S02 coefficient3
  (ym yr"1)      (ym yr'Vyg m~l
                 Number
                   of
                  data
                  sets
Field Studies

CAMP (Haynie and Upham
  1970)

ISP (Cavender et al.
1971)

Guttman 1968

Guttman and Sereda 1968

St. Louis (Mansfield
1980)
  1.15 +_ 0.60


  1.05 +_ 0.96


  1.79

  2.47 4-0.86

  2.36 + 0.13
0.081 +_ 0.005       37


0.073^0.007      173


0.024            < 400

0.037 +_ 0.008      136

0.022 + 0.004      153
Chamber Study

Haynie et al. 1976
  1.53 + 0.39
0.018 + 0.002
96
al ppm S02 = 2620 yg m'3 S02.
                                  7-26

-------
      Few quantative  studies of air pollution damage to stone have been
 reported although the Increased rate of erosion for marble tombstones in
 the  urban environment of Edinburgh was observed as early as 1880 (Geike
 1880).  A study of tombstones in U.S. National  Cemeteries (Baer and
 Berman 1981) has developed methodology for measuring damage to marble
 headstones exposed to the environment for 1 to 100 years.  Their data
 base consists of measurements of some 3,900 stones in 21 cemeteries
 distributed throughout the United States.  The factors affecting damage
 rates include grain  size, total precipitation,  and local air quality.

      In the United States, measured rates of marble deterioration have
 generally been small, on the order of 2.0 mm per 100 years (Winkler
 1982).  This is substantially below values reported for stones exposed
 in urban areas in Europe although direct comparison is difficult because
 the  stones exposed in Europe are generally more reactive.

      Comparing the condition of similar samples of sandstone exposed  in
 different areas of Germany for about 100 years,  Luckat associated large
 differences in observed deterioration with trends in local  air quality
 (Luckat 1981, Schreiber 1982).  These results presented in  Table 7-6
 describe stones openly exposed to the environment.   For similarly
 reactive test stone specimens protected from the direct action of rain
 and  placed at 20 locations in West Germany, the following functions
 correlating reaction with S02 immission rate were  obtained:

      Baumberg sandstone             U = 0.54 D;  r2  = 0.92        [7-13]

     Krehnsheim shell limestone     U = 0.22 D;  r2  = 0.72.        [7-14]

 When  similar test samples were exposed to the rain  the following damage
 functions were obtained:

     Baumberg sandstone        L =  0.03 D + 0.5;  r2  =  0.36        [7-15]

     Krensheim shell  limestone L =  0.018 D + 0.6;  r2 = 0.80      [7-16]

where:

     U = SOo immission rate (uptake rate)  of the stone in (mg
         m~2 d~l)  by weight gain,

     D = by  weight gain  S0£ immission rate,  IRMA  measured value  (mg
         itr2 d-1),

     L = loss in  weight,  and

     r = correlation coefficient.

     The high contribution of the non-SOx  factors  for  stones exposed
to rainfall  suggests that damage functions for  stone must specifically
address such variables as other pollutants and  rainfall.
                                 7-27

-------
     TABLE 7-6.   DETERIORATION OF SCHLAITDORF SANDSTONE EXPOSED FOR
            100  YEARS  IN WEST GERMANY (AFTER SCHREIBER 1982)
     Monument
Location
   Relative S02
immission Rate,3
    mg m~2-day
Deterioration
Neuschwan stein
Castle
Ulm Cathedral
Cologne Cathedral
Fussen
Ulm
Cologne
6
48
111
Practically
Moderate
Very severe
none


Relative immission  or uptake  rate of $03, annual average (August
 1973 - July 1974) measured by IRMA method.  (See Baer et al. 1983 for
 details of the technique.)
                                  7-28

-------
      A series of measurements made at St. Paul's Cathedral, London on
 the Portland stone  (biosparite limestone) balustrade, demonstrate a high
 rate of weathering  (Sharp, et al. 1982).  Using lead plugs filled in
 1718 in openings in  the  stone as base level references, a mean rate of
 lowering of 0.078 mm yr-1 was obtained for the period 1718-1980.   The
 balustrades represent conditions of exposed rain flow.  Similarly, by
 use of a micro-erosion meter (dial micrometer gauge mounted on reference
 studs)  a current erosion rate of 0.139 mm yr'1 was obtained for six
 sites on the cathedral.  These sites represented drip erosion zones.
 Though the two sets  of data are not strictly comparable, both represent
 substantially higher rates of loss than observed for marble in the
 United States.

 7.3.2.2  Ceramics and 61ass--Although enamels and glasses are quite
 resistant to chemical attack by air pollutants, in certain circumstances
 damage has been observed.  In a three year exposure study on porcelain
 enamels placed in seven  U.S. cities, some change in surface condition of
 the enamel  was observed  although the base metal was protected (Moore and
 Potter 1962).

      Flourides, especially HF, are capable of attack on a wide variety
 of ceramic materials and glasses.  Restrictive legislation on flouride
 emissions has, for the most part, eliminated fluoride induced damage.

      Perhaps the most serious glass damage problems is that associated
 with the decay of medieval stained glass windows.  The unique
 composition of these glasses combined with their open exposure to the
 atmosphere makes them particularly susceptible to deterioration.   This
 problem is discussed in  detail below.

      Properly fired  brick is highly resistant to attack by air
 pollutants while poorly  fired brick is highly susceptible to chemical
 attack.  Acidic solutions accelerate such damage, increasing the  rate of
 reaction 10-fold over water alone.  Residual  sulfates from decay  of
 mortars can combine  with other salts to produce failure in brick
 (Robinson 1982).

 7.3.2.3  Concrete—World production of concrete amounts to some 3
 billion cubic meters per year.  Many important structures, e.g.,  bridge
 decjJk  highways,  military installations, and naval  shore structures
 sufWr from severe durability problems (NMAB 1980).  Similarly, concern
 has been expressed over  leaching of possibly toxic components of  cement
 culverts transporting acidified water.

      The highly alkaline nature of cement/concrete leaves such surfaces
 vulnerable to acidic deposition.  Spedding (1969b), reporting on  the
 contamination/decontamination of laboratory surfaces accidentally
 exposed to sulfur-35/sulfur dioxide, observed that good decontamination
 was obtained by simple water washing.  This suggests that the reaction
 products of the deposition of S02 on concrete are water soluble.   The
 high volume of water flow through rain collecting and distribution


                                  7-29
409-262 0-83-22

-------
culverts in drinking water systems  raises questions about the  possible
release of toxic materials leached  from  the concrete matrix.

     Similar concerns have been  expressed over the errosive effects of
acidified streams on concrete bridge  piers.  The literature reveals only
limited research on the effects  of  acidified water runoff on concrete
durability.

     Cements used in dams and culverts require a special formulation for
sulfate resistance when exposed  to  concentrations in excess of 200 ppm
in water (Nriagu 1978).

     Specialized concretes in which sulfur replaces cement as  the
binding agent have been developed by  the Bureau of Mines for resistance
to acid and salt attack and damage  to freeze-thaw cycling (Sulphur
Institute 1979).

7.3.3  Paint

     Paint consists of pigment and  vehicle.  Pigments,  such as titanium
dioxide and zinc oxide, provide  color, hiding power, and durability.
Sometimes fillers such as calcium carbonate or inorganic silicates are
also added.  The vehicle provides the film-forming properties  of the
paint and contains resin binders, solvents, and additives.  Together,
the pigment (along with fillers) and  vehicle protect the underlying
surface and enhance the appearance  of the exposed surface.  Air
pollution may limit both of these functions by damaging the protective
coating, thus exposing the underlying surface to attack and/or spoiling
the appearance of the surface.  The most important potential effects of
S02 on paints are interference with the  drying process  and
acceleration of the normal erosion  process.

     The primary effect of particulate matter on paint  is soiling.
Soluble salts such as iron sulfate  contained in deposited particles can
also produce staining.  Chemically  active large particles such as acid
smut (or soot) from oil-fired boilers, mortar dust near building
demolition sites, or iron particles from grinding operations can
severely damage automotive paint (Yocom  and Upham 1977).  The  effects
range from discoloration of the  paint film to corrosion of the
underlying metal in the vicinity of individual particles.  Large
particles becoming imbedded in a freshly painted surface can act as
wicks to transfer moisture and corrosive pollutants such as S02 to the
underlying material's surface.

     Hoi brow (1962) has reported a  number of experiments to determine
effects of sulfur dioxide on newly  applied paints.  Drying time for
various oil-based paints exposed to extremely high concentrations of
S02 (1 to 2 ppm) was increased 50 to  100 percent.  Thus far no
experiments have been carried out on  the effect of S02  on drying time
of water-based latex paints.
                                  7-30

-------
     Campbell  et al.  (1974)  carried out an  extensive study of paint
erosion for a  variety of paint  types and exposure conditions (including
S02 and 03).  Both chamber and  field experiments were conducted.
The researchers evaluated four  important types of paint:

     1.  Acrylic latex and oil-based house  paints,

     2.  Urea-alkyd coil  coating  for sheet  metal in coil form,

     3.  Nitrocellulose-acrylic automotive  refinishing paint, and

     4.  Alkyd industrial  maintenance coating.


Table 7-7 presents the principal  findings of  this work.

     Generally, exposures to high concentrations (1 ppm of both $63
and ozone) produced statistically significant erosion rate increases
compared to clean air (zero pollution)  conditions.  Oil-based house
paint experienced the largest erosion rate  increases.  The greater
susceptibility of oil-based house paint to  S02 was attributed to the
use of extenders such as CaC03  or metal  silicates.  Latex and coil
coatings experienced moderate increases, and  the industrial maintenance
coating and automotive refinish experienced the smallest increases.  In
general, exposures to sulfur dioxide produced higher erosion rates than
ozone.  Unshaded panels eroded  more than shaded panels.  Exposures to
0.1 ppm pollutants did not produce significant erosion rate increases
over clean air exposures.   It should be noted that even these lower
concentrations are high when compared with  average concentrations found
in the ambient air of urban areas.

     In the field portion of this same  study, painted panels were
exposed at four locations with  different environments:

     1.  Rural - clean air (Leeds, North Dakota),

     2.  Suburban (Valparaiso,  Indiana),

     3.  Urban - sulfur dioxide-dominant (Chicago, Illinois), and

     4.  Urban - oxidant-dominant (Los  Angeles, California).

In most cases, southern exposures produced  somewhat larger erosion
rates, which agreed with the unshaded versus  shaded results of the
laboratory study.  Oil-based house paint again experienced by far the
largest erosion rate increases, followed in order by the urea-alkyd coil
coating, latex house paint, industrial  maintenance paint, and automotive
refinish.  Generally, the field exposures showed that the relative paint
erosion rate was about the same for the sulfur dioxide-dominant as for
the oxidant-dominant location which appeared  to contradict the chamber
studies.  However, the authors  believed that  differences in the
pollutant mix  at the two locations and  especially the presence of


                                  7-31

-------
                                            TABLE 7-7.
             PAINT EROSION  RATES AND PROBABILITY DATA (T-TEST) FOR CONTROLLED ENVIRONMENTAL
                        LABORATORY EXPOSURES (ADAPTED FROM  CAMPBELL ET AL. 1974)
Type of Paint
House paint
oil
latex
Coil coating
Automotive refinish
Industrial maintenance
Mean Erosion Rate (nm hr~* with 95
confidence limits) for unshaded
Clean air S02
control (1.0 ppn)
5.11 +_ 1.8 35.8 +_4.83a
0.89 + 0.38 2. 82 +_ 0.253
3.01 _+ 0.58 8.66 +_ 1.193
0.46 +_ 0.02 0.79+^0.66
4.72 +_ 1.30 5. 69 +_ 1.78
percent
panels
03
(1.0 ppn)
11.35 +_ 2.67a
2.16 jf 1.50°
3.78 +_ 0.64b
1.30 +_ 0.333
7.14 jv 3.56
                              PAINT  EROSION RATES AND PROBABILITY DATA (T-TEST)
                           FOR FIELD EXPOSURES (ADAPTED FROM CAMPBELL  ET AL. 1974)
Mean Erosion Rate (nm hr'l with 95 percent confidence
limits) for panels facing south
Type of Paint
House paint:
oil
latex
Coil coating
Automotive refinish
Industrial maintenance
Rural
(clean air)

109
46
53
23
91

+
+
+
+
+

191
13
20
28
41
Suburban

376+^
76 +_
254 +_
58 +_
208 +_

1243
183
483
18b
361b
Urban
(S02 dominant),
- 60 yg m"J

361
97
241
41
168

+_ 124b
±8»
+_ 203
± 10
+_ 99
Urban
(oxidant dominant),
- 40 wi m"3

533 +_
165 jf
223 +_
43 +_
198 +_

157a
142
433
10
613
Significantly different from control  at  p = 0.01.

^Significantly different from control  at  p = 0.05.

Note:   1  ppn S02 = 2620 yg m'3
                                                7-32

-------
nitrogen oxides at the oxidant-dominant site could have enhanced the
erosion rate at this location,  bringing it up to the level of damage at
the sulfur dioxide-dominated location  (Campbell et al. 1974).

     It is noteworthy that the  oil-based house  paint and urea-alkyd coil
coating experienced the largest erosion rate increases in both the field
and laboratory sulfur dioxide exposures.  These coatings were the only
ones that contained a calcium carbonate extender—a substance sensitive
to attack by acidic materials.

     Spence et al. (1975)  summarized the results of paint exposure to
several gaseous pollutants from the  full-scale chamber studies reported
by Haynie et al. (1976) and discussed  earlier in relation to metal
exposures.  Four classes of painted  surfaces were evaluated:  oil-based
house paint, vinyl-acrylic latex house paint, vinyl coil coating, and
acrylic coil coating.  A strong correlation was found between paint
erosion for the oil-based  house paint  and S02 and humidity.  The vinyl
and acrylic coil coating were unaffected, but blistering was noted on
the latex house paint.   It was  not certain if the blistering was the
result primarily of S02 or moisture.

     A multiple regression relationship was developed for the joint
influence of S02 and relative humidity on the oil-based house paint:

     E = 14.3 + 0.0151 S02 + 0.388 RH                           [7-17)

Where

       E = erosion rate of pm yr-1,
     S02 = concentration of S02 in v9  n>~3»
      RH = means annual relative humidity in percent.

     This relationship indicates that  paint erosion is significantly
more sensitive to changes  in humidity  than to S02-  However, one must
be careful in using models based on  accelerated chamber tests for actual
exposures because Equation 7-17 would  predict that in an atmosphere with
no S02 present, with an average relative humidity of 50 percent, the
paint erosion rate would be about 34 ym yr'1.   Assuming a typical
paint thickness of 50 ym,  the paint  film would  be completely eroded
away within 1.5 years.

     The present understanding  of damage to paint  from air  pollution  is
based  primarily upon two sets of chamber studies and one  set of  field
exposures.  Since the field studies  were carried out in the early 197'0s,
further laboratory and field studies are needed to determine the
importance of paint damage from present levels  of  sulfur oxides.
Furthermore, these studies should include present  formulations
(especially water-based paints) that may have a different response to
air pollutants than those  used  earlier.
                                  7-33

-------
7.3.4  Other Materials

     In addition to coatings,  a wide range  of organic materials  are
found to be susceptible to attack  by atmospheric  pollutants.  These
materials, including paper,  photographic  materials,  textiles and leather
were not considered in the EPA's Criteria Assessment Documents,  so they
are considered here although the indoor locations in which  they  are
normally found dictate gaseous transport  mechanisms  for  deposition.
         Paper—The role of sulfur dioxide  in  the  deterioration of  paper
         accepted since the 1930's.   Early  experiments  (Langwell  1952,
7.3.4.1
has been
1953) relied on unrealistically  high  S02  concentrations of 5,000  ppm
interacting with damp paper.   Hudson  and  Milner  (1961) used  sulfur-35 as
a radioactive tracer to demonstrate that  measureable  amounts of S02
were rapidly deposited in paper.   Working with concentrations of  10 ppn,
Grant (1963) showed that S02  deposition increased with increasing
aluminum sulfate/resin sizing of the  paper.

     A comparative study of identical  copies  of  twenty-five  17th  and
18th century books in two English libraries,  one in an unpolluted
atmosphere in Chatsworth, the other in the badly polluted urban
atmosphere of Manchester, revealed a  significant increase in paper
acidity in the Manchester library (Hudson 1967). This acidity  was
greatest at the page edges and decreased  greatly toward the  center of
the page, which might be considered the initial  sheet acidity.

     Wallpapers form an important part of the indoor  surface area
available for S02 sorption.  Spedding and Rowlands  (1970) measured the
sorption characteristics of PVC  and conventional wallpapers  on  exposure
to maximum initial S02 concentrations of  150  yg  nr3.  Sorption
depended largely on surface finish and design pattern, with  greater
sorption by conventional wallpapers.   The researchers suggested that
S02 sorption accelerated the deterioration of wallpaper.

7.3.4.2  Photographic Materials--Under normal conditions of  temperature
and relative humidity, paper, acetate film, and  other photographic
materials are oxidized at a very slow rate.   One of the most serious
factors in the preservation of photographic materials is the presence of
large quantities of oxidizing gases:   hydrogen  sul fide,  sulfur  dioxide,
and to a lesser extent NOX, peroxides, formaldehyde,  and ozone
(Eastman Kodak 1979).

     The effect of these pollutants is usually yellowing and fading of
the silver of the image.  The paper base  may  also be  degraded and
stained.  Acidic gases will degrade gelatin,  paper, and  the  film  base of
negatives (Eastman Kodak 1979).

     Agfa produces a colloidal silver test strip which is 8  to  10 times
more sensitive to gaseous pollutants  than ordinary  photographic
materials.  In a survey of major libraries and  archives  using this
technique many examples of significant air quality  problems  were
observed (Weyde 1972).
                                  7-34

-------
7.3.4.3  Textiles and Textile Dyes—Sulfur oxides are capable of causing
deterioration to natural  and synthetic fibers.  Cotton, like paper,  a
cellulosic fiber, is weakened by  sulfur dioxide.  Under circumstances
where sulfuric acid comes in contact with a cellulosic surface,  the
product of reaction is water soluble with little tensile strength
(Petrie 1948).  In field  tests  in St. Louis, cotton duck exposed to
varying SOX levels showed a direct relationship between loss in
tensile strength and increasing SOX concentration (Brysson et al.
1967).  Zeronian (1970) exposed cotton and rayon fabrics under
accelerated aging conditions of light and water spray with and without
0.1 pprc S02.   Loss in strength  was 13 percent in the absence of  S02
and 22 percent in the presence  of S02-  In a study of nylon fabrics
exposed to 0.2 ppm Sfy under similar conditions, he found that nylon
fabrics lost 40 percent of their  strength under the S02 free
conditions and 80 percent of their strength in the presence of S02
(Zeronian et al. 1971).

     The degradation of nylon 66 by exposure to light and air is
increased by the addition of 0.2  ppm of S02 to the air.  Chemical
properties, and yarn tensile properties both reflect this damage
(Zeronian et al. 1973).   Results demonstrated that the mode of
degradation is not changed although S02 accelerates the rate of
reaction.

     Among proteinecous textiles, silk is most vulnerable to the effects
of light, acidity, and sulfur dioxide, demonstrating much greater loss
in strength than wool (Leene et al. 1975).

     Damage to textiles has been  attributed to NOX (Harrison 1975).
Such damage has been caused both  by loss of fiber strength and fading of
textile dyes.  Significant reduction in breaking strength and increase
in cellulose fluidity were observed for combed cotton yarns exposed  in
Berkeley, California, to  unfiltered air when compared to exposure to
carbon filtered air (Morris et  al. 1964).  Both sets of samples  were
unshaded and exposed at a 45° angle facing south.  Though the authors
did not isolate the effects of  individual pollutants, they implied that
compounds associated with photochemical smog, especially NOX, were the
probable cause of increased damage.

     In an EPA chamber study of the effects of individual pollutants on
20 dyed fabrics, it was demonstrated that N02 at 0.1 to 1.0 ppm
produced appreciable dye  fading, and S02 at 0.1 to 1.0 ppm caused
visible fading on wool fabrics  (Beloin 1973).  It was also concluded
that higher temperatures  and relative humidities increase dye fading and
that the rate of fading as a function of exposure time appears to be
nonlinear.

7.3.4.4  Leather—Michael  Farady  is attributed (Parker 1955) with having
established in 1843 a link between the rotting of leather armchairs  in
the London Atheneum Club  and sulfur dioxide emitted by its gas
illumination.  Plenderleith (1946), Innes (1948), and Smith (1964)
                                  7-35

-------
describe the sequence of chemical deterioration for leather and consider
possible mitigative actions.

     It has been observed that leather initially free of sulfuric acid
will accumulate up to one percent acid by weight per year if exposure to
an atmosphere containing S02«  The mechanism is thought to involve the
metal ion catalyzed conversion to sulfuric acid of the 863 absorbed by
the collagen of the leather.  Using sulfur-35 labelled SOg. Spedding
et al. (1971) showed that it  is  sorbed evenly over the leather surface,
with the limiting factor in uptake being gas-phase diffusion to the
surface.

7.3.5  Cultural  Property

     It has been estimated that  the United States has over 6,000
museums, historical  societies, and related institutions; more than
10,000 entries on the National Register of Historic Places, and in
excess of 26,000 libraries and archives of substantial size (NCAC 1976).
Light, oxidation, fluctuations in humidity, and chemical pollutants
threaten this precious cultural  heritage.

     Damage to cultural  property cannot be quantified in simple-dose-
response terms.   Just as an electrical component may require replacement
due to corrosion of a fraction of its mass or stress-corrosion fracture
may lead to failure of a mechanical system, damage to the texture of
sculpture or the surface of a fresco exposed to the environment
diminishes their aesthetic importance far in excess of the amount of
material damage.  Still  more critical is the circumstance that, for most
cultural property, replacement is impossible.  What is lost is lost.

7.3.5.1  Architectural  Monuments—Hi storic and artistic structures
represent the single most visible aspect of our history and culture.
For the United States,  legislation providing a mandate for preservation
began with the Antiquities Act of 1906, followed most recently by the
Historic Preservation Act Amendments of 1980.  In Canada, the
Archaeological Sites Protection  Act and the Historic Sites and Monuments
Act were adopted in 1953.

     Architectural monuments  are universally threatened by the effects
of pollution and urbanization as well as by weathering cycles and other
natural phenomena (CCMS 1979).   Although damage to these monuments is
frequently attributed to acid precipitation, no clear evidence providing
a cause and effect relationship  between acid precipitation and damage to
a specific monument exists.   In  general, it appears that while acidic
deposition can effect significant damage to cultural property, the
sources are predominantly of local origin.

7.3.5.2  Museums, Libraries and  Archives—As discussed above, the
sorption of SOX and NOX by organic materials in the indoor
environment is well  established.  In some cases, as in paper and leather
embrittlement, dye fading, and "red-ox" blemishes on microfilm, a direct
                                  7-36

-------
relationship between pollutant sorption and damage has been established.
This has led major museums,  libraries, and archives to install scrubbers
for the removal  of acid gases.

     Among the systems in use are  activated charcoal and Purafil
(activated alumina impregnated with KMnCty) dry scrubbers and alkaline
wash wet scrubbers.  Such systems  have been introduced as part of new
construction or retrofitted  at the National Gallery (London), the
Library of Congress (Washington, D.C.), the Newbury Library (Chicago),
and the National  Gallery (Washington, D.C.).  Many other collections of
cultural artifacts are preparing for the eventual retrofitting of their
air handling systems to use  scrubbers for removing air pollutants.

     The universal nature of concern for the effects of polluted air on
cultural property is reflected in  a Japanese study of ambient and indoor
SOX and NOX concentrations for buildings where important screen and
panel paintings are housed (Kadokura and Emoto 1974).  Six sites in
Kyoto were investigated.   Average  concentrations for SOx and NOx
were found to be about one-third of those in Tokyo.  Seasonal
concentrations for SOX peaked in winter and were highest for a site
near a dyeing factory whose  liquid wastes emitted $03.  The NOX
concentrations were found to be more evenly distributed throughout the
city.  Tight buildings showed higher NOX levels indoors than were
found for ambient conditions.   Although they did not cite specific
examples of damage, the authors called for protective measures to
prevent air pollution damage to paintings.

7.3,5.3  Medieval Stained Glass--Some evidence exists that medieval
stained glass exposed to the atmosphere has deteriorated more rapidly
since World War II than in previous centuries.  This accelerated
deterioration has been attributed  to the effects of air pollution
(Frenzel 1971, Froedel-Kraft 1971, Korn 1971) because gypsum and
syngenite (CaS04-K2S04-H20)  are found in the weathering
crust.  However,  such crusts are found even in locations with low S02
concentrations,  suggesting that background SOg levels are sufficient
to produce the sulfates observed.  An alternative mechanism of decay
suggests that storage of the windows under damp conditions during the
war permitted the formation  of a fissured hydrated layer that led to
enhanced corrosion after reinstallation of the windows.  The sulfates
found in the weathering crusts are thought to be by-products of the
deterioration process (Newton 1973).

     A broad range of preservation techniques has been employed,
including lamination, coating with inorganic and organic materials, and
"isothermal glazing."  In the latter process, the ancient glass is moved
just inside the building and modern glass is placed in the grooves in
the stone.

7.3.5.4  Conservation and Mitigation Costs--Some indication of the
problem's magnitude is given by cost for mitigative actions taken for
cultural property in West Germany  (Table 7-8).  Similar cost estimates
exist for national preservation programs in the United Kingdom, Greece,
                                  7-37

-------
                        TABLE 7-8.   ESTIMATED COSTS ASSOCIATED WITH AIR POLLUTION DAMAGE TO
                             CULTURAL PROPERTY IN WEST GERMANY (AFTER SCHREIBER 1982)
co
CO
Location
Federal
Republic
of Germany
Objects
All municipal bronze
monuments and sculptures
All metal sculptures in
Measures
Desirable
cleaning
Desirable
Period
Annual
Annual
Costs DM
4,000,000
1,000,000
     Munster



     Cologne


     Cologne



     Freiburg


     Ulm
                        museums  and open air


                        All  medieval stained
                        glass

                        Artifacts  in museuns
Castle facade
Cathedral stained glass
windows

Cathedral facade
Cathedral  stained  glass
windows

Cathedral  stained  glass
windows
cleaning,
conservation

Desirable
conservation

Air condition-
ing with air
improvement

Cleaning,
restoration,
conservation

Conservation
Cleaning,
restoration,
conservation

Restoration,
conservation

Desirable
restoration
                                                  10 year cost
                                                  estimate

                                                  During
                                                  construction
                     1965-1973
                     1978
Annually
1977-1997
1978


Total cost
                                       200,000,000-
                                       300,000,000

                                       15% of construc-
                                       tion costs
                     1,000,000
                       448,000
                                          3,000,000-
                                         60,000,000
                                         (estimated)

                                            150,000
                                          3,000,000

-------
 France,  Italy,  and  the United States.  For example,  the Italian
 Parllment designated $200,000,000 In 1980 for a 5-year program to
 restore  and maintain the ancient monuments 1n Rome (Hofinann  1981)  and  It
 1s  estimated by a British Parliamentary Committee that restoration of
 the fabric of the Houses of Parliament will  cost up to £5,000,000
 (International  Herald Tribune 1980).

 7.4  ECONOMIC IMPLICATIONS

      The possibility of determining the economic costs of air
 pollution's damaging effects has long attracted environmental  policy
 makers.  If reliable cost estimates could be developed for such effects
 1n  relation to  the  pollutant levels that produced them,  It then might be
 possible to compare the costs for achieving  various  levels of  air
 quality control  through emission control  with the cost savings from
 reduced damage--a significant step toward developing cost-benefit
 relationships for air pollution control.   The many attempts  to estimate
 costs associated with air pollution-induced  material damage  have
 recently been summarized by Yocom and Stankunas (1980).   Without
 exception, all  of the generalized estimates  of material  damage costs
 related to all  types of air pollution existing at the  time of  this
 review are of questionable value.   The reasons for this  include the
 following:

    0   As was pointed out earlier,  it is usually  not possible  to isolate
       the specific portion of damage and therefore  the  associated costs
       created by a given air pollution effect.

    0   Improper assumptions and inaccurate estimates of  the  quantities
       of materials in  place and exposed  to  pollutants.

    «   Unrealistic or improper scenarios  of  use,  repair, and replacement
       of materials susceptible to air pollution  damage,  together with
       improper  or Inaccurate assignment  of  costs to the  scenarios.

    °   Incomplete knowledge of substitution  scenarios  where more
       expensive material  systems may replace more susceptible
       materials.

    •   Inadequate knowledge of the exposure  conditions of  susceptible
       materials,  for example,  coexistance of pollutants with  other
       environmental effects such  as  moisture and  temperature, and the
       physical  aspects of exposure such  as  orientation  and  degree of
       sheltering.

     A recent study by  Stankunas et al.  (1981)  has addressed many of the
above difficulties.   In  this study  the quantities of potentially
susceptible  materials were determined within  357  randomly  selected 100  x
100 foot square  areas covering the Boston metropolitan area.   Teams of
observers using  survey  techniques  determined  the  areas of various types
of exposed painted surface,  bare metal  of several  types, brick, stone,
and concrete, and  several  other  types of  surfaces.   Of the 357 areas
                                  7-39

-------
selected,  183 were found to  contain  manmade  structures.  The total areas
of each material  found  at the  survey sites were extrapolated to the
entire Boston metropolitan area.   Then, using air quality records for
S02 in the Boston area, together with humidity data and published air
pollution damage functions for given materials, the researchers computed
the total  damage to a given  material  for  the entire area.  In the case
of painted surfaces,  assumptions were made on the average thickness of
typical paint films.  Then costs were assigned to the increase in
painting frequency, based on the S02-related increase in paint
erosion, to arrive at a total  S02-related damage cost to paint in the
Boston metropolitan area. The excess painting costs for the Boston
metropolitan area attributable to  S02 damage for the year 1978 were
estimated to be $31.3 million. This is equivalent to a per capita cost
between $11 and $12.  Costs  for damage to zinc coated materials were two
orders of magnitude lower.

     Haynie (1982) estimated costs for damage to zinc-coated
transmission towers and to galvanized roofing, siding, and guttering.
Different approaches were used for transmission towers than for the
other materials.  Costs for  transmission  tower damage were based on a
single group of towers  serving the Colbert Steam Plant in the TVA
system. Measurements were made by  TVA of  the thickness of the zinc
coating at several points on 19 towers likely to be affected by S02
from the plant in question.   Using S02/moisture damage functions for
zinc corrosion and an estimate of  how height above ground would affect
S02 deposition velocity (based primarily  on  changes in wind speed with
height), estimates were made of change in zinc thickness with time for
the group of towers.  Then,  using  several scenarios of painting, repair,
and replacement, researchers estimated annualized costs for mitigating
the effects of the damage, based on  local S02 and humidity levels.
Since TVA owned the towers,  such costs could be internalized and were
estimated to be 0.0028  mills/Kwh +_ 0.0011 to be added to customers'
electric bills.  These  estimates were based  on an S02 concentration of
17 yg m~3.  if $02 levels were allowed to reach the ambient air
quality standard of 80  yg m~3, the annualized extra maintenance cost
would rise to an estimated 0.0132  mills/Kwh  +_ 0.0052.

     Cost estimates for damage to  galvanized roofing, siding, and
guttering required estimating the  relative quantity of these materials
in place.  One of the complicating factors in making this determination
was the trend in recent years of  replacing bare galvanized materials
exposed to the outdoor  atmosphere  with coil  coated galvanized steel or
bare aluminum.  Various models were used  to  convert data on shipments of
the materials in question and anticipated use of alternate materials to
a realistic picture of the amount  of bare galvanized materials in these
categories in 1980.  Damage  functions for the effects of S02 and
moisture on zinc, together with estimates for the thickness of zinc
coatings and various maintenance  scenarios and their costs were used to
estimate per capita costs.  These  costs were computed to be in the range
of $0.60 to $1.50 with  the best estimate  being $1.05 at an annual
average S02 concentration of 30 yg nr3.   At  the  primary standard
of 80  yg nr3, the best  estimate of per capita costs would be $1.80.
                                  7-40

-------
     Such approaches as these should  be  refined and extended so that
realistic estimates may be made  of the total costs of damage from acidic
deposition.

7.5  MITIGATIVE MEASURES

     Assuming that acidic deposition  produces  significant damage to
materials, the primary mitigative  measure  is to reduce the
concentrations of acidic deposition components in the ambient
atmosphere.  However, as has been  implied  in the foregoing discussion,
the relative amount of damage and  associated costs from acidic
deposition is not known.  Therefore,  at  present there is no basis for a
given degree of control of acidic  deposition components and precursors
that will eliminate or bring to  acceptable levels the potential damage
from this source.

     Given that there is some degree  of  damage to materials from acidic
deposition, a wide range of mitigative actions may be taken in response
to damage.  Table 7-1 listed several  of  these  in relation to various
material categories.  The particular  mitigative measure and whether it
will be implemented will depend  on many  factors, including

   0   Physical and chemical nature of the material,

   0   Age and state of repair of  the materials system,

   0   Availability and cost of  substitute materials,

   0   Feasibility of isolating  the object or  surface of concern from
       the ambient environment,

   o   The importance of aesthetics in the appearance of the materials,

   o   The impact of damage on structural  integrity, and

   °   The attitudes of those responsible  for  the objects made of the
       materials in question regarding the relative importance of the
       damage.

      As stated earlier, material  damage from  acidic deposition is
generally indistinguishable from damage  caused by the natural
environment.  However, chemical  analysis of corrosion or damage products
can often distinguish various damage  mechanisms.  In general,
superimposing acidic deposition  on these natural phenomena only tends to
shorten the time before some mitigative  measure must be considered.  It
does not change the mitigative actions themselves.  Thus mitigative
measures taken to protect, replace, repair, and maintain materials
exposed to the ambient environment will  generally not change whether any
acidic deposition has an effect.   Only the frequency of implementing
these measures will  change.
                                 7-41

-------
7.6  CONCLUSIONS

     From a review of the  available literature on the effects of acidic
deposition on materials  the  following conclusions are drawn:

  °    Several  senarios  and  mechanisms exist for damage to materials
       from acidic deposition  including both long-range transport and
       local  source emissions  (Section 7.1).

  o    Without question  acidic deposition causes significant incremental
       damage to materials beyond  that caused by natural environmental
       phenomena (Section  7.1).

  o    Because very few  research efforts have attempted to isolate the
       effects of specific acidic  deposition scenarios, it is presently
       impossible to determine quantitatively if any one scenario is
       more important than another in causing material damage.  However,
       based on the juxtaposition  of primary acidic pollutant (e.g.,
       $02) sources and  large  quantities of susceptible material
       surfaces in urban areas, damage to materials from primary
       pollutants directly or  in oxidized form together with surface
       moisture (e.g., condensed dew) is believed to be more due to
       acidic deposition than  to acidified rain produced from long-range
       transport of pollutants and their reaction products (Section 7.2)

  o    Reliable cost estimates for material damage from acidic
       deposition are at present fragmentary because they deal with only
       selected material systems or linked geographical areas.
       Available estimates of  total material damage costs on a
       nationwide basis  are  unreliable.  There is a need for improved
       inventories of materials in place in various parts of the
       country (Sections 7.3 and 7.4).

  o    Damage to cultural  property from acidic deposition is a complex
       problem because of  the  high value placed upon such objects, their
       often irreplaceable nature, and the wide range of material types
       represented.  Highest priority should be placed on identifying
       and quantifying actual  and  potential damage to such artifacts and
       developing methods  to prevent damage (Section 7.3.5).

  0    Further research  directed at isolating damage caused by  specific
       acidic deposition processes and identifying those processes that
       are most important  and/or amenable to control is needed  (Section
       7.3).

  o    Studies that accurately assess damage costs associated with
       acidic deposition are needed  (Section 7.4).

  0    Further research  is needed  in the development of mitigative
       measures such as  reliable surface protection systems when damage
       has already been  observed and when protection cannot wait for
       improvement in air  quality  (Section 7.5).


                                  7-42

-------
7.7  REFERENCES

Baer, N. S. and S. M.  Berman.   1981.  Acid  rain material damage in
stone.  Final  Report to National  Atmospheric Deposition Program, North
Carolina State University,  Raleigh, NC.

Baer, N. S., G. M. Helms,  and  R.  A. Livingston.  1983.  The
Conservation/Restoration of Stone Monuments: A Progress Report on the
NATO/CCMS Pilot Study.  In  Proceedings of the Fourth International
Congress on Deterioration  and  Preservation  of Stone Objects.  In Press.

Barton, K.  1976.   Protection  Against Atmospheric Corrosion:  Theories
and Methods.  John Wiley and Sons, New York, NY.

Beloin, N. J.   1973.  Fading of dyed fabrics exposed to air pollutants.
Textile Chemist and Colorist.  5:128-133.

Browne, R. D.   1981.  Corrosion of steel  in reinforced concrete in
marine and other chloride  environments.   2nd International Conference on
the Durability of Building  Materials and  Components.  Preprints.  NBS,
Washington, D.C. pp. 210-224.

Brysson, R. J., B. J.  Trask, J. B. Upham, and S. G. Borras.  1967.
Effects of air pollution on exposed cotton  fabrics.  J. Air Pollution
Control Assoc. 17:294.

Campbell, G. G., G. G. Shurr,  D.  E. Slawikowski, andj. W. Spence.
1974.  Assessing air pollution damage to  coatings.  J. Paint Tech.
46:59-71.

Cavendar, J. H., W. M. Cox, M. Georgevich,  N. Huey, G. A. Jutze, and C.
E. Zimmer.  1971.    Interstate surveillance project:  measurement of air
pollution using static monitors.   APTD-0666, U.S. Environmental
Protection Agency, Air Pollution  Control  Office, Research Triangle Park,
NC.

CCMS.  1979.  Proposal for a CCMS Pilot Study on the
Conservation/Restoration of Monuments, pp.  341-345.  In Conservation of
Historic Stone Buildings and Monuments.   National AcaTJemy Press,
Washington, D.C. (1982).

Douglas, R. W. and J.  0. Isard.  1949.  The action of water and sulphur
dioxide on glass surfaces.   Soc.  Glass Tech. 33:289-335.

Eastman Kodak.  1979.   Preservation of Photographs, Kodak Publication
No. F-30.  pp. 25-26.

Fink, F. W., F. H. Buttner, and W. K. Boyd.  1971.  Technical-economic
evaluation of air-pollution corrosion costs on metals in the U.S.
APTD-0654.  U.S. Environmental Protection Agency, Research Triangle
Park, NC.
                                  7-43

-------
 Frenzel,  G.   1971.  Environmental dangers threaten medieval  stained
 glass.  Kirche and Kunst 49:58-60.

 Froedel-Kraft, E.  1971.  Restoration and conservation.   Osteraeichische
 Zeitschrift  fur Kunst und Denkmalpflege 25:70-73.

 Geike,  A.  1880.  Rock Weathering as Illustrated in Edinburg
 Churchyards.  Proc. of the Royal Society of Edinburg,  1879-1880.   pp.
 518-532.

 Grant,  R. L.  1963.  Some factors affecting the attack on paper by
 atmospheric  sulfur dioxide.   Ph.D. Thesis.   The Victoria  University of
 Manchester,  England.

 Guttman,  H.   1968.  Effects of atmospheric  factors on  the corrosion of
 rolled zinc,  pp. 223-239.  In Metal  Corrosion in the Atmosphere. ASTM
 STP 435,  American Society for Testing and Materials, Philadelphia.

 Guttman,  H.  and P. J.  Sereda.  1968.  Measurement of atmospheric  factors
 affecting the corrosion of metals,  Jh^ Metal  Corrosion in the
 Atmosphere.  ASTM STP  435, American Society for Testing and  Materials,
 Philadelphia.

 Haagenrud, S. E.  1978.  The effect of long-range transport  of  air
 pollutants on the atmosphere corrosion rate of carbon  steel  and zinc as
 measured  by weight loss.   Presented at 8th  Scandinavian Corrosion
 Conference, Helsinki,  Finland:  August.

 Haagenrud, S. E., R. W. Lanting, and G.  Santomauro.  1982.   Draft report
 on effects of sulphur  compounds on materials, including historical  and
 cultural monuments.   United  Nations Economic  and Social Council, Interim
 Executive Body for Convention on Long-Range Transboundary Air Pollution,
 Geneva, Switzerland.  August 30-September 3,  1982.

 Harker, A. B., F. B. Mansfield, D.  R.  Strauss,  and D.  D.  Landis.  1980.
 Mechanism of S02 and H2S04 Aerosol  Zinc  Corrosion.   EPA-600/3-80-018.

 Harrison, D.   1975.  Who  Pays for Clean  Air:   The Cost and Benefit
 Distribution of Federal  Automobile Emission Controls.  Ballinger
Publishing Co.,  Cambridge, MA.

Haynie, F. H.  1982.   Economic  assessment of  pollution related  to
corrosion damage.   In  Admospheric Corrosion.   William  Ail or, ed.  John
 Wiley and Sons,  NewTork,  NY.

Haynie, F. H.  1983.   Effects of small  particles  on  materials.  In  Small
Particles Monograph.   A.  P.  Altshuller,  ed.   To  be  published.

Haynie, F. H. and J. B. Upham.   1970.   Effects of atmospheric sulfur
dioxide on the corrosion  of  zinc.  Mater. Prot. Perform.  9(8):35-40.
                                 7-44

-------
Haynie, F. H.  and J.  B.  Upham.   1971.   Effects of atmospheric pollutants
on corrosion behavior of steels.   Mater. Prot. Perform. 10:18-21.

Haynie, F. H.  and J.  B.  Upham.   1974.   Correlation between corrosion
behavior of steel and atmospheric  pollution data, pp. 33-51.  In
Corrosion in Natural  Environments.  ASTM STP 558, American Society for
Testing and Materials.

Haynie, F. H., J. W.  Spence,  and J. B.  Upham.  1976.  Effects of gaseous
pollutants on  materials—a  chamber study.  EPA-600/3-76-015, U.S.
Environmental  Protection Agency, Research Triangle Park, NC.

Hofinann, P.  1981.  Rome's  Monuments Go Under Wraps.  New York Times,
February 15, 1981.

Holbrow, G. L. 1962.   Atmospheric  pollution:  Its measurement and some
effects on paint.  J. Oil Color Chem. Assoc. 45:701-718.

Hudson, F. L.   1967.   Acidity of 17th and 18th century books in two
libraries.  Paper Technology  8:189-190.

Hudson, F. L.  and W.  D.  Milner.  1961.  The permanence of paper.  Paper
Technology 2:155.

Innes, F. R.  1948.  Smokeless Air No.  68.  p. 23.

International  Herald Tribune.  1980.  September 2, 1980.  p. 5.

Johnson, J. B., P. Elliot,  M. A. Winterbottom, and G. C. Wood.  1977.
Short-term atmospheric corrosion of mild steel at two weather and
pollution monitored sites,  pp.  691-700. J_n_ Corrosion Science, Vol. 17.
Pergamon Press, New York, NY.

Judeikis, H. S.  1979.  Use of building surfaces in  the passive
abatement of gaseous air pollutants.  J. Architectural Res. 7(l):28-33.

Kadokura, T. and Y. Emoto.   1974.   The  effect of polluted air as a
factor of the enviornment of  screen and panel paintings.  Science for
Conservation 12:33-34.

Korn, U. D.  1971.  Causes  and symptoms of the deterioration of medieval
stained glass.  Deutsche Kunst and Denkmalpflege.  19:58-74.

LaBastille, A.  1981. Acid rain.   How  Great A Menace?  National
Geographic 160:652-680.   November  1981.

Langwell, W. H.  1952.  The permanence  of  papers, Part II.  Technical
Bulletin 29:21.

Langwell, W. H.  1953.  The permanence  of  papers, Part III.  Technical
Bulletin 30:2.
                                  7-45

-------
Leene, J. E., L.  Demeny,  R. J. Elema, A. J. de Graaf, and J. J. Surtel.
1975.  Artificial  aging of yarns in  presence as well as in absence of
light and under different atmospheric conditions: Condensed final
report.  ICOM Committee on Conservation.  4th Triennial Meeting.
Preprints.  75/10/2:1-11.

Livingston, R. A.  and N.  S. Baer.  1983.  Mechanisms of air
pollution-induced  damage  of stone.   6th World Congress on Air Quality.
Paris, France.  In Press.

Luckat, S.  1981.   Quantitative Investigations of the Influence of Air
Pollution on the Corrosion of Stone.  Staub-Reinholt Luft.
41:440-442.

Mansfeld, F. B.  1980. Regional air pollution study.  Effects of
airborne sulfur pollutants on materials.  EPA-600/4-80-007, U.S.
Environmental Protection  Agency, Research Triangle Park, NC.

Moore, D. G. and A. Potter.   1962.   National Bureau of Standards
Monograph 44:13.

Morris, M. A., M.  A. Young, and T. A-W. Molvig.  1964.  The effect of
air  pollutants on cotton.  Textile Research J. 34:563-564.

NCAC.  1976.  Conservation of cultural  property  in the United States.
National Conservation Advisory Council, Washington, D.C. pp. 3-4.

National Materials Advisory Board.   1980.  The status of cement and
concrete R&D in the United States.   NMAB Report  - 361.  p. 23.

New York Times.  1978a.   The  Temple  and the Obelisk.  Editorial November
17, 1978.

New York Times.  1978b.   Oh,  Egypt!   December 13, 1978.

Newton, R. G.  1973.  Bibliography of studies on the deterioration and
oconservation of stained  glass.  Art and Archaeology Technical Abstracts
10:131-178.

Nriagu, J. 0.  1978.  Deteriorative  effects of sulfur  pollution on
materials,  pp 1-59.  In  Sulfur  in the  Environment, Part II: Ecological
Impacts.  John Wiley anS Sons, Toronto.

Parker, A.  1955.   The Destructive Effects of Air Pollution on
Materials.  National Smoke Abatement Society, London, England.

Petrie, T.C.  1948.  Smokeless Air No. 67.   p. 62.

Plenderleith, H. J.  1946.  The Preservation of  Leather Bookbindings.
British Museum, London, England.
                                 7-46

-------
Plenderleith,  H.  J.  and A.  E.  A.  Werner.  1971.  The Conservation of
Antiquities and Works of Art.   Second Edition.  Oxford University Press,
London,  pp. 316-317.

Pourbaix, M.  1966.   Atlas  of Electrical Equilibria in Aqueous
Solutions.  Translated by J.A.  Franklin.  Pergamon Press, London.

Rice, D. W., R. J.  Cappell,  P.  B.  P. Phipps, and P. Peterson.  1982.
Indoor atmospheric  corrosion of copper,  silver, nickel, cobalt, and
iron, pp. 651-666.   In Atmospheric Corrosion.  W. H. Ailor, ed.
Wiley-Interscience  PuFlications,  New York.

Robinson, G. C.  1982.  Characterization of bricks and their resistance
to deterioration mechanisms.  In:   Conservation of Historic Stone
Buildings and Monuments. National  Academy Press, Washington, DC.  pp.
145-162.

Sanyal, B. and G. K. Singhania.   1956.   Atmospheric corrosion of metals:
Part I.  J. Sci.  Indus. Res. (India) 158:448-55.

Schreiber, H.   1982.  Air pollution effects on materials.  In:  Impact
of Air Pollutants on Materials.   Report  of Panel 3 of the NATO/CCMS
Pilot Study on Air  Pollution Control Strategies and Impact Modeling.
pp. 26-27.  English  Translation.

Sharp, A. D.,  S.  T.  Trudgill,  R.  U. Cooke, C. A. Price, R. W. Crabtree,
A. M. Pickles, and  D. L. Smith.   1982.   Weathering of the balustrade on
St. Paul's Cathedral, London.   Earth Surface Processes and Landforms
7:387-389.

Skoulikidis, T. N.   1982.  Atmospheric corrosion of concrete
reinforcements, limestones,  and marbles, pp. 807-825.  In Atmospheric
Corrosion.  W. H. Ailor, ed.  Wiley-Interscience Publications, New York.

Smith, R. D.  1964.   The Preservation of leather bookbindings from
sulfuric acid deterioration.  MLA Research Paper, University of Denver,
Denver, Colorado.

Spedding, D. J.  1969a.  Sulphur  dioxide uptake by limestone.  Atmos.
Environ. 3:683.

Spedding, D. J.  1969b.  The fate of sulphur-35/sulphur dioxide released
in a laboratory.  Atmos. Environ.  3:341-346.

Spedding, D. J. and R. P. Rowlands. 1970.  Sorption of sulphur dioxide
by indoor surfaces.   I. Wallpaper.  J. Applied Chemistry 20:143-146.

Spedding, D. J.,  R.  P. Rowlands,  and J.  E. Taylor.  1971.  Sorption of
sulphur dioxide by  indoor surfaces III.  Leather.  J. Applied Chem.
Biotechnol. 21:68-70.
                                  7-47

-------
Spence,  J.  W.,  F.  H.  Haynie  and J.  B. Upham.  1975.  Effects of gaseous
pollutants  on paints:   A  chamber  study.  J. Paint Tech. 47:57-63.

Stankunas,  A. R.,  D.  F. Unites, and E. F. McCarthy.  1981.  Air
pollution damage to manmade  materials:   physical and economic estimates.
Final Report to the Electric Power  Research Institute, Palo Alto, CA,
EPRI Project No. RP 1004.  (to be published).

Stevens, R. K., T. G.  Dzubay, R.  W. Shaw, W. A. McClenny, C. W. Lewis,
and W. E. Wilson.   1980.   Characterization of the aerosol in the Great
Smokey Mountains.   Environ.  Sci.  Tech. 14:1491-1498.

Sulphur Institute. 1979.  Sulphur  concrete.  Sulphur  Research and
Development 2:2-14.

Sydberger,  T.  1976.   Influence of  sulphur pollution on the atmospheric
corrosion of steel.  University of Gothenburg, Department of Inorganic
Chemistry,  Sweden.

Sydberger,  T. and N.  G. Vannenberg.  1972.  The influence of the
relative humidity and corrosion  products on the adsorption of sulfur
dioxide on  metal surfaces.  Corros. Sci.  12:775-84.

Tombach, I.  1982. Measurement   of local climatological and air
pollution factors affecting stone decay,  pp. 197-210.  J^n_ Conservation
of Historic Stone Buildings and Monuments.  National Academy Press,
Washington, DC.

U.S.-Canada Memorandum of Intent  on Transboundary Air  Pollution.  1982.
Group I.  Phase III.   Draft. June 11,  1982.

U.S. Environmental Protection Agency.  1981.  Air quality criteria for
particulate matter and sulfur oxides.  Vol. Ill, Welfare Effects,
External Review Draft No. 2. Env.  Criteria and Assess. Off., Research
Triangle Park, NC.

Upham, J. B.  1967.  Atmospheric  corrosion  studies  in  two metropolitan
areas.  J.  Air. Poll. Control Assoc. 17:398-402.

Vernon, W.  H. J.  1935.  A laboratory study of atmospheric corrosion  of
metals, Part II:  Iron, the primary oxide film, Part  III.  The  secondary
product of rust (influence of sulfur dioxide, carbon  dioxide, and
suspended particles on the rusting of iron).  Trans.  Faraday Soc.
31:1668-1700.

Yolkwein, A. and R. Springenschmid.  1981.  Corrosion  of Reinforcement
in Concrete Bridges at Different Age Due to Carbonation  and Chloride
Penetration.  2nd International  Conference on the Durability of  Building
Materials and Components.  Preprints.  NBS, Washington, DC 199-209.

Weyde, C.  1972.  A simple test to identify gases which  destroy  silver
images.  Photographic Science and Engineering 16:283-286.
                                  7-48

-------
Winkler, E. M.  1980.   Historical  Implications  in the Complexity of
Destructive Salt Weathering  -  Cleopatra's Needle.  New York.  APT
Bulletin 12:94-101.

Winkler, E. M.  1982.   Problems  in the  deterioration of stone, pp.
108-119.  lr± Conservation  of Historic Stone Buildings and Monuments,
National Academy Press,  Washington,  D.C.

Yocom, J. E.  1979.   Air pollution damage to buildings on the Acropolis.
J. Air Poll. Control  Assoc.  29:333-38.

Yocom, J. E. and N.  Grappone.  1976.  Effects of power plant emissions
on materials.  Report No.  ED-139.   Electric Power Research Institute,
Palo Alto, CA.

Yocom, J. E. and A.  R. Stankunas.   1980.  A review of air pollutant
damage to materials.   A Report to  EPA,  Environmental Criteria and
Assessment Office (NATO-CCMS Program),  Research Triangle Park, NC.

Yocom, J. E. and J.  B. Upham.  1977.  Effects on economic materials and
structures, pp. 65-116.   In  A. C.  Stern, ed.  3rd Edition, Vol. II.
Academic Press, Inc.,  NewTork.

Zeronian, S. H.  1970.  Reaction of cellulosic  fabrics to air
contaminated with sulfur dioxide.   Textile Research Journal 40:695-698.

Zeronian, S. H., K.  W. Alger,  and  S. T. Omaye.  1971.  Reaction of
fabrics made from synthetic  fibers to air contaminated with nitrogen
dioxide, ozone, or sulphur dioxide,  pp. 468-476.  In Proceedings of the
Second International  Clean Air Congress.  H.M.  EngTund and W.T. Beery,
eds.  Academic Press,  New York.

Zeronian, S. H., K.  W. Alger,  and  S. T. Omaye.  1973.  Effect of sulfur
dioxide on the chemical  and  physical properties of Nylon 66.  Textile
Research Jour. 43:228-237.
                                  7-49

-------
                       Review Comment  Guidelines


     Reviewers are encouraged to  submit  comments regarding the material
presented in this document.   To  facilitate  the  revision  process, we
request that certain guidelines be  followed.  Both general and specific
comments on each of the chapters  will  be useful to us.   However, we
cannot emphasize enough the  importance of making your comments specific,
to the extent of indicating  the  section  and page numbers of  the material
in question.  This will greatly  increase the  likelihood  that the editors
and authors will be able to  adequately address  your  comments within a
reasonable period of time.   We also hope your comments will  include your
concurrence with the information  as well as any disagreements.

     Please carefully review the  following  guidelines prior  to providing
your written comments.  Since these comments  may be  included, as
provided, possibly in a separate  volume  at  a  later date, please have
your comments neatly presented;  typing is preferred.

Guidelines to Reviewers -

 1.  Chapters should be reviewed  using a separate form for each.

 2.  General comments, impressions, or opinions of the chapter should  be
     presented on the comment form, separate  from specific comments
     referencing sections,  paragraphs, or statements.

 3.  Comments regarding concepts  or discussions presented should be
     addressed by noting the section at  the beginning of the reviewer's
     comments.  Specific comments on the text within sections should be
     identified with page number, paragraph (P) and  line (L), if
     necessary.

Example -

                    Page
                             Section 2.2.3
                     15      P2       Discussion on  ... does  not
                                     consider studies by ... in
                                     making conclusions

                     20      P3L5    "Susceptible  to change" would be
                                     more appropriate  than  "sensitive"


 4.  When commenting on specific sections/discussion and paragraphs/
     statements, some  suggested reviewers responses  that may be  useful
     include  (with additional appropriate details  by reviewer):

  -  information/statement inaccurate
  -  information/statement incorrectly  interprets  available
          data and/or  literature
  -  information/statement is based on  insufficient data and/or
          literature

-------
Commonly, reviewer's comments note "this  information  is  not correct!"
You should be aware that unless  specific  evidence is  provided
demonstrating that the information is,  in fact,  inaccurate, no action
can be taken.

     During your review of each  chapter,  it  will be necessary to keep in
mind the intended purpose and scope of  this  document  as  outlined in the
Preface.  This is not a criteria document and  is not  designed to address
policy option issues.  The Acidic Deposition Phenomenon  and Its Effects:
Critical Asssessment RevieWPapers j_s an  authored document.  Comments
will be addressed by the authors in cooperation  with  the editors as
deemed appropriate.  These authors are  charged with producing a
scientifically sound and accurate document and are not expected to make
recommendations for the use of the material  in future decision-making
processes.  They will therefore  address comments on technical accuracy
only.

     Only one copy of the review form follows.  This  form should be
copied and used for all written  comments.

-------
                               MANUSCRIPT REVIEW FORM
Title/Draft
         The Acidic Deposition Phenomenon and Its
         Effects:  Critical Assessment Review Papers
         Public Review, Draft, May 1983
             Review Coordinator/Return  to:
               Ms.  Betsy A.  Hood
               CAD  Coorainator
               NCSU Acid Precipitation  Program
               1509 Varsity Drive
               Raleigh. HC  27606	
Reviewer/Organization/Address
Chapter No./Title
\
SUMMARY RATING
Please rate the manuscript as follows: Satisfactory Unsatisfactory
Content 3nd scope
Organization and presentation
Soundness of conclusions based on
material presented
Tprhniral arrurary

Coverage of extant literature .
Editorial quality
Other (specify)

RECOMMENDATIONS
O (1) Acceptable as is
D (2) Acceptable after minor revision
D (3) Acceptable after major revision
D (4) Not acceptable
If you have checked either 3 or 4,
the comments space below.
•



Reviewer's Signature Date
           General Comments  (Use extra  pages  if needed but  please be concise)

-------
Page
Specific  Comments  (by Section with P  and L)
                                          U.S. GOVERNMENT PRINTING OFFICE :  1983 0 -  409-262

-------
United States                                                                                                                              Postage and Fees Paid
Environmental Protection                                                                                                                    Environmental Protection Agency
Agency                                                                                                                                   FPA-335
Washington DC 20460


Official Business                                                                                                                                                           Special
Penalty for Private Use $300                                                                                                                                                 Fourth-Class
                                                                                                                                                                          Rate
                                                                                                                                                                          Book

-------