Skip to main content

Full text of "A Screening procedure to evaluate air pollution effects on Class I wilderness areas"

See other formats


Historic,  Archive  Document 

Do  not  assume  content  reflects  current 
scientific  knowledge,  policies,  or  practices. 


United  States 
Department  of 
Agriculture 

Forest  Service 


Rocky  Mountain 
Forest  and  Range 
Experiment  Station 


A  Screening  Procedure  to  Evaluate 

Air  Pollution  Effects  on 
Class  I  Wilderness  Areas 


Fort  Collins, 
Colorado  80526 


General  Technical 
Report  RM-168 


u4s( 


Douglas  G.  Fox,  Ann  M.  Bartuska, 
James  G.  Byrne,  Ellis  Cowling, 
Richard  Fisher,  Gene  E.  Likens, 
Steven  E.  Lindberg,  Rick  A.  Linthurst, 
Jay  Messer,  and  Dale  S.  Nichols 


Base  cations  -  Ca  +  Mg  +  K  +  Na  (adjusted  for  marine  influence) 
jieq/l 


917257 


Fox,  Douglas  G.;  Bartuska,  Ann  M.;  Byrne,  James  G.;  and  others. 
1989.  A  screening  procedure  to  evaluate  air  pollution  effects  on 
Class  I  wilderness  areas.  Gen.  Tech.  Rep.  RM-168.  Fort  Collins, 
CO:  U.S.  Department  of  Agriculture,  Forest  Service,  Rocky 
Mountain  Forest  and  Range  Experiment  Station.  36  p. 

This  screening  procedure  is  intended  to  help  wilderness 
managers  conduct  "adverse  impact  determinations"  as  part  of 
Prevention  of  Significant  Deterioration  (PSD)  applications  for  sources 
that  emit  air  pollutants  that  might  impact  Class  I  wildernesses.  The 
process  provides  an  initial  estimate  of  susceptibility  to  critical 
loadings  for  sulfur,  nitrogen,  and  ozone.  It  also  provides  a  basis  for 
requesting  necessary  additional  information  where  potential  adverse 
impacts  are  identified. 

Keywords:  Prevention  of  Significant  Deterioration,  air  pollution 


On  the  Cover: 

Foreground:  The  screening  graph  for  determining  effects  of 
atmospheric  deposition  on  aquatic  ecosystems  (fig.  1,  page  6). 
Background:  West  Glacier  Lake,  part  of  the  Glacier  Lakes 
Ecosystem  Experiments  Site  (GLEES),  a  high-elevation  area  that, 
while  not  a  designated  wilderness,  is  being  used  for  research  to 
quantify  atmospheric  effects  on  wilderness.  GLEES  is  instrumented 
for  meteorological,  aerometric,  deposition,  snowmelt,  and  streamflow 
measurements  as  part  of  a  holistic  ecosystem  monitoring  program 
conducted  by  the  air  polution  research  unit  at  the  Rocky  Mountain 
Station.  GLEES  is  located  on  the  Medicine  Bow  National  Forest, 
approximately  15  km  west  of  Centennial,  Wyoming,  in  the  Snowy 
Range  Mountains. 


USDA  Forest  Service 

General  Technical  Report  RM-168 


January,  1989 


A  Screening  Procedure  to  Evaluate 
Air  Pollution  Effects  on 
Class  I  Wilderness  Areas 


Douglas  G.  Fox1,  Ann  M.  Bartuska, 
James  G.  Byrne,  Ellis  Cowling, 
Richard  Fisher,  Gene  E.  Likens, 
Steven  E.  Lindberg,  Rick  A.  Linthurst, 
Jay  Messer,  and  Dale  S.  Nichols 


'Rocky  Mountain  Forest  and  Range  Experiment  Station.  The  Station's  headquarters  is  in 
Fort  Collins,  in  cooperation  with  Colorado  State  University.  Supervision  was  provided  by 
Douglas  G.  Fox,  Chief  Meteorologist  and  Project  Leader  for  The  Research  Work  Unit, 
Effects  of  Atmospheric  Deposition  on  Natural  Ecosystems  in  the  Western  United  States. 


Contents 


Page 


INTRODUCTION   1 

Workshop  Organization  and  Participants   1 

Federal  Land  Managers'  Responsibilities  Concerning  Protection  of 

Class  I  Area  Wilderness   2 

Wilderness  Act   2 

Clean  Air  Act   3 

WORKSHOP  RESULTS   4 

The  Green-Yellow-Red  Screening  Model   4 

Terrestrial  Green  and  Red  Line  Screening  Numbers   5 

Aquatic  Green  and  Red  Line  Screening  Graph   5 

Implementing  the  Screening  Technique   7 

Information  Needs   7 

Monitoring  Considerations   8 

Meteorology   8 

Ambient  Air  Concentration   8 

Deposition   8 

General  Considerations  for  Data  Collection   9 

Quality  Assurance  and  Quality  Control   9 

SPECIFIC  FACTORS  AND  CONSIDERATIONS  IN  DEVELOPING  THE 

MODEL    9 

Terrestrial  Systems   9 

Rationale  Used  in  Selecting  Ozone  Values   10 

Rationale  Used  in  Selecting  Sulfur  Values   11 

Rationale  Used  in  Selecting  Nitrogen  Values   11 

Aquatic  Systems   12 

Concept  of  Surface  Water  Sensitivity   12 

Acidification  Response  Levels   13 

S  and  N  Loadings   14 

Illustration  of  Graph  Use   14 

Information  Needs   14 

Cautions   15 

LITERATURE  CITED   15 

APPENDIX  A.  MAP  OF  FOREST  SERVICE  CLASS  I  AREAS   18 

APPENDIX  B.  BACKGROUND  OF  SAMPLE  CLASS  I  AREAS   20 

Alpine  Lakes  and  Glacier  Peak  Wilderness  -  Washington   20 

Hoover  and  Dome  Land  Wilderness  -  California   20 

San  Gorgonio  Wilderness  -  California   21 

Bob  Marshall  Wilderness  -  Montana   22 

Bridger  Wilderness  -  Wyoming   22 

Superstition  Wilderness  -  Arizona   23 

Joyce  Kilmer-Slickrock  Wilderness  -  North  Carolina,  Tennessee   23 

Otter  Creek  -  West  Virginia  and  Great  Gulf  -  New  Hampshire 

Wildernesses   24 

Boundary  Waters  Canoe  Area  Wilderness  -  Minnesota   25 

APPENDIX  C.  OTHER  AQUATIC  MEASUREMENT  METHODS   27 

Loading/Response  Relationships   27 

Graph  Construction   27 

Detailed  Information  Needs   30 

Conversion  of  Deposition  Values   34 

APPENDIX  D.  PARTICIPANTS  AND  THEIR  AFFILIATIONS   35 


PREFACE 


ACKNOWLEDGMENTS 


A  group  of  scientists  and  land  managers  held  a 
cooperative  workshop  to  help  the  Forest  Service 
develop  a  screening  process  tor  evaluating  Prevention 
of  Significant  Deterioration  (PSD)  applications  for 
sources  that  might  impact  Class  I  area  wildernesses. 
The  process  described  in  this  document  provides  an 
initial  estimate  of  the  susceptibility  of  different  Class  I 
areas  to  critical  loadings  for  sulfur,  nitrogen,  and 
ozone.  Results  should  help  Forest  Service  land 
managers  when  conducting  "adverse  impact 
determinations"  of  PSD  permit  applications  and  provide 
a  ready  basis  for  requesting  necessary  additional 
information  where  potential  adverse  impacts  are 
identified. 

This  document  was  prepared  by  the  authors  and 
participants  at  the  Workshop  on  Air  Pollution  Effects  on 
Wilderness,  held  May  2-5,  1988,  at  the  Institute  of 
Ecosystem  Studies,  Millbrook,  New  York. 


Dr.  Gene  Likens  and  his  co-workers  at  the  Institute 
of  Ecosystems  Studies,  New  York  Botanical  Gardens, 
hosted  the  workshop  at  the  Mary  Flagler  Cary 
Arboretum  in  Millbrook,  NY.  The  participants  at  the 
May  1988  meeting  in  Millbrook  developed  the  concept 
of  this  document,  and  the  authors  wrote  the  first  draft. 
All  the  participants  reviewed  a  second  draft.  A  final 
step  involved  the  review  of  8  scientific  peers  who  were 
not  at  the  meeting,  but  by  virtue  of  both  their  research 
and  their  positions  with  government,  industry,  and 
interested  groups,  were  able  to  substantially  improve 
the  document.  Finally,  scientists  at  the  Rocky  Mountain 
Station  conducting  research  on  effects  of  atmospheric 
deposition  on  natural  ecosystems,  particularly  Frank 
Vertucci,  Robert  Musselman,  and  Anna  Schoettle 
added  significantly  to  the  final  report  by  evaluating  and 
incorporating  reviewers'  comments,  correcting 
references,  and  providing  the  benefit  of  their 
substantial  knowledge  and  experience  to  the  final 
report. 


USER  NOTES 


When  implementing  the  PSD  review  process,  line 
officers  and  staff  must  understand  the  assumptions 
and  variables  used  to  construct  the  screening  model. 
The  model  will  help  in  PSD  review  only  if  the 
assumptions  and  logic  involved  are  fully  understood.  It 
is  critical  that  the  user  recognize  the  development 
methodologies  and  limitations.  For  instance, 
participating  scientists  and  managers  agreed  on  similar 
numerical  loadings  for  a  pollutant  in  seemingly  different 
Class  I  areas.  This  agreement  resulted  because 
similarly  sensitive  ecosystems  occur  in  many  different 
Class  I  areas,  although  not  to  the  same  extent.  For 
example,  alpine  is  the  dominant  ecosystem  in  Alpine 
Lakes  Wilderness  in  northern  Washington,  but  a  minor 
portion  of  the  San  Gorgonio  Wilderness  in  southern 
California.  However,  the  loading  values  for  these  two 


wildernesses  are  the  same  because  the  alpine 
ecosystem  was  considered  most  sensitive,  and  the 
loadings  were  established  to  protect  the  most  sensitive 
ecosystems. 

It  should  also  be  recognized  that  the  loadings 
suggested  by  this  screening  technique  are  likely  to 
overestimate  potential  impacts.  As  such,  they  may  be 
applicable  for  PSD  permit  review  of  effects  on 
designated  Class  I  air  quality  areas,  but  are  not 
intended  to  suggest  target  loadings  on  ecosystems  in 
general. 

Users  should  recognize  that  this  document 
represents  the  state  of  understanding  in  Spring  1988. 
Science  is  very  productive  in  this  field,  and  it  is 
anticipated  that  this  document  will  be  upgraded 
periodically. 


A  Screening  Procedure  to  Evaluate 
Air  Pollution  Effects  on 
Class  I  Wilderness  Areas 


Douglas  G.  Fox,  Ann  M.  Bartuska, 
James  G.  Byrne,  Ellis  Cowling, 
Richard  Fisher,  Gene  E.  Likens, 
Steven  E.  Lindberg,  Rick  A.  Linthurst, 
Jay  Messer,  and  Dale  S.  Nichols 


INTRODUCTION 

Forest  Service  land  managers  need  information 
about  the  effects  of  air  pollution  on  wilderness  areas 
that  have  been  formally  designated  as  Class  I  by  the 
Clean  Air  act  (Public  Law  95-95).  Managers  of  Class  I 
areas  are  responsible  for  the  review  of  preconstruction 
applications  termed  "Prevention  of  Significant 
Deterioration"  (PSD)  permits.  Forest  Service  managers 
must  review  PSD  permits  for  major  new  emission 
sources  (more  than  100  tons  of  a  pollutant  per  year)  or 
the  modification  of  an  existing  source  that  may  cause 
possible  effects  on  Class  I  areas. 

This  introductory  section  describes  a  workshop  of 
Forest  Service  management  leaders  and  prominent 
scientists  studying  the  biological  effects  of  air  pollution 
and  acid  deposition.  They  worked  together  to  identify 
how  best  to  merge  the  current  state  of  science  with 
needs  of  Class  I  area  managers.  (Work  group 
participants  are  also  identified.)  It  then  briefly  describes 
Forest  Service  responsibilities  under  the  Clean  Air  Act 
and  the  Wilderness  Act. 

The  second  section  summarizes  the  major  results 
of  the  workshop.  Results  are  stated  as  proposed 
maximum  acceptable  pollutant  loadings  on  specific 
ecosystems.  These  maximum  loadings  are  intended 
for  use  by  federal  land  managers  screening  PSD 
permits.  The  proposed  screening  process  suggests 
one  of  three  decisions:  recommend  permit  approval 
(new  pollutants  will  lead  to  loadings  below  Green  Line), 
recommend  permit  denial  (new  pollutants  will  lead  to 
loadings  above  Red  line),  and  an  intermediate  zone 
(Yellow  Zone),  where  more  data  are  needed  before 
deciding  on  a  course  of  action. 

The  screening  concept  uses  numerical  values  of 
sulfur  and  nitrogen  deposition  and  ozone 
concentrations  in  nine  different  wildernesses 
considered  representative  of  the  diversity  of  wilderness 
ecosystems. 


The  third  and  fourth  sections  provide  detailed 
explanations,  justifications,  and  cautions  regarding  the 
screening  approach  as  applicable  to  aquatic  and 
terrestrial  ecosystems  in  wilderness  landscapes. 


Workshop  Organization  and  Participants 

A  partnership  between  scientists  and  managers  is 
needed  to  protect  air-quality-related  values  in  Class  I 
area  wildernesses.  The  form  of  such  a  partnership  was 
developed  and  approved  by  some  70  distinguished 
scientists  at  the  1987  Cary  Conference^  which 
focused  on  long-term  studies  of  ecosystems: 

"Ecological  understanding  is  required  to  develop 
environmental  policies  and  to  manage  resources  for 
the  benefit  of  humankind.  Sustained  ecological 
research  is  one  of  the  essential  approaches  for 
developing  this  understanding,  and  for  predicting 
the  effects  of  human  activities  on  ecological 
processes.  Sustained  research  is  especially 
important  for  understanding  ecological  processes 
that  vary  over  long  periods  of  time.  However,  to 
fulfill  its  promise,  sustained  ecological  research 
requires  a  new  commitment  on  the  part  of  both 
management  agencies  and  research  institutions. 
This  new  commitment  should  include  longer  funding 
cycles,  new  sources  of  funding,  and  increased 
emphasis  and  support  from  academic  and  research 
institutions.  Because  they  have  common  long-term 
goals,  we  propose  a  new  partnership  between 
scientists  and  resource  managers.  Elements  of  this 
partnership  include: 

1.    Agreement   by    scientists   to   answer  the 
questions  asked  by  managers,  while  making 
clear  the  level  of  uncertainty  that  exists  and 
 what  additional  research  needs  to  be  done. 

^Statement  adopted  at  the  Cary  Conference  in  Millbrook,  New 
York,  on  May  1 3,  1987:  revised  July  4,  1987  (Likens  in  press). 


1 


2.  Agreement  by  managers  to  give  serious 
consideration  to  these  answers  and  to  support 
the  continuing  research  toward  better  answers. 

Sustained  ecological  research  supported  by  this 

new  partnership  can  contribute  significantly  to  the 

resolution  of  critical  environmental  problems." 

Such  partnerships  are  essential  to  use  scientific 
information  in  an  orderly  and  efficient  manner  for  the 
management  of  complex  natural  resources. 

Organizers  of  this  workshop  invited  a  group  of 
prominent  scientists,  knowledgeable  in  the  areas  of 
effects  of  air  pollution  (sulfur  and  nitrogen  deposition 
and  ozone  exposure)  on  ecosystems,  to  interact  with  a 
group  of  Forest  Service  managers  who  have  air 
resource  management  responsibilities.  The  objectives 
of  this  workshop  were  to  establish  communication 
between  these  two  groups  of  individuals,  and  to 
develop  a  screening  process  for  evaluating  PSD 
applications.  This  relationship  was  fostered  by  a  3-day 
workshop  at  the  Institute  of  Ecosystem  Studies  of  the 
New  York  Botanical  Garden  in  Millbrook,  New  York. 

The  May  1988  workshop  was  to  develop  an  air 
pollution  screening  process  for  managers  of  Class  I 
areas.  The  participants  decided  that  a  screening 
process  that  considered  only  the  impacts  of  the 
deposition  of  sulfur  and  nitrogen  and  ozone 
concentration  on  specific  ecosystems  would  be 
appropriate.  Other  pollutants  can  adversely  affect 
ecosystems,  but  the  chosen  pollutants  are  those  most 
commonly  of  concern.  Pollutant  loadings  are 
determined  by  using  air  dispersion  models  and 
estimates  of  deposition  velocity  to  project  the  worst 
case  deposition  of  S  and  N  from  proposed  industrial 
emissions. 

Four  teams  of  scientists  and  managers  (see  table 
1)  were  formed  to  determine  independently  the  sulfur, 
nitrogen,  and  ozone  values  to  be  used  in  answering 
the  following  questions: 

1.  Below  what  magnitude  of  sulfur  and  nitrogen 
deposition  and  ozone  concentration,  resulting  from 
proposed  air  pollution  emissions,  for  each  of  the  nine 
Class  I  area  wildernesses,  can  a  land  manager  have  a 
high  degree  of  confidence  that  no  air-quality-related 
values  (AQRV's)  would  be  adversely  affected? 

2.  Above  what  magnitude  of  sulfur  and  nitrogen 
deposition  and  ozone  concentration  for  each  of  the 
nine  Class  I  area  wildernesses  can  a  land  manager 
have  a  high  degree  of  confidence  that  at  least  one  of 
the  selected  air-quality-related  values  would  be 
adversely  affected  by  the  proposed  air  pollution 
emissions? 

The  Forest  Service  managers  present  at  the 
Workshop  picked  tentative  AQRV's  (or  reported  those 
already  developed  in  Forest  Plans)  for  the  selected 
wildernesses  in  their  Regions.  These  AQRV's  were 
then  used  by  the  teams  and  working  groups  in  the 


development  of  their  loading  estimates.  Also,  each 
Class  I  area  wilderness  was  described.  Appropriate 
site  data  and  first-hand  knowledge  were  used  to 
estimate  numerical  loadings  and  identify  problems  in 
applying  these  numbers  to  specific  areas.  Values  were 
chosen  to  protect  the  current  condition  of  the  selected 
AQRV's  in  each  Class  I  area. 

Visibility  is  the  only  AQRV  specifically  mentioned  in 
the  Clean  Air  Act,  and  it  has  been  determined  to  be  an 
important  AQRV  in  all  class  I  areas  except  Bardwell 
Bay  (FL)  and  Rainbow  Lake  (Wl).  However,  this 
workshop  did  not  address  visibility.  The  scientists, 
known  for  their  expertise  in  air  pollution  effects  on 
biotic  systems,  were  invited  to  this  workshop  to 
develop  screening  guidelines  for  only  the  terrestrial 
and  aquatic  components  of  the  ecosystem.  The 
absence  of  comments  on  visibility  should  not  be 
construed  as  a  judgment  of  its  relative  value  compared 
to  biotic  systems.  In  fact,  in  some  areas,  visibility  might 
be  considered  adversely  affected  by  air  pollution 
concentrations  that  were  not  considered  adverse  to  the 
biotic  systems.  For  more  discussion  of  visibility,  the 
Forest  Service  Air  Resource  Management  Manual 
(USDA  1987)  should  be  consulted. 


Federal  Land  Managers'  Responsibilities 
Concerning  Protection  of  Class  I  Area 
Wildernesses 

Wilderness  Act 

The  Wilderness  Act  of  1964  (Public  Law  88-557) 
established  the  National  Wilderness  Preservation 
System  "to  secure  for  the  American  people  an 
enduring  resource  of  wilderness."  The  Act  states: 

"A  wilderness...  is  an  area  where  the  earth  and 
community  of  life  are  untrammeled  by  man,  where 
man  himself  is  a  visitor  who  does  not  remain... 
Wilderness  is. ..undeveloped  Federal  land  retaining 
its  primeval  character  and  influence,  without 
permanent  improvements  or  human  habitation, 
which  is  protected  and  managed  so  as  to  preserve 
its  natural  conditions  and  which  generally  appears 
to  have  been  affected  primarily  by  the  forces  of 
nature,  with  the  imprint  of  man's  work  substantially 
unnoticeable..." 

Wilderness  is  a  distinct  resource  with  inseparable 
parts.  When  possible,  natural  processes  are  allowed  to 
operate  within  wilderness;  for  example,  lightning- 
caused  fires  are  allowed  to  burn  under  prescribed 
conditions.  Wilderness  is  managed  to  make  it  as  wild 
and  natural  as  possible,  including  closing  old  roads, 
restoring  damaged  trails  and  campsites,  and  removing 
most  structures.  Managers  use  primitive  tools  to  do  the 


2 


Table  1  .-Work  group  assignments  for  participants. 


Team  1  Aquatic  Ecosystems 

Gene  E.  Likens  (Chairperson) 

Peter  Dillon  (Combined group  chair) 

Thomas  Frost 

Dale  W.  Johnson 

Dale  Nichols 

Ed  Brannon 

Tom  Thompson 

Bill  Carothers 

Dave  Unger 

Jay  Messer  (Note  Taker) 

Team  2  Aquatic  Ecosystems 

Rick  A.  Linthurst  (Chairperson) 

Mike  Pace 

Richard  Wright 

Steve  Mealey 

Mike  Edrington 

Gray  Reynolds 

Anne  Fege 

Richard  Fisher  (Note  Taker) 

Team  3  Terrestrial  Ecosystems 

Ann  M.  Bartuska  (Chairperson) 
Jan  Nilsson 
John  Reuss 
Bill  Mattson 

Steve  Lindberg  (Combined group  chair) 

David  F  Karnosky 

Chuck  Wildes 

John  Butruille 

Clif  Benoit 

Bob  Loomis 

Douglas  G.  Fox  (Note  Taker) 

Team  4  Terrestrial  Ecosystems 

Ellis  Cowling  (Chairperson) 

Gary  M.  Lovett 

Dave  Peterson 

J.  R.  N.  Jeffers 

Peter  B.  Reich 

Dave  Radloff 

Dick  Stauber 

Steve  Harper 

James  G.  Byrne  (Note  Taker) 


1  Affiliations  of  participants  are  given  in  appendix  D. 


job.  As  with  other  National  Forest  resource 
management  efforts,  public  involvement  is  sought  in 
planning  for  wilderness  management  and  use. 

Many  management  activities  and  uses  are 
prohibited  in  wilderness:  roads,  motorized  equipment 
and  mechanical  transport,  landing  of  aircraft,  most 
commercial  enterprises,  and  permanent  structures  and 
installations.  The  Wilderness  Act  allows  certain 
activities  within  wilderness,  as  long  as  the  wilderness 
character  is  preserved.  These  uses  include  livestock 
grazing,  hunting,  fishing,  exercising  water  rights,  and 
existing  mineral  claims.  Special  exceptions  are  made 
in  some  wilderness  legislation  that  permit  mineral 
exploration  and  exploitation,  access  to  private  land, 
maintenance  and  use  of  airstrips,  and,  in  Alaska, 
native  use  for  subsistence. 

The  scientific  value  of  wilderness  is  recognized  in 
the  1964  Act.  A  decade  or  a  century  in  the  future, 
wildernesses  will  serve  as  baseline  or  "control"  areas, 
since  they  are  managed  to  preserve  natural  conditions 
and  generally  will  have  been  affected  primarily  by  the 
forces  of  nature.  Permission  to  conduct  scientific 
studies  is  granted  only  if  the  studies  require  a 
wilderness  environment,  and  cannot  be  accomplished 
outside  the  wilderness.  Motorized  equipment  or 
mechanical  transport  cannot  be  justified  on  the  basis  of 
cost  or  efficiency,  and  are  allowed  only  if  a 
comprehensive  analysis  shows  there  are  no 
alternatives. 


Clean  Air  Act 

The  Clean  Air  Act  (CAA)  Amendments  of  1977 
included  a  program  for  prevention  of  significant 
deterioration  of  air  quality,  generally  referred  to  as  the 
"PSD"  program.  This  PSD  program  is  to  prevent  areas 
currently  having  clean  air  from  becoming  too  polluted. 
Certain  wilderness  areas  and  National  Parks 
established  before  August  1977  were  designated  as 
Class  I  areas.  A  Class  I  designation  allows  only  very 
small  increments  of  new  pollution  above  already 
existing  air  pollution  levels  within  the  area. 
Wildernesses  established  since  August  7,  1977,  are 
Class  II  areas.  Class  II  areas  have  a  larger  increment, 
which  is  about  25  percent  of  the  national  ambient  air 
quality  standard.  Class  I  areas  in  the  National  Forest 
System  are  identified  in  figure  A-1  in  the  appendix  to 
this  report. 

The  CAA  charges  the  federal  land  manager  (FLM) 
of  Class  I  areas  with  an  affirmative  responsibility  to 
protect  the  air-quality-related  values  (AQRV's)  of  these 
areas  from  adverse  air  pollution  impacts.  AQRV's  are 
those  values  within  the  Class  I  area  that  could  be 
affected  by  air  pollution  such  that  the  purpose  for  which 
the  area  was  established  (biological  diversity,  water 


3 


quality,  fish)  would  be  adversely  affected.  Within  the 
Forest  Service,  the  Regional  Forester  has  been 
delegated  this  affirmative  responsibility.  Managers 
must  minimize  the  conflicting  human  impacts  of  air 
pollution,  much  as  they  manage  other  uses  to  limit  their 
impacts  on  the  wilderness  resource. 

The  PSD  program  is  a  preconstruction  review  and 
permitting  process  for  major  new  or  expanding  sources 
of  pollution.  Any  major  facility  seeking  a  new  source 
permit  for  location  or  expansion  in  a  clean  air  area 
must  meet  several  requirements:  Class  I  and/or  II 
increments,  the  AQRV  impact  analysis,  and  the  Best 
Available  Control  Technology  (BACT)  evaluation.  In 
the  PSD  permitting  process,  the  FLM  determines 
whether  a  proposed  source's  emissions  will  have  an 
adverse  impact  on  Class  I  area  AQRV's. 

New  source  permit  applicants  submit  plans  to  the 


permitting  authority,  who  examines  the  proposed 
location  of  the  facility,  its  general  design,  projected  air 
pollution  emissions,  and  potential  impacts.  When  a 
proposed  source's  emissions  may  have  an  impact  on  a 
Class  I  area,  the  permitting  authority  (EPA,  or  the 
State,  if  EPA  has  delegated  PSD  authority  to  that 
State)  alerts  the  FLM.  The  FLM  then  determines  the 
impact  of  the  projected  pollution  level  increases  on  the 
Class  I  area  AQRV's  and  recommends  approval, 
denial,  or  modification  of  the  preconstruction  permit. 
When  the  air  regulatory  authority  certifies  that  a  permit 
application  is  complete,  the  FLM  might  have  as  little  as 
30  days  to  review  the  permit  application  and  respond 
to  the  regulatory  authority.  The  FLM's  determination  of 
adverse  impact  must  be  completed  within  this  period. 
This  reply  is  included  in  the  required  public 
participation  phase  of  the  PSD  program. 


WORKSHOP  RESULTS 


The  Green- Yellow-Red  Screening  Model 

A  conceptual  framework  was  developed  to 
implement  the  partnership  between  scientists  and 
managers  to  help  evaluate  the  potential  impact  of 
proposed  new  air  pollution  sources  on  Class  I  areas. 
This  framework  includes  the  idea  of  acceptable  (Green 
Line),  unacceptable  (Red  Line),  and  intermediate 
(Yellow  Zone)  levels  of  pollution.  It  is  very  important  to 
keep  in  mind  that  this  framework  represents  a 
screening  tool.  As  such,  it  is  intended  to  simplify  the 
decision  process  by  providing  guidelines  for  general 
use  rather  than  formulas  for  specific  application.  In  all 
circumstances,  the  magnitude  of  these  screening 
values,  both  Red  and  Green,  are  subject  to  change 
based  on  better  site  specific  information.  In  the 
absence  of  such  data,  use  of  screening  values  should 
advance  the  evaluation  of  PSD  permits. 

Pollutant  doses  less  than  the  Green  Line  value 
might  be  judged  permissible  by  managers,  and  the 
application  recommended  for  approval  without 
additional  data.  Conversely,  doses  above  the  Red  Line 
value  are  likely  to  cause  at  least  one  AQRV  to  be 
adversely  affected.  Thus  they  would  result  in  a 
recommendation  for  denial  unless  additional  site- 
specific  data  are  provided  to  prove  that  the  identified 
AQRV  of  the  Class  I  area  would  not  be  adversely 
affected.  Doses  falling  between  the  Green  and  Red 
Lines  (the  Yellow  Zone)  would  be  evaluated  on  the 
basis  of  additional  information  provided  or  gathered  by 
the  applicant  or  the  USDA  Forest  Service. 


It  is  prudent  for  the  Class  I  manager  to  have 
AQRV's  clearly  identified,  their  current  status 
monitored,  and  specific  limits  of  impact  defined.  To 
avoid  challenges,  such  information  must  be  based 
upon  or  include  multiyear  data,  and  scientific  peer 
review.  Use  of  these  screening  techniques  is  also 
based  on  the  availability  of  accurate  deposition  and 
concentration  data  at  or  near  the  Class  I  areas.  These 
data  also  should  be  quality  assured.  Suggestions  from 
long-term  sustained  ecological  research  will  be  useful 
in  this  context. 

Specifically,  the  Green  Line  denotes  a  total  loading 
(current  deposition  plus  predicted  additional  deposition 
from  the  new  source)  of  sulfur  and  nitrogen  and  the 
total  dose  of  ozone  that  predicts,  with  a  very  high 
degree  of  certainty,  that  no  AQRV  will  be  adversely 
affected.  The  Red  Line  denotes  a  total  loading  of  sulfur 
and  nitrogen  and  the  total  dose  of  ozone  that  predicts, 
with  a  very  high  degree  of  certainty,  that  at  least  one 
AQRV  will  be  adversely  affected.  Sustained  ecological 
research,  part  of  the  partnership  between  managers 
and  scientists,  will  refine  and  modify  these  decision 
points  with  new  or  better  data. 

Participants  agreed  that  Green  and  Red  Line 
numbers  need  to  be  ecosystem-specific.  The  selected 
numbers  reflect  the  effects  of  pollutants  on  the  AQRV's 
identified  within  the  nine  example  Class  I  areas. 
Terrestrial  and  aquatic  systems  were  considered 
separately  because  the  understanding  of  combined 
impacts  is  not  sufficiently  developed  to  set  numerical 
levels.  Ozone  was  considered  only  to  affect  terrestrial 


4 


ecosystems.  Aquatic  impacts  were  estimated  by  the 
sensitivity  of  surface  waters  as  measured  by  the 
combined  concentrations  of  calcium,  magnesium, 
potassium,  and  sodium  (corrected  for  marine 
influences)  expressed  in  microequivalents  per  liter 
(u.eq/1).  Green  and  Red  Line  values  for  aquatic  impacts 
are  presented  graphically. 


Terrestrial  Green  and  Red  Line  Screening  Numbers 

Participating  scientists  familiar  (to  varying  degrees) 
with  detailed  data  applicable  to  these  Class  I  area 
wildernesses  agreed  to  the  values  in  table  2.  The 
Green  Line  represents  the  total  pollution  loadings 
(current  plus  proposed  new  source  contribution) 
pollution  loadings  below  which  a  land  manager  can 
recommend  a  permit  be  issued  for  a  new  source 
unless  data  are  available  to  indicate  otherwise.  The 
Red  Line  represents  an  estimate  of  the  total  pollutant 
loadings  that  each  wilderness  can  tolerate.  Total 
loadings  above  these  values  suggest  the  land  manager 
recommend  reduction  of  emissions  from  a  new  source 
unless  data  are  available  to  indicate  that  no  AQRV  of 
the  Class  I  area  is  likely  to  be  adversely  affected. 

Pollutant  loadings  between  these  values  require  the 
gathering  of  enough  valid  data  to  determine  whether  or 
not  a  permit  for  a  new  source  should  be 
recommended.  General  ideas  for  dealing  with  loadings 
that  fall  between  the  values  are  described  in  the  next 
section. 


Aquatic  Green  and  Red  Line 
Screening  Graph 

Green  and  Red  Line  screening  values  associated 
with  effects  on  aquatic  ecosystems  are  most 
appropriately  displayed  graphically.  The  sensitivity  of 
aquatic  ecosystems  to  S  and  N  deposition  is  measured 
by  their  acid-neutralizing  capacity  (ANC).  The  ANC 
may  already  be  reduced,  however,  in  systems 
subjected  to  significant  deposition  loading.  A  good 
measure  of  sensitivity  for  fresh  surface  waters  is  the 
sum  of  the  concentrations  of  base  cations  (calcium, 
magnesium,  potassium  and  sodium  ions)  in  the  water. 
Since  Class  I  areas  contain  a  diversity  of  lakes  and 
streams,  the  participants  felt  that  Green  and  Red  Line 
values  should  be  presented  as  a  function  of  the  ion 
concentration.  The  manager  will  need  loadings  based 
on  knowledge  of  the  surface  waters  in  the  Class  I  area 
as  well  as  the  deposition  environment. 

The  graph  for  aquatic  systems  shows  Green  and 
Red  Line  values  with  total  deposition  loading  (in  kg  of 
S/ha-yr)  on  the  vertical  axis  and  concentration  of 
(nonmarine)  Ca+Mg+K+Na  (in  u.eq/1)  on  the  horizontal 
axis.  The  significance  of  these  concentrations  is  based 
on  the  relative  amount  of  water  that  is  exported  from 
the  watershed.  Green  and  Red  Line  values  are 
presented  in  figure  1  for  runoff  estimated  to  be  about 
60-70%  of  the  precipitation,  and  for  40-50%  runoff. 
Green  and  Red  Line  values  for  additional  runoff 
percentages  are  presented  in  appendix  C  in  figures  C- 
1  and  C-2. 


Table  2.-Terrestrial  Green  and  Red  Line  screening  values. 


Nitrogen  deposition^ 


Ozone  concentrations 


Wilderness  areal 

Green  Ln 

Red  Line 

Green  Ln 

Red  Line 

Green  Ln 

Red  Ln 

— kg  N/ha-y- — 

— kg  S/ha-y— 

ppb-  

Alpine  Lakes,  WA 

5-7 

15 

3-5 

20 

35/75 

55/110 

Hoover,  CA 

3-5 

10 

3-5 

20 

35/75 

55/110 

San  Gorgonio,  CA 

5 

15 

3-5 

20 

35/75 

55/110 

Bob  Marshall,  MT 

3-5 

10-15 

5 

20 

35/75 

55/110 

Bridger,  WY 

3-5 

10 

5 

20 

35/75 

55/110 

Superstition,  AZ 

3-5 

15 

5-7 

20 

35/75 

55/110 

Joyce  Kilmer,  NC/Slick  Rock,  TN 

7-10 

15 

5-7 

20 

35/75 

55/110 

Otter  Creek,  WV 

7 

10-15 

5 

20 

35/75 

55/110 

Boundary  Waters  Canoe  Area,  MN 

3-5 

10 

5 

20 

35/75 

55/110 

7  See  appendix  B  for  description  of  wildernesses. 

^Nitrogen  and  sulfur  deposition  are  total  values  including  all  forms,  wet,  dry,  NH4-N  and  NOx-N,  SO4-S, 
SO2-S,  etc. 

^Growing  season  average/second  highest  1  hour  average  value  in  a  year. 


5 


60-70%  Runoff 


14 


40-50% 
Runoff 


12 


CO 


c 

o 

'•4— ' 

CO 

o 

Q. 
0 
"D 


10 


8 


iS  6 


Red  values  -  acidification 
ikely 

Green  values  -  no 
acidification  likely 

Yellow  values  -  uncertain 
whether  or  not  acidification 
occurs 


40 

Base  cations 


160 


200 


80  120 
Ca  +  Mg  +  K  +  Na  (adjusted  for  marine  influence) 
jLteq/l 

Figure  1,-Green  and  Red  Line  va.ues  for  effects  of  deposition  on  freshwater  systems. 


Total 


^deposition  Is  Z  su7u,  oep.s„,.„  —p.  -  ss^ed  ,.oa,,o„s  s,  no.sd  >„  ,Ko  «  whoro  ,5% 
of  total  Nitrogen  deposition  should  be  included. 

6 


Part  of  the  water  that  falls  on  a  watershed  as 
precipitation  is  lost  as  water  vapor  through  evaporation 
or  through  transpiration  by  plants.  Depending  on 
geologic  conditions,  some  may  seep  deep  below 
surface  and  be  lost  from  the  immediate  watershed  as 
ground  water.  The  rest  leaves  as  surface  runoff.  High 
mountain  areas  with  cool  temperatures,  large  amounts 
of  rain  and  snow,  steep  slopes,  and  thin  soils  have 
high  runoff  percentages.  Warm  temperatures,  deep 
soils,  level  topography,  and  vigorous  plant  growth  all 
favor  evapotranspiration  and  reduce  runoff. 

Participants  considered  that,  with  a  few  complex 
exceptions,  effects  of  N  deposition  on  aquatic 
resources  are  not  likely  to  be  significant  because  the  N 
is  taken  up  by  the  watershed  terrestrial  and  aquatic 
biota  and  does  not  contribute  to  acidification. 
Exceptions  are  very  sensitive  lakes  and  watersheds, 
primarily  at  high-elevation  sites  in  the  western  United 
States,  with  base  cation  concentrations  below  50  ueq/l. 
Such  systems  can  be  acidified  by  addition  of  N 
(Grennfelt  and  Hultberg  1986).  For  such 
circumstances,  we  recommend  adding  25%  of  the  total 
N  deposition  to  the  S  deposition  for  use  in  the  aquatic 
graph.  Thus,  if  the  total  deposition  projected  for  a 
western  Class  I  area  containing  low  base  saturation 
waters  is  2  kg  S/ha-yr  and  4  kg  N/ha-yr,  the  value  of  2 
+  .25x4  =  3  should  be  used  in  determining  the  Green 
and  Red  Line  loadings  on  the  Graph. 

Below  a  total  deposition  of  3  kg  S/ha-yr,  there  are 
no  field  data  to  develop  the  Green  and  Red  lines. 
Particularly  in  Class  I  areas  in  the  western  United 
States,  deposition  levels  are  low  and  surface  waters 
have  low  ionic  concentrations  (10-40  ueq/l).  No 
evidence  of  chronic  acidification  has  been  reported. 
However,  snow  melt  has  the  potential  to  seasonally 
acidify  these  surface  waters.  Another  potential  effect  of 
episodic  snowmelt  loading  in  lakes  in  the  west  is 
eutrophication,  a  nutrient  fertilization  effect  leading  to 
increased  organic  productivity.  This  effect  would  also 
require  additional  study.  Thus,  these  systems  fall  in  the 
Yellow  zone. 

Implementing  the  Screening  Technique 

Information  Needs 

Listed  below  are  six  types  of  data  helpful  to 
managers  for  using  the  Green/Yellow/Red  screening 
technique.  These  data  can  be  obtained  from  published 
sources,  or  local  scientists  who  may  have  access  to 
additional  sources  of  information.  It  is  also  prudent  for 
managers  to  formulate  recommendations  for  additional 
research  or  assessment  efforts  which  should  be 
undertaken  by  the  permittee,  Forest  Service,  state,  or 
by  other  organizations  before  or  as  a  condition  to  the 
permit. 


Managers  are  encouraged  to  develop  working 
relationships  with  local  university,  state,  federal,  and 
industrial  research  personnel  to  assist  in  identifying 
already  existing  sources  of  information  or 
recommendations  for  further  research.  This  workshop 
report  should  be  useful  in  initiating  such 
communication. 

Data  needed  to  responsibly  evaluate  a  PSD  permit 
include: 

1.  Deposition  and  air  concentrations  to  estimate 
current  loadings. --Current  loading  and  exposure 
conditions  at  wilderness  sites  must  be  estimated  to 
assess  the  impact  of  new  deposition  increments. 
Measurements  should  take  into  account  expected 
higher  fluxes  at  higher  elevations.  Some  protocols  for 
these  measurements  have  been  established  (Fox  et  al. 
1987). 

Ozone.  Determine  maximum  hourly  average  values 
and  growing  season  average  concentrations. 

Sulfur.  Determine  total  deposition  by  wet,  dry,  and 
cloudwater  processes.  For  some  forest  systems  it 
has  been  shown  that  measurements  of  throughfall 
plus  stemflow  fluxes  provide  a  simple  but  accurate 
estimate  of  total  deposition  of  the  major  S 
components. 

Nitrogen.  Determine  total  deposition  from 
precipitation,  cloudwater,  and  air  chemistry 
measurements  (including  HNO3  vapor  and 
ammonium  ion)  and  appropriate  dry  deposition 
models.  Characterization  of  meteorologic  and 
climatologic  parameters  should  also  be  considered. 
These  can  be  used  to  determine  potential 
climatologic  stresses  and  to  evaluate  dry  deposition 
and  cloudwater  deposition. 

2.  Expected  deposition  and  air  concentrations 
due  to  proposed  source. --Predicted  loading  and 
exposure  at  each  site  must  be  estimated  to  assess  the 
change  in  current  loading  or  air  concentrations  for 
comparison  with  Red  and  Green  Line  values. 
Estimates  must  account  for  elevational  effects  and  for 
important  nearby  sources  that  contribute  to 
background  loading  and  concentration.  The  expected 
worst  case  ambient  loading  or  concentration  should  be 
predicted.  Modeling  is  generally  conducted  by  the 
proponent  and/or  the  regulatory  agency.  Managers 
should  be  aware  that  ozone  is  a  secondary  pollutant 
(generated  in  the  atmosphere)  and  must  be  predicted 
with  a  model  incorporating  photo-chemical  reactions. 
Sulfur  modeling  should  include  any  increased  loading 
due  to  all  important  sulfur  species  (SO2,  particle 
sulfate,  cloudwater  sulfate).  Nitrogen  modeling  should 
consider  all  species  of  N  available  for  plant  uptake 
(HNO3  vapor,  nitrate  and  ammonium  ions  in  rain  and 
cloudwater,  NH3). 


7 


3.  Inventory  of  biological  resources  associated 
with  the  identified  AQRV's  of  the  Class  I  area. -A 
description  of  the  vegetation  communities  (type,  cover) 
is  needed  to  assess  the  relative  response  of  the 
ecosystem(s)  to  pollutants.  Include  in  a  general 
assessment  the  identification  of  unique  communities 
(such  as  small  bog  in  an  otherwise  forested  system), 
and  the  percentage  cover  of  major  ecosystem  types. 

Periodic  remeasurement  of  stand  composition  and 
integrity.  Linkage  to  developing  long-term  monitoring 
programs  (EPA,  FS)  will  assist  in  an  evaluation  of 
change.  A  species  list  including  relative  frequencies  of 
occurrence  is  needed.  An  estimate  of  the  biomass 
increment  for  assessment  of  nitrogen  demand  and  use 
(for  instance  Douglas-fir  require  more  than  alpine 
plants)  must  be  made.  Percent  cover  by  major 
vegetation  type  is  useful  for  this  purpose. 

A  full  inventory  of  aquatic  resources,  including 
water  column  and  benthic  sampling  to  determine 
phytoplankton,  zooplankton,  and  macroinvertebrates 
as  well  as  associated  water  chemistry  is  needed.  A 
quantitative  sampling  procedure  for  macroinvertebrates 
and  flowing  waters  should  be  followed.  Fish 
abundance,  condition,  age  class,  and  other  aspects  of 
community  composition  should  be  measured  (Fox  et 
al.  1987). 

4.  Species  response/biological  effects  data.- 
Following  the  vegetation  survey,  the  response  of  key 
species  to  pollution  loading  must  be  evaluated.  The 
FLM  should  coordinate  these  needs  with  FS  Research 
and  other  research  activities  in  the  area  of  air  pollution, 
plant  response,  acid  deposition,  and  aquatic  resources. 
One  outcome  of  this  might  be  the  development  of 
bioindicators  and  key  sensitive  organisms  in  Class  I 
areas. 

5.  Lake,  stream,  and  soil  survey/geological 
assessment. -Data  needs  for  lake  and  stream  water 
chemistry  are  identified  above.  Information  is 
necessary  to  understand  the  relative  ability  of  soil  and 
bedrock  to  buffer  pollutant  inputs  for  all  subsystems 
within  the  wilderness. 

Lake  and  stream  water.  Care  should  be  exercised 
to  ensure  that  appropriate  guidelines  are  followed 
(see  Fox  et  al.  1987). 

Soil  survey.  Identify  major  soil  series,  followed  by 
more  detailed  chemical  characterizations  of  the 
important  series.  (See  for  example  the  description 
and  protocols  in  the  recent  EPA  Soil  Survey;  Fox  et 
al.  1987.) 

Geological  assessment.  Parent  material  can  be 
assigned  to  one  of  several  weathering  as  described 
in  the  Swedish  critical  load  document  (Nilsson 
1986).  Ecosystem  sensitivity  to  S  inputs  can  then 
be  related  to  the  percentages  of  the  various  classes 
within  the  wilderness. 


6.  Snowpack  chemistry  and  hydrologic 
characteristics  of  the  area.-Snowpacks  in  high- 
elevation  wilderness  have  large  surface  area  to 
capture  S  and  N  compounds.  Thus  pollution  may 
accumulate  in  the  snowpack.  Careful  measurement  of 
snowpack  chemistry  (Fox  et  al.  1987)  can  provide 
good  deposition  loading  information.  Snowmelt  causes 
a  significant  pulse  of  water  which  initially  can  release 
concentrated  chemicals  to  the  ecosystem.  Since 
pollutants  are  not  soluble  in  ice,  they  reside  on  the 
surface  of  the  ice.  As  the  snowpack  warms,  these 
chemicals  are  removed  by  the  initial  meltwater.  This 
may  result  in  a  chemical  pulse  more  concentrated  in 
the  initial  runoff  than  in  the  snowpack  itself.  Managers 
should  assess  the  potential  and  the  likely  effects  of  this 
process  of  pollutant  storage  and  delivery. 

Monitoring  Considerations 

A  major  consideration  for  all  ecosystems  is  the 
current  condition  of  the  atmospheric  environment.  This 
requires  measurement  of  meteorology  and  air  quality  in 
sites  representative  of  the  wilderness. 

Meteorology 

Meteorological  instrumentation  can  be  operated 
with  battery  power  using  microprocessors  to  record 
and  process  the  data.  The  details  of  these  systems  are 
available  in  Fox  et  al.  (1987). 

Ambient  Air  Concentration 

Air  quality  measurement  is  more  problematical. 
Ozone  measurement  requires  a  major  investment  in  an 
air-conditioned  instrument  shelter.  The  shelter  and  the 
ozone  monitor  require  line  power,  frequent  calibration, 
and  standardization.  Such  instrumentation  cannot  be 
put  in  a  wilderness.  Rather,  the  site  selected  for 
monitoring  ozone  must  be  carefully  selected  to  be 
representative  in  exposure,  elevation,  and  ground 
cover  (canopy,  etc.)  of  the  wilderness  being  monitored 
(Fox  et  al.  1987). 

Ambient  air  concentrations  of  SO2,  NOx,  and  NH4 
can  be  measured  using  filter  packs.  These  filter  packs 
also  collect  aerosol  SO4  and  NO3.  These  instruments 
also  require  power,  although  they  need  not  be 
sheltered.  Again,  siting  must  be  representative  of  the 
ecosystems  being  monitored.  These  techniques  are 
described  in  Fox  et  al.  (1987). 

Deposition 

A  major  concern  in  assessing  impacts  is  the 
measurement  of  deposition.  Dry  deposition  cannot 
easily  be  measured  except  with  research  quality 
instruments.  However,  it  can  be  approximated  by 
measurements  of  surrogates,  for  example  snowpack  in 
alpine  areas.  The  snowpack  can  be  monitored  by 


8 


carefully  digging  a  snowpit  and  collecting  snow 
samples  along  its  depth  (Fox  et  al.  1987).  In  a  forest, 
throughfall  and  streamflow  together  have  proven  a 
useful  measure  of  dry  deposition  of  some  chemical 
elements. 

Wet  deposition  should  be  measured  using  NADP- 
type  collectors  and  protocols  for  consistency  and 
comparison  within  the  large  national  network.  Other 
data  required  should  be  collected  using  the  guidelines 
for  wilderness  measurements  (Fox  et  al.  1987). 

Cloud  and  fog  water  interception  in  certain  locations 
can  add  considerably  to  the  total  deposition.  They 
should  be  considered  in  mountain  locations  where 
such  events  occur. 

General  Considerations  for  Data  Collection 

The  land  manager  should  be  aware  of  the  degree  of 
uncertainty  in  the  numbers  obtained,  including  those 
used  to  establish  Red  and  Green  Line  values.  The 
FLM  should  accept  that  some  uncertainty  is 
unavoidable  and  does  not  negate  use  in  decision 
making.  The  following  points  are  relevant. 

1.  The  variance  of  certain  measurements  can  be 
quite  high,  increasing  the  uncertainty  in  estimates,  and 
decreasing  confidence  in  prediction. 

2.  The  level  of  resolution  can  increase 
uncertainty.  Finer  temporal  resolution  of  data  (such  as 
hourly  ozone  averages)  may  be  quite  variable  and 
difficult  to  interpret,  but  when  averaged  over  a  longer 
period  of  time  (weekly),  values  are  more  stable  and, 
hence,  certain. 

3.  Temporal  patterns  in  water  and  soil  chemistry 
may  or  may  not  be  greater  than  the  magnitude  of 
differences  among  soil  types  in  the  same  watershed. 
The  focus  of  the  monitoring  is  on  the  most  sensitive 
component  of  the  ecosystem,  rather  than  any  average 
or  representative  condition. 


4.  Most  wildernesses  are  comprised  of  several 
ecosystems  (such  as  alpine  at  high  elevation;  Douglas- 
fir  at  lower  elevation).  The  manager  needs  to  evaluate 
the  sensitivity  and  the  importance  of  identified  AQRV's 
in  each  ecosystem. 

5.  There  may  be  mismatches  between  data  sets. 
For  example,  air  quality  data  may  be  provided  for  a 
region  or  large  parcel  of  land  (especially  if  derived  from 
a  model);  however,  the  soil  chemistry  or  vegetation 
type  may  be  specific  to  a  location.  Also,  microclimatic 
effects  might  alter  an  air-quality  and/or  deposition 
effect  locally,  reducing  the  representativeness  of  a 
measurement. 

Quality  Assurance  and  Quality  Control 

The  need  for  quality  assurance  and  quality  control 
is  implicit  in  the  need  for  data  upon  which  decisions 
can  be  upheld  in  an  appeal.  The  following  items  reflect 
this  need: 

1.  Utilize  standardized  quality  assurance/quality 
control  guidelines  where  available;  in  particular,  EPA 
procedures  and  the  QA  Methods  Manuals  of  the  Forest 
Response  Program  (Blair  1986). 

2.  Implement  standard  protocols  across  regions. 
For  soils,  coordination  with  the  Soil  Conservation 
Service  is  recommended. 

3.  For  chemical  analyses,  evaluate  laboratory 
capability  and  performance  prior  to  selecting  a 
laboratory.  Evaluation  is  especially  relevant  for  many 
state  and  university  laboratories  where  procedures 
may  be  appropriate  for  agricultural  but  not  forest  soils, 
and  for  lake  and  stream  water  but  not  necessarily 
dilute  surface  waters  and  precipitation.  Water 
chemistry  is  particularly  expensive  and  demanding  on 
laboratory  resources.  Laboratory  procedures  should  be 
carefully  evaluated  and  monitored  both  prior  to 
receiving  samples  and  during  sample  analysis. 


SPECIFIC  FACTORS  AND  CONSIDERATIONS  IN 
DEVELOPING  THE  MODEL 


Terrestrial  Systems 

Effects  of  direct  air  pollution  on  terrestrial  resources 
have  been  the  subject  of  considerable  research  over 
the  past  50  years.  Many  plant  species  have  been 
tested  for  direct  phytotoxicity  due  to  the  so-called 
criteria  pollutants  (03,  S02,  NOx)  as  well  as  other 
reactive  hydrocarbons.  Concentrations  necessary  to 
cause  a  noticeable  impact  are  generally  well  above  the 
current  loadings  in  many  Class  I  areas,  although  ozone 
routinely  occurs  at  phytotoxic  levels  in  California  and 
the  eastern  United  States.  The  major  problems 
associated  with  ozone  toxicity  are:  (1)  plants  respond 


almost  immediately  to  low  concentrations  of  ozone,  but 
their  response  is  not  likely  to  be  significant  until 
concentrations  are  somewhat  higher  than  the  response 
level  (Reich  1987),  and  (2)  generally  only  economically 
important  plants  have  been  studied.  Other  species  may 
or  may  not  respond  in  the  same  manner. 

In  the  1980's  there  has  been  a  growing  awareness 
of  so  called  forest  declines:  large-scale  reductions  in 
the  health  and  vigor  of  trees.  Declines  are  likely 
associated  with  a  host  of  interacting  stress  factors; 
direct  causes  are  hard  to  pin-point.  Research  has  been 
focused  recently  on  determining  the  role  of  acidic 
deposition  in  forest  decline.  This  program  is  rapidly 


9 


accumulating  quantitative  and  qualitative  information 
about  the  effects  of  addition  of  S,  N,  and  associated 
pollutants  on  forest  health.  Considerable  research  is 
addressing  the  mechanisms  of  how  S  and  N  affects 
forests,  including  soil  influences,  foliar  leaching,  carbon 
allocation,  winter  injury,  reproduction  and  regeneration, 
and  insect  and  pathogen  influences.  Finally,  direct 
dose-response  relationships  are  being  determined. 

Workshop  scientists  considered  the  current  state  of 
this  rapidly  moving  field  in  developing  the  numerical 
values  in  the  Green  and  Red  Line  tables.  In  addressing 
ozone  impacts  they  needed  to  address  the  critical 
question  of  what  constitutes  an  ecosystem-level 
impact,  given  that  most  experiments  have  dealt  with 
single  species.  An  exception  may  be  studies  in 
California  by  Miller  and  his  coworkers  (Miller  1973). 

When  considering  the  levels  of  S  and  N  deposition, 
the  scientists  focused  on  soil  effects  because  soils 
were  presumed  to  be  a  very  sensitive  ecosystem 
component,  and  clearly  soil  effects  are  an  ecosystem- 
level  impact.  Dealing  with  N  in  this  context  was  difficult, 
however,  because  most  Class  I  area  ecosystems  are 
likely  to  be  N  limited.  In  this  case  any  increment  of  N  is 
likely  to  cause  some  effect.  Scientists  had  to  estimate 
the  significance  of  anticipated  effects  at  an  ecosystem 
level  in  order  to  develop  numerical  values. 

Rationale  Used  in  Selecting  Ozone  Values 

It  has  been  well  established  that  exposure  of  plant 
leaves  to  air  containing  ozone  results  in  a  number  of 
quantifiable  effects,  including  visible  injury,  reduced 
photosynthetic  capacity,  increased  respiratory  rate, 
briefer  leaf  retention  time,  and  reduced  growth  (Barnes 
1972,  Hayes  and  Skelly  1977,  Pye  1988).  The 
magnitude  of  these  effects  depends  on  several  factors, 
including  the  concentration  of  the  pollutant,  the 
duration  of  exposure,  and  other  environmental  factors 
(USEPA  1986).  Sensitivity  to  ozone  varies  among  and 
within  species  because  of  inherent  differences  in 
uptake  rates  (Reich  1987)  and  also  because  of  other 
unknown  genetic  factors  (Karnosky  and  Steiner  1981). 
Despite  differences  at  the  leaf  level,  responses  of  a 
wide  variety  of  species  types  can  be  effectively 
characterized  by  taking  into  consideration  exposure 
dynamics  and  uptake  characteristics  (Reich  1987). 

The  immediate  effect  of  elevated  ozone  levels  in 
wilderness  areas  would  be  decreased  leaf  longevity, 
reduced  net  carbon  gain  of  foliage,  reduced  growth  of 
individual  plants,  and  foliar  injury.  Other  adverse 
effects  could  include  alteration  of  plant  allocation  of 
carbon;  greater  susceptibility  to  insects,  pathogens, 
water  stress,  winter  injury,  or  other  stress  agents; 
possible  changes  in  species  composition  of  plant 
communities;  and  possible  loss  of  genetic  resources  of 
sensitive  genotypes  within  a  species. 


The  Green  Line  values  for  ozone  for  all  wilderness 
terrestrial  plant  ecosystems  are  set  at  75  ppb  (peak  1- 
hour  average)  or  35  ppb  (growing  season  average).3 
We  follow  regulatory  procedures  established  by  the 
Environmental  Protection  Agency,  which  define  a  peak 
as  the  second-highest  one  hour  average  concentration 
in  a  year  (EPA  1986).  Estimates  of  average  ozone 
concentrations  in  clean  air  range  from  15  to  30  ppb. 
However,  estimates  of  background  ozone 
concentration  are  very  difficult  because  measurements 
do  not  exist,  and  models  show  complex  nonlinear 
interactions  where  ozone  production  depends  on  NOx 
concentration,  nonmethane  hydrocarbon  (NMHC) 
concentration,  and  seasonality  (Liu  et  al.  1987).  NOx 
background  concentrations  range  from  less  than  1 
ppbv  (remote  locations  in  the  western  United  States)  to 
about  7  ppbv  (remote  locations  in  the  eastern  United 
States)  and  about  twice  that  in  Europe  (Fehnsenfeld  et 
al.  1988).  Modeling  estimates  (Liu  et  al.  1987)  would 
then  project  background  ozone  concentrations  of 
approximately  20  ppb  in  the  western  United  States  and 
70  ppb  in  the  eastern  United  States.  Of  course,  NOx 
and  ozone  concentrations  in  the  vicinity  of  urban  areas 
(such  as  Los  Angeles,  Phoenix,  and  Denver)  are  often 
higher  than  the  eastern  background. 

The  Green  Line  values  were  chosen  to  give 
reasonable  certainty  that  no  significant  damage  will 
occur  to  the  ecosystem.  Based  on  available 
information  about  plant  response  to  ozone,  we 
conclude  that  any  increase  in  ozone  levels  above 
background  (clean  air)  will  have  some  adverse  effect 
on  individual  leaves  of  at  least  some  species. 
However,  we  believe  that  the  integrity  of  the  ecosystem 
can  be  maintained  with  the  slight  amount  of  stress  on 
either  sensitive  individuals  and/or  sensitive  species 
that  might  occur  below  Green  Line  levels. 

The  Red  Line  values  for  ozone  are  set  at  110  ppb 
(peak  1-hour  average)  or  55  ppb  (growing  season 
average).  Species  from  all  plant  types  suffer  reduced 
net  photosynthesis  and  growth  if  exposed  to  55  ppb  for 
the  daylight  hours  every  day  of  the  growing  season. 
Although  some  of  the  data  used  in  the  development  of 
this  value  are  based  on  average  concentration  during 
daylight  hours  only  (12  hours),  the  loading  value 
seasonal  averages  use  24  hours  per  day.  While  it  is  an 
area  of  scientific  controversy  (Musselman  et  al.  1988) 
whether  a  12-hour  or  a  24-hour  based  ozone  season 
average  is  better  correlated  with  effects,  12-hour  data 
are  not  available  from  regulatory  agencies.  Thus,  24- 
hour  data  are  recommended  to  calculate  seasonal 
averages.  

3Growing  season  average  may  not  be  available  in  many 
locations,  and  determination  of  growing  season  will  be  specific  to 
each  species.  Thus,  it  is  likely  that  the  peak  values  will  be  more 
useful  than  the  growing  season  average  values  (USEPA  1986, 
Musselman  et  al.  1988). 


10 


Ambient  ozone  levels  in  Class  I  areas  should  not 
exceed  peak  annual  1-hour  average  values  of  110  ppb. 
Data  from  numerous  ozone  monitoring  stations 
suggest  that  exceeding  110  ppb  for  the  peak  1-hour 
period  of  the  year  would  be  accompanied  by  15  to  50 
(or  more)  hours  of  exposure  to  ozone  levels  greater 
than  80  ppb.  Adverse  effects  are  greater  at  higher 
ozone  concentrations. 

Ozone  effects  are  cumulative  for  each  individual 
plant,  but  the  chemical  itself  is  ephemeral  and  does  not 
accumulate  in  the  plant  or  ecosystem.  Also,  ozone 
does  not  enter  the  soil  in  sufficient  quantities  to  be  of 
any  significance.  Finally,  we  conclude  that  some 
individuals  and  species  will  be  damaged  in  all 
wilderness  ecosystems  at  ozone  levels  between  the 
Red  and  Green  Line  values.  In  such  Yellow  Zones, 
predicted  damage  must  be  evaluated  on  a  case  by 
case  basis.  The  PSD  recommendation  may  depend  on 
the  relative  value  of  the  plant  community  as  an  AQRV 
within  that  particular  wilderness  area. 

Rationale  Used  in  Selecting  Sulfur  Values 

Two  criteria  or  effects  have  been  considered  to  set 
the  Green  and  Red  Line  levels  of  deposition  for  sulfur: 
(1)  removal  of  base  cation  from  soils  in  association 
with  the  SO42-  anion,  a  "capacity"  effect,  and  (2)  the 
"intensity"  effects  resulting  from  the  changes  in  soil 
solution  composition.  This  distinction  becomes 
important  in  areas  affected  by  marine  air  masses 
where  natural  SO42-  levels  may  be  well  above  our 
proposed  Green  Line  values.  An  approximate 
correction  can  be  made  by  subtracting  the  marine 
component  based  on  the  S0427CI"  ratio  in  seawater. 
Marine  sulfate  is  generally  not  considered  deleterious 
because  it  is  normally  accompanied  by  base  cations, 
particularly  Na  and  to  some  extent  Mg,  and  thus  does 
not  contribute  to  acidification  of  the  system.  There  may 
be  episodic  exceptions  to  this. 

For  our  basic  capacity  comparisons,  we  have 
assumed  a  soil  depth  of  30  cm  with  a  bulk  density  of 
1.1  kg/liter.  At  a  loading  of  3  kg  S/ha,  it  would  require 
approximately  175  years  to  achieve  a  reduction  of  1 
meq  of  base  cations  per  100  g  soil.  This  reduction 
would  be  at  least  partially  offset  by  weathering  of 
primary  minerals.  Somewhat  higher  deposition  levels 
would  be  acceptable  in  areas  where  soils  are  deep  or 
are  well  supplied  with  bases,  and  these  considerations 
are  reflected  in  the  proposed  values  for  some  of  the 
particular  ecosystems. 

Given  these  assumptions,  the  maximum  allowable 
(Red  Line)  values  of  20  kg/ha  of  S  could  achieve  the 
reduction  of  1  meq  base  cation  within  about  26  years. 
This  base  cation  reduction  would  generally  be 
unacceptable  unless  the  system  contains  free  CaC03. 
However,  with  the  possible  exception  of  the 
Superstition  Wilderness,  all  of  the  specific  ecosystems 


considered  here  contain  considerable  areas  of  non- 
calcareous  soils. 

For  our  evaluation  of  intensity  effects,  we  have 
assumed  1  m  precipitation  in  excess  of 
evapotranspiration.  The  Green  Line  value  of  3  kg/ha 
would  increase  solution  concentrations  by  about  19 
u.eq/1,  which  is  near  the  natural  background  for  surface 
waters  in  areas  that  do  not  contain  significant  amounts 
of  readily  oxidizable  sulfur-bearing  minerals. 
Furthermore,  this  concentration  would  be  unlikely  to 
result  in  significant  mobilization  of  soluble  inorganic 
forms  of  aluminum.  The  corresponding  increase  for  the 
maximum  value  of  20  kg/ha  would  increase  solution 
concentrations  by  about  125  u.eq/1.  This  concentration 
is  in  the  range  where  Al  mobilization  might  occur  in 
acid  soils,  and  with  the  possible  exception  of  the 
Sonoran  systems,  would  probably  not  be  acceptable. 

Rationale  Used  in  Selecting  Nitrogen  Values 

The  basic  features  of  N  cycling  in  forest 
ecosystems  are  fairly  well  understood,  and  can  provide 
a  broad  conceptual  outline  for  arriving  at  deposition 
loadings  to  wilderness  areas.  The  following  is  a  brief 
summary  of  some  of  the  important  features  of  N  cycles 
relevant  to  loading  considerations. 

Nitrogen  is  the  only  major  plant  nutrient  that  does 
not  accumulate  to  any  significant  extent  in  inorganic 
forms  in  the  soil.  Although  ammonium  is  strongly 
adsorbed  to  soil  cation  exchange  sites,  ammonium 
almost  never  significantly  accumulates  because  of 
biological  uptake  by  plants,  grazers,  decomposers,  and 
nitrifying  bacteria.  Thus,  forest  ecosystems  can 
accumulate  atmospherically  deposited  N  only  by 
biological  mechanisms;  specifically  through 
incorporation  into  plants,  plant  feeders  (herbivores), 
and  decomposers  such  as  soil  microorganisms  and 
invertebrates.  Because  N  is  the  nutrient  most 
commonly  limiting  growth  of  forests  in  North  America, 
forested  ecosystems  usually  show  a  net  accumulation 
of  atmospherically  deposited  N. 

Any  increase  in  N  deposition  as  nitrate  or 
ammonium  ion  to  N-limited  wilderness  areas  will  most 
probably  result  in  some  increase  in  growth,  and  may 
actually  improve  the  health  of  the  ecosystem.  Species 
adapted  to  low  N  conditions  might  be  replaced  as  a 
result  of  fertilization.  It  is  also  possible  that  chronic  N 
enrichment  may  eventually  predispose  plants  to 
outbreaks  of  plant-feeding  insects  and  fungal 
pathogens  because  of  changes  in  the  plants'  carbon 
allocation  to  growth  and  defensive  processes. 

N  deposited  in  excess  of  biological  need  almost 
invariably  leads  to  nitrification,  microbially  mediated 
nitrate  and  nitrite  formation  in  the  soil,  and  increased 
leaching  of  nitrate  and  associated  cations.  The  nitrate 
so  produced  may  lead  to  surface-  or  groundwater 


degradation  unless  it  encounters  anaerobic  conditions. 
Under  these  conditions,  it  may  be  microbially  reduced 
to  N2O  gas  (denitrification),  thus  decreasing  the 
potentially  deleterious  effects  of  excessive  N 
deposition  on  water  quality.  These  processes  may  still 
leave  the  potential  increases  in  soil  acidification  to  be 
considered. 

The  Green  Line  values  (3-10  kg/ha-yr)  for  nitrogen, 
across  all  the  ecosystems  considered,  were  selected  to 
give  reasonable  certainty  that  no  significant  change  in 
the  forest  ecosystem  will  occur  below  this  amount  of 
nitrogen  deposition. 

The  Red  Line  values  (10-15  kg/ha-yr)  for  nitrogen, 
across  all  the  ecosystems  considered,  were  selected  to 
give  reasonable  certainty  that  these  amounts  of 
nitrogen  deposition  will  result  in  significant  cnanges  in 
the  accumulation  of  nitrogen  and  in  the  species 
composition  or  other  important  features  of  the 
ecosystem. 

While  the  fundamental  elements  of  forest  N  cycles 
are  reasonably  well  understood,  quantitative  data  on  N 
cycling  in  wilderness  areas  is  quite  scarce  at  best,  and 
in  many  areas  completely  lacking.  Therefore,  the  Red 
and  Green  Line  loadings  for  N  deposition  in  wilderness 
areas  are  judgments  based  on  a  very  limited  database. 
We  strongly  urge  that  relevant  N  cycling  parameters  be 
measured  in  those  wilderness  areas  for  which  there  is 
a  potential  concern  about  increased  N  deposition.  It  is 
also  important  to  note  that  atmospheric  deposition  at 
the  chosen  target  loadings  may  well  have  some  effect 
upon  wilderness  areas  in  terms  of  stimulating  growth; 
thus,  there  is  no  assertion  that  these  levels  will  protect 
the  wilderness  areas  from  all  effects.  In  our  judgment, 
however,  the  Green  Line  levels  are  sufficiently  low  that 
perceptible  deleterious  effects  upon  plant  health, 
changes  in  species  composition,  or  degradation  of 
water  quality  are  unlikely. 


Aquatic  Systems 

Aquatic  resources  are  important  Air  Quality  Related 
Values  in  most  Class  I  areas.  Determining  how  best  to 
prevent  significant  deterioration  by  atmospheric 
pollutants,  however,  is  not  as  straightforward  as 
establishing  their  importance. 

This  section  provides  general  guidelines  for  all 
surface  waters  relative  to  the  amount  of  sulfur  and 
nitrogen  that  can  be  deposited  on  an  annual  basis. 

Green  Line  values  indicate  levels  below  which  it  is 
highly  unlikely  that  the  most  sensitive  aquatic 
resources  will  be  significantly  affected,  while  Red  Line 
values  indicate  levels  above  which  it  is  highly  //7<e/ythat 
the  most  sensitive  aquatic  resources  will  be 
significantly  affected. 

The  guidelines  use  the  concentrations  of  base 


cations:  calcium  (Ca),  magnesium  (Mg),  potassium  (K), 
and  sodium  (Na),  as  a  measure  of  sensitivity.  The 
sulfur  and  nitrogen  loadings  above  which  adverse 
change  is  likely  and  below  which  change  is  unlikely  are 
based  on  the  most  sensitive  waters. 


Concept  of  Surface  Water  Sensitivity 

Lakes  and  streams  differ  in  their  inherent  sensitivity 
to  inputs  of  acidifying  compounds  from  the 
atmosphere.  A  number  of  factors  affect  lake  sensitivity; 
bedrock  geology,  soil  and  vegetation  type,  hydrologic 
characteristics,  lake  chemistry  and  biology,  and 
precipitation  volume  are  among  the  important  factors. 
Maps  of  bedrock  geology  are  often  used  to  indicate 
areas  with  sensitive  lakes  and  streams.  Seepage 
lakes,  lakes  which  have  no  visible  outlet,  are  likely  to 
be  dominated  by  precipitation,  while  drainage  lakes  are 
likely  to  be  influenced  by  watershed  base  cation 
supply.  Seepage  lakes,  all  other  things  being  equal, 
will  be  more  sensitive  to  acidification.  The  lake  or 
stream  chemistry  itself  provides  a  convenient  measure 
of  sensitivity.  The  lake  water  integrates  many 
watershed  factors  that  may  be  difficult  to  measure  or 
estimate  in  the  field. 

Any  of  several  water  chemistry  parameters  may  be 
used  to  estimate  sensitivity.  In  pristine  areas  receiving 
little  or  no  acid  deposition,  acid  neutralizing  capacity 
(ANC)  provides  a  useful  measure-the  lower  the  ANC, 
the  more  sensitive  is  the  water  body.  In  areas  receiving 
acid  deposition,  however,  ANC  may  have  decreased. 
Since  ANC  changes  with  acid  deposition,  it  cannot  be 
used  directly  to  assess  sensitivity.  Acid  neutralizing 
capacity  can  be  defined  as  the  sum  of  the  base  cations 
minus  the  sum  of  the  strong  acid  anions  (SO4-,  NO3-, 
Ch)  in  a  water  sample  if  concentrations  of  organic 
acids  and  aluminum  are  insignificant.  Because  of  the 
principle  of  electroneutrality,  changes  in  base  cations 
and/or  acid  anion  concentrations  must  affect  the  ANC 
of  the  sample. 

Calcium  and  magnesium  concentrations  have  been 
used  widely  as  a  measure  of  inherent  sensitivity. 
Henriksen's  (I979)  empirical  nomograph  for  lake 
acidification  uses  Ca+Mg  concentrations  as  a  measure 
of  sensitivity  and  SO4  concentration  (or  alternatively 
pH  of  precipitation)  as  a  measure  of  acid  deposition  to 
determine  whether  a  given  lake  will  be  acidic  (pH<4.7), 
transitional  (4.7-5.3),  or  bicarbonate  dominated 
(p(H>5.3).  This  empirical  approach  developed  on  the 
basis  of  several  hundred  Norwegian  lakes  has  been 
shown  to  be  of  general  applicability  to  lakes  in  many 
regions  of  Europe  and  North  America  (Wright  and 
Henriksen  1983,  Henriksen  and  Brakke  1988,  Wright 
1988,  Reuss  et  al.  1986). 

While  Ca  and  Mg  are  the  major  cations  usually 


12 


associated  with  alkalinity,  the  weathering  of  minerals 
containing  K  and  Na  can  also  contribute  significantly  to 
ANC.  Given  the  geological  diversity  of  the  Class  I 
areas,  we  used  the  sum  of  the  four  major  base  cations 
(adjusted  to  subtract  any  marine  influences)  as  the 
principal  measure  of  inherent  sensitivity. 

The  relationships  between  lake  ANC,  anions,  base 
cations  (Ca+Mg+K+Na),  pH,  and  conductivity  can  be 
derived  either  empirically  or  from  basic  water  chemistry 
theory.  The  figures  used  for  our  screening  technique 
were  constructed  showing  the  relationships  between 
non-marine  base  cations  and  total  S  or  S+N  deposition 
for  sensitive  lakes.  Measured  lake  chemistry  data  from 
the  1984  Eastern  Lake  Survey  (Linthurst  et  al.  1986) 
and  the  1985  Western  Lake  Survey  (Landers  et  al. 

1987)  were  used.  Total  deposition  for  S  and  N  were 
estimated  for  eastern  lakes  from  analysis  of  wet 
deposition  data  done  by  Husar  (1986)  with  30  percent 
added  to  account  for  dry  deposition.  For  the  western 
lakes  we  averaged  data  from  nearby  high-elevation 
National  Atmospheric  Deposition  Program  (NADP) 
sites  (NADP  1988)  with  data  from  high-elevation  snow 
chemistry  studies  (Brown  and  Skau  1975,  Melack  et  al. 
1982,  Laird  et  al.  1986,  Loranger  1986,  Loranger  and 
Brakke  1988,  Reddy  and  Classen  1985,  Vertucci  in 
press).  No  additional  correction  was  made  for  dry 
deposition  because  not  enough  information  is  available 
to  estimate  the  potential  contribution  (Young  et  al. 

1988)  . 


Acidification  Response  Levels 

The  effects  of  O3,  N,  and  S  can  be  assessed 
directly  for  aquatic  ecosystems.  Ozone  has  no  known 
direct  effects  on  aquatic  systems,  and  therefore  does 
not  warrant  further  consideration.  For  aquatic  systems, 
pollutant  loadings  by  N  and  S  exert  their  influence  on 
biotic  communities  primarily  by  changing  pH  conditions 
rather  than  by  a  direct  influence  due  to  the  chemical 
species  of  N  or  S.  Our  focus,  therefore,  is  on  defining 
threshold  levels  for  N  and  S  loading  based  on  their 
influence  on  pH.  Again  in  very  dilute,  high-elevation  N- 
limited  lakes,  the  addition  of  N  can  initiate 
eutrophication. 

Changes  in  lake  or  stream  pH  due  to  atmospheric 
inputs  of  N  and  S  can  have  a  variety  of  direct  and 
indirect  effects  on  aquatic  communities  and  ecosystem 
processes.  Increased  hydrogen  ion  concentrations  can 
have  a  direct,  toxic  effect  on  organisms.  Such  direct 
effects  on  one  or  a  group  of  organisms  may  exert, 
subsequently,  an  indirect  influence  on  the  occurrence 
of  other  organisms,  primarily  through  food  web 
interactions.  Changing  pH  may  also  influence  the 
solubility  of  nutrients  or  toxic  compounds  and  elements 
(such  as  aluminum)  which  in  turn  may  affect  the 


occurrence  of  organisms  either  directly  or  indirectly.  It 
is  important  to  note  that  small-scale  changes  in 
chemical  conditions  are  likely  to  affect  physiological 
processes  or  a  particular  life  stage  of  an  organism  prior 
to  the  disappearance  of  a  taxon. 

Information  on  the  effects  of  a  particular  decrease 
in  pH  on  a  lake  or  stream  can  be  derived  from  four 
types  of  sources  (EPRI  1986):  (1)  laboratory 
bioassays,  (2)  synoptic  surveys  of  the  distribution  of 
organisms  across  systems  with  a  range  of  pH  values 
(Eilers  et  al.  1984,  Confer  et  al.  1983,  Haines  1981), 
(3)  manipulations  of  pH  in  mesocosms,  and  (4)  whole- 
system  experimental  manipulations  of  pH  (Schindler  et 
al.  1985,  Brezonik  et  al.  1986,  Hall  et  al.  1980,  Hall  and 
Likens  1981).  Each  of  these  sources  can  provide 
useful  information  on  the  effects  of  changing  pH 
conditions.  However,  whole-system  experiments 
provide  the  best  detailed  information  on  the  response 
of  aquatic  systems  to  acid  stress  because  they  involve 
a  direct,  controlled  manipulation  of  pH  conditions,  and 
they  are  conducted  at  a  scale  that  encompasses  a  full 
range  of  population  and  ecosystem  processes 
(Schindler  1988,  Hall  and  Likens  1984).  Specifically, 
results  from  these  studies  indicate  effects  that  could 
not  have  been  discovered  with  other  approaches. 

In  general,  considering  information  drawn  from  all  of 
the  sources  listed  above,  it  is  possible  to  conclude  that 
pH  changes  of  less  than  0.5  units  are  capable  of 
producing  considerable  change  in  the  biotic 
communities  of  either  lakes  or  streams.  In  many  cases, 
fish  populations  would  be  expected  to  respond  to  a  0.5 
unit  pH  change.  Shifts  of  1  pH  unit  can  lead  to  major 
changes  in  the  occurrence  of  other  organisms, 
particularly  sensitive  ones  such  as  mollusks.  Workshop 
participants  suggested  a  0.5  pH  unit  change  as  a  Red 
Line  projection  and  a  0.1-0.2  unit  projected  change  for 
the  Green  Line,  in  sensitive  systems  with  pH  of  order  6 
or  very  low  ANC. 

Because  many  wilderness  areas  contain  a  diversity 
of  lakes  and  streams,  it  is  important  to  target  a  subset 
of  lakes  and  streams  as  primary  AQRV's.  Generally, 
lakes  and  streams  with  low  base  cation  concentrations 
(BCC)  or  acid  neutralizing  capacity  (ANC)  are  most 
likely  to  be  affected  by  the  lowest  level  of  pollutant 
input.  The  federal  land  manager  should  therefore 
target  the  lowest  BCC  and  ANC  systems  within  a 
wilderness  area  for  evaluation.  An  inventory  of  the 
BCC  and  ANC  of  aquatic  resources  in  an  area  would 
provide  extremely  valuable  and  cost-effective  baseline 
information.  Among  the  low  BCC  lakes  and  streams, 
those  with  pH  values  of  around  6.0  may  be  the  most 
likely  to  change  with  an  increased  S  or  N  loading,  and 
should  be  given  the  most  detailed  attention.  Typical 
symptoms  of  acidification  for  lakes  and  streams  include 
the  development  of  extensive  mats  of  filamentous 
green  algae,  increased  water  clarity,  and/or  changes  in 


13 


the  proportional  occurrence  of  macroinvertebrate 
species  (Schindler  et  al.  1985,  Hall  et  al.  1985,  1987). 
It  is  also  important  to  note,  however,  that  a  shift  in  the 
pH  of  a  lake  or  stream  with  a  current  value  of  7.0  is 
also  likely  to  cause  changes  in  the  biota. 

In  some  wilderness  areas,  lakes  or  streams  may 
already  have  pH<6.0.  In  many  cases  these  could  be 
naturally  acid  rather  than  anthropogenically  altered 
systems.  Natural  acidification  is  often  the  case  where 
sphagnum  bogs  occur  and  runoff  waters  are  yellow- 
brown  stained.  These  waters  can  have  high  organic 
carbon  concentrations,  and  therefore  the  natural 
contributions  to  acidity  may  be  high.  Although  such 
naturally  acid  systems  may  contain  assemblages  of 
species  that  are  adapted  to  low  pH  conditions,  they 
may  still  be  sensitive  to  the  effects  of  increased  N  and 
S  loading.  These  colored  water  systems  require  more 
detailed  consideration. 

Although  the  graphs  presented  are  based  on  S 
deposition,  N  may,  in  some  circumstances,  also  affect 
lake  acidification.  To  account  for  the  acidifying  effect  of 
N  deposition,  we  again  used  an  empirical  approach. 
Generally,  most  N  inputs  are  retained  in  the  terrestrial 
ecosystem.  The  fraction  that  leaks  out  to  surface  water 
depends  on  a  variety  of  site  factors  such  as  vegetation 
type,  stage  of  ecosystem  development,  hydrology,  and 
history  of  acid  deposition.  Large  leaks  of  N  often  result 
from  vegetation  disturbance  such  as  clearcutting,  fire, 
and  windthrow  (Likens  et  al.  1970,  Bormann  and 
Likens  1979). 

Henriksen  and  Brakke  (1988)  have  shown  from 
empirical  data  for  surface  waters  in  Norway  that  the 
percent  of  incoming  N  retained  by  the  terrestrial 
system  is  generally  75-100%.  Many  of  these  lakes  and 
streams  are  comparable  chemically  and  biologically  to 
mountainous  areas  in  the  United  States.  Some 
acidified  areas  have  shown  an  increase  in  NO3-,  while 
unacidified  areas  have  very  low  concentrations  of 
NO3-  in  runoff  (Henriksen  and  Brakke  1988). 

While  there  will  be  unusual  situations  where  N  can 
be  released  from  ecosystems,  a  general  exception  is 
extremely  sensitive  high  mountain  lake  watersheds. 
For  such  high  elevation  systems  (BCC<50  |ieq/l), 
adding  25%  N  deposition  to  the  S  deposition  is  merely 
a  guideline  because  the  uptake  will  vary  from  site  to 
site,  and  also  over  time  at  a  given  site.  Our  approach 
here  is  based  on  current  situations  measured  at  lakes 
and  streams  of  varied  sensitivity  and  receiving  varied 
amounts  of  acid  deposition  both  as  N  and  S. 


S  and  N  Loadings 

Current  S  and  N  loadings  are  necessary  to  locate 
the  lake(s)/stream(s)  on  the  nomograph  (fig.  1).  The  S 
loading  should  be  total  S  (wet  SO4  +  dry  particulate 


SO4  +  SO2  gas)  and  can  probably  be  best  estimated 
from  the  NADP  wet  deposition  fields  +  measured  SO2 
levels,  combined  with  best  estimates  of  deposition 
velocity.  N  loading  (NO3-  +  NH4+)  can  be  calculated  in 
an  analogous  manner  for  cases  (BCC<50  u.eq/1)  where 
N  is  to  be  considered. 


Illustration  of  Graph  Use 

For  example,  assume  that  lake  water  quality  has 
been  identified  as  an  AQRV,  and  that  pH  was  chosen 
as  a  measurement  to  be  monitored.  Lake  pH,  identified 
as  an  indicator  of  the  health  of  the  aquatic  ecosystem, 
needs  to  be  maintained  above  5.8.  This  is  equivalent  to 
maintaining  an  ANC  over  10  u.eq/1  in  water.  Data  have 
been  collected  that  identify  a  particular  lake  whose 
base  cation  concentration  is  80  u.eq/1.  The  screening 
concept  in  the  Aquatic  Graph  (fig.  1)  is  to  be  applied  to 
this  lake. 

Since  the  lake  in  this  example  has  a  measured  non- 
marine  base  cation  sum  of  80  ueq/l,  the  results  are  3 
kg  S/ha/yr  and  5.5  kg  S/ha/yr  for  Green  and  Red  Line 
deposition  loading,  respectively,  if  runoff  is  between 
40-50%  of  precipitation.  If  runoff  is  60-70%  of 
precipitation,  the  Green  Line  deposition  is  6  kg  S/ha/yr 
and  Red  Line  is  1 1  kg  S/ha/yr.  That  is,  if  the  low  runoff 
lake  would  receive  a  total  of  <3  kg  S/ha  including 
deposits  from  the  new  source,  the  pH  of  the  lake  would 
not  likely  drop  below  5.8.  The  AQRV  would  not  be 
adversely  impacted,  and  the  recommendation  to  the 
state  regulatory  agency  would  be  for  permit  approval.  If 
the  low  runoff  lake  would  receive  a  total  of  >5.5  kg 
S/ha/yr,  including  deposits  from  the  new  source,  the 
pH  of  the  lake  would  certainly  drop  below  5.8  and 
probably  below  5.0.  The  AQRV  would  be  adversely 
impacted,  and  the  recommendation  to  the  state 
regulatory  agency  would  be  for  permit  denial  or  permit 
modification,  to  reduce  deposition  from  the  new  source 
to  levels  that  would  not  adversely  affect  the  lake. 

Total  deposition,  including  that  from  the  proposed 
new  source,  between  3  and  5.5  kg  S/ha-yr  would  have 
uncertain  effects  on  the  lake  pH.  The  assessment 
would  then  require  additional  site-specific  information 
indicating  physical,  chemical,  and/or  biological 
response  to  sulfur  input. 


Information  Needs 

To  use  the  Aquatic  Graph  (fig.  1),  the  following 
information  is  needed: 

o  Distribution  of  cations,  anions,  ANC,  and  pH  in 
wilderness  lakes  and  streams,  collected  after  spring 
runoff  has  receded. 


14 


o   Estimates    of    annual    runoff    for  watersheds 
containing  low  base  cation  (sensitive)  systems. 

o   Estimates  of  background  annual  average  (wet  plus 
dry)  sulfur  and  nitrogen  deposition. 

o   Estimated  total  S  and  N  deposition  based  on 
modeling  of  the  proposed  new  source  emissions. 


Cautions 

The  aquatic  Green/Red  model  developed  at  this 
workshop  is  based  on  an  empirical  relationship 
involving  a  large  number  of  lakes  and  streams.  These 
aquatic  systems  differ  in  watershed  biogeochemistry 
and  hydrology,  and  in  their  specific  response  to 
incremental  additions  of  sulfur  or  nitrogen  loading.  The 
loadings  themselves  are  based,  in  most  cases,  on 
estimated  atmospheric  values  developed  from  models 
that  use  regional  assumptions  and  "rules  of  thumb." 
Empirical  models  are  best  used  as  a  screening 
technique  to  estimate  the  probability  of  a  water  body  or 
group  of  water  bodies  responding  to  a  given  sulfur  or 
nitrogen  deposition  rate  in  cases  where  minimal  data 
are  available.  The  terrestrial  Green/Red  Line  model  is 
equally  approximate. 

Empirical  models  are  not  able  to  predict  exact 
results  in  any  specific  ecosystem.  Because  all  of  the 
assumptions  in  this  report  are  conservative,  a  loading 
value  below  the  Green  Line  has  a  low  probability  of 
causing  negative  effects  on  AQRV.  However,  there 
remain  some  sources  of  error  that  would  cause  an 
underestimate  of  the  potential  for  wilderness 
acidification: 

0  Failure  to  include  the  most  sensitive  lakes  or 
streams. 

°  Overestimation  of  average  annual  runoff. 

0   Underestimation  of  background  S  or  N  deposition. 

0  Underestimation  of  nitrogen  assimilation  and 
storage  by  watershed  vegetation  and  litter. 

°  Episodic  acidification  due  to  acidic  snowmelt  or 
storm  events  (primarily  in  streams). 

0  Higher  than  normal  initial  concentrations  of  SO4  in 
lakes  from  natural  sources  other  than  marine 
sources. 

0  Failure  to  correct  lake  cation  values  for  marine 
influence  or  for  other  geological  sources  of  CI  and 
associated  anions. 

The  nitrogen  loading  data  also  do  not  consider 
possible  effects  of  increased  nitrogen  on  eutrophication 
(algal  growth)  and  consequent  low  dissolved  oxygen 
content  in  lakes. 


If  a  loading  appears  above  the  Green  Line,  the 
graph  indicates  that  the  lake  or  stream  may  experience 
a  pH  below  5.8.  The  following  factors  lead  to  an 
overestimation  of  the  effects  of  the  predicted  future 
loadings. 

o  Overestimation  of  background  sulfur  loadings  due 
to: 

--a  large  component  of  alkaline  sulfate  dust, 
--overestimation  of  background  dry  deposition 
rates. 

0   Overestimation  of  background  nitrogen  loading  due 
to: 

-overestimation  of  dry  deposition  rates, 
-underestimation  of  nitrogen  assimilation  by 
watershed  vegetation. 

o   Delayed  response  to  loadings  because  of: 

-high  sulfate  adsorption  capacity  of  watershed 
soils. 

-higher  than  average  background  weathering 
rates. 

If  a  loading  falls  above  the  Green  Line  criterion  value, 
the  manager  should  request  data  from  a  proponent  as 
part  of  the  PSD  permit  to  determine  if  one  or  more  of 
the  above  cases  may  apply.  Determinations  would 
involve  deposition  chemistry  measurements  (including 
dry  deposition),  watershed  element  budgets,  analyses 
of  watershed  soils,  and  watershed  simulation  models. 
Such  studies  should  be  suggested  or  approved 
following  consultation  with  scientists. 


LITERATURE  CITED 

Barnes,  R.  L.  1972.  Effects  of  chronic  exposure  to 
ozone  on  photosynthesis  and  respiration  of  pines. 
Environ.  Pollut.  18:  45-50. 

Blair,  R.  1986.  Forest  response  program  QA  methods 
manuals.  Manual  1:  Exposure  systems  and 
physiological  measurements  by  L.  S.  Evans  and  P. 
Dougherty.  Manual  2:  Site  classification  and  field 
measurements  by  S.  M.  Zedaker  and  N.  S. 
Nicholas.  Manual  3:  Laboratory  analytical 
techniques  by  W.  P.  Robarge  and  I.  Fernandez. 
Manual  4:  Experimental  design  and  data 
management  by  C.  F.  Baes,  R.  J.  Olson,  K.  Ritters. 
EPA/600/x-86/193abcd,  U.S.  EPA/Corvallis, 
Oregon,  Internal  Report. 

Bormann,  F.  H.;  Likens,  G.  E.  1979.  Pattern  and 
process  in  a  forested  ecosystem.  New  York: 
Springer-Verlag.  253  p. 

Brezonik,  P.  L.;  Baker,  L.  A.;  Eaton,  J.  R.;  and  others. 
1986.  Experimental  acidification  of  Little  Rock  Lake, 
Wisconsin.  In:  Martin,  H.  O,  ed.  Acidic  precipitation: 


15 


Proceedings  of  the  international  symposium  on 
acidic  precipitation,  1985  September  15-20; 
Muskoga,  Ontario:  115-122. 

Brown,  J.  C;  Skau,  C.  M.  1975.  Chemical  composition 
of  snow  in  the  East-Central  Sierra  Nevada.  Center 
for  Water  Resources  Research,  Desert  Research 
Inst.,  Reno,  Nevada.  Publ.  AG-1 .  13  p. 

Confer,  J.;  Kaaret,  T.;  Likens,  G.  1983.  Zooplankton 
diversity  and  biomass  in  recently  acidified  lakes. 
Can.  J.  Fish.  Aquat.  Sci.  40:  36-42. 

Eilers,  J.  M.;  Kanciruk,  P.;  McCord,  R.  A.;  and  others. 
1987.  Characteristics  of  lakes  in  the  western  United 
States.  Volume  II:  Data  compendium  for  selected 
physical  and  chemical  variables.  EPA-600/3- 
86/054b,  U.S.  Environmental  Protection  Agency, 
Washington,  D.C.  425  p. 

Eilers,  J.  M.;  Lien,  G.  J.;  Bag,  R.  G.  1984.  Aquatic 
organisms  in  acidic  environments:  a  literature 
review.  Tech.  Bulletin  150,  Dep.  Nat.  Resour.,  P.O. 
Box  7921 ,  Madison,  Wl  53707.  18  p. 

EPRI.  1986.  An  evaluation  and  compilation  of  the 
reported  effects  of  acidification  on  aquatic  biota. 
EA-4825.  Research  Proj.  2346-1.  EPRI,  3412 
Hillview  Ave.,  Palo  Alto,  CA  94304. 

Fehsenfeld,  F.  C;  Parrish,  D.  D.;  Fahey,  D.  W.  1988. 
The  measurement  of  NOx  on  the  non-urban 
troposphere.  In:  Isaksen,  J.S.A.  (ed.).  Tropospheric 
Ozone.  D.  Reidel:  185-215. 

Fox,  D.  G.;  Bernabo,  J.  C;  Hood,  B.  1987.  Guidelines 
for  measuring  the  physical,  chemical,  and  biological 
condition  of  wilderness  ecosystems.  USDA  Forest 
Service,  General  Technical  Report  RM-146.  Rocky 
Mountain  Forest  and  Range  Experiment  Station. 
Fort  Collins,  CO.  48  p. 

Grennfelt,  P.;  Hultberg,  H.  1986.  Effects  of  nitrogen 
deposition  on  the  acidification  of  terrestrial  and 
aquatic  ecosystems.  Water,  Air,  and  Soil  Pollut.  30: 
945-963. 

Haines,  T.  A.  1981.  Acidic  precipitation  and  its 
consequences  for  aquatic  ecosystems:  a  review. 
Trans.  Am.  Fish.  Soc.  110(6):  669-707. 

Hall,  R.  J.;  Driscoll,  C.  T.;  Likens,  G.  E.  1987. 
Importance  of  hydrogen  ions  and  aluminium  in 
regulation  the  structure  and  function  of  stream 
ecosystems:  an  experimental  test.  Freshwater  Biol. 
18:  17-43. 

Hall,  R.  J.;  Driscoll,  C.  T.;  Likens,  G.  E.;  Pratt,  J.  M. 
1985.  Physical,  chemical,  and  biological 
consequences  of  episodic  aluminum  additions  to  a 
stream.  Limnol.  Oceanogr.  30(1):  212-220. 

Hall,  R.  J.;  Likens,  G.  E.  1981.  Chemical  flux  in  an 
acid-stressed  stream.  Nature  292:  329-331. 

Hall,  R.  J.;  Likens,  G.  E.  1984.  Effect  of  discharge  rate 
on  biotic  and  abiotic  chemical  flux  in  an  acidified 
stream.  Can.  J.  Fish  Aquat.  Sci.  41:  1132-1138. 


Hall,  R.  J.;  Likens,  G.  E.;  Fiance,  S.  B.;  Hendry,  G.  R. 
1980.  Experimental  acidification  of  a  stream  in  the 
Hubbard  Brook  Experimental  Forest.  Ecology  61(4): 
976-989. 

Hayes,  E.  M.;  Skelly,  J.  M.  1977.  Transport  of  ozone 

from  the  northeastern  United  States  into  Virginia 

and  its  effect  on  eastern  white  pines.  Plant  Dis. 

Rep.  61 :  778-782. 
Hem,  J.  D.  1970.  Study  and  interpretation  of  the 

chemical     characteristics     of     natural  water. 

Geological  Survey   Water-Supply   Paper  1473, 

USDI,  Geological  Survey,  Washington,  D.C.  1 1  p. 
Henriksen,  A.  1979.  A  simple  approach  for  identifying 

and  measuring  acidification  of  freshwater.  Nature 

278:  542-545. 
Henriksen,   A.    1984.    Changes    in    base  cation 

concentrations  due  to  freshwater  acidification.  Verb. 

Internat.  Verein.  Limnol.  22:  692-698. 
Henriksen,  A.;  Brakke,  D.  F.  1988.  Sulfate  deposition 

to  surface  waters:  estimating  critical  loads  for 

Norway  and  the  eastern  United  States.  Environ. 

Sci.  Technol.  22(1):  8-14. 
Husar,  R.  B.  1986.  Emissions  of  sulfur  dioxide  and 

nitrogen  oxides  and  trends  for  eastern  North 

America.  In:  Acid  deposition:  long-term  trends. 

Washington,  D.C:  National  Academy  Press:  48-92. 
Kanciruk,  P.;  Eilers,  J.  M.;  McCord,  R.  A.;  and  others. 

1986.  Characteristics  of  lakes  in  the  eastern  United 
States.  Volume  III:  Data  compendium  of  site 
characteristics  and  chemical  variables.  EPA/600/4- 
86/007c,  U.S.  Environmental  Protection  Agency, 
Washington,  D.C.  374  p. 

Karnosky,  D.  F.;  Steiner,  K.  C.  1981.  Provenance  and 
family  variation  in  response  of  Fraxinus  americana 
and  F.  pennsylvanica  to  ozone  and  sulfur  dioxide. 
Photopath  71 : 804-807. 

Laird,  L.  B.;  Taylor,  H.  E.;  Kennedy,  V.  C.  1986.  Snow 
chemistry  of  the  Cascade-Sierra  Nevada 
Mountains.  Environ.  Sci.  Technol.  20:  275-290. 

Landers,  D.  H.;  Eilers,  J.  M.;  Brakke,  D.  F.;  and  others. 

1987.  Characteristics  of  lakes  in  the  western  United 
States.  Volume  I:  Population  descriptions  and 
physico-chemical  relationships.  EPA-600/3- 
86/054a.  U.S.  Environmental  Protection  Agency, 
Washington,  D.C.  176  p. 

Likens,  G.  E.  (ed.).  [In  press].  Long-term  studies  in 
ecology.  New  York:  Springer-Verlag.  In  Press. 

Likens,  G.  E.;  Bormann,  F.  H.;  Johnson,  N.  M.;  and 
others.  1970.  Effects  of  forest  cutting  and  herbicide 
treatment  on  nutrient  budgets  in  the  Hubbard  Brook 
watershed  ecosystem.  Ecol.  Monogr.  40(1):  23-47. 

Linthurst,  R.  A.,  Landers,  D.  H.;  Eilers,  J.  M.;  and 
others.  1986.  Characteristics  of  lakes  in  the  eastern 
United  States.  Volume  I:  Population  descriptions 
and  physico-chemical  relationships.  EPA  600/4-86- 


16 


007A,  U.S.  Environmental  Protection  Agency, 
Washington,  D.C.  136  p. 

Liu,  S.  C;  Trainer,  M.;  Fehsenfeld,  F.  C;  and  others. 
1987.  Ozone  production  in  the  rural  troposphere 
and  the  implications  for  regional  and  global  ozone 
distribution.  J.  Geophys.  Res.  92(D4):  4191-4207. 

Loranger,  T.  J.  1986.  Temporal  variability  of  water  and 
snowpack  chemistry  in  the  North  Cascade 
Mountains.  M.S.  Thesis,  Western  Washington 
University,  Bellingham,  WA.  140  p. 

Loranger,  T.  J.;  Brakke,  D.  F.  1988.  The  extent  of 
snowpack  influence  on  water  chemistry  in  a  North 
Cascades  Lake.  Water  Resour.  Res.  24(5):  723- 
726. 

Melack,  J.  M.;  Stoddard,  J.  L;  Ochs,  C.  A.  1985.  Major 
ion  chemistry  and  sensitivity  to  acid  precipitation  of 
Sierra  Nevada  lakes.  Water  Resour.  Res.  21:  27- 
32. 

Miller,  P.  R.  1973.  Oxidant-induced  community  change 
in  a  mixed  conifer  forest.  In:  Advan.  Chem.  Ser.  No. 
122:  101-117. 

Musselman,  R.C.;  McCool,  P.M.;  Younglove,  T.  1988. 
Selecting  ozone  exposure  statistics  for  determining 
crop  yield  loss  from  air  pollutants.  Environ.  Pollu. 
53:  63-78. 

National  Atmospheric  Deposition  Program.  1988. 
NADP/NTN  annual  data  summary,  precipitation 
chemistry  in  the  United  States  1987.  NADP, 
Colorado  State  University,  Ft.  Collins,  CO  80523. 
340  p. 

Pye,  John.  1988.  Impact  of  ozone  on  the  growth  and 
yield  of  trees:  a  review.  J.  Environ.  Qual.  17(3): 
347-360. 

Reddy,  M.  M.;  Classen,  H.  C.  1985.  Estimates  of 
average  major  ion  concentrations  in  bulk 
precipitation  at  two  high-altitude  sites  near  the 
continental  divide  in  southwestern  Colorado.  Atmos. 
Environ.  19(7):  1 199-1203. 

Reich,  P.B.  1987.  Quantifying  plant  response  to  ozone: 
a  unifying  theory.  Tree  Physiol.  3:  63-91. 

Reuss,  J.  O.;  Christopherson,  N.;  Seip,  H.  M.  1986.  A 
critique  of  models  for  freshwater  and  soil 
acidification.  Water,  Air  and  Soil  Pollut.  30:  909- 
930. 


Riggan,  P.J.;  Lockwood,  R.N.;  Lopez,  E.N.  1985. 
Deposition  and  processing  of  airborne  nitrogen 
pollutants  in  Mediterranean-type  ecosystems  of 
Southern  California.  Environ.  Sci.  Technol.  19(9): 
781-789. 

Schindler,  D.  W.  1986.  The  significance  of  in-lake 
production  of  alkalinity.  Water,  Air  and  Soil  Pollut. 
30:931-944. 

Schindler,  D.  W.  1988.  Effects  of  acid  rain  on 
freshwater  ecosystems.  Science  239: 149-157. 

Schindler,  D.  W.;  Mills,  K.  H.;  Malley,  D.  F.;  and  others. 
1985.  Long-term  ecosystem  stress:  the  effects  of 
experimental  acidification  on  a  small  lake.  Science 
228:  1395-1401. 

Small,  M.  J.;  Sutton,  M.  C;  Milke,  M.  W.  1988. 
Parametric  distributions  of  regional  lake  chemistry: 
fitted  and  derived.  Environ.  Sci.  Tech.  22: 196-204. 

U.S.  Department  of  Agriculture,  Forest  Service.  1987. 
Air  Resource  Management  Handbook,  FSH 
2509.19.  Washington,  D.C. 

U.S.  Environmental  Protection  Agency.  1986a.  Air 
quality  criteria  for  ozone  and  other  photochemical 
oxidants  (Vols.  I-V).  Environmental  Protection 
Agency,  Research  Triangle  Park,  NC.  EPA-600/8- 
84-20aF.  Vol.  1  =  235  p. 

U.S.  Environmental  Protection  Agency.  1986b. 
Guideline  on  air  quality  models  (revised).  EPA- 
450/2-78-027R.  Office  of  Air  Quality  Planning  and 
Standards,  Research  Triangle  Park,  NC  27711. 
290p. 

Vertucci,  F.  A.  [In  press].  Methods  of  detecting  and 
quantifying  lake  acidification.  In:  Proceedings, 
international  mountain  watershed  symposium,  1988 
June  8-10,  Lake  Tahoe,  NV. 

Wright,  R.  F.  1988.  Acidification  of  lakes  in  the  eastern 
United  States  and  southern  Norway:  a  comparison. 
Environ.  Sci.  Technol.  22:  178-182. 

Wright,  R.  F.;  Henriksen,  A.  1983.  Acid  precipitation: 
how  many  lakes  in  Norway  will  be  restored  by  a 
30%  reduction  in  sulfur  deposition?  Nature  305: 
422-424. 

Young,  J.R.;  Ellis,  E.L.;  Hidy,  G.M.  1988.  Deposition  of 
air-borne  acidifiers  in  the  water  and  environment.  J. 
Environ.  Qual.  17(1):  1-26. 


17 


18 


WASHINGTON 


1  Pasayten  505,524 

2  Glacier  Peak  464,258 

3  Alpine  Lakes  303,508 

4  Goat  Rocks  82,680 

5  Mt.  Adams  32,356 

OREGON 

6  Mt.  Hood  14,160 

7  Hells  Canyon  See  ID 

8  Eagle  Cap  293,476 

9  Mt.  Jefferson  100,208 

10  Strawberry  Mtn.  33,003 

11  Mt.  Washington  46,116 

12  Three  Sisters  199,902 

13  Diamond  Peak  36,637 

14  GearhartMtn.  18,709 

15  Kalmiopsis  76,900 

16  Mountain  Lakes  23,071 

CALIFORNIA 

17  Marble  Mtn.  213,743 

18  South  Warner  68,507 

19  Thousand  Lakes  15,695 

20  Caribou  19,080 

21  Yolla-Bolly-Middle  Eel  109,091 

22  Desolation  63,469 

23  Mokelumne  50,400 

24  Emigrant  104,311 

25  Hoover  47,916 

26  Minarets  (Now  Ansel  Adams)  109,484 

27  Kaiser  22,500 

28  John  Muir  484,673 

29  Ventana  95,152 

30  Dome  Land  62,206 

31  San  Rafael  142,722 

32  Cucamonga  9,022 

33  San  Gabriel  36,137 

34  San  Gorgonio  34,644 

35  San  Jacinto  20,564 

36  Agua  Tibia  15,934 

NEVADA 

37  Jarbridge  64,667 

IDAHO 

7  Hells  Canyon  193,840 

38  Sawtooth  216,383 

39  Selway-Bitterroot  1,240,618 

MONTANA 

39  Selway-Bitterroot  See  ID 

40  Anaconda-Pintler  157,803 

41  Gates-of-the  Mtn.  28,562 

42  Scapegoat  239,295 

43  Mission  Mountains  73,877 

44  Bob  Marshall  950,000 

45  Cabinet  Mountains  94,272 

WYOMING 

46  North  Absaroka  351,104 

47  Washakie  686,584 

48  Teton  557,31 1 

49  Fitzpatrick  191,103 

50  Bridger  392,160 

COLORADO 

51  Rawah  26,674 

52  Mt.  Zirkel  72,472 


53  Flat  Tops  235,230 

54  Eagles  Nest  133,910 

55  Maroon  Bells-Snowmass  71,060 

56  West  Elk  61,412 

57  La  Garita  48,486 

58  Weminuche  400,907 

ARIZONA 

59  Sycamore  Canyon  47,757 

60  Pine  Mtn.  20,061 

61  Mazatzal  205,137 

62  Mt.  Baldy  6,975 

63  Sierra  Ancha  20,850 

64  Superstition  124,117 

65  Galiuro  52,717 

66  Chiricahua  18,000 

NEW  MEXICO 

67  Gila  433,690 

68  White  Mtn.  31,171 

69  Pecos  167,416 

70  San  Pedro  Parks  41,132 

71  Wheeler  Park  6,027 

ARKANSAS 

72  Caney  Creek  14,344 

73  Upper  Buffalo  9,912 

MISSOURI 

74  Hercules-Glades  12,315 

MINNESOTA 

75  Boundary  Waters  Canoe  Area  747,840 

WISCONSIN 

76  Rainbow  Lake  6,388 

NEW  HAMPSHIRE 

77  Great  Gulf  5,552 

78  Presidential  Range-Dry  River  20,000 

VERMONT 

79  Lye  Brook  12,430 

W.  VIRGINIA 

80  Dolly  Sods  10,215 

81  Otter  Creek  20,000 

VIRGINIA 

82  James  River  Face  8,703 

N.  CAROLINA 

83  Linville  Gorge  7,575 

84  Joyce  Kilmer-Slickrock  14,033 

85  Shining  Rock  13,350 

TENNESSEE 

84       Joyce-Kilmer-Slickrock  See  NC 
GEORGIA 

86  Cohutta  33,776 

ALABAMA 

87  Sipsey  12,646 

FLORIDA 

88  Bradwell  Bay  23,432 


19 


APPENDIX  B.  BACKGROUND  OF  SAMPLE  CLASS  I 
AREAS 


These  descriptions  were  prepared  jointly  by 
scientists  and  managers  at  this  workshop.  The 
precision  of  these  descriptions  varies  because  the 
amount  of  information  available  to  workshop 
participants  was  different.  These  Class  I  areas  show 
the  breadth  of  AQRV's  considered  and  the  diversity  of 
approaches  suggested  by  participants  working 
together.  Terrestrial  values  given  each  area  may  be 
the  same  because  the  different  areas  contain 
ecosystems  with  similar  sensitivities,  such  as  alpine 
areas.  It  is  essential  to  consider  the  details  of  a  specific 
Class  I  area  being  screened  when  applying  information 
contained  here. 

Although  visibility  is  an  important  AQRV  in  all  these 
Class  I  areas,  this  workshop  focused  on  the  effects  of 
air  pollution  on  biotic  systems  and  did  not  address 
physical  impacts  on  visibility.  In  no  way  should  the 
absence  of  visibility  as  an  AQRV  be  construed  as  a 
judgment  of  its  relative  value  compared  to  biological 
components. 


Alpine  Lakes  and  Glacier  Peak  Wildernesses  - 
Washington 

Brief  Description 

The  Alpine  Lakes  and  Glacier  Peak  Wildernesses 
are  typical  of  the  North  Cascade  mountains.  The 
vegetation  is  fir,  Douglas-fir,  and  hemlock,  and 
precipitation  is  high.  Soils  are  diverse  in  origin  with 
modest  fertility  and  moisture.  This  is  a  high  mountain 
area  with  general  elevation  above  6,000  feet.  Lakes 
and  large  perennial  snow  fields  are  common  at  the 
higher  elevations.  Streams  peak  during  snowmelt 
runoff,  but  abundant  year  round  stream  flow  persists. 

Air  Quality  Related  Values 

Water  flowing  from  the  Cascade  crest  has 
significant  value.  The  hydrologic  system  includes 
snowfields,  glaciers,  high  mountain  streams,  small 
cirque  lakes,  cascading  waterfalls,  and  larger  streams 
and  rivers  in  lower  systems.  The  water  and  aquatic 
biota  systems  contribute  greatly  to  these  wildernesses. 
Maintaining  these  systems  and  their  natural  water 
clarity  depends  upon  little  chemical  degradation  or 
change. 

These  wildernesses  include  a  wide  variety  of 
diverse  plant  communities  and  species  typical  of  the 
northern  Cascade  range.  Maintaining  natural  diversity 


is  a  critical  component  of  general  health  and  balance  of 
the  ecosystem.  Any  significant  change  in  plant 
communities  due  to  the  effects  of  air  pollution  would 
not  only  change  the  quality  of  wilderness  experience, 
but  also  ecosystem  interrelationships  which  contribute 
to  wilderness  values. 

Much  of  the  wilderness  experience  in  the  Cascades 
is  influenced  by  sights,  sounds,  feelings,  experiences, 
and  even  the  smells  the  visitor  encounters.  In  areas 
close  to  metropolitan  areas,  one  of  the  significant 
changes  for  the  city  resident  is  the  smell  of  the  great 
out-of-doors.  Whether  it  is  a  whiff  of  pine  forest,  an 
aroma  of  rain  forest,  or  briskness  of  the  clean,  crisp 
high  mountain  air  rising  up  and  over  the  Cascade 
range,  natural  smell  is  a  value  only  truly  appreciated 
when  it's  replaced  with  the  odor  of  civilization. 


Hoover  and  Dome  Land  Wildernesses  -  California 
Brief  Description 

Hoover.  This  Wilderness  lies  along  the  eastern 
slope  of  the  Sierra  Nevada  Range  in  Mono  County, 
California.  It  is  bounded  on  the  west  by  Yosemite 
National  Park,  which  lies  on  the  western  slope  of  this 
Range.  The  area  is  characterized  by  recently  glaciated 
canyons,  composed  of  granitic  and  metavolcanic 
rocks.  The  vegetation  is  scattered  among  the  rocky 
flats  and  ledges  of  the  Sierra  granite  batholith. 

Most  soils  are  derived  from  granitic  rock,  are  weakly 
developed,  and  are  low  in  productivity.  They  are 
typically  shallow  over  granitic  parent  material  on  the 
sloping  areas,  and  are  deeper  in  the  canyon  bottoms. 
They  are  sandy  textured  and  low  base  saturation. 
Water  quality  is  good  to  excellent,  with  low  sediment 
loads  in  the  streams.  The  many  lakes  in  the  high 
country  act  as  natural  sinks  in  absorbing  sediments 
from  the  canyon  uplands. 

Scattered  stands  of  timber  grow  on  approximately 
11  percent  of  the  wilderness.  Timber  types  generally 
are  mixed  conifer,  with  Jeffery  pine  and  white  fir 
dominating  the  lower  elevations  and  lodgepole  pine 
and  limber  pine  dominating  the  higher  elevations. 
Subalpine  meadows  occur  throughout  the  area; 
riparian  areas  exist  along  the  stream  courses,  springs, 
and  other  water  influence  zones.  The  higher  elevations 
are  dominated  by  subalpine  and  alpine  shrubs  and 
herbaceous  vegetation,  while  elevations  below  the 
forest  zones  are  dominated  by  sagebrush,  bitterbrush, 
and  mountainmahogany. 


20 


Air  quality  within  the  wilderness  is  excellent.  Among 
potential  threats  is  the  possibility  for  NOx,  SOx,  and 
ozone  to  drift  up  the  Toulumne  valley,  flow  over  the 
crest,  and  influence  the  AQRV  of  the  area  and  acidify 
precipitation. 

Dome  Land:  The  Dome  Land  Wilderness  is  located 
on  the  southeastern  slopes  of  the  Sierra  Nevada 
Range  at  the  southern  end  of  the  Kern  Plateau. 
Granite  domes  and  unique  geologic  formations  are  the 
dominant  features.  Climatic  conditions  range  from 
montane  to  semi-arid  to  desert  with  elevations  from 
3,000  to  9,700  feet  above  sea  level.  The  South  Fork  of 
the  Kern  River  flows  through  the  wilderness. 

Dome  Land  geography  has  been  primarily 
influenced  by  the  South  Fork  Kern  River  drainage.  The 
wilderness  is  rimmed  by  high  elevation  peaks  in  a 
horseshoe  configuration.  Pollution  may  be  transported 
up  the  Kern  River  drainage  from  the  Bakersfield  area 
of  the  San  Joaquin  Valley. 

The  Dome  Land  is  covered  mainly  by  mixed  conifer 
forest.  The  higher  elevations  support  primarily 
lodgepole  and  Jeffrey  pine,  red  and  white  fir,  and  small 
amounts  of  oak  and  various  shrubs.  Small  stands  of 
limber  and  foxtail  pine  are  also  found  at  the  higher 
elevations.  A  unique  association  of  limber  and  foxtail 
pine  at  the  southern  most  ends  of  their  ranges  has 
resulted  in  the  establishment  of  a  research  natural 
area.  The  lower  elevations  support  mainly  pinyon, 
digger  pine,  oak,  and  shrubs. 

Wildlife  in  the  Dome  Land  is  abundant.  The 
wilderness  provides  summer  range  for  the  Monache 
and  Kern  River  deer  herds.  A  comprehensive  species 
list  is  lacking,  but  other  wildlife  observed  include  quail, 
squirrels,  chickaree,  chipmunk,  marten,  marmot,  black 
bear,  mountain  lion,  and  bobcat.  In  the  1970's 
California  condors  were  sighted  on  several  occasions. 

There  is  light  to  moderate  fishing  within  the 
wilderness.  The  heaviest  fishing  area  is  located  on  the 
South  Fork  of  the  Kern  River.  The  rainbow  trout  found 
in  the  wilderness  are  introduced. 

The  Dome  Land  contains  six  major  tributary 
streams  of  the  South  Fork  of  the  Kern  River.  The 
general  character  of  the  wilderness  is  dry  during  the 
normal  season  of  use,  so  these  streams  are  very 
important  to  visitors,  livestock,  and  wildlife.  Below  the 
wilderness,  water  from  the  South  Fork  is  used  for 
agriculture,  recreation,  electrical  power,  and  domestic 
supplies.  All  of  the  Kern  tributaries  in  the  wilderness 
drop  to  a  low  level  or  become  dry  in  late  summer. 
Water  becomes  quite  warm  in  creeks  still  flowing. 

Soils  within  the  Dome  Land  are  derived  form 
weathered  granite.  Most  of  the  soil  consists  of  coarse, 
sandy  materials  that  have  weathered  from  the  barren, 
exposed  rock  that  dominates  the  wilderness.  These 
soils  are  very  young,  and  lack  the  development 
characteristic  of  older  soils.  They  are  very  infertile  due 


to  coarseness,  shallowness,  and  lack  of  capacity  to 
store  water.  The  soils  are  also  susceptible  to  erosion. 


Air  Quality  Related  Values 

Jeffery  and  ponderosa  pines  are  prevalent  above 
5,000  feet.  These  sensitive  tree  species  are  subject  to 
damage  by  ozone,  and  can  be  used  as  indicators  of 
changes  in  plant  communities.  Needle  retention  and 
natural  color  are  needed  to  maintain  the  aesthetics  of 
these  wildernesses.  Limber,  foxtail,  and  pinyon  pine 
are  also  important  vegetative  species. 

The  buffering  effect  of  meadows  on  water  quality 
and  quantity  makes  these  areas  very  valuable  for 
protecting  the  Class  I  areas'  aquatic  systems, 
especially  the  rainbow  trout  fisheries.  Meadow 
condition  and  water  quality  should  be  maintained  within 
current  biological  variability.  Water  quality  needs  to  be 
maintained  in  the  river  and  its  tributaries. 

The  selected  loadings  for  these  wildernesses  are 
lower  than  in  some  other  wildernesses  because  of  the 
presence  of  alpine  ecosystems  with  limited  ability  to 
buffer  additional  S  and  N.  The  desire  to  maintain 
current  ecosystem  structure  and  function  were  primary 
considerations  in  the  selection  of  threshold  values. 

The  granitic  domes  characteristic  in  this  wilderness 
should  be  protected. 


San  Gorgonio  Wilderness  -  California 

Brief  Description 

Geology  of  this  wilderness  is  highly  diverse  and 
typical  of  southern  California  mountains.  Climate  is 
Mediterranean,  and  the  soils  are  dry  with  base 
saturations  about  50%.  Water  resources  are  scarce. 
Vegetation  varies  with  elevation  from  chaparral  through 
a  pinyon-juniper  and  pine  forest  to  alpine. 


Air  Quality  Related  Values 

Ponderosa  pine  and  Jeffery  pine  are  dominant 
species  known  to  be  susceptible  to  air  pollutants, 
especially  O3.  Symptoms  of  ozone  injury  (needle 
chlorosis  and  premature  needle  senescence)  can  be 
readily  identified.  The  ponderosa  pine  forests  in 
southern  California  often  have  high  concentrations  of 
pollutants  present.  West  of  the  San  Gorgonio 
Wilderness,  ozone  concentrations  are  high  from  May 
through  September,  with  moderate  concentrations  at 
other  times.  Nitrogen  deposition  is  very  high,  and 
although  poorly  quantified,  may  be  an  important 
component  of  the  ecosystem  (Riggan  et  al.  1985). 


21 


Water  quality  is  important  as  it  relates  to  pH,  ANC, 
and  productivity.  The  pH  and  ANC  values  are  related 
to  changes  in  acidity,  which  affect  chemical  processes 
and  ultimately  biological  processes.  Productivity  is  a 
general  term  that  refers  to  the  amount  of  carbon  fixed 
on  an  annual  basis;  more  N-rich  systems  are  generally 
more  productive.  Productivity  should  be  maintained. 

Meadows  are  critical  areas  to  maintain  in  subalpine 
and  alpine  ecosystems. 

Bob  Marshall  Wilderness  -  Montana 
Brief  Description 

The  Bob  Marshall  Wilderness  is  nearly  one  million 
acres  in  size,  located  in  northwest  Montana  in  the 
Rocky  Mountain  Province.  The  bedrock  is  mostly 
precambrian  meta-sedimentary  argillites,  quartzites, 
and  limestones.  Glaciation  influenced  the  shape  of  the 
land  and  the  composition  of  the  soil.  Soils  are  cool, 
moist,  with  base  saturation  of  25  to  50%  with  a 
volcanic  ash  surface  ranging  from  4  to  8  inches.  The 
terrain  has  been  influenced  by  glaciation,  which  formed 
high  alpine  basins  and  broad  u-shaped  valleys. 

Precipitation  ranges  from  16  inches  in  the  valley 
bottoms  to  more  than  100  inches  on  the  mountain 
peaks.  Snow  comprises  over  80  percent  of  the 
precipitation  at  the  higher  elevations  and  50  percent  in 
the  valley  bottoms.  Elevation  within  the  area  varies 
from  3,000  feet  in  the  valleys  to  nearly  10,000  feet  at 
the  highest  peaks. 

Habitat  types  range  from  warm-dry  ponderosa 
pine/bunchgrass  to  cool-moist  whitebark  pine. 
Subalpine  fir  is  the  dominant  habitat  type.  The  country 
is  known  for  a  mixture  of  big,  open  meadows  and 
dense  forest.  Uncontrolled  natural  fire  played  a  large 
part  in  producing  a  mosaic  of  different  even-aged 
communities. 

About  250  wildlife  species  and  22  fish  species  are 
found  in  the  wilderness  and  surrounding  national  forest 
lands.  Native  fish  species  include  bull  trout,  west  slope 
cutthroat  trout,  mountain  whitefish,  and  several  non- 
game  species.  Big  game  species  include  elk,  mule 
deer,  white-tailed  deer,  moose,  Rocky  Mountain  goat, 
grizzly  bear,  black  bear,  and  cougar.  Endangered 
species  include  the  gray  wolf,  bald  eagle,  and 
peregrine  falcon.  Threatened  species  include  the 
grizzly  bear. 

Lakes  and  streams  are  common  and  dependent  on 
snowmelt.  The  Middle  Fork  and  South  Fork  of  the 
Flathead  River  flow  out  of  the  Bob  Marshall.  These 
rivers  are  designated  Wild  and  Scenic  and  are 
important  for  rafting,  fishing,  photography,  and 
domestic  and  other  consumptive  needs.  Water  quality 
is  considered  excellent,  although  water  quality  has  not 
been  extensively  sampled. 


Air  Quality  Related  Values 

Grizzly  bear  and  west  slope  cutthroat  trout  are  the 
key  air  quality  related  values  in  this  wilderness.  Effects 
of  air  pollutants  on  forage  species  and  other  critical 
grizzly  habitat  plant  communities  and  on  meadow 
vegetation  that  could  change  trout  habitat  must  to  be 
determined.  This  wilderness  provides  one  of  only  two 
major  grizzly  bear  population  centers  in  the  lower  48 
states.  The  cutthroat  is  classified  as  a  species  of 
special  concern  in  Montana  because  of  declines  in 
abundance  and  distribution.  It  is  important  to  the 
wilderness  visitor  for  both  consumptive  and  non- 
consumptive  uses. 

Alpine  and  subalpine  plant  communities  were 
thought  to  be  the  most  sensitive  to  increases  in  N, 
because  they  are  naturally  stressed  ecosystems  and 
are  likely  to  be  naturally  low  N-consuming  systems. 
The  Bob  Marshall  Wilderness  is  in  a  very  clean  air 
region.  Current  deposition  rates  for  N  and  S  are 
probably  1  kg/ha-yr  or  less  for  each  of  these  elements. 
Therefore,  increases  of  N  and  S  could  represent  large 
percentage  increases  in  the  quantity  of  these 
elements.  This  implies  that  tolerable  increase  levels 
are  likely  to  be  in  the  low  end  of  the  Yellow  Zone. 


Bridger  Wilderness  -  Wyoming 

Brief  Description 

The  Bridger  Wilderness  is  located  on  the  west  side 
of  the  continental  divide  in  the  Wind  River  Mountain 
Range.  The  elevation  ranges  from  about  8,000  to  over 
13,000  feet  on  Gannet  Peak,  with  most  of  wilderness 
above  9,000  feet.  Almost  all  of  the  area  is  precambrian 
crystalline  granite  except  for  a  small  section  of 
sedimentary  rock  in  the  northwest  part.  The  area  was 
glaciated  in  the  past,  and  still  contains  the  largest 
glaciers  in  the  continental  United  States.  Lakes  are 
very  common  (roughly  1 ,300)  and  have  been  stocked 
since  1907  with  all  major  species  of  trout  found  in 
North  America.  Since  a  large  portion  of  the  wilderness 
is  above  timberline,  the  vegetation  is  primarily  alpine 
and  subalpine  in  character.  Precipitation  is  primarily 
snow,  and  the  annual  snow  pack  is  deep.  The  soils  are 
cold,  wet,  and  shallow  with  base  saturation  below  25%. 
Granite  or  quartz  rock  outcrops  and  talus  slopes  are 
common.  Perennial  streams  are  fed  by  snowmelt. 
Groundwater  flow  is  minimal. 


Air  Quality  Related  Values 

This  wilderness  was  originally  designated  as  a 
Primitive  Area  in  1930  because  of  its  unique  alpine 


22 


ecosystem,  with  numerous  cirque  lakes.  These  lakes 
are  the  primary  AQRV  needing  protection. 

Because  of  the  large  amount  of  alpine  vegetation, 
this  area  is  potentially  very  sensitive  to  the  effects  of 
increased  nitrogen  deposition.  The  harsh  climate, 
shallow  soils,  and  presumed  low  nitrogen  uptake  rates 
of  the  alpine  plants  suggest  significant  changes  in 
growth  rates  and  species  composition  under  conditions 
of  even  small  atmospheric  N  deposition.  The  problem 
would  be  exacerbated  because  the  exposed  bedrock  in 
the  watersheds  will  focus  large  amounts  of  deposition 
into  small  areas  of  alpine  meadow. 

The  effects  of  air  pollution  on  alpine  vegetation  are 
not  well  known,  and  the  interaction  of  pollutants  with 
the  other  severe  stresses  acting  on  alpine  vegetation 
make  the  problem  especially  complex.  To  improve  the 
knowledge  base,  the  response  of  species 
characteristic  of  this  wilderness  to  ozone  exposure  and 
N  and  S  loading  should  be  determined.  Such 
determinations  should  be  performed  in  natural  field 
settings  which  incorporate  the  rigor  of  the  alpine 
environment.  Plant  communities  of  special  concern  are 
the  primary  successional  plant  communities  near 
glacial  margins.  The  chemistry  and  hydrology  of  the 
snowpack  needs  special  attention  because  of  its 
crucial  role  in  maintaining  the  diverse  plant 
communities. 


palo  verde,  saguaro  cactus,  cholla  cacti,  ocotillo, 
catclaw,  beargrass,  agave,  yucca,  mesquite, 
mountainmahogany,  hopbush,  turbinella  and  Emory 
oaks,  pinyon  pine,  and  junipers. 


Air  Quality  Related  Values 

Water  is  scarce  in  the  Superstitions  and  its 
availability  and  quality  are  critical  to  sustaining  the 
diverse  faunal  populations,  as  well  as  providing  water 
for  recreationists. 

Riparian  species  are  important  to  the  visual  quality 
of  this  unique  wilderness,  as  well  as  furnishing  perhaps 
one  of  the  most  valuable  wildlife  habitats  found  in  the 
Upper  Sonoran  Desert.  The  ability  of  the  wilderness  to 
support  the  diverse  wildlife  species  found  here  would 
be  greatly  diminished  without  riparian  areas.  Riparian 
species  include  cottonwood,  willow,  sycamore,  and 
numerous  others. 

Both  the  number  and  uniqueness  of  Upper  Sonoran 
Desert  plants  give  this  wilderness  its  special  character. 
To  lose  any  of  these  species  would  be  a  serious  loss  to 
the  wilderness.  Vegetation  in  this  type  is  thought  to  be 
quite  resistant  to  environmental  stress,  except  that 
vegetation  growing  in  riparian  areas  may  be  more 
sensitive  to  O3,  SO2,  and  other  pollutant  effects. 


Superstition  Wilderness  -  Arizona 
Brief  Description 

The  Superstition  Wilderness  is  located  south  of  the 
Mazatzal  Mountains  about  65  miles  east  of  Phoenix. 
Elevation  is  approximately  1,000  to  4,000  feet.  The 
rugged,  dissected  landscape  that  rises  spectacularly 
out  of  the  desert  has  deep  canyons  with  steep  sonoran 
relief.  Streams  are  ephemeral,  and  there  are  no  lakes. 
Hydrographs  are  storm-dominated. 

Climate  is  warm  semi-arid  and  arid,  with  summer 
convection  storms  and  occasional  winter  rain.  Annual 
precipitation  is  10  to  20  inches,  but  can  vary  as  much 
as  40  percent  annually.  The  growing  season  is  about 
280  days.  Average  annual  temperature  is  60  to  75°F. 

Soils  are  deep,  dry  forest  soils  with  base 
saturations  above  50%.  The  geology  includes  highly 
diverse  rock  types  and  complex  geological  structures, 
including  metamorphic,  sedimentary,  and  intrusive  and 
extrusive  igneous  rocks.  The  east  half  is  proterozoic 
rocks  that  have  been  pervasively  faulted.  The  west  half 
is  tertiary  volcanic  rocks  of  many  different  types. 

Vegetation  is  typically  open,  with  sonoran  desert 
shrubs  at  lower  elevations  to  interior  chaparral  and 
juniper  woodland  at  higher  elevations.  Upland  plants 
include  grama  grasses,  creosote  bush,  yellow  and  blue 


Joyce  Kilmer  -  Slickrock  Wilderness  -  North 
Carolina,  Tennessee 

Brief  Description 

Cove  and  upland  hardwoods  are  the  dominant 
forest  types  typical  of  this  warm,  humid  climate  that 
has  abundant,  uniform  precipitation.  Soils  range  from 
deep,  moist,  and  well-developed  to  shallow,  poorly 
developed,  and  with  base  saturation  less  than  25%. 
The  low  mountains  underlain  by  sedimentary  geology 
vary  from  2,000  to  5,300  feet  in  elevation.  Intermittent 
and  perennial  mountain  streams  are  common. 


Air  Quality  Related  Values 

Flora,  water  quality,  and  trout  fisheries  all  had 
important  roles  in  the  designation  of  the  Joyce  Kilmer- 
Slickrock  Wilderness,  and  continue  to  be  important 
characteristics  of  the  wilderness,  and  the  experiences 
valued  by  visitors. 

The  floral  diversity  is  great,  with  more  than  60 
species  of  trees.  Most  of  the  flora  is  of  tertiary  origin, 
and  a  number  of  plant  species  are  close  relatives  to 
species  in  eastern  Asia.  Wildflowers  are  abundant 
throughout  the  wilderness.  The  Joyce  Kilmer  Memorial 


23 


Forest  portion  of  the  wilderness  has  many  trees  over 
300  years  old,  some  more  than  20  feet  in 
circumference  and  100  feet  tall.  This  part  of  the 
wilderness  is  a  remnant  virgin  forest  preserved  by  the 
Forest  Service  since  1935.  Approximately  30%  of  the 
Slickrock  portion  of  the  wilderness  is  also 
representative  of  a  forest  in  its  primeval  condition. 

The  high-quality  mountain  streams  found  in  the 
wilderness,  free  of  sediment  with  clean  bottoms,  cool 
and  clear,  with  deep  pools  and  numerous  riffles,  are 
rare  in  this  part  of  the  country.  These  streams  have 
been  rated  by  the  State  of  North  Carolina  as  type  "C 
Trout"  and  have  met  State  standards  for  use  as  a 
public  water  supply.  Slickrock  Creek  is  a  highly 
productive  trout  stream  yielding  about  twice  as  much 
poundage  per  acre  as  neighboring  streams.  "Native" 
(reproducing  naturally  in  the  stream)  brook  trout  are 
abundant  in  the  upper  reaches  of  Slickrock  Creek. 
Brown  and  rainbow  trout  are  prominent  in  the  lower 
reaches.  Little  Santeelah  Creek  and  its  tributaries  are 
habitat  for  brown,  brook,  and  rainbow  trout.  The  trout 
fisheries  in  these  streams  represent  a  major  recreation 
opportunity  in  the  wilderness. 

The  location  of  the  Slickrock  area  in  the  southern 
Appalachians  and  the  locations  of  some  portions  of  the 
area  at  elevations  above  4,800  feet  suggest  that  parts 
of  this  system  currently  receive  relatively  high  loadings 
of  S  and  N,  as  well  as  high  concentrations  of  O3 
(NADP  1988).  Any  added  loading  due  to  new  sources 
will  move  pollutant  levels  into  the  Yellow  Zone  or 
above  the  Red  Line  where  granting  of  PSD  Permits  is 
not  automatic.  At  the  high-elevation  sites,  minor 
loadings  may  move  pollutant  levels  into  values  above 
the  Red  Line. 

Slickrock's  diverse  forest  systems  of  high  and  low 
elevations  require  that  the  target  loading  values  (Green 
Line  values)  cover  a  somewhat  higher  range  than  other 
wilderness  areas.  Over  90%  of  the  area  is 
characterized  by  a  capacity  to  utilize  higher  loadings  of 
N  because  of  deep,  well-developed  soils  of  moderate 
sulfate  absorption  capacity  that  can  tolerate  higher  S 
loadings.  The  remaining  10%  of  the  area  is  boreal 
forest,  which  receives  higher  loadings  due  to 
elevational  effects,  and  has  soils,  tree  species,  and 
age  classes  of  trees  sensitive  to  low  loadings  of  S  and 
N.  The  effect  of  the  loss  of  the  relatively  small  area  of 
high-elevation  boreal  forest  on  downslope  ecosystems 
is  currently  unknown,  but  hydrologic  and  chemical 
disturbances  might  result. 

Otter  Creek  -  West  Virginia  and  Great  Gulf  -  New 
Hampshire  Wildernesses 

Brief  Description 

Otter  Creek:  This  20,000  acre  wilderness  is  located 


in  northeast  West  Virginia  in  an  area  with  a  cool,  humid 
climate  and  abundant,  uniform  precipitation  averaging 
50  to  55  inches  annually.  It  is  located  in  the 
unglaciated  Allegheny  Plateau;  mountainous,  with 
elevations  ranging  between  1,800  feet  near  the  mouth 
of  Otter  Creek,  to  3,900  feet  on  McGowan  Mountain. 
Otter  Creek,  a  perennial  stream,  bisects  the  wilderness 
and  a  number  of  perennial  tributaries  occur  throughout 
the  area. 

Waters  are  generally  acid  and  low  in  productivity. 
There  is  a  small  native  brook  trout  population  in  the 
upper  reaches  of  Otter  Creek,  made  possible  by  a 
demonstration  project  being  conducted  by  the  West 
Virginia  Department  of  Natural  Resources  in  which 
ground  limestone  is  continuously  added  to  Otter  Creek 
just  outside  the  wilderness  boundary.  Trout  may  also 
occur  in  the  lower  reaches  of  Otter  Creek,  below  its 
confluence  with  Turkey  Run,  where  limestone  bedrock 
borders  the  stream.  Otter  Creek  drains  from  a  6-acre 
open  acid  fen,  and  some  tributaries  (Yellow  Creek  and 
Moore  Run)  drain  from  open  sphagnum/sedge/spruce 
swamps. 

Soils  are  moderately  deep  to  deep,  with  base 
saturation  less  than  35%  (Utisols),  and  35  to  50% 
(Inceptosols).  They  are  high  in  iron  and  aluminum. 
Forest  vegetation  is  a  mixture  of  northern  hardwoods 
and  Allegheny  mixed  hardwoods,  including  a 
component  of  yellow  poplar,  and  with  red  spruce  at  the 
higher  elevations.  Rhododendron  makes  up  understory 
vegetation  over  extensive  areas.  Ground  and 
herbaceous  vegetation  is  somewhat  depauperate  over 
most  of  the  wilderness  compared  to  other  low-elevation 
mesic  sites,  with  the  exception  of  limited  areas  of 
limestone  bedrock  in  which  vegetation  richness 
increases  and  spring  wildflowers  may  be  abundant. 
The  area  was  logged  between  1890  and  1915,  and 
200  acres  of  Norway  spruce  was  later  planted  on  the 
top  of  Green  Mountain. 

Great  Gulf.  This  gulf  and  its  tributary  gulfs  were 
hollowed  out  by  the  action  of  glaciers  before  the  last 
ice  age.  One  of  the  distinctive  features  of  the  eastern 
slopes  of  the  Presidential  Range,  this  glacial  valley 
between  Mount  Washington  and  the  Northern  Peaks  is 
from  1,100  to  1,600  feet  deep.  It  extends  easterly  from 
Mount  Washington  some  3-1/2  miles  as  a  narrow, 
steep-sided  gulf  before  broadening  gradually  to  more 
open  terrain.  It  contains  a  number  of  remarkable 
cascades,  and  the  views  from  the  walls  and  from 
points  on  the  floor  are  among  the  best  in  New  England. 

Many  of  the  older  trees  have  been  damaged  by 
hurricanes,  but  a  few  scattered  stands  of  large  virgin 
spruce  remain.  A  small  portion  of  the  eastern  part  of 
the  area,  in  the  lower  slope  type,  was  cut  over  for  large 
spruce  late  in  the  19th  century.  Northern  hardwoods 
are  the  typical  forest  at  lower  elevations.  Alpine  plants 
and  lichens  abound  above  treeline.  Stunted  spruce  and 


24 


fir  provide  the  transition  between  the  alpine  and 
forested  areas. 

A  weather  observatory  on  the  6,262-foot  summit  of 
Mount  Washington,  the  highest  point  in  the  Northeast, 
just  outside  the  southwest  corner  of  the  wilderness, 
has  recorded  the  highest  wind  speed  (231  mph)  of  any 
weather  station  world-wide.  Wind  speeds  in  excess  of 
100  mph  are  not  uncommon.  The  weather  is  severe 
most  of  the  year,  and  approximates  conditions 
encountered  at  a  much  higher  latitude.  The  summit  of 
Mount  Washington  is  in  the  clouds  approximately  55 
percent  of  the  time.  The  effects  and  extent  of  acid 
cloud/fog  water  (pH  generally  less  than  3.0)  are 
currently  being  studied. 


Air  Quality  Related  Values 

Otter  Creek:  Three  isolated  freshwater  wetlands 
occur,  with  some  sphagnum  vegetation  most 
commonly  associated  with  more  northern  wetland 
areas.  A  59-acre  stand  of  virgin  red  spruce  and 
hemlock  remains  on  Shavers  Mountain.  Spring 
ephemerals,  especially  on  the  more  productive  sites, 
provide  some  very  desirable  diversity  and  richness.  A 
large  number  of  tree,  shrub,  and  herbaceous  species 
within  the  area  have  known  sensitivity  to  various  air 
pollutants  (black  cherry,  yellow  poplar,  red  spruce, 
etc.).  Change  in  the  native  plant  communities  and 
associated  fauna  resulting  from  air  pollution  would  be 
undesirable. 

Water  quality  is  important  for  drinking  water  by 
wilderness  users,  and  for  the  limited  cold  water  fishery 
in  Otter  Creek.  However,  water  quality  in  Otter  Creek  is 
being  artificially  improved  for  fishery  purposes  by  the 
continuous  addition  of  ground  limestone,  raising  the  pH 
and  alkalinity  of  the  stream.  Therefore,  water  quality 
measurements  in  Otter  Creek  are  not  representative  of 
natural  conditions.  Without  such  limestone  treatment, 
Otter  Creek  water  in  its  natural  condition  is  acid  (pH 
5.0)  and  very  low  in  productivity  (alkalinity  much  less 
than  2  milligrams  per  liter,  and  conductivity  25  u,S/cm) 
where  it  enters  the  wilderness.  Water  quality  further 
deteriorates  going  downstream  due  to  even  poorer 
quality  tributary  inputs,  until  the  neutralizing  influence 
of  limestone  bedrock  is  encountered  near  the  mouth  of 
Otter  Creek,  where  water  quality  improves  somewhat 
for  a  fairly  short  reach  of  the  stream  (pH  5.8  to  6.0, 
alkalinity  3.1  mg/l  and  conductivity  31).  Tributaries  to 
Otter  Creek  have  pH  less  than  4.0  and  alkalinities  less 
than  0.2  mg/l. 

Great  Gulf:  Water  quality  is  important  for  trout 
fisheries,  and  for  hikers  to  drink  and  enjoy  its  scenic 
quality.  Alpine  flowers  are  rare  in  the  Northeast  and 
they  exist  in  a  harsh  environment  probably  susceptible 
to  damage  from  changes  in  soil  chemistry. 


Boundary  Waters  Canoe  Area  Wilderness  - 
Minnesota 

Brief  Description 

This  second  largest  wilderness  area  (798,458 
acres)  under  Forest  Service  administration  sits  astride 
the  border  between  Minnesota  and  Ontario,  Canada. 
The  elevation  averages  1,150  feet  above  sea  level. 

The  climate  is  continental  polar,  with  long  cold 
winters  and  cool  summers  that  provide  only  95  frost- 
free  days  per  year.  Annual  precipitation  varies  from  20 
to  30  inches  per  year. 

The  bedrock  underlying  the  Boundary  Waters 
Canoe  Wilderness  is  precambrian  metamorphic  and 
intrusive  igneous  rock,  which  has  been  glaciated  only 
recently.  The  bedrock  is  overlaid  with  very  thin, 
nutrient-poor  spodosol  soils  of  low  cation  exchange 
capacity,  high  in  iron  and  aluminum,  with  moderately 
low  acid  neutralizing  capacity,  and  are  essentially 
neutral  in  pH. 

This  wilderness  contains  over  1,000  lakes  larger 
than  10  acres.  Three-fourths  of  these  lakes  are  slightly 
to  heavily  stained  a  brown  color  from  organic  materials 
draining  from  the  abundant  peatlands  in  the 
wilderness.  The  pH  of  most  of  the  lakes  falls  in  the  6.6 
to  8.3  range,  with  a  mean  of  about  7.3.  A  few  of  the 
highly  stained  lakes  have  pH's  as  low  as  5.6.  Many  of 
the  lakes  are  sensitive,  and  could  become  acidified  if 
acid  deposition  were  to  increase.  About  half  have 
ANC's  less  than  130  u.eq/1  and  base  cation 
concentrations  below  215  |ieq/l.  About  5%  of  the  lakes 
in  this  wilderness  have  ANC's  less  than  50  and  base 
cation  concentrations  below  140  jieq/l. 

Air  quality  at  present  is  very  good  since  the  BWCA 
sits  on  the  eastern  fringe  of  the  Canadian  and  United 
States  Great  Plains,  which  have  so  far  sustained  little 
industrial  development.  Also,  the  air  quality  standards 
of  the  State  of  Minnesota  are  substantially  more 
stringent  than  the  United  States  federal  ambient  air 
quality  standards.  As  a  result,  in  1985  the  Boundary 
Waters  Canoe  Wilderness  sustained  a  measured  wet 
deposition  of  only  1.4  to  2.3  kg  of  nitrate  nitrogen,  0.2 
to  0.3  kg  of  ammonium  nitrogen,  2.3  to  3.5  kg  of 
sulfate,  and  a  hydrogen  ion  deposition  of  generally  less 
than  0.1  kg  per  hectare  per  year.  The  annual  average 
pH  of  precipitation  was  about  5.0.  The  average  ozone 
concentration  during  the  growing  season  is  about  35 
ppb. 


Air  Quality  Related  Values 

o  High-quality  waters  that  support  a  highly  diverse 
fishery. 


25 


o  Coniferous  and  mixed  coniferous  forests  that 
provide  the  critical  habitat  for  one  of  the  last 
remaining  and  viable  eastern  timber  wolf 
populations  in  the  continental  United  States. 

o  Bird  populations,  especially  bald  eagles  and  loons. 

o  Native  American  pictographs  and  buried  sites. 

The  relatively  lower  than  usual  Green  Line  and  Red 
Line  values  recommended  for  total  sulfur  and  nitrogen 
deposition  in  this  Wilderness  are  justified  because  of 
the  substantial  sensitivity  of  the  shallow  soils,  the  hard 
crystalline  bedrock,  and  the  low  alkalinity  of  the  surface 
waters. 

Suggested  factors  to  be  considered  in  making  a 
determination  of  Green  Line  (or  better)  conditions: 

o  Based  on  current  knowledge  of  species  sensitivity, 
modelled  increases  in  pollutant  loads  will,  with  a 


high  degree  of  certainty,  result  in  no  reduction  in 
distribution  of  known  pollutant  sensitive  tree  or 
lesser  vegetation  species. 

o  With  a  high  degree  of  certainty,  modelled  increases 
in  pollutant  loads  will  have  negligible  or  no  impact 
on  acid  neutralizing  capacity  of  any  BWCA  lake. 

Suggested  factors  to  be  considered  in  making  a 
determination  of  Red  Line  (or  worse)  conditions: 

o  Based  on  current  knowledge  of  species  sensitivity, 
modelled  increases  in  pollutant  loads  will  result  in 
either  complete  elimination  or  reduce  distribution  of 
at  least  one  tree  or  lesser  vegetation  species. 

o  Modelled  increases  in  sulfate  or  nitrate  deposition 
will  result  in  complete  elimination  of  acid 
neutralizing  capacity  from  one  or  more  lakes. 


26 


APPENDIX  C.  OTHER  AQUATIC  MEASUREMENT 
METHODS 


Loading/Response  Relationships 

The  acidification  of  lakes  has  been  considered  to  be 
analogous  to  the  titration  of  a  bicarbonate  solution 
(acid  neutralizing  capacity,  ANC)  with  acidic  (sulfuric 
acid)  atmospheric  deposition  (Henriksen  1979).  When 
additions  of  acid  consume  ANC,  pH  decreases  slowly 
at  first,  then  more  markedly  as  ANC  is  depleted  (Small 
et  al.  1988).  Acidic  deposition  may  also  increase  the 
weathering  of  base  cations  and  not  result  in  an 
equivalent  consumption  of  ANC  for  each  equivalent  of 
acid  deposited  (Henriksen  1984).  Lake  ANC  is 
produced  from  watershed  weathering  and  exchange 
reactions.  These  reactions  generate  equivalent 
amounts  of  bicarbonate  and  base  cations.  Lake  ANC 
can  also  be  produced  from  in-lake  processes,  such  as 
Ca  exchange  with  sediments,  and  biologically 
mediated  removal  of  nitrate  and  sulfate  (Schindler 
1986).  The  relative  importance  of  in-lake  versus 
watershed  sources  of  alkalinity  and  the  relationship 
between  acid  deposition  and  enhanced  weathering  of 
base  cations  is  known  for  only  a  few  ecosystems. 
These  additional  mechanisms,  which  act  to  reduce  the 
effect  of  acidic  deposition,  cannot  be  included  in  a 
conservative  estimate  of  the  relationship  between 
deposition  amount  and  ecosystem  impacts. 

Our  approach  is  analogous  to  the  Henriksen 
empirical  model  where  deposition  amount,  lake 
sensitivity  (sum  of  base  cations),  and  the  results  of 
lakes  surveys  are  used  to  empirically  derive,  for  lakes 
of  a  given  sensitivity  and  deposition  level,  where  the 
system  will  experience  ANC  decline  and  pH 
depression.  We  assume  that  in-lake  alkalinity 
generation  and  enhanced  weathering  of  base  cations 
is  negligible. 


Graph  Construction 

Figures  C-1  and  C-2  show,  for  various  deposition 
levels,  the  concentrations  of  non-marine  Ca+Mg+K+Na 
in  lakes  having  ANC  of  1 0  to  25  |ieq/l  and  pH  values  of 
about  5.9  to  6.2,  and  in  lakes  with  ANC  between  -20 
and  -5  and  pH  of  about  4.8  to  5.2.  Figure  C-1  shows 
the  Green  Lines  for  these.  The  Green  Lines  indicate 
the  deposition  level  below  which  lakes  with  various 
base  cation  concentrations  should  maintain  ANC  of  at 
least  10  to  25  u.eq/1  and  pH  of  at  least  5.8  to  6.2.  Figure 
C-2  shows  Red  Lines.  If  subjected  to  a  particular 
deposition  level,  lakes  with  base  cation  concentrations 


less  than  those  indicated  by  the  Red  Lines  can  be 
expected  to  become  acidic  with  ANC  falling  below  zero 
and  pH  reaching  5.2  or  less. 

The  graphs  are  based  on  the  assumption  that,  while 
lake  HCO3  decreases  and  is  replaced  by  SO4  in 
response  to  increasing  S  deposition,  base  cation 
concentrations  do  not  change.  In  fact,  as  deposition 
increases,  mineral  weathering  of  base  cations  from  the 
watershed  may  increase  somewhat  (Henriksen  1984). 
As  a  result,  lakes  with  a  given  base  cation 
concentration  may  be  able  to  withstand  a  somewhat 
greater  increase  in  S  deposition  than  indicated  by  the 
nomographs.  In-lake  alkalinity  production  may  also 
reduce  the  impact  of  S  deposition  (Schindler  1986). 
Because  these  effects  are  uncertain  in  magnitude  and 
probably  do  not  occur  in  all  lakes,  we  have  taken  the 
conservative  approach  of  protecting  the  lakes  and 
have  assumed  that  these  processes  are  not  significant. 

The  amount  of  runoff  relative  to  the  amount  of 
precipitation  on  a  watershed  affects  how  lake 
chemistry  responds  to  acid  loadings.  As  more 
precipitation  is  lost  to  evapotranspiration  and  less  is 
yielded  as  runoff,  acid  deposition  is,  in  effect, 
concentrated,  and  its  impact  on  a  lake  is  greater. 
Consequently,  the  graphs  show  several  Red  and 
Green  Lines  for  various  amounts  of  runoff,  expressed 
as  percentages  of  annual  precipitation. 

N  deposition  is  included  for  very  low  ANC  (<50 
u.eq/1)  waters  in  the  western  United  States.  The 
rationale  for  including  N  is  based  on  observations  that 
most  N  deposited  on  a  watershed  is  retained  in  the 
watershed.  At  most,  about  20%  can  be  seen  in  surface 
waters.  This  is  explained  in  the  text.  Deposition  loading 
was  determined  as  outlined  in  the  surface  water 
sensitivity  section,  page  12.  Essentially,  total 
deposition  was  determined  by  combining  wet 
deposition  data  with  dry  deposition  estimates.  In  the 
east,  dry  deposition  was  estimated  to  be  30%  of  wet 
deposition,  while  in  the  west  dry  deposition  was 
assumed  to  be  zero. 

Within  about  125  miles  of  the  sea  coast, 
precipitation  contains  significant  amounts  of  sodium, 
magnesium,  chloride,  and  sulfate  and  lesser  amounts 
of  other  ions  of  marine  origin.  These  ions  increase  the 
base  cation  concentration  of  lakes  in  these  areas 
without  adding  HCO3  or  ANC.  To  correct  for  this  effect, 
we  assume  that  all  chloride  (CI)  in  such  lake  water  is 
from  marine  sources,  and  subtract  from  the  base 
cation  concentration  an  amount  in  proportion  to  the 
relative  concentrations  of  CI  and  base  cation  in 
seawater  (Hem  1970). 


27 


60-70%  Runoff 
/ 


14 


12 


40-50% 
Runoff 

/ 

/ 


10 


8 


4  - 


2  - 


Adirondack  Mountains,  runoff=60-70% 
of  precipitation 
o  ANC  =  10  to  25^eq/l 

•  ANC  =  -5  to  -20 

Pocono  and  Catskill  Mountains,  runoff=:60-70% 
O  ANC  =  10  to  25 

•  ANC  =  -5  to  -20 

New  England,  includes  southern  New  England, 
central  New  England,  and  Maine,  runoff=60-70% 
a  ANC  =  10  to  25 
a  ANC  =  -5  to  -20 

Northern  Wisconsin  and  upper  Michigan, 
runoff=40-50% 
□  ANC  =  10  to  25 
■  ANC  =  -5  to  -20 

Western  Mountains,  ANC  =  10-25  ^eq/l 
o  Runoff  =  80-90% 

•  Runoff  =  50-70% 
®  Runoff  =  25-35% 

 I  s  


40 


80 


120 


160 


200 


Base  cations,  /ieq/l  (adjusted  for  marine  influence) 

Figure  C-1.--Base  cations/deposition  relationship  for  Green  Lines.  Lakes  to  the  right  of  the 
appropriate  runoff  lines  are  not  considered  to  be  acidic.  (Data  from  Kanciruk  et  al.  1986  and  Eilers 
et  al.  1987.) 


28 


80-90%  Runoff 


60-70%  Runoff 


40-50%  Runoff 


14 


12 


10 


8 


2  - 


25-35% 
Runoff 


-  Adirondack  Mountains,  runoff=60-70% 
of  precipitation 
o  ANC  =  10  to  25Meq/l 

•  ANC  =  -5  to  -20 

-  Pocono  and  Catskill  Mountains,  runoff=60-70% 
O  ANC  =  10  to  25 

•  ANC  =  -5  to  -20 

-  New  England,  includes  southern  New  England, 
central  New  England,  and  Maine,  runoff=60-70% 
a  ANC  =  10  to  25 
a  ANC  =  -5  to  -20 

-  Northern  Wisconsin  and  upper  Michigan, 
runoff=40-50% 
o  ANC  =  10  to  25 
■  ANC  =  -5  to  -20 

-  Western  Mountains,  ANC  =  10-25  ^eq/l 
o  Runoff  =  80-90% 

•  Runoff  =  50-70% 
®  Runoff  =  25-35% 


1 


1 


1 


40  80  120  160  200 

Base  cations,  pteq/l  (adjusted  for  marine  influence) 

Figure  C-2.--Base  cations/deposition  relationship  for  Red  Lines.  Lakes  to  the  left  of  the  appropriate 
runoff  lines  are  likely  acidic.  (Data  from  Kanciruk  et  al.  1986  and  Eilers  et  al.  1987.) 


29 


Inland  lakes  generally  have  low  concentrations  of  CI 
and  Cl-associated  base  cations,  but  natural  geologic 
sources  and  contamination  by  road  salt  may  increase 
them.  The  base  cation  concentration  of  these  lakes 
also  must  be  corrected  for  this  influence.  This  is  done 
by  assuming: 

Non-marine  Ca+Mg+K+Na  =  Total 
Ca+Mg+K+Na  -  (CI  x  1.115) 

Concentrations  of  all  ions  are  expressed  in  u.eq/1. 

Lake  SO4  generally  increases  with  increasing  S 
deposition,  but  natural  geologic  sources  also  contribute 
variable  amounts  of  SO4  to  lakes  (Loranger  and 
Brakke  1988,  Eilers  et  al.  1987).  Base  cations 
associated  with  geologic  SO4  add  significant  variability 
to  the  ANC:base  cation  relationship.  In  figures  C-1  and 
C-2,  much  of  the  scattering  of  data  points  for  any  given 
deposition  level  and  runoff  percentage  can  be 
attributed  to  geologic  SO4  and  associated  base 
cations. 

This  approach  is  similar  to  the  so-called  Hendriksen 
model,  which  has  been  shown  to  have  limitations 
(Reuss  et  al.  1986,  Vertucci  in  press).  However,  the 
Red  Line  values  used  here  are  based  on  an  empirical 
fit  to  the  data  on  acidified  lakes.  The  Green  Line  values 
essentially  represent  a  simple  balance  of  increased 
sulfate  (and  nitrate)  against  ANC.  This  is  an 
approximation  that  Wright  (1988),  among  others, 
suggests  can  be  improved  by  the  introduction  of  a 
factor  "f"  representing  the  ratio  of  change  in  base 
cation  concentration  to  net  sulfate  (that  attributable  to 
anthropogenic  sources).  The  model  here  assumes  f  to 
be  zero.  A  nonzero  f  would  make  the  Green  Lines 
steeper.  Since  this  is  intended  as  a  worst  case 
screening  technique  that  errs  on  the  side  of 
conservatism,  and  actual  f  values  are  unknown,  we  felt 
it  was  appropriate  to  use  an  f  factor  of  zero. 

In  cases  where  little  cation  data  exist,  conductivity 
was  considered  as  an  alternate  measure  of  sensitivity. 
Since  conductivity  is  easily  and  cheaply  measured  in 
the  field,  conductivity  data  may  be  more  widely 
available  for  surface  waters  than  are  cation  data.  Easy 
and  cheap  don't  necessarily  equate  with  accurate, 
however,  and  field  conductivity  data  must  be  closely 
screened  to  ensure  reliability.  Electrical  conductivity 
(usually  expressed  as  u.Siemens/cm  at  25  degrees  C) 
is  a  measure  of  the  total  amount  of  ions  dissolved  in 
the  water.  Consequently,  conductivity  is  related  to  the 
sum  of  the  base  cations  and  anions.  Waters  that  are 
inherently  sensitive  to  acid  deposition  have  little 
buffering  or  acid  neutralizing  capacity,  low 
concentrations  of  base  cations,  and  low  conductivity. 

Figures  C-3  and  C-4  are  similar  to  C-1  and  C-2,  but 
show  lake  conductivity  in  place  of  base  cation 
concentration  as  a  measure  of  lake  sensitivity. 
Because  conductivity  is  an  indicator  of  the  total 


concentration  of  ions  dissolved  in  the  waters,  it  is  used 
as  a  substitute  for  the  base  cation  concentration.  Since 
the  contribution  of  SO4  and  HCO3  to  conductivity  are 
about  the  same,  we  assume  that  the  conductivity  of  a 
lake  does  not  change  with  increasing  S  deposition.  As 
with  figures  C-1  and  C-2,  we  ignore  the  fact  that  base 
cations,  and  consequently  conductivity,  may  increase 
somewhat  at  greater  deposition  levels,  due  to 
increased  mineral  weathering. 

As  with  base  cations,  conductivity  must  be 
corrected  for  marine  influences.  This  correction  is  even 
more  critical  for  conductivity,  because  conductivity  is 
influenced  not  only  by  the  ocean-derived  base  cations, 
but  also  by  the  CI  and  SO4.  In  addition  to  the 
adjustments  for  marine  contributions,  conductivity  must 
also  be  corrected  for  hydrogen  ion  (pH)  influences.  In 
acidic  waters,  hydrogen  contributes  heavily  to 
conductivity,  and  this  contribution  must  be  subtracted. 
All  the  lake  conductivity  data  in  figures  C-3  and  C-4  are 
corrected  for  both  pH  and  marine  influences.  This  is 
done  by  assuming: 

Non-marine  conductivity  =  measured 
conductivity  -  (CI  x  0.1422) 

with  CI  expressed  in  u.eq/1  and  conductivity  in 
liSiemens. 

Adjusted  conductivity  =  measured 
conductivity  -  (H  x  0.34965) 

where  H  is  the  hydrogen  ion  concentration  in  jieq/l 
(H=10  raised  to  the  -pH  power,  then  that  quantity 
multiplied  by  1,000,000). 

Differences  in  natural  sources  of  SO4  add  much 
variation  to  the  ANC:conductivity  relationship.  This 
effect  is  greater  than  that  on  ANC:base  cations 
because  both  SO4  and  the  associated  base  cations 
contribute  to  conductivity.  For  any  particular  deposition 
level  and  runoff  percentage,  the  scatter  among  the 
data  points  and  the  overlap  between  groups  of  acidic 
and  non-acidic  lakes  is  greater  in  figures  C-3  and  C-4 
than  in  C-1  and  C-2.  Therefore,  base  cations  rather 
than  conductivity  should  be  used  as  a  measure  of  lake 
sensitivity  where  cation  data  are  available.  Conductivity 
is  useful  as  a  rough  tool  to  separate  lakes  into  non- 
sensitive  and  possibly  sensitive  groups. 


Detailed  Information  Needs 

Surface  water  chemistry  data  can  be  collected  by 
means  of  special  purpose  surveys,  by  census,  or  by 
estimating  values  based  on  previous  surveys  in  similar 
geographic  terrains.  As  an  example  of  the  last 
approach,  an  approximate  characterization  of  the 
surface  water  chemistry  of  seven  of  the  nine 
wilderness  ecosystem  types  is  presented  in  table  C-1. 


30 


80-90%  Runoff 


~  14 

CD 
O 

c 

0) 


12 


10 


8 


6  - 


Adirondack  Mountains,  runoff=60-70% 

of  precipitation 

o  ANC  =  10  to  25ixeq/\ 

•  ANC  =  -5  to  -20 

Pocono  and  Catskill  Mountains,  runoff=60-70% 
O  ANC  =  10  to  25 

•  ANC  =  -5  to  -20 

New  England,  includes  southern  New  England, 
central  New  England,  and  Maine,  runoff=60-70% 
a  ANC  =  10  to  25 
a  ANC  =  -5  to  -20 

Northern  Wisconsin  and  upper  Michigan, 
runoff=40-50% 
o  ANC  =  10  to  25 
■  ANC  =  -5  to  -20 

Western  Mountains,  ANC  =  10-25  j^eq/l 
o  Runoff  =  80-90% 

•  Runoff  =  50-70% 
©  Runoff  =  25-35% 


1 


1 


1 


1 


5  10  15  20  25 

Conductivity,  juS/cm  (adjusted  for  marine  influences  and  H+) 

Figure  C-3. --Conductivity/deposition  relationships  for  Green  Lines.  Lakes  to  the  right  of  the 
appropriate  runoff  lines  are  not  considered  acidic.  (Data  from  Kanciruk  et  al.  1986  and  Eilers  et  al. 
1987.) 

31 


80-90%  Runoff 


60-70%  Runoff 


-77T  14 


12 


10 


^  8 


4  - 


2  - 


40-50% 
Runoff 


25-35% 
Runoff 


Adirondack  Mountains,  runoff=60-70% 
of  precipitation 
o  ANC  =  10  to  25^eq/l 

•  ANC  =  -5  to  -20 

Pocono  and  Catskill  Mountains,  runoff=60-70% 
O  ANC  =  10  to  25 

•  ANC  =  -5  to  -20 

New  England,  includes  southern  New  England, 
central  New  England,  and  Maine,  runoff=60-70% 
a  ANC  =  10  to  25 
a  ANC  =  -5  to  -20 

Northern  Wisconsin  and  upper  Michigan, 
runoff=40-50% 
□  ANC  =  10  to  25 
■  ANC  =  -5  to  -20 

Western  Mountains,  ANC  =  10-25  ^eq/l 
o  Runoff  =  80-90% 

•  Runoff  =  50-70% 
©  Runoff  =  25-35% 


1 


1 


1 


5  10  15  20  25 

Conductivity,  nS/cm  (adjusted  for  marine  influences  and  H+) 

Figure  C-4.~Conductivity/deposition  relationships  for  Red  Lines.  Lakes  to  the  left  of  the  appropriate 
runoff  lines  are  likely  acidic.  (Data  from  Kanciruk  et  al.  1986  and  Eilers  et  al.  1987.) 


32 


Table  C-1  .--Conductivity  and  neutralizing  capacity,  and  pH  statistics  for  geographic  regions  represented  by  seven  wilderness 

ecosystem  types. 


Wilderness  area 


Chemical  factor 


and 
NSWS  region 

Conductivity 

ANC 

pH 

Min. 

Q1 

Med. 

Min. 

Qi 

Med. 

Min. 

Q1 

Med. 

liS/cm 

\xeq/l 

Alnino  I  al^oc 
AM|JIIIC  L_cmUo, 

1 AIA 

OH/ \JC. 

fi  0, 

r^U^Ior  Danle  /DM\A/1\ 

oiacier  reals  ^riNvv  1 ) 

Hoover,  Dome  Land 

>2 

5 

8 

7/13 

34 

60 

>5.8 

6.6 

7.0 

(SNM2) 

Bob  Marshall  (NR3) 

>3 

9 

39 

72 

77 

342 

6.3 

6.9 

7.1 

Bridger  (ALP4) 

>7 

12 

15 

39 

74 

109 

>5.8 

6.9 

7.1 

Joyce  Kilmer, 

10 

14 

21 

16 

87 

120 

6.4 

6.8 

7.0 

Slickrock  (EHW5) 

Otter  Creek,  Great 

19 

33/22 

69/35 

-48 

20/52 

110/119 

4.4 

5.7/6.3 

6.6/6.8 

Gulf  (NH6) 

Boundary  Waters  Canoe 

18 

22 

30 

34 

98 

185 

5.6 

6.6 

6.9 

Water  Area  (C?) 


1 WLS  Pacific  NW,  Middle  Washington  and  Wenatchee  Mtns. 

2WLS  California,  Sierra  Nevada. 

3WLS  Northern  Rockies,  Lewis  Range. 

4WLS  Central  Rockies,  Wind  River. 

5NSS  Southern  Blue  Ridge. 

6NSS  N.  Appalachians,  ELS  C.  New  England. 

7ELS  NE  Minnesota. 


WLS  =  EPA  Western  Lake  Survey. 
ELS  =  EPA  Eastern  Lake  Survey. 
NSS  =  National  Stream  Survey. 


These  data  are  based  on  the  National  Surface  Water 
Survey  (NSWS),  which  measured  the  chemistry  of  a 
large  statistical  sample  of  lakes  and  streams  in  regions 
of  the  United  States,  expected  to  have  surface  water 
with  low  acid  neutralizing  capacity.  In  several  cases 
(such  as  the  Bridger  Wilderness),  many  lakes  were 
sampled.  In  most  cases,  however,  only  a  few  lakes  or 
streams  were  actually  included.  As  an  approximation, 
the  chemical  data  in  table  C-1  were  aggregated  to 
include  the  geographic  units  nearest  to  exact 
wilderness  that  approximate  the  geology  of  the 
corresponding  regions,  based  on  the  NSWS  data.  No 
data  were  collected  in  southern  California  nor  in 
Arizona. 

Half  of  the  lakes  or  streams  in  each  region  are 
expected  to  fall  below  the  median  value  for  each 
region.  Twenty  percent  of  the  systems  are  expected  to 
fall  below  the  first  quintile  (Q-|).  Minimum  values 
represent  the  lowest  value  observed  in  the  sample, 
and  do  not  necessarily  represent  the  lowest  lake  or 


stream  in  the  region.  Lake  chemistry  was  measured 
following  fall  overturn.  Stream  chemistry  was  sampled 
in  the  spring  between  snowmelt  and  leaf-out,  avoiding 
rain  storms.  Streams  affected  by  acid  mine  drainage 
and  polluted  lakes  were  avoided. 

These  data  are  statistically  valid  randomly  selected 
samples  of  water  quality  in  all  areas.  To  obtain  a  better 
estimate  of  the  true  chemistry  distributions  in  a 
particular  wilderness,  a  random  sample  of 
approximately  50  lakes  can  generally  give  acceptable 
confidence  bounds  if  the  area  is  not  too 
heterogeneous.  If  50  represents  less  than  5%  of  the 
total  population  of  lakes,  or  if  the  area  is  highly  diverse, 
a  larger  sample  size  may  be  needed  to  reduce 
uncertainties  in  the  estimates.  Field  sampling,  while 
inexpensive,  must  follow  protocols  for  wilderness  areas 
(Fox  et  al.  1987). 

Annual  runoff  can  be  calculated  from  estimated 
precipitation  and  evapotranspiration  measurements  on 
site,  or  measured  at  a  gauged  stream  site  in  the  region 


33 


of  interest.  In  the  absence  of  such  data,  published 
values  of  mean  annual  runoff  from  state  and  federal 
agencies  in  state  water  atlases  and  other  publications 
can  be  used.  Annual  variations  in  runoff  are  not  a 
significant  concern  in  using  figure  1 ,  provided  long  term 
data  are  available. 

Dry  deposition  of  sulfate  and  nitrate  are  often 
estimated  from  obtaining  wet  deposition  data  from  the 
nearest  National  Acid  Deposition  Program  (NADP)  site. 
As  a  rule  of  thumb,  in  rural  areas  removed  from  point 
sources  of  pollution,  dry  deposition  of  sulfur  can  be 
assumed  to  equal  30%  of  the  wet  deposition  value.  Dry 
nitrogen  deposition  may  be  somewhat  greater  than  the 
wet  value.  These  factors  are  subject  to  considerable 
local  variations,  including  impaction  of  particles  on  dry 
surfaces,  and  adsorption  of  gaseous  species  (SO2  and 
HNO3)  by  moist  surfaces,  including  lakes  and  the  open 
stomata  of  vegetation.  If  air  concentration  data  of  S 
and  N  are  available,  dry  deposition  can  be  calculated 
using  assumed  values  of  deposition  velocity  taken  from 
the  Air  Resource  Handbook.  Still  more  desirable  are 
dry  deposition  estimates  from  a  nearby  NDDN 
(National  Dry  Deposition  Network)  site.  These  are 
currently  being  installed  throughout  the  United  States. 


Conversion  of  Deposition  Values 

Deposition  loadings  are  presented  in  kg/ha-yr  of  S 
and  N.  Deposition  measurements  are  often  reported  as 
deposition  of  SO4  and  NO3  in  mg/m2/yr.  Land 
managers  may  also  be  familiar  with  applications  of  S 
and  N  in  Ib/A/yr.  The  following  conversion  factors  may 
be  useful: 

Multiply  S  deposition  by  3.0  to  determine  SO4 
deposition 

Multiply  N  deposition  by  4.43  to  determine  NO3 
deposition 

Multiply  kg/ha  by  0.89  to  determine  lb/A 
Multiply  kg/ha  by  100  to  determine  mg/m2 

Multiply  kg/ha  by  0.1  to  determine  g/m2 

To  convert  from  mg/l  to  u.eq/1,  multiply  mg/l  of  Ca  by 
49.90,  Mg  by  83.26,  K  by  25.57,  Na  by  43.50,  CI  by 
28.21,  and  SO4  by  20.82. 


34 


APPENDIX  D.  PARTICIPANTS  AND  THEIR 
AFFILIATIONS 


Ann  M.  Bartuska 
Research  Plant  Physiologist 
USDA-Forest  Service 
NC  State  University 
1509  Varsity  Drive 
Raleigh,  NC  27607 

Clif  R.  Benoit 

Regional  Air  Resource  Specialist 
USDA/FS  R-4  RWM 
324  25th  Street 
Ogden,  UT  84401 

Edgar  B.  Brannon 
Forest  Supervisor 
Flathead  National  Forest 
1935  3rd  Ave.  E. 
Kalispell,  MT  59901 

John  Butruille 

Director,  Recreation  Management 
USDA/Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

James  G.  Byrne 
Air  Resource  Program  Manager 
USDA/Forest  Service 
Watershed  and  Air  Management 
Rm.  1210,  RPE 
Washington,  DC  20090-6090 

William  A.  Carothers 
Regional  Air  Resource  Specialist 
USDA/Forest  Service,  R-8,  SW&A 
1720  Peachtree  Rd.  N.W. 
Atlanta,  GA  30367 

Ellis  Cowling 

Associate  Dean 

North  Carolina  State  University 

1509  Varsity  Drive 

Raleigh,  NC  27606 

Peter  Dillon 

Supervisor,  Limnology  Unit 
Ontario  Ministry  of  the  Environment 
P.O.  Box  39 
Dorset,  Ontario 
Canada  POA  1EO 

Michael  Edrington 
Forest  Supervisor 
Williamette  National  Forest 
P.O.  Box  10607 
Eugene,  OR  97440 


Anne  Fege 

Wilderness  Program  Manager 
USDA/Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Richard  Fisher 

Air  Resource  Management  Specialist 
USDA/Forest  Service 

Rocky  Mountain  Forest  and  Range  Experiment  Station 
240  West  Prospect 
Fort  Collins,  CO  80526 

Douglas  G.  Fox 
Chief  Meteorologist 

Rocky  Mountain  Forest  and  Range  Experiment  Station 
240  West  Prospect 
Fort  Collins,  CO  80526 

Professor  Thomas  Frost 
Center  for  Limnology 
University  of  Wisconsin-Madison 
608  N.  Park  Street 
Madison,  Wl  53706 

Stephen  C.  Harper 
Forest  Supervisor 

Green  Mountain  and  Finger  Lakes  National  Forest 
P.O.  Box  519 
Rutland,  VT  05701 

Professor  J.  R.  N.  Jeffers 
Institute  of  Terrestrial  Ecology 
Ellerhow,  Lindale 
Grange-Over-Sands 
Cumbria  LA1 1  6JU 
United  Kingdom 

Dale  W.  Johnson 

Research  Ecologist 

Environmental  Sciences  Division 

Oak  Ridge  National  Lab 

P.O.  Box  X 

Oak  Ridge,  TN  37831 

Professor  David  F.  Karnosky 
Michigan  Technological  University 
School  of  Forestry  and  Wood  Products 
A11030 

Houghton,  Ml  49931 

Gene  E.  Likens,  Director 
Institute  of  Ecosystem  Studies 
The  New  York  Botanical  Gardens 
Mary  Flagler  Cary  Arboretum,  Box  AB 
Millbrook,  NY  12545 


35 


Steve  Lindberg 

Research  Ecologist 

Oak  Ridge  National  Laboratory 

Environmental  Sciences  Division 

Oak  Ridge,  TN  37831 

Rick  A.  Linthurst 

Director,  EPA  Aquatics  Effects  Research 
MD-39,  EPA/EMSL  (Annex) 
Research  Triangle  Park,  NC  2771 1 

Robert  C.  Loomis 

Ecologist,  Forest  Pest  Management 
USDA/Forest  Service,  FPM 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Gary  M.  Lovett 

Research  Ecologist 

Institute  of  Ecosystem  Studies 

The  New  York  Botanical  Gardens,  Box  AB 

Millbrook,  NY  12545 

William  J.  Mattson 
Research  Entomologist 
USDA,  Forest  Service 
Michigan  State  University 
1407  S.  Harrison  Road 
East  Lansing,  Ml  48823 

Steve  Mealey,  Assistant  Chief 
USDA,  Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Jay  Messer,  Ecologist 
EPA  Aquatic  Effects  Research 
MD-39,  EPA/EMSL  (Annex) 
Research  Triangle  Park,  NC  2771 1 

Dale  Nichols 

Research  Forester 

USDA,  Forest  Service 

NC  Station,  Forestry  Sciences  Lab 

1831  Highway  169  E. 

Grand  Rapids,  MN  55744 

Jan  Nilsson,  Director 

Department  of  Research  &  Development 

SNV  (National  Environmental  Protection) 

Box  1302 

S-171  25  Solna 

Sweden 

Dave  Peterson 
Research  Forester 
USDA,  Forest  Service 
Forest  Fire  Lab 
4955  Canyon  Crest  Drive 
Riverside,  CA  92507 


David  L.  Radloff 

Assistant  Director,  Forest  Fire  and  Atmospheric 
Sciences  Research 
USDA,  Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Professor  Peter  B.  Reich 
Department  of  Forestry 
University  of  Wisconsin 
121  Russell  Lab,  1630  Linden  Drive 
Madison,  Wl  53706 

Professor  John  Reuss 
Department  of  Agronomy 
Colorado  State  University 
Fort  Collins,  CO  80521 

Gray  F.  Reynolds 

Director,  Watershed  and  Air  Management 
USDA,  Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

William  T.  Sommers 

Director,  Forest  Fire  and  Atmospheric  Sciences 
Research 

USDA,  Forest  Service 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Richard  L.  Stauber 

Forest  Supervisor 

San  Bernardino  National  Forest 

1824  S.  Commercenter  Circle 

San  Bernardino,  CA  92408-3430 

Tom  L.  Thompson 
Forest  Supervisor 
Siuslaw  National  Forest 
P.O.  Box  1148 
Corvallis,  OR  97339 

David  G.  Unger,  Associate  Deputy  Chief 
USDA,  Forest  Service,  National  Forest  System 
P.O.  Box  96090 
Washington,  DC  20090-6090 

Charles  C.  Wildes 
Deputy  Forest  Supervisor 
Tonto  National  Forest 
P.O.  Box  5348 
Phoenix,  AZ  85010 

Richard  Wright 
Limnologist 

NIVA,  Norwegian  Institute,  Water  Res. 
P.O.  Box  333,  Blindern 
Oslo  3,  Norway 


36 


Rocky 
Mountains 


Great 
Plains 


U.S.  Department  of  Agriculture 
Forest  Service 

Rocky  Mountain  Forest  and 
Range  Experiment  Station 


The  Rocky  Mountain  Station  is  one  of  eight 
regional  experiment  stations,  plus  the  Forest 
Products  Laboratory  and  the  Washington  Office 
Staff,  that  make  up  the  Forest  Service  research 
organization. 

RESEARCH  FOCUS 

Research  programs  at  the  Rocky  Mountain 
Station  are  coordinated  with  area  universities  and 
with  other  institutions.  Many  studies  are 
conducted  on  a  cooperative  basis  to  accelerate 
solutions  to  problems  involving  range,  water, 
wildlife  and  fish  habitat,  human  and  community 
development,  timber,  recreation,  protection,  and 
multiresource  evaluation. 


RESEARCH  LOCATIONS 

Research  Work  Units  of  the  Rocky  Mountain 
Station  are  operated  in  cooperation  with 
universities  in  the  following  cities: 


Albuquerque,  New  Mexico 

Flagstaff,  Arizona 

Fort  Collins,  Colorado* 

Laramie,  Wyoming 

Lincoln,  Nebraska 

Rapid  City,  South  Dakota 

Tempe,  Arizona 


'Station  Headquarters:  240  W.  Prospect  St.,  Fort  Collins,  CO  80526