Skip to main content

Full text of "Management of amphibians, reptiles, and small mammals in North America : proceedings of the symposium, July 19-21, 1988, Flagstaff, Arizona"

See other formats


Historic,  Archive  Document 


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


■1 


sou 

United  States  ^ 
;  Department  of 
Agriculture 

Forest  Service 


Rocky  Mountain 
Forest  and  Range 
Experiment  Station 


Fort  Collins, 
Colorado  80526 


General  Technical 
Report  RM-166 


Management  of 
Amphibians,  Reptiles, 
and  Small  Mammals  ir 
North  America 

Proceedings  of  the  Symposium 


July  19-21,1988 
Flagstaff,  Arizona 


■■■■■■  "-T^^j^-^  .■■^'■i^^<'r~^ 


ACKNOWLEDGEMENTS 


This  meering  owes  its  success  to  sev- 
eral organizations  and  individuals. 
First,  we  thank  the  sponsoring  or- 
ganizations (listed  on  the  title  page) 
whose  financial  support  and  encour- 
agement helped  make  the  conference 
a  reality.  The  local  committee  on  ar- 
rangements, J.  Kevin  Aitkin,  Marga- 
ret Bailey,  Tom  Britt,  Roxanne  Britt, 
Charles  BuUington,  Glen  Dickens, 
and  Katherine  Holly  did  a  superb  job 
of  handling  room  setup,  registration, 
providing  rides,  and  running  the 
slide  projector. 

We  are  especially  grateful  to  the 
session  chairman,  K.  Bruce  Jones, 
George  Dalrymple,  Robert 
M,Closkey,  David  Germano, 
Winifred  Sidle,  Constantine  Slobod- 
chikoff,  Michael  Morrison,  Gregory 
Adler,  Martin  Raphael,  and  Ray- 
mond Dueser,  for  their  help  and  for 
keeping  the  meeting  on  schedule. 
Our  thanks  to  those  who  attended 
for  their  enthusiastic  participation. 
We  thank  Randall  Babb  for  the  line 
drawings  in  the  proceedings  and  De- 
borah Johnson  and  J.  Kevin  Aitkin 
for  their  help  in  organizing  manu- 
script files  and  standardizing  word 
processing  formats. 


We  would  like  to  extend  our  sin- 
cere thanks  to  the  following  peer  re- 
viewers who  generously  gave  their 
time  to  improve  the  quality  of  this 
proceedings:  Gregory  H.  Adler, 
Stanley  H.  Anderson,  Michael  J. 
Armbruster,  David  M.  Armstrong, 
Walter  Auffenberg,  Keith  B.  Aubry, 
Gary  C.  Bateman,  Ronald  E. 
Beiswenger,  Kristin  H.  Berry,  Wil- 
liam M.  Block,  Michael  A.  Bowers, 
Richard  C.  Bruce,  James  H.  Brown, 
K.  A.  Buhlmann,  Russell  Burke,  R. 
Bruce  Bury,  Ronald  K.  Chesser,  Ste- 
ven P.  Christman,  Tim  W.  Clark, 
Tames  P.  Collins,  Stephen  Corn, 
Stephen  P.  Cross,  George  Dalrymple, 
Joan  E.  Diemer,  James  G.  Dickson,  C. 
Kenneth  Dodd,  Jr.,  Raymond  D. 
Dueser,  Gary  M.  Fellers ,  Henry  S. 
Fitch,  Jerran  Flinders,  Vagn  F.  Flyger, 
Kenneth  Feluso,  Richard  Fitzner, 
David  J.  Germano,  Lowell  L.  Getz, 
William  E.  Grant,  Patrick  T.  Gregory, 
Marc  P.  Hayes,  Clyde  Jones,  K.  Bruce 
Jones,  Donald  W.  Kaufman,  Brian  J. 
Klatt,  Thomas  Kunz,  J.  Larry  Lan- 
ders, James  N.  Layne,  Harvey  B.  Lil- 
lywhite,  Raymond  Linder,  William 
Mannan,  S.  Clark  Martin,  Robert  T. 
M'Closkey,  David  A.  McCullough, 


Gary  K.  Meefe,  Joseph  C.  Mitchell, 
Paul  E.  Moler,  Henry  R.  Mushinsky, 
Thomas  J.  O'Shea,  William  S.  Parker, 
Kenneth  H.  Pollock,  Mary  V.  Price, 
Martin  G.  Raphael,  O.  J.  Reichman, 
Fred  B.  Samson,  D.  J.  Schmidly,  Nor- 
man Scott,  Steven  W.  Seagle,  Ray- 
mond D.  Semlitsch,  Henry  L.  Short, 
Lee  H.  Simons,  Graham  W.  Smith, 
Hobart  M.  Smith,  Dan  Speake,  James 
R.  Spotila,  Judy  A.  Stamps,  Thomas 
P.  Sullivan,  Daniel  W.  Uresk,  Laurie 
J.  Vitt,  Peter  D.  Weigl,  Gary  C.  White, 
Daniel  F.  Williams,  Richard  G. 
Zweifel. 

Finally,  we  thank  the  speakers  for 
following  our  schedule  for  submit- 
ting the  various  stages  of  their  manu- 
scripts and  providing  us  with  excel- 
lent manuscripts  in  computer  format 
to  expedite  and  enhance  the  publica- 
tion of  the  proceedings.  The  opinions 
expressed  in  these  papers  are  the  au- 
thors' and  do  not  necessarily  reflect 
those  of  the  U.S.  Department  of  Agri- 
culture. 


30301  Baltimore  SVd 
Beltsv/lle.  MO  20705.2351 


ral  Lfbrary 


USDA  Forest  Service 

General  Technical  Report  RM-166 


November  1988 


Management  of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America 

Proceedings  of  the  Symposium 

July  19-21, 1988 
Flagstaff,  Arizona 

Robert  C.  Szaro,  Kieth  E.  Severson,  and  David  R.  Patton 

technical  coordinators^ 

Sponsored  by: 
Arizona  Chapter  of  the  Wildlife  Society 
Arizona  Game  and  Fish  Department 
Northern  Arizona  University,  School  of  Forestry 
USDA  Forest  Service,  Rocky  Mountain  Forest  and  Range  Experiment  Station 
USDA  Forest  Service,  National  Wildlife  and  Fish  Ecology  Program 
USDA  Forest  Service,  Southwestern  Region 


'Szaro  and  Severson  are  with  the  USDA  Forest  Service,  Rocky  Mountain  Forest  and  Range 
Experiment  Station,  at  the  Station's  Research  VJork  Unit  in  fempe,  in  cooperation  with  Arizona 
State  University.  Patton  is  with  the  School  of  Forestry.  Northern  Arizona  University,  Flagstaff. 


The  Management  of  Amphibians,  Reptiles  and  Small  Mammals 
in  North  America:  Historical  Perspective  and  Objectives 

Robert  C.  Szaro   1 

The  Management  of  Amphibians,  Reptiles  and  Small  Mammals 
in  North  America:  The  Need  for  an  Environmental  Attitude 

J.  Whitfield  Gibbons  4 

Douglas-Fir  Forests  in  the  Oregon  and  Washington  Cascades: 
Relation  of  the  Herpetofauna  to  Stand  Age  and  Moisture 

R.  Bruce  Bury  and  Paul  Stephen  Corn  11 

Long-Term  Trends  in  Abundance  of  Amphibians,  Reptiles,  and 
Mammals  in  Douglas-Fir  Forests  of  Northwestern  California 

Martin  G.  Raphael  23 

Use  of  Woody  Debris  by  Plethodontid  Salamanders  in  Douglas- 
Fir  in  Washington 

Keith  B.  Aubry,  Lawrence  L  C.  Jones,  and  Patricia  A.  Hall   32 

Forestry  Operations  and  Terrestrial  Salamanders:  Techniques  in 
a  Study  of  the  Cow  Knob  Salamander,  Plethodon 
punctafus 


Kurt  A.  Buhlmann,  Christopher  A.  Pague,  Joseph  C.  r\/litchelL 


and  Robert  B.  Glasgow  38 

Conserving  Genetically  Distinctive  Populations:  The  Case  of  the 
Huachuca  Tiger  Salamander  (Ambystoma  tighnum 
sfebbinsi  Lowe) 

James  P.  Collins,  Thomas  R.  Jones,  and  Howard  J.  Berna  45 

Habitat  Requirements  of  New  Mexico's  Endangered 
Salamanders 

Cynthia  A.  Ramotnik  and  Norman  J.  Scott,  Jr.   54 

Utilization  of  Abandoned  Mine  Drifts  and  Fracture  Caves  By  Bats 
and  Salamanders:  Unique  Subterranean  Habitat  in  the 
Ouachita  Mountains 

David  A.  Saugey,  Gary  A.  Heidt,  and  Darrell  R.  Heath  64 

The  Herpetofauna  of  Long  Pine  Key,  Everglades  National  Park, 
in  Relation  to  Vegetation  and  Hydrology 

George  H.  Dalrymple   72 

The  Herpetofaunal  Community  of  Temporary  Ponds  in  North 
Florida  Sandhills:  Species  Composition,  Temporal  Use,  and 
Management  Implications 

C.  Kenneth  Dodd,  Jr.  and  Bert  G.  Charest  87 


(Continued) 


Management  of  Amphibians,  Reptiles,  and  Small  Mammals  in 
Xeric  Pinelands  of  Peninsular  Florida 

/.  Jock  Stout,  Donold  R.  Richordson,  ond  Richord  E.  Roberts  98 

Distribution  and  Habitat  Associations  of  Herpetofauna  in 
Arizona:  Comparisons  by  Habitat  Type 

K.  Bruce  Jones  109 

Multivariate  Analysis  of  thie  Summer  Habitat  Structure  of  Rana 
pipiens  Sctireber,  in  Lac  Saint  Pierre  (Quebec,  Canada) 

N.  Beouregord  ond  R.  Lecloir  Jr.  1 29 

Habitat  Correlates  of  Distribution  of  the  California  Red-Legged 
Frog  (Rona  aurora  draytonii)  and  the  Foothill  Yellow- 
Legged  Frog  (Rana  boylii):  Implications  for  Management 

More  P.  Hoyes  ond  Mork  R.  Jennings  144 

Integrating  Anuran  Amphibian  Species  into  Environmental 
Assessment  Programs 

Ronold  E.  Beiswenger  159 

PrelimirKjry  Report  on  Effect  of  Bullfrogs  on  Wetland 
Herpetofaunas  in  Southeastern  Arizona 

Cecil  R.  Schwolbe  ond  Philip  C.  Rosen   166 

Developing  Management  Guidelines  for  Snapping  Turtles 

Ronold  J.  Brooks,  Dovid  A.  Golbroith,  E.  Grohom  Noncekiveli, 

ond  Christine  A.  Bishop   1 74 

Spatial  Distribution  of  Desert  Tortoises  (Gopherus  agassizii)  at 
Twentynine  Palms,  California:  Implications  for  Relocations 

Ronold  J.  Boxter  180 

Changes  in  a  Desert  Tortoise  (Gopherus  agassizii)  Population 
After  a  Period  of  High  Mortality 

Dovid  J.  Germono  ond  Michele  A.  Joyner  190 

A  Survey  Method  for  Measuring  Gopher  Tortoise  Density  and 
Habitat  Distribution 

Doniel  M.  Spillers  ond  Don  W.  Speoke  199 

Evaluation  and  Review  of  Field  Techniques  Used  to  Study  and 
Manage  Gopher  Tortoises 

Russell  L  Burke  ond  Jomes  Cox   205 

Talus  Use  by  Amphibians  and  Reptiles  in  the  Pacific  Northwest 

Robert  E.  Herrington   216 


Comparison  of  Herpetofounas  of  a  Natural  and  Altered  Riparian 


Ecosystem 

K.  Bruce  Jones  222 

Critical  Habitat,  Predator  Pressures,  and  the  Management  of 
Epicrates  monoensis  (Serpentes:  Boidae)  on  the  Puerto 
Rico  Bank:  A  Multivariate  Analysis 

Peter  J.  Tolson  228 

The  Use  of  Timed  Fixed-Area  Plots  and  a  Mark-Recapture 

Technique  in  Assessing  Riparian  Garter  Snake  Populations 

Robert  C.  Szaro,  Scott  C.  Belfit,  J.  Kevin  Aitkin,  and 

Randall  D.  Babb   239 

Design  Considerations  for  the  Study  of  Amphibians,  Reptiles  and 
Small  Mammals  in  California's  Oak  Woodlands:  Temporal 
and  Spatial  Patterns 

William  M.  Block,  Michael  L  Morrison,  John  C.  Slaymaker, 

and  Gwen  Jongejan  247 

The  Importance  of  Biological  Surveys  in  Managing  Public  Lands 
in  the  Western  United  States 

Michael  A.  Began,  Robert  B.  Finley,  Jr,  and 

Stephen  J.  Petersburg  254 

Sampling  Problems  in  Estimating  Small  Mammal  Population  Size 

George  E.  Menkens,  Jr.  and  Stanley  H.  Anderson   262 

The  Design  and  Importance  of  Long-Term  Ecological  Studies: 
Analysis  of  Vertebrates  in  the  Inyo-White  Mountains, 
California 

Michael  L.  Morrison  267 

An  Ecological  Problem-Solving  Process  for  Managing  Special- 
Interest  Species 

Henry  L.  Short  and  Samuel  C.  Williamson  276 

Comparative  Effectiveness  of  Pitfalls  and  Live-Traps  in 
Measuring  Small  Mammal  Community  Structure 

Robert  C.  Szaro,  Lee  H.  Simons,  and  Scott  C.  Belfit  282 

The  Role  of  Habitat  Structure  in  Organizing  Small  Mammal 
Populations  and  Communities 

Gregory  H.  Adier  289 

Microhabitat  as  a  Template  for  the  Organization  of  a  Desert 
Rodent  Community 

Michael  A.  Bowers  and  Christine  A.  Flanagan   300 


(Continued) 


Response  of  Small  Mammal  Communities  to  Silvicultural 

Treatments  in  Eastem  Hardwood  Forests  of  West  Virginia 
and  Massachusetts 

Robert  T.  Brooks  and  William  M.  Healy  313 

Habitat  Structure  and  ttie  Distribution  of  Small  Mammals  in  a 
Northiern  Hardwoods  Forest 

JefferyA.  Gore  319 

Thie  Value  of  Rocky  Mountain  Juniper  (Juniperus  scopulorum) 
Woodlands  in  Soutti  Dakota  as  Small  Mammal  Habitat 

Carolyn  Hull  Sieg  328 

Postfire  Rodent  Succession  Following  Prescribed  Fire  in  Southern 
California  Chaparral 

William  O.  Wirfz,  II,  David  Hoekman,  John  R.  Muhm,  and 

Sherrie  L  Sauza   333 

Douglas-Fir  Forests  in  the  Cascade  Mountains  of  Oregon  and 
Washington:  Is  the  Abundance  of  Small  Mammals  Related 
to  Stand  Age  and  Moisture? 

Paul  Stephen  Corn,  R.  Bruce  Bury,  and  Thomas  A.  Spies  340 

Evaluation  of  Small  Mammals  as  Ecological  Indicators  of  Old- 
Growth  Conditions 

Kirk  A.  Nordyke  and  Steven  W.  Buskirk  353 

Habitat  Associations  of  Small  Mammals  in  a  Subalpine  Forest, 
Southeastern  Wyoming 

Martin  G.  Raphael  359 

Differences  in  the  Ability  of  Vegetation  Models  to  Predict  Small 
Mammal  Abundance  in  Different  Aged  Douglas- Fir  Forests 

Cathy  A.  Taylor,  C.  John  Ralph,  andArlene  T.  Doyle  368 

Small  Mammals  in  Streamside  Management  Zones  in  Pine 
Plantations 

James  G.  Dickson  and  J.  Howard  Williamson  375 

Patterns  of  Relative  Diversity  Within  Riparian  Small  Mammal 
Communities,  Platte  River  Watershed,  Colorado 

Thomas  E.  Olson  and  Fritz  L.  Knopf  379 

Estimated  Carrying  Capacity  for  Cattle  Competing  with  Prairie 
Dogs  and  Forage  Utilization  in  Western  South  Dakota 

Daniel  W.  Uresk  and  Deborah  D.  Paulson  387 


(Continued) 


Cattle  Grazing  and  Small  Mannmals  on  the  Sheldon  National 
Wildlife  Refuge,  Nevada 

John  L  Oldemeyer  and  Lydia  R.  Allen-Johnson  391 

Effect  of  Seed  Size  on  Removal  by  Rodents 

William  G.  Standley  399 

Habitat  Use  by  Gunnison's  Prairie  Dogs 

C,  N.  Slobodchikoff,  Anthony  Robinson,  and  Clark  Schaack  403 

Environmental  Contaminants  and  the  Management  of  Bat 
Populations  in  the  United  States 

Donald  R.  Clark,  Jr.   409 

Habitat  Structure,  Forest  Composition  and  Landscape 

Dimensions  as  Components  of  Habitat  Suitability  for  the 
Delmarva  Fox  Squirrel 


Raymond  D.  Dueser,  James  L.  Dooley,  Jr.,  and  Gary  J.  Taylor  ....414 


Effects  of  Treating  Creosotebush  with  Tebuthiuron  on  Rodents 

William  G.  Standley  and  Norman  S.  Smith  422 

Foraging  Patterns  of  Tassel-Eared  Squirrels  in  Selected 
Ponderosa  Pine  Stands 

Jacks  States,  William  S.  Gaud,  W.  Sylvester  Allred,  and 

William  J.  Austin   425 

Small  Mammal  Response  to  the  Introduction  of  Cattle  into  a 
Cottonwood  Floodplain 

Fred  B.  Samson,  Fritz  L.  Knopf,  and  Lisa  B.  Mass  432 

Old  Growth  Forests  and  the  Distribution  of  the  Terrestrial 
Herpetofauna 

HartwellH.  Welsh,  Jr.  and  Amy  L.  Lind  439 


The  Management  of 
Amphibians,  Reptiles  and 
Small  Mammals  in  North 
America:  Historical 
Perspective  and  Objectives^ 

Robert  C.  Szaro^ 


Historically  the  management  of  pub- 
lic lands  from  a  multiple  use  perspec- 
tive has  led  to  a  system  that  empha- 
sizes those  habitat  components  or 
faunal  elements  that  primarily  re- 
sulted in  some  sort  of  definable  eco- 
nomic value.  While  this  often  benefit- 
ted other  species  that  were  not  even 
considered  in  the  original  prescrip- 
tions, it  also  negatively  impacted  oth- 
ers. We  no  longer  can  afford  to  take 
this  simplistic  view  of  ecosystem 
management.  We  need  to  use  a  more 
holistic  approach  where  ecological 
landscapes  are  considered  as  units, 
and  land  management  practices  in- 
corporate all  elements  into  an  inte- 
grated policy.  This  includes  examin- 
ing the  impacts  of  proposed  land 
uses  on  amphibian,  reptile,  and  small 
mammal  populations. 

With  the  passage  of  the  National 
Forest  Management  Act  of  1976,  the 
monitoring  of  all  renewable  natural 
resources  became  law.  Even  with  this 
legislation,  most  emphasis  by  Na- 
tional Forests  in  the  United  States  has 
been  placed  on  big  game,  other  game 
species,  or  threatened  and  endan- 
gered species.  Yet,  the  act  lists  five 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff, AZ,  July  19-21  1988). 

^Robert  C.  Szaro  is  Researct)  Wildlife  Bi- 
ologist, USDA  Forest  Service,  Rocky  Moun- 
tain Forest  and  Range  Experiment  Station, 
at  thie  Station's  Research)  Work  Unit  in 
Tempo,  in  cooperation  with  Arizona  State 
University.  Station  Headquarters  is  in  Fort 
Collins,  in  cooperation  with  Colorado  State 
University. 


categories  of  management  indicator 
species:  (1)  endangered  and  threat- 
ened plants  and  animals;  (2)  species 
with  special  habitat  needs;  (3)  species 
commonly  hunted,  fished,  or 
trapped;  (4)  nongame  species  of  spe- 
cial interest;  and  (5)  plant  and  animal 
species  selected  because  their  popu- 
lation changes  are  believed  to  indi- 
cate the  effects  of  management  activi- 
ties on  other  species  of  selected  ma- 
jor biological  communities  or  on  wa- 
ter quality. 

Nongame  birds  have  been  the  first 
group  to  benefit  from  changing  man- 
agement practices  and  public  con- 
cern. The  management  of  nongame 
birds  within  the  National  Forest  Sys- 
tem received  a  big  boost  from  the 
"Symposium  on  Management  of  For- 
est and  Range  Habitats  for  Nongame 
Birds"  held  in  Tucson  in  May  1975 
(Smith  1975).  Since  that  initial  sym- 
posium, four  regional  workshops 
were  held  emphasizing  the  manage- 
ment of  nongame  birds  in  forest  and 
range  habitats  (Degraaf  1978a,  1978b; 
Degraaf  and  Evans  1979;  Degraaf 
and  Tilghman  1980).  There  have  also 
been  Forest  Service  sponsored  sym- 
posia targeting  specific  bird  groups 
such  as  owls  (Nero  et  al.  1987)  and 
birds  using  specific  habitat  features 
such  as  snags  (Davis  et  al.  1983). 

Only  recently  has  the  management 
of  other  nongame  species  gained  in- 
creased recognition.  The  landmark 
symposium  on  "Herpetological 
Communities"  held  in  Lawrence, 
Kansas,  August  1977,  as  part  of  the 


joint  meeting  of  the  Herpetologists' 
League  and  the  Society  for  the  Study 
of  Amphibians  and  Reptiles,  was  the 
first  attempt  to  organize  a  vehicle  for 
the  incorporation  of  papers  dealing 
with  herpetological  communities 
(Scott  1982).  Yet,  as  Gibbons  (this 
volume)  clearly  shows,  little  progress 
has  been  made  in  the  recognition  of 
amphibians,  reptiles,  and  small 
mammals  as  being  important  focal 
points  for  research  and  management 
efforts.  It  is  encouraging  that  recent 
comprehensive  symposia  have  incor- 
porated papers  dealing  with  these 
groups.  There  was  an  entire  session 
on  Amphibians  and  Reptiles  in  the 
symposium  "Riparian  Ecosystems 
and  Their  Management"  (Johnson  et 
al.  1985),  and  almost  30%  of  the 
Southern  Evaluation  Project  Work- 
shop reports  work  on  amphibians, 
reptiles,  and  small  mammals  (Pear- 
son et  al.  1987). 

The  intent  of  this  symposium  was 
to  bring  scientists  and  managers  to- 
gether to  exchange  knowledge  and 
ideas  on  habitat  requirements,  man- 
agement needs,  and  other  informa- 
tion on  these  often  overlooked  com- 
ponents of  North  American  fauna. 
Another  purpose  was  to  summarize 
the  state-of-the-science  of  habitats 
and  habitat  requirements  of  species 
within  these  groups.  Of  particular 
interest  were  papers  emphasizing 
habitat  models,  habitat  requirements, 
sampling  techniques  and  problems, 
community  dynamics,  and  manage- 
ment recommendations. 


1 


The  overwhelming  response  to 
our  announcement  for  papers  was 
unexpected.  More  than  60  abstracts 
were  originally  submitted  for  presen- 
tation. In  order  to  overcome  recent 
criticism  concerning  so-called  "gray" 
literature  (Bart  and  Anderson  1981, 
Capen  1982,  Finch  et  al.  1982,  Scott 
and  Ralph  1988),  we  made  every  ef- 
fort to  improve  the  quality  of  the 
symposium  and  its  subsequent  pro- 
ceedings. All  authors  were  required 
to  submit  their  first  drafts  5  months 
prior  to  the  meeting  in  order  to  en- 
sure adequate  time  for  peer  review 
and  editing.  Each  manuscript  was 
reviewed  by  two  experts  familiar 
with  the  topic,  and  edited  for  style 
and  content  by  one  of  the  sympo- 
sium editors. 

We  found  the  meeting  itself  to  be  a 
fertile  exchange  of  ideas  and  tech- 
niques between  managers  and  re- 
searchers from  all  over  the  country. 
Those  attending  found  the  meeting 
extremely  enlightening  both  for  re- 
searchers and  managers  because  of 
their  exposure  to  new  viewpoints.  It 
is  a  testament  to  those  attending  and 
the  quality  of  the  presentations  that 
very  little  discussion  occurred  out- 
side the  meeting  hall  when  papers 
were  in  progress.  Virtually  all  partici- 
pants were  present  throughout  the 
symposium,  from  the  first  session  to 
the  last. 

We  hope  this  symposium  will 
prove  to  be  the  boost  that  these  fau- 
nal  groups  need  to  get  increased  re- 
search and  management  recognition. 
For  only  with  an  adequate  data  base 
can  models  be  developed  that  predict 
diversity  in  relation  to  natural  or 
man-made  disturbance  of  ecosys- 
tems. These  holistic  models  are  of  the 
utmost  importance  for  the  mainte- 
nance of  worldwide  biodiversity 
(Wilson  and  Peters  1988).  Ecosystem 
diversity  is  a  key  correlate  with  bio- 
logical productivity  and  has  recently 
attracted  considerable  interest  both 
from  theoreticians  and  from  profes- 
sionals concerned  with  management 
of  land  and  water  systems  (Suffling 
et  al.  1988).  We  feel  that  amphibians. 


reptiles,  and  small  mammal  popula- 
tions may  prove  to  be  the  ultimate 
indicators  of  habitat  quality  and 
health,  because  of  their  sedentary 
characteristics  which  make  them 
much  more  susceptible  to  manage- 
ment activities  than  do  highly  mobile 
bird  species  and  ubiquitous  species 
such  as  deer  and  turkey. 

Literature  Cited 

Bart,],  and  D.  R.  Anderson.  1981. 
The  case  against  publishing  sym- 
posia proceedings.  Wildlife  Soci- 
ety Bulletin  9:201-202. 

Capen,  David  E.  1982.  Publishing 
symposia  proceedings:  another 
viewpoint.  Wildlife  Society  Bulle- 
tin 10:183-184. 

Davis,  Jerry  W.,  Gregory  A.  Good- 
win, and  Richard  A.  Ockenfeis 
(Technical  Coordinators).  1983. 
Snag  habitat  management:  Pro- 
ceedings of  the  symposium. 
USDA  Forest  Service  General 
Technical  Report  RM-99.  Rocky 
Mountain  Forest  and  Range  Ex- 
periment Station,  Ft.  Collins,  Colo. 
226  p. 

Degraaf,  Richard  M.  (Technical  Coor- 
dinator). 1978a.  Proceedings  of  the 
workshop  on  nongame  bird  habi- 
tat management  in  the  coniferous 
forests  of  the  western  United 
States.  USDA  Forest  Service  Gen- 
eral Technical  Report  PNW-64. 
Pacific  Northwest  Forest  and 
Range  Experiment  Station,  Port- 
land, Oregon.  100  p. 

Degraaf,  Richard  M.  (Technical  Coor- 
dinator). 1978b.  Proceedings  of  the 
workshop:  Management  of  south- 
ern forests  for  nongame  birds. 
USDA  Forest  Service  General 
Technical  Report  SE-14.  Southeast- 
ern Forest  Experiment  Station, 
Asheville,  North  Carolina.  176  p. 

Degraaf,  Richard  M.  and  Keith  E. 
Evans  (Proceedings  Compilers). 
1979.  Management  of  north  central 
and  northeastern  forests  for 
nongame  birds.  USDA  Forest 
Service  General  Technical  Report 


NC-51.  North  Central  Forest  Ex- 
periment Station,  St.  Paul,  Minn. 
268  p. 

Degraaf,  Richard  M.  and  Nancy  G. 
Tilghman  (Proceedings  Compil- 
ers). 1980.  Workshop  proceedings: 
Management  of  western  forests 
and  grasslands  for  nongame  birds. 
USDA  Forest  Service  General 
Technical  Report  INT-86.  Inter- 
mountain  Forest  and  Range  Ex- 
periment Station,  Ogden,  Utah. 
535  p. 

Finch,  Deborah  M.,  A.  Lauren  Ward, 
and  Robert  H.  Hamre.  1982.  Com- 
ments in  defense  of  symposium 
proceedings:  response  to  Bart  and 
Anderson.  Wildlife  Society  Bulle- 
tin 10:181-183. 

Johnson,  R.  Roy,  Charles  D.  Ziebel, 
David  R.  Patton,  Peter  F.  Ffolliott, 
and  Robert  H.  Hamre  (Technical 
Coordinators).  1985.  Riparian  eco- 
systems and  their  management: 
reconciling  conflicting  uses.  First 
North  American  Riparian  Confer- 
ence. USDA  Forest  Service  Gen- 
eral Technical  Report  RM-120. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Ft.  Collins, 
Colo.  523  p. 

Nero,  Robert  W.,  Richard  J.  Clark, 
Richard  J.  Knapton,  and  R.  H. 
Hamre  (Editors).  1987.  Biology 
and  conservation  of  northern  for- 
est owls.  USDA  Forest  Service 
General  Technical  Report  RM-142. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Ft.  Collins, 
Colo.  309  p. 

Pearson,  Henry  A.,  Fred  E.  Smeins, 
and  Ronald  E.  Thill  (Proceedings 
Compilers).  1987.  Ecological, 
physical,  and  socioeconomic  rela- 
tionships within  southern  national 
forests:  Proceedings  of  the  south- 
ern evaluation  workshop.  USDA 
Forest  Service  General  Technical 
Report  SO-68.  Southern  Forest  Ex- 
periment Station,  New  Orleans, 
Louisiana.  293  p. 

Scott,  J.  Michael  and  C.  John  Ralph. 
1988.  Quality  control  of  symposia 
and  their  published  proceedings. 
Wildlife  Society  Bulletin  16:68-74. 


2 


Scott,  Norman  J.,  Jr.  1982.  Herpeto- 
logical  communities.  USDI  Fish 
and  Wildlife  Service,  Wildlife  Re- 
search Report  13.  239  p. 

Smith,  Dixie  R.  (Technical  Coordina- 
tor). 1975.  Proceedings  of  the  sym- 
posium on  management  of  forest 
and  range  habitats  for  nongame 
birds.  USDA  Forest  Service  Gen- 
eral Technical  Report  WO-1. 
Washington,  D.C.  343  p. 

Suffling,  Roger,  Catherine  Lihou,  and 
Yvette  Morand.  1988.  Control  of 
landscape  diversity  by  cata- 
strophic disturbance:  a  theory  and 
a  case  study  in  a  Canadian  Boreal 
Forest.  Environmental  Manage- 
ment 12:73-78. 

Wilson,  E.  O.  (Editor)  and  Frances  M. 
Peter  (Associate  Editor).  1988.  Bio- 
diversity. National  Academy 
Press,  Washington,  D.C.  521  p. 


■-. ;.. 


3 


The  Management  of 
Amphibians,  Reptiles  and 
Small  Mammals  in  North 
America:  The  Need  for  an 
Environmental  Attitude 
Adjustment^ 


Abstract.—  Amphibians,  reptiles,  and  smail 
mammals  need  special  consideration  in 
environmental  management  and  conservation 
because  (1)  they  are  significant  biotic  components 
in  terrestrial  and  freshwater  habitats;  (2)  research 
and  management  efforts  have  lagged  behind  those 
on  other  vertebrates;  (3)  a  stronger  understanding 
of  their  ecology  and  life  history  is  needed  to  guide 
management  decisions;  and  (4)  their  importance 
has  not  been  promoted  satisfactorily  to  develop  the 
proper  public  attitude. 


J.  Whitfield  Gibbons^ 

My  objective  is  to  provide  an  over- 
view and  perspective  of  the  amphibi- 
ans, reptiles,  and  small  mammals  of 
North  America  as  a  group  that  de- 
serves more  careful  consideration 
from  an  environmental  management 
and  conservation  standpoint.  The 
justification  of  the  need  for  and  time- 
liness of  a  careful  examination  of  am- 
phibian, reptile,  and  small  mammal 
assemblages  is  based  on  the  premises 
stated  below.  One  intent  is  to  bring 
the  problem  into  focus  so  that  both 
scientists  and  managers  can  identify 
problem  areas  and  conjoin  in  an  ef- 
fort that  will  result  in  the  manage- 
ment of  these  animals  in  North 
America  in  a  prudent  and  far-sighted 
manner. 

I  offer  four  premises  to  support 
the  contention  that  amphibians,  rep- 
tiles, and  small  mammals  deserve 
special  attention  with  regard  to  man- 
agement considerations: 

1.  Amphibians,  rephles,  and 
small  mammals  are  a  signifi- 
cant and  important  wildlife 
component  of  the  fauna  in 
most  terrestrial  and  freshwa- 
ter habitats  in  North  Amer- 
ica. 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff, AZ,  July  19-21,1988). 

'J.  Whitfield  Gibbons,  Head,  Division  of 
Stress  and  Wildlife  Ecology,  Savannah  River 
Ecology  Laboratory,  Drawer  E,  Aiken,  SC 
29801. 


2.  Research  and  management 
publication  efforts  as  well  as 
funding  have  lagged  behind 
those  of  many  of  the  more 
obvious  faunal  components 
(e.g.,  game  species  of  large 
mammals,  birds  and  fishes, 
and  many  insects,  because  of 
their  importance  as  pests). 

3.  The  direct  empirical  meas- 
urements of  habitat  require- 
ments, species  interactions, 
and  life  history  patterns 
needed  for  proper  manage- 
ment are  often  lacking  for 
amphibians,  reptiles,  and 
small  mammals. 

4.  An  attitude  that  amphibians, 
reptiles,  and  small  mammals 
should  be  of  concern  in  envi- 
ronmental management  deci- 
sions has  not  been  satisfacto- 
rily instilled  among  some 
managers,  the  general  public, 
and  political  officials. 


Support  for  Premises 

Premise  1  —Amphibians,  reptiles, 
and  small  mammals  are  a 
significant  and  important  wildlife 
component  in  North  American 
ecosystems. 

One  way  for  a  taxonomic  group  or 
species  assemblage  to  qualify  as  im- 


portant to  an  environmental  manager 
is  to  be  identified  as  making  a  major 
contribution  to  biological  complexity 
in  terms  of  species  diversity,  trophic 
dynamics,  and  interactions  within 
communities.  Some  groups  clearly 
have  the  potential  for  overall  com- 
munity influence  by  virtue  of  abun- 
dance. Salamanders  at  Hubbard 
Brook  were  demonstrated  to  have  a 
higher  biomass  than  other  vertebrate 
groups  (Burton  and  Likens  1975). 
The  capture  of  as  many  as  88,000 
amphibians  in  one  year  (SREL  Report 
1980)  and  large  numbers  in  most 
years  (Pechmann  et  al.  1988)  at  a 
1  ha  temporary  pond  in  South  Caro- 
lina suggest  that  they  dominate  the 
higher  trophic  level  in  some  habitats. 
Other  studies  support  the  postula- 
tion  that  amphibians  are  often  the 
top  predators  in  some  aquatic  sys- 
tems (Taylor  et  al.  in  press).  Freshwa- 
ter turtles  represent  the  majority  of 
vertebrate  biomass  in  many  aquatic 
habitats  (Congdon,  Greene,  and  Gib- 
bons 1986),  and  their  potential  sig- 
nificance as  vectors  for  seeds  and 
parasites  among  temporary  aquatic 
habitats  has  been  suggested  (Cong- 
don and  Gibbons  1988).  Box  turtles 
(Terrapene  Carolina)  have  also  been 
implicated  as  seed  vectors  (Braun 
and  Brooks  1987).  Small  rodents  are 
noted  for  their  impact  on  plant  com- 
munities under  certain  environ- 
mental conditions  (Hayward  and 
Phillipson  1979);  desert  granivores 
affect  the  density,  biomass,  and  com- 
position of  annual  plants  (Brown  et 


4 


Table  1  .—Publications  on  different  taxonomic  groups  in  major  North  American 
journals  in  general  ecology  and  wildlife  ecology.  Issues  from  1983-1988  were 
selected  at  random  until  200  titles  were  chosen.  Assignment  to  taxonomic 
categories  was  based  on  the  appearance  of  study  organism  names  in  the 
titles.  Not  all  papers  used  in  tabulation  were  based  on  North  American  fauna. 
The  definition  of  small  mammals  is  that  used  in  this  Symposium. 


JOURNAL 
(total) 

A 

R 

S 

L 

F 

B 

1 

(218)  AMN 

9 

12 

43 

14 

43 

18 

61 

(201)  ECOL 

10 

22 

28 

10 

24 

50 

58 

(213)  CJZ 

8 

9 

36 

39 

34 

53 

34 

(614)  Total 

27 

42 

107 

63 

101 

121 

153 

% 

4 

7 

17 

10 

16 

20 

25 

(139)  Hsr 

1 

4 

4 

13 

50 

67 

(204)  JWM 

0 

2 

6 

103 

93 

(343)  Total 

1 

6 

110 

126 

50 

160 

% 

<1 

2 

3 

37 

15 

47 

A=  Amphibians 

R  =  Reptiles 

S  =  Small  mammals 

L  =  Large  mammals 

F  =  Fishes 

B  =  Birds 

I  =  Insects 


General  Ecology 

AMN  =  American  Midland  Naturalist 
ECOL  =  Ecology 

CJZ    =  Canadian  Journal  of  Zoology 
Wildlife  Ecology 

HSI     =  U.S.  Fish  and  Wildlife  Service  Habitat 

Suitability  Index  Models 
JWM  =  Journal  of  Wildlife  Management 


*Only  139  fifles  were  available. 


al.  1986).  These  represent  only  a  few 
of  the  available  examples  for  am- 
phibians, reptiles,  and  small  mam- 
mals; however,  many  more  studies 
are  needed  that  document  the  role 
and  importance  of  species  in  these 
groups  in  enhancing  biological  com- 
plexity. 

Another  way  for  a  group  to  as- 
sume importance  is  for  it  to  have  a 
direct,  measurable  economic  value  or 
impact.  Several  examples  can  be 
given  of  the  importance  of  amphibi- 
ans, reptiles,  and  small  mammals 
from  the  economic  perspective,  but 
their  impact  has  been  trivial  in  com- 
parison to  large  game  mammals  or 
insect  pests,  and  controls  and  regula- 
tions have  been  comparatively  loose. 
The  limited  economic  importance  of 
most  small  terrestrial  or  semi-aquatic 
vertebrates  is  presumably  one  expla- 
nation for  their  being  given  minimal 
attention  in  many  management 
schemes.  A  few  species  such  as 
American  alligators  (Joanen  and 
McNease  1987),  bullfrogs  (Shifter 
1987),  and  snapping  turtles  (Bushey, 
no  date)  are  commercially  important 
as  human  food  items.  Other  species 
assume  an  economic  value  in  the  le- 
gal pet  trade  (Conant  1975)  or  as  re- 
search animals  sold  by  biological 
supply  houses  (Carolina  Biological 
Supply  1987).  Some  venomous 
snakes,  especially  eastern  (Crotalus 
adamanteus)  and  western  (C.  atrox) 
diamondback  rattlesnakes,  are  an 
economic  irony  in  that  the  venom  is 
necessary  to  make  antivenin  (Parrish 
1980).  Of  course,  such  species 
achieve  some  level  of  importance 
simply  by  being  potentially  injurious. 
Small  mammals  have  been  indicted 
in  a  variety  of  situations  for  negative 
economic  impacts,  such  as  prairie 
dog  damage  (Walker  1983),  rabies  in 
bats  (Constantine  1970),  and  grain- 
eating  by  rodents  (Rowe  1981). 

Another  measure  of  importance  of 
some  species  or  groups  is  the  intan- 
gible aesthetic  value  that  some 
people  place  on  them.  Many  species 
assume  an  undeniable  importance  to 
many  people  and  may  ultimately  ac- 


quire protected  status.  Legal  protec- 
tion of  "the  species"  often  provides 
protection  to  certain  habitats.  This 
circle  of  protection  is  a  factor  that  can 
work  to  great  advantage  for  those 
persons  interested  in  preservation — 
the  species  is  protected  because  it  is 
important  (aesthetic)  and  becomes 
even  more  important  (legal)  because 
it  is  protected  and  results  in  preser- 
vation of  the  habitat.  For  example, 
the  legal  status  offered  the  desert  tor- 
toise {Xerobates  agassizi;  Luckenbach 
1982)  and  the  Morro  Bay  kangaroo 
rat  (Dipodomys  heermanni  morroensis; 
USDI  1980)  in  California  or  the 
American  crocodile  (Crocodylus 
acutus;  Kushlan  and  Mazzotti  1986) 
in  Rorida  serves  to  provide  some 
level  of  environmental  protection  for 
the  entire  community  where  they  oc- 
cur. The  protection  given  the  black 
footed  ferret  has  resulted  in  protec- 
tion of  its  prey.  The  World  Wildlife 


Fund  recognizes  this  effect  in  its  con- 
servation programs  by  designating 
"flagship"  species  such  as  great  apes 
or  monkeys,  for  which  funds  are 
more  easily  raised,  in  order  to  pro- 
tect entire  communities  or  ecosys- 
tems. 


Premise  2— Ecological  research  on 
herpetofouna  and  small 
mammals  has  lagged  behind  that 
of  other  animal  groups. 

Support  for  the  contention  that  the 
level  of  ecological  research  on  am- 
phibians, reptiles,  and  small  mam- 
mals is  lower  than  that  of  certain 
other  animal  groups  can  be  given  in 
several  ways.  These  include  annual 
publications  on  particular  groups 
(table  1)  and  the  proportion  of 
funded  grants  that  fall  into  each  cate- 
gory (table  2). 


5 


The  reasons  for  the  lower  levels  of 
publication  and  funding  in  research 
on  amphibians,  reptiles,  and  small 
mammals  are  varied  and  in  part  con- 
jectural. One  seemingly  obvious  rea- 
son is  that  most  species  in  these 
groups  have  low  profile  in  health, 
hunting,  agricultural,  or  other  eco- 
nomic issues  and  therefore  receive 
minimal  attention  from  some  quar- 
ters. The  comparatively  low  level  of 
attention  given  to  small,  non-game 
terrestrial  and  semi-aquatic  verte- 
brates by  certain  sectors  of  society  is 
reflected  in  lower  overall  funding 
and  subsequently  in  fewer  general 
publications. 

Research  funding  is  inequitable 
because  of  the  emphasis  on  species 
that  have  important  economic  status; 
thus,  the  life  history  and  ecology  of 
even  moderately  abundant  herpe- 
tofaunal  or  small  mammal  species 
are  seldom  understood  at  a  level  that 
would  permit  prudent  management. 
Even  those  with  potential  economic 
importance  receive  less  emphasis 
than  many  birds,  large  mammals, 
and  fish.  As  an  example,  the  Ameri- 
can alligator  represents  a  reptile  spe- 
cies of  vital  concern  from  a  manage- 


ment standpoint,  yet  the  number  of 
publications  that  focus  on  the  life  his- 
tory, ecology,  behavior,  and  genetics 
of  the  sp)ecies  is  limited  (see  Brisbin 
et  al.  1985)  compared  to  the  hun- 
dreds on  large  mammal  game  species 
such  as  white-tailed  deer  (Halls  1984; 
Johns  and  Smith  1985). 


Premise  3— The  basic  ecological 
and  life  history  information 
necessary  to  make  thoughtful 
environmental  management 
decisions  is  often  absent  for  many 
of  the  amphibians,  reptiles,  and 
small  mammals  in  a  community. 

As  indicated  above,  the  research  ef- 
fort directed  toward  amphibians, 
reptiles,  and  small  mammals  by 
ecologists  appears  to  be  below  that 
for  other  vertebrate  groups.  Al- 
though difficult  to  measure,  it  would 
also  be  expected  that  the  fundamen- 
tal data  bases  necessary  for  thought- 
ful management  decisions  would  ex- 
ist in  lower  proportions  for  herpe- 
tofaunal  and  small  mammal  species. 
One  reason  is  that,  compared  to 
many  large  mammals,  birds,  and 


Table  2.— Number  of  grant  proposals  funded  by  selected  U.S.  granting  agen- 
cies on  particular  groups  of  animals. 


A 

R 

S 

L 

F 

B 

1 

NSF(1987) 

1 

3 

4 

2 

11 

7 

24 

Sigma  Xi 

4 

8 

9 

7 

9 

24 

10 

(March  1987) 

National  Geograpl-iic 

0 

3 

2 

28 

8 

15 

14 

(1988) 

World  Wildlife  Fund 

0 

15 

0 

49 

0 

14 

1 

(1987-1988) 

Total 

5 

29 

15 

86 

28 

60 

49 

%  2 

11 

6 

32 

10 

22 

18 

A=Amphibians 

R=Reptiles 

S=Small  mammals 

L=Large  mammals 

F=Fishes 

B=Birds 

l=lnsects 


fishes,  certain  aspects  of  field  studies 
on  many  of  the  amphibians,  reptiles, 
and  small  mammals  are  sometimes 
perceived  as  being  more  difficult  be- 
cause of  factors  such  as  small  body 
or  population  sizes,  fossorial  or  cryp- 
tic habits,  patchy  distribution,  and 
unpredictable  seasonality.  Conse- 
quently, fewer  papers  are  likely  to  be 
published  in  general  ecology  journals 
that  expect  quantitative  ecological 
and  life  history  research  results 
rather  than  ones  that  are  descriptive 
and  qualitative.  An  exception  to  this 
may  be  manipulative  field  experi- 
ments in  which  small  rodents  have 
been  used  in  almost  half  of  the  stud- 
ies involving  vertebrates. 

The  actual  or  apparent  rarity  or 
unpredictability  of  occurrence  of 
many  amphibian,  reptile,  and  small 
mammal  species  makes  it  difficult  or 
impossible  for  the  research  ecologist 
to  gather  useful  data  without  a  fund- 
ing base  that  is  accepting  of  the  un- 
certainty of  whether  data  will  actu- 
ally be  forthcoming  in  a  particular 
year.  The  environmental  manager  in 
turn  cannot  incorporate  such  species 
into  a  management  plan,  and  thus 
their  perceived  imp>ortance  is  dimin- 
ished. The  unpredictability  of  occur- 
rence of  some  species  can  be  demon- 
strated with  amphibians  and  reptiles 
on  the  Savannah  River  Plant  (SRP)  in 
South  Carolina.  In  spite  of  more  than 
a  quarter  of  a  century  of  field  studies 
and  the  capture  of  more  than  half  a 
million  reptiles  and  amphibians 
across  all  available  habitats,  species 
previously  unreported  from  the  SRP 
continue  to  be  discovered  (Gibbons 
and  Semlitsch  1988;  Young  1988).  Or, 
some  species  have  gone  for  intervals 
as  long  as  one  decade  (e.g.,  pickerel 
frog.  Ram  palustris)  or  two  (e.g., 
glossy  water  snake,  Regina  rigida)  be- 
tween sightings  (Gibbons  and  Sem- 
litsch 1988).  Clearly,  developing  a 
basic  ecological  field  study  on  such 
species  in  a  region  is  not  feasible  un- 
der typical  funding  situations. 

Resolutions  to  the  problem  of  gar- 
nering information  about  rare  sp>ecies 
include  intensifying  survey  efforts  in 


6 


geographic  regions  of  interest  by 
supporting  long-term  research  pro- 
grams that  can  ultimately  reveal  the 
presence  of  rare  or  fossorial  species. 
Once  a  species  is  identified  to  be 
present  in  a  habitat,  the  decision 
should  be  made  on  whether  an  eco- 
logical research  effort  is  warranted. 

Long-term  studies  may  be  neces- 
sary to  reveal  certain  life  history 
traits,  even  about  common  species, 
because  of  the  inherent  variability  in 
some  life  history  features  that  can 
result  from  natural  environmental 
variation  (Semhtsch  et  al.  1988).  Such 
studies  may  be  essential  to  identify 
the  extent  of  variability  due  to  an- 
nual weather  patterns  and  climatic 
variation  (Semlitsch  1985;  Pechmann 
et  al.  1988).  Long-term  research  pro- 
grams may  be  needed  because  some 
species  are  long-  lived,  or  in  the  case 
of  many,  because  the  potential  lon- 
gevity is  great  but  unknown  (Gib- 
bons 1987). 

For  many  species  that  have  eco- 
nomic value  (e.g.,  snapping  turtle, 
Chelydra  serpentina;  Congdon  et  al. 
1987),  the  impact  of  harvesting  has 
not  been  properly  assessed.  Because 
of  the  limited  baseline  ecological  and 
life  history  data  for  most  species,  a 
priority  goal  should  be  the  establish- 
ment of  a  moratorium  on  the  whole- 
sale removal  of  all  native  species  of 
amphibians,  reptiles,  and  small 
mammals  until  it  can  be  verified  that 
regional  populations  can  sustain  the 
removal  rate.  State  permits  should  be 
required  of,  and  possession  limits 
should  be  set  for,  all  commercial  col- 
lectors for  all  species  of  amphibians, 
reptiles,  and  small  mammals. 

Today's  emphasis  should  be  on 
protection  of  each  species  until  con- 
vincing evidence  is  supplied  that  har- 
vesting has  no  long-term  impact, 
rather  than  placing  the  burden  on 
herpetologists  and  mammalogists  to 
demonstrate  population  irrecovera- 
bility  before  harvesting  is  discontin- 
ued. The  negative  consequences  of 
the  latter,  and  current,  approach  (i.e., 
demonstrating  the  impact  of  removal 
while  harvesting  is  in  progress)  is 


that  some  populations  will  be  re- 
duced to  the  point  of  no  recovery  be- 
fore the  necessary  evidence  can  be 
collected.  Each  species  should  be 
protected  until  proven  harvestable. 
The  appropriate  basic  research 
should  be  conducted  by  scientists 
with  no  economic  or  emotional  in- 
vestment in  the  outcome.  Research 
support  should  be  provided  by  state 
or  federal  agencies  and  by  special 
interest  groups  that  have  no  influ- 
ence over  the  final  management  deci- 
sions. The  ideal  approach  is  that  sci- 
entists would  gather  the  facts  and 
that  environmental  managers  would 
interpret  them  in  the  context  of  har- 
vesting quotas.  The  development 
and  use  of  predation  (Holling  1966) 
or  harvest  (Ricker  1975)  models  may 
be  effective  approaches  for  address- 
ing the  issue  of  human  predation 
(i.e.,  harvestability  by  man). 

One  area  that  deserves  attention  in 
strengthening  the  study  of  small  ter- 
restrial or  semi-aquatic  vertebrates  is 
the  use  of  innovative  techniques  to 
address  physiological,  ecological, 
and  behavioral  questions  under  natu- 
ral conditions.  Non-destructive  field 
sampling  techniques  are  critical  in 
the  study  of  both  rare  and  endan- 
gered species  but  are  also  important 
for  preserving  the  integrity  of  any 
study  population.  These  include 
techniques  for  capture,  field  identifi- 
cation of  individuals,  non-disruptive 
handling  or  observation,  recapture, 
and  the  acquisition  of  non-  destruc- 
tive physiological,  genetic,  behav- 
ioral, and  life  history  data.  Some  ex- 
amples include  radiography  (Gib- 
bons and  Greene  1978)  or  sonogra- 
phy for  determination  of  clutch  sizes, 
blood  sampling  for  genetic  and 
hormonal  analyses  (Scribner  et  al. 
1986),  and  cyclopropane  for  measur- 
ing lipid  levels  (Peterson  1988).  A 
broader  use  of  such  techniques  in 
field  studies  could  strengthen  the 
foundation  of  ecological  and  life  his- 
tory understanding  that  is  necessary 
for  environmental  management. 

A  direct  contribution  to  environ- 
mental managers  could  be  achieved 


by  attempts  to  verify  the  several  am- 
phibian, reptile,  and  small  mammal 
Habitat  Suitability  Index  models  of 
the  U.S.  Fish  and  Wildlife  Depart- 
ment. The  concept  has  the  potential 
value  of  providing  an  initial  quanti- 
tative approach  that  gives  a  tangible 
product.  However,  to  be  of  greatest 
value,  the  HSI  models  must  be  evalu- 
ated and  modified  as  appropriate.  It 
is  perhaps  noteworthy  that  the  HSI 
models  prepared  for  amphibians  (1), 
reptiles  (4),  and  small  mammals  (4) 
collectively  represent  only  6%  of  the 
139  that  have  been  completed  on  ver- 
tebrates (table  1).  For  these  to  be- 
come an  effective  tool  in  manage- 
ment of  herpetofauna  and  small 
mammals,  more  herpetologists  and 
mammalogists  need  to  volunteer  to 
develop  HSI  models  for  these 
groups. 

A  distinction  must  be  made  be- 
tween (1)  problem  oriented  applied 
research  on  specific  systems  that  re- 
lies on  qualitative  assessments  or  in- 
direct measurements  of  variables 
with  minimal  inference  power  and 
(2)  basic  research  that  is  founded  on 
quantitative  or  direct  measurements 
of  variables,  has  a  conceptual  or 
theoretical  base  or  orientation,  and 
can  be  strongly  inferential  through 
general  field  or  laboratory  experi- 
ments. The  latter  approach  will  be 
necessary  if  environmental  managers 
are  to  have  a  reliable  data  base  that  is 
founded  on  broad  applicability,  lev- 
els of  predictability,  and  clear  direc- 
tions for  future  research. 


Premise  4— The  attitude  of  most 
people  in  North  America  toward 
most  amphibians,  reptiles,  and 
small  mammals  is  either  negative 
or  neutral,  in  part  because  efforts 
to  develop  an  attitude  change 
have  been  insufficient  or 
ineffective. 

Although  documentation  is  difficult, 
it  would  appear  that  in  North  Amer- 
ica we  are  far  from  a  suitable  accep- 
tance level  toward  these  groups  of 


7 


organisms.  People  still  try  to  run 
over  snakes  on  highways,  have  little 
awareness  that  many  conspicuous 
predators  rely  on  small  mammals  for 
their  basic  diets,  and  give  no  thought 
to  how  many  small  vertebrates  will 
be  eliminated  by  the  draining  of  a 
swamp  or  damming  of  a  stream.  I 
think  the  situation  is  an  embarrassing 
one  for  the  scientists  and  general 
public  of  a  nation  that  espouses  edu- 
cation and  knowledge. 

Evidence  that  a  more  positive  atti- 
tude and  less  environmental  leniency 
has  developed  over  the  last  several 
years  is  the  recent  federal  listings  of 
snakes  (e.g.,  indigo  snake,  Dry- 
marchon  corias;  San  Francisco  garter 
snake,  Thamnophis  sirtalis  tetrataenia) 
and  small  rodents  (e.g.,  Utah  prairie 
dog,  Cynomys  parvidens;  salt  marsh 
harvest  mouse,  Reithrodontomys 
raviventris;  Key  Largo  cotton  mouse, 
Peromyscus  gossypinus  allapaticola)  as 
protected  species.  However,  many  of 
the  listings  involving  amphibians, 
reptiles,  and  small  mammals  have 
been  hard  fought  ones  against  public 
and  political  opinions  that  such  spe- 
cies hardly  deserve  such  concessions. 
The  failed  efforts  at  protection  far 
outnumber  the  successful  ones.  The 
attitude  that  these  animals  are  unim- 
portant is  pervasive  throughout  the 
general  public,  politicians,  and  even 
some  environmental  managers.  The 
basic  responsibility  for  eliminating 
ignorance  and  effecting  the  proper 
environmental  attitude  adjustment 
must  start  with  the  scientist. 

It  is  my  firm  opinion  that  many 
scientists  have  lost  sight  of  who  their 
patrons  are  (for  most  of  us,  the  U.S. 
taxpayers)  and  of  their  responsibility 
to  communicate  findings  to  all  levels 
of  society.  This  communication  proc- 
ess entails  a  level  of  cooperation  and 
an  educational  spirit  that  allows  each 
individual  to  contribute  in  the  most 
effective  manner.  However,  we  must 
all  accept  and  work  toward  the  com- 
mon goals  of  establishing  a  thorough 
and  general  foundation  of  ecological 
information  for  amphibians,  reptiles, 
and  small  mammals  and  of  being 


generous  in  the  distribution  of  the 
findings  in  a  form  palatable  to  and 
usable  by  the  intended  audience. 

Conclusions 

An  environmental  attitude  adjust- 
ment model  must  be  developed  and 
promoted  that  considers  where  we 
want  to  end  up,  who  we  must  edu- 
cate and  influence,  and  what  we 
must  know  and  do  to  achieve  the 
goal  of  education  in  a  convincing 
manner.  The  desired  end  point  is  a 
nationwide  attitude  among  scientists, 
managers,  politicians,  and  the  public 
that  amphibians,  reptiles,  and  small 
mammals  are  critical  wildlife  compo- 
nents. Each  species  population  and 
community  must  be  identified  as 
having  an  intrinsic  value  in  maintain- 
ing the  integrity  of  the  natural  eco- 
systems of  North  America. 

Scientists  have  a  responsibility  for 
collecting  extensive  and  intensive  in- 
formation on  the  life  history  patterns 
and  habitat  requirements  of  native 
amphibians,  reptiles,  and  small 
mammals.  The  required  data  must  be 
collected  in  a  rigorous  experimental 
manner  that  promotes  an  under- 
standing of  these  species  and  com- 
munities through  strong  inferences 
and  syntheses. 

Politicians  have  a  responsibility  to 
assure  that  the  approval  of  a  govern- 
ment project  is  as  contingent  on  envi- 
ronmental consequences  as  on  budg- 
etary considerations.  Our  attitude 
must  graduate  to  become  one  of  ac- 
ceptance of  a  proposed  project  only 
after  environmental  impact  determi- 
nations have  led  to  an  objective  deci- 
sion that  the  gain  from  the  project 
warrants  the  loss  to  the  environment. 

Managers  have  a  responsibility  for 
promoting  basic  research,  for  apply- 
ing the  findings  to  habitat  manage- 
ment, and  for  having  the  patience  to 
wait  for  the  completion  of  long-term 
studies  as  required.  In  situations 
where  removal  of  animals  or  elimina- 
tion of  habitat  is  an  issue,  the  burden 
of  proof  should  be  borne  by  the  har- 


vester or  developer,  and  not  by  the 
scientist  or  manager.  The  status  of  a 
species  should  be  determined  before 
the  decision  to  proceed  is  made,  cer- 
tainly not  after  harvesting  begins  or 
during  the  physical  development  of  a 
project.  This  assessment  should  be 
made  and  evaluated  before  the  proj- 
ect is  approved.  Each  species  should 
be  protected  until  proven  harves- 
table. 

Both  scientists  and  managers  have 
a  responsibility  to  inform  the  public 
and  political  arena  that  the  protec- 
tion and  ecological  understanding  of 
inconspicuous  and  non-game  species 
are  vital  to  proper  ecosystem  man- 
agement and  to  the  preservation  and 
maintenance  of  North  America's 
natural  heritage. 

Acknowledgments 

I  thank  Justin  D.  Congdon,  Nat  B. 
Frazer,  Trip  Lamb,  William  D. 
McCort,  Joseph  H.  K.  Pechmann, 
David  E.  Scott  and  Raymond  D. 
Semlitsch  for  commenting  on  the 
original  manuscript.  I  appreciate  the 
efforts  of  Marianne  Reneau,  Marie 
Fulmer,  Jeff  Lovich,  Tony  Mills,  and 
Tim  Owens  in  manuscript  prepara- 
tion. Manuscript  preparation  was 
aided  by  Contract  DE-  AC09- 
76SR00819  between  the  U.S.  Depart- 
ment of  Energy  and  the  University  of 
Georgia's  Savannah  River  Ecology 
Laboratory. 

Literature  Cited 

Brisbin,  I.  Lehr,  Charles  A.  Ross, 
M.C.  Downes,  Mark  A.  Staton  and 
Brad  R.  Gammon.  1986.  A  bibliog- 
raphy of  the  American  alligator 
(Alligator  mississippiensis).  Savan- 
nah River  National  Environmental 
Research  Park  Publ.  Aiken,  SC. 

Brown,  J.  H.,  D.  W.  Davidson,  J.  C. 
Munger,  and  R.  S.  Inouye.  1986. 
Experimental  community  ecology: 
The  desert  granivore  system, 
pp.  41-61.  In:  Community  Ecol- 


8 


ogy.  J.  Diamond  and  T.  J.  Case, 
editors.  Harper  and  Row.  NY. 

Brown,  J.  and  G.  R.  Brooks,  Jr.  1987. 
Box  turtles  (Terrapene  Carolina)  as 
potential  agents  for  seed  dispersal. 
Am.  Midi.  Nat.  117:312-318. 

Burton,  T.M.  and  G.E.  Likens.  1975. 
Salamander  populations  and  bio- 
mass  in  the  Hubbard  Brook  Ex- 
perimental Forest,  New  Hamp- 
shire. Copeia  1975:541-546. 

Bushley,  S.  Date  of  publication  un- 
known. How  to  profit  from  snap- 
ping turtles.  Reptiles  Unlimited, 
York,  Pennsylvania. 

Carolina  Biological  Supply  Com- 
pany. 1987.  Biology /Science  mate- 
rials. Catalog  58.  Burlington,  NC. 

Conant,  Roger.  1975.  A  field  guide  to 
reptiles  and  amphibians  of  east- 
ern/central North  America. 
Houghton-Mifflin:  NY. 

Congdon,  Justin  D.  and  J.  Whitfield 
Gibbons.  1988.  Biomass  productiv- 
ity of  turtles  in  freshwater  wet- 
lands: A  geographic  comparison. 
In  Freshwater  Wetlands  and  Wild- 
life. Rebecca  R.  Sharitz  and  J. 
Whitfield  Gibbons,  editors.  Office 
of  Scientific  and  Technical  Infor- 
mation, U.S.  Department  of  En- 
ergy. Oak  Ridge,  TN. 

Congdon,  Justin  D.,  Gary  L.  Breiten- 
bach  and  Richard  C.  van  Loben 
Sels.  1987.  Reproduction  and  nest- 
ing ecology  of  snapping  turtles 
(Chelydra  serpentina)  in  southeast- 
ern Michigan.  Herpetologica 
43:39-54. 

Congdon,  Justin  D.,  Judith  L.  Greene, 
and  J.  Whitfield  Gibbons.  1986. 
Biomass  of  freshwater  turtles:  A 
geographic  comparison.  Am. 
Midi.  Nat.  115:165-173. 

Constantine,  D.  G.  1970.  Bats  in  rela- 
tion to  the  health,  welfare,  and 
economy  of  man.  pp.  320-449.  In: 
W.  A.  Wimsatt,  editor.  Biology  of 
bats.  Vol.  II.  Academic  Press.  New 
York. 

Gibbons,  J.  Whitfield.  1987.  Why  do 
turtles  live  so  long?  Bioscience 
37:262-269. 

Gibbons,  J.  Whitfield  and  Judith  L. 
Greene.  1979.  X-ray  photography: 


A  technique  to  determine  repro- 
ductive patterns  of  freshwater 
turtles.  Herpetologica  35:86-89. 

Gibbons,  J.  Whitfield  and  Raymond 
D.  Semlitsch.  1988.  Guide  to  the 
reptiles  and  amphibians  of  the 
Savannah  River  Plant.  SRO-NERP 
18.  U.S.  Department  of  Energy's 
Savannah  River  National  Environ- 
mental Research  Park.  Aiken,  SC. 

Halls,  L.  K.  1984.  White-tailed  deer: 
Ecology  and  management.  Stack- 
pole  Books.  Harrisburg,  PA. 

Hayward,  G.  F.  and  J.  Phillipson. 
1979.  Community  structure  and 
functional  role  of  small  mammals 
in  ecosystems,  pp.  135-211.  In: 
Ecology  of  small  mammals.  D.  M. 
Stoddart,  editor.  John  Wiley  and 
Sons.  New  York. 

Holling,  C.S.  1966.  The  functional  re- 
sponse of  invertebrate  predators 
to  prey  density.  Mem.  Entomol. 
Soc.  Can.  48:1-86. 

Joanen,  Ted  and  Larry  McNease. 
1987.  The  management  of  alliga- 
tors in  Louisiana,  USA.  p.  33-42. 
In:  Wildlife  Management.  Croco- 
diles and  Alligators.  Grahame  J.W. 
Webb,  S.  Charlie  Manolis  and  Pe- 
ter J.  Whitehead,  editors.  Surrey 
Beatty  &  Sons  Pty.  Limited.  Chip- 
ping Norton,  NSW,  Australia. 

Johns,  Paul  E.  and  Michael  H.  Smith. 
1985.  Bibliography  for  the  white- 
tailed  deer  on  the  Savannah  River 
Plant.  University  of  Georgia,  Insti- 
tute of  Ecology  and  Savannah 
River  Ecology  Laboratory.  Aiken, 
SC. 

Kushlan,  James  A.  and  Frank  J.  Maz- 
zotti.  1986.  Population  biology  and 
status  of  the  American  crocodile  in 
South  Florida,  pp.  188-194.  In: 
Proceedings  of  the  7th  working 
meeting  of  the  crocodile  specialist 
group.  lUCN.  Caracas,  Venezuela. 

Luckenbach,  R.  A.  1982.  Ecology  and 
management  of  the  desert  tortoise 
(Gopherus  agassizii)  in  California, 
pp.  1-37.  In:  R.  B.  Bury  (ed.).  North 
American  tortoises:  Conservation 
and  Ecology.  USDI,  Fish  and 
Wildlife  Service.  Wildlife  Resource 
Department.  12. 


Parrish,  Henry  M.  1980.  Poisonous 
snakebites  in  the  United  States. 
Vantage  Press,  Inc.  New  York. 

Pechmann,  Joseph  H.K.,  David  E. 
Scott,  J.  Whitfield  Gibbons  and 
Raymond  D.  Semlitsch.  1988.  In- 
fluence of  wetland  hydroperiod  on 
diversity  and  abundance  of  meta- 
morphosing juvenile  amphibians. 
Wetlands  Ecology  and  Manage- 
ment 1:1-9. 

Peterson,  C.  1988.  Presentation  at 
Desert  Tortoise  Council  Meeting, 
Laughlin,  NV. 

Ricker,  N.E.  1975.  Computation  and 
interpretation  of  biological  statis- 
tics of  fish  populations.  Bull.  Fish. 
Res.  Board  Can.  191:1-382. 

Rowe,  P.P.  1981.  Wild  house  mouse 
biology  and  control.  Symposium 
of  the  Zoological  Society  of  Lon- 
don. 47:575-589. 

Scribner,  Kim  T.,  Joseph  E.  Evans, 
Stephen  J.  Morreale,  Michael  H. 
Smith  and  J.  Whitfield  Gibbons. 
1986.  Genetic  divergence  among 
populations  of  the  yellow-bellied 
slider  turtle  (Pseudemys  scripta) 
separated  by  aquatic  and  terres- 
trial habitats.  Copeia  1986:691-700. 

Semlitsch,  Raymond  D.  1985.  Analy- 
sis of  climatic  factors  influencing 
migrations  of  the  salamander  Am- 
bystoma  talpoideum.  Copeia 
1985:477-489. 

Semlitsch,  Raymond  D.,  David  E. 
Scott  and  Joseph  H.K.  Pechmann. 
1988.  Time  and  size  at  metamor- 
phosis related  to  adult  fitness  in 
Ambystoma  talpoideum.  Ecology 
69:184-192. 

Shiffer,  Clark  N.  1987.  The  bullfrog. 
Pennsylvania  Angler.  August 
1987. 

SREL  Report.  1980.  Savannah  River 
Ecology  Laboratory'  Annual  Re- 
port, FY-1980.  A  biological  inven- 
tory of  the  proposed  site  of  the 
Defense  Waste  Processing  Facility 
on  the  Savannah  River  Plant  in 
Aiken,  South  Carolina.  NTIS  Publ. 
SREL-7UC-66e.  Aiken,  SC. 

Taylor,  Barbara  E.,  Ruth  A.  Estes,  Jo- 
seph H.  K.  Pechmann  and  Ray- 
mond D.  Semlitsch.  Trophic  rela- 


9 


tions  in  a  temporary  pond:  Larval 
salamanders  and  their  microin- 
vertebrate  prey.  Can.  J.  Zool.  In 
Press. 

USDI  (United  States  Department  of 
the  Interior).  1980.  Republication 
of  the  lists  of  endangered  and 
threatened  species  and  correction 
of  technical  errors  in  final  rules. 
Federal  Register.  45:33768-33781. 

Walker,  E.  P.  1983.  Mammals  of  the 
world.  Vol.  I.  John  Hopkins  Uni- 
versity Press,  Baltimore.  568  p. 

Young,  David  P.  1988.  Rhadinaea  flav- 
ilata  (Pine  woods  snake).  Herp. 
Rev.  19:20. 


Douglas-fir  Forests  in  tlie 
Oregon  and  Washington 
Cascades:  Relation  of  the 
Herpetofauna  to  Stand  Age 
and  Moisture^ 

R.  Bruce  Bury^  and  Paul  Stephen  Corn^ 


Abstract.— Pitfall  traps  effectively  sampled 
amphibians  but  not  reptiles  in  Douglas-fir 
(Pseudofsugo  menz/es/O  forests.  The  abundance  of 
only  one  amphibian  species  varied  across  an  age 
gradient  or  a  moisture  gradient.  Salamanders  and 
frogs  that  breed  in  ponds  or  streams  vjere  captured 
in  large  numbers  in  some  stands,  likely  due  to  the 
presence  of  nearby  breeding  habitat  rather  than 
forest  conditions.  Lizards  occurred  mostly  in  dry 
stands  and  clearcuts.  Time-constrained  searches 
showed  different  use  of  dov^ned  v\/oody  debris 
among  terrestrial  salamanders.  The  occurrence  and 
abundance  of  species  in  naturally  regenerated 
forests  markedly  differed  from  clearcut  stands. 


The  value  of  old-growth  forests  for 
wildlife  is  highly  debated  (Fosburg 
1986,  Harmon  et  al.  1986,  Harris 
1984,  Kerrick  et  al.  1984,  Ruggiero 
and  Carey  1984,  Salwasser  1987, 
Wilcove  1987).  Most  attention  has 
been  directed  toward  the  spotted  owl 
(Strix  occidentalis),  which  is  one  of 
several  hundred  vertebrate  species 
occurring  in  the  Pacific  Northwest 
(Bruce  et  al.  1985).  Franklin  and  Spies 
(1984)  distinguished  old-growth  for- 
ests of  Douglas-fir  (Pseudotsuga  men- 
ziesii)  as  having  a  wide  range  of  tree 
sizes  and  ages,  a  deep  mulhlayered 
crown  canopy,  large  individual  trees, 
and  accumulations  of  coarse  woody 
debris  (CWD),  including  snags  and 
downed  logs  of  large  dimension. 
They  reported  that  these  forests  are 
productive,  diverse  ecosystems,  and 
highly  specialized  habitats. 

We  need  to  evaluate  sampling 
techniques  continually  to  better  de- 
scribe, understand  and  predict  the 
species  richness,  abundance  and  bio- 
mass  of  herpetological  assemblages. 
However,  few  herpetological  com- 
munities or  their  habitats  have  been 

^  Roper  presented  at  symposium,  Mon- 
agement  of  AmphibioDS.  Reptiles,  and 
Smoll  Mammals  in  North  America  (Flagstaff, 
AZ.July  17-21,  1988). 

Bruce  Bury  is  Zoologist  (Research)), 
USDA  Fishi  and  Wildlife  Sen/ice,  National 
Ecology  Research  Center,  1300  Blue  Spruce 
Drive,  Fort  Collins,  CO  80524. 

^Paul  Stephen  Corn  is  Zoologist,  USDA 
Fish  and  Wildlife  Service,  National  Ecology 
Research  Center,  1300  Blue  Spruce  Drive, 
Fort  Collins.  CO  80524. 


sampled  using  more  than  one  quanti- 
tative technique. 

Recently,  field  techniques  for  the 
study  of  herpetological  communities 
have  improved  (Scott  1982).  Some  of 
the  most  promising  methods  employ 
pitfall  traps  and  drift  fences  to  cap- 
ture amphibians  and  reptiles.  Several 
promising  pitfall  designs  have  been 
developed  for  varied  habitats  in  Aus- 
tralia (Friend  1984,  Webb  1985)  and 
in  North  America  (Bennett  et  al. 
1980,  Bury  and  Corn  1987,  Bury  and 
Raphael  1983,  Campbell  and  Christ- 
man  1982,  Enge  and  Marion  1986, 
Gibbons  and  Semlitsch  1981,  Jones 
1981, 1986,  Raphael  1984,  Raphael 
and  Rosenberg  1983,  Rosenberg  and 
Raphael  1986,  Vogt  and  Hine  1982). 
Pitfall  traps  are  effective  for  capture 
of  commmon  terrestrial  species  and 
they  are  particularly  valuable  in  sam- 
pling secretive  or  rare  forms. 

Searches  by  hand  (either  based  on 
specific  areas  or  time  of  collecting)  or 
observation  are  used  to  sample  her- 
petofaunas  (see  reviews  by  Bury  and 
Raphael  1983,  Jones  1986,  Rough  et 
al.  1987).  Campbell  and  Christman 
(1982)  suggested  that  time-con- 
strained collecting  (searching  within 
a  specific  period  of  time  by  trained 
collectors)  can  sample  terrestrial  spe- 
cies that  are  under-sampled  or  not 
taken  in  pitfall  traps. 

The  first  year  of  our  old-growth 
study  (1983)  was  partly  devoted  to 
refining  field  techniques.  A  compari- 
son of  different  pitfall  designs  is  re- 
ported elsewhere  (Bury  and  Corn 


1987).  Here,  we  employ  a  standard- 
ized pitfall  array  and  time-con- 
strained searches  to  determine  the 
occurrence  and  abundance  of  the  ter- 
restrial (upland)  herpetofauna  in  the 
Cascade  Mountains  of  the  Pacific 
Northwest. 

The  current  work  on  small  mam- 
mals (Anthony  et  al.  1987,  Com  et  al. 
1988,  West  1985),  birds  (Carey  1988, 
Manuwal  and  Huff  1987),  and  bats 
(Thomas  in  press)  are  part  of  an  inter- 
disciplinary effort  to  better  under- 
stand the  relationship  of  nongame 
wildlife  in  old-growth  forest  stands 
(Ruggiero  and  Carey  1984).  Our 
study  is  the  first  to  attempt  to  iden- 
tify which  species  of  the  herpe- 
tofauna, if  any,  are  associated  with 
age  and  moisture  gradients  in  forests 
of  the  Cascade  Mountains. 

Our  specific  objectives  were  (1)  to 
compare  effectiveness  and  relative 
merits  of  time-constrained  collecting 
versus  pitfall  trapping,  (2)  to  com- 
pare the  species  richness  and  relative 
abundance  of  amphibians  and  rep- 
tiles between  different  forest  stands, 
and  (3)  to  examine  the  association  of 
the  herpetofauna  with  old-growth 
forest  conditions. 


DESCRIPTION  AND 
CLASSIFICATION  OF  STUDY  SITES 

We  sampled  30  sites:  18  in  or  near 
the  H.  J.  Andrews  Experimental  For- 
est in  eastern  Linn  and  Lane  coun- 
ties, Oregon,  and  12  stands  in  the 


11 


Figure  1  .—Conducting  time-constrained  searches  in  an  old-growtti  stand,  Oregon.  Note 
large  amounts  of  downed  woody  debris. 


Wind  River  Experimental  Forest, 
Skamania  County,  Washington.  All 
sites  are  on  the  western  slopes  of  the 
Cascade  Mountains.  Specific  loca- 
tions, stand  classification,  elevations 
and  other  details  are  provided  in 
Corn  et  al.  (this  volume). 

Study  sites  represent  a  range  of 
forest  development  across  a 
chronosequence  (principally  age) 
and,  for  old-growth,  a  moisture  gra- 
dient. These  stands  were  independ- 
ently selected  and  assessed  by  Spies 
et  al.  {in  press).  They  were  all  in  natu- 
rally regenerated  forest  caused  by 
wildfire.  There  were  three  develop- 
ment stages  in  moderate  moisture 
conditions:  young  (30-76  years  old), 
mature  (105-150  years)  and  old- 
growth  (195-450  years).  Clearcut  sites 
represent  recent  timber  harvest  (<10 
years  old).  For  old-growth  stands 
only,  there  were  representative  mois- 
ture conditions:  wet,  moderate  and 
dry  sites.  Stand  classification  was 
based  on  age  determined  by  incre- 
ment boring  of  trees  or  other  meth- 
ods, characteristic  plant  species  in 
the  understory,  physiography,  and 
soils.  These  methods  and  other  para- 
meters are  described  by  Corn  et  al. 
(this  volume).  Franklin  et  al.  (1981) 
and  Spies  et  al.  {in  press). 

Following  the  initial  stand  selec- 
tion, there  were  minor  adjustments 
in  assignment  of  stand  classification 
(Corn  et  al.,  this  volume).  We  re- 
jected a  few  sites  that  were  either  not 
continually  accessible  for  our  weekly 
checking  of  pitfall  traps  or  were 
being  actively  logged. 

MATERIAL  AND  METHODS 

Time-Constrained  Searches  (TCS) 

Details  of  this  technique  are  pro- 
vided elsewhere  (Campbell  and 
Christman  1982,  Bury  and  Raphael 
1983,  Raphael  and  Rosenberg  1983). 
A  team  of  3-8  people  intensively 
searched  each  stand  for  8  person-hrs 
in  the  spring  (8-25  April  1983  in  Ore- 
gon and  3-12  May  1983  in  Washing- 


ton). We  turned  over  moveable  sur- 
face objects  (twigs  to  logs  <1  m  dia- 
mater),  dug  into  decayed  wood,  and 
removed  bark  from  downed  wood  or 
the  bases  of  standing  snags  by  hand 
or  with  potato  rakes  (fig.  1). 

Collectors  remained  within 
boundaries  of  habitat  typical  of  the 
stand,  avoiding  conspicuous  special- 
ized habitats  such  as  ponds,  creeks 
or  rock  outcrops.  Further,  we 
searched  4  sites  in  each  state  again 
during  warm  weather  (July- Aug 
1983).  These  surveys  were  performed 
for  4  hrs  per  plot.  We  recorded  infor- 
mation on  exact  position  of  capture 
for  each  animal,  including  vertical 
position  (e.g.,  on  or  under  litter;  on, 
under  or  in  log;  etc.),  identification  of 
cover  object,  length  and  diameter  of 
object,  time  of  capture,  total  length, 
and  mass  of  animal. 

We  determined  the  decay  class  of 
coarse  woody  debris  occupied  by 
animals  on  the  forest  floor.  Large 
woody  debris  or  felled  trees  (logs) 
occur  in  five  progressive  broad  decay 
classes  (Bartels  et  al.  1985,  Franklin  et 
al.  1981,  Harmon  et  al.  1986,  Maser  et 
al.  1979,  Maser  and  Trappe  1984):  (1) 


intact,  recently  downed  trees;  (2)  logs 
with  loose  bark;  (3)  loss  of  bark  and 
stem  partly  rotted;  (4)  invasion  of 
roots  and  deep  decomposition  of 
stem;  and  (5)  hummocks  of  wood 
chunks  and  organic  material.  Once 
fallen,  a  large  tree  might  require  200 
or  more  years  to  progress  from  class 
1  to  5  (Spies  et  al.  in  press),  providing 
habitat  for  many  generations  of  resi- 
dent wildlife. 


Pitfall  Arrays 

We  installed  a  pitfall  array  at  each 
site  in  Oregon  and  Washington  (de- 
tails in  Bury  and  Corn  1987).  Each 
array  had  two  triads  with  their  cen- 
ters 25  m  apart.  Each  triad  was  com- 
posed of  three  drift  fences  5  m  long 
and  0.5  m  tall;  about  0.3  m  of  fence 
was  above  ground.  Fences  radiated 
at  120°  angles,  beginning  3  m  from 
the  center  point.  The  compass  direc- 
tions of  the  arms  depended  on  open- 
ings between  trees  or  large  logs  on 
the  forest  floor.  Pitfall  traps  were 
constructed  from  two  stacked  #10  tin 
cans  (3.2  1  volume)  connected  with 


12 


Table  1  .—Numbers  of  amphibians  and  reptiles  captured  during  time-constrained  searches  (TCS)  corKiucted  8-25 
April  1983  at  the  H.  J.  Andrews  Experimental  Forest  In  Oregon.  Old-growth  stands  are  arranged  in  order  of  increasing 
dryness. 


Old  growth 


Wet 


Moderate 


Dry 


Mature 


Young 


Clearcut 


Species                 Stand  No. 

15 

03 

24 

«02 

17 

33  25 

29 

11 

35 

42 

39  47 

48 

75 

55 

291 

391 

Clouded  Salamander 

3 

8 

6 

9 

3 

11 

17 

4 

2 

1 

2 

12 

2 

Oregon  Slender  Salamander 

2 

6 

4 

12 

9 

n 

5 

1 

9 

1 

1 

Oregon  Ensatina 

4 

3 

1 

9 

5 

7  22 

2 

10 

6 

4 

5  3 

9 

8 

9 

4 

1 

Dunn's  Salamander 

2 

1 

Rough-skinned  Newt 

2 

1 

1 

Pacific  Tree  Frog 

] 

4 

1 

1 

1 

Western  Skink 
Norhtern  Alligator  Lizard 
Western  Fence  Lizard 


°Two  surveys  were  conducted  in  this  stand  and  the  results  are  combined  here. 


duct  tape.  A  pit  trap  was  placed 
fl'jsh  with  the  ground  surface  at  each 
end  of  the  fence.  Funnel  traps  were 
constructed  of  aluminum  screening, 
rolled  into  a  tube  1  m  long  by  0.1  m 
diameter,  with  inward  funnels 
stapled  at  each  end  of  the  trap.  A 
funnel  trap  was  placed  midway  on 
either  side  of  the  fence.  No  water  or 
preservatives  were  added  to  the 
traps.  A  wooden  shingle  was 
propped  over  each  pitfall  and 
funnnel  trap,  but  water  entered  pit- 


falls during  heavy  rains.  We  rou- 
tinely removed  water  from  traps 
with  scoops  or  a  hand-operated  aq- 
uarium siphon. 

We  operated  pitfall  traps  conti- 
nously  for  180  days,  from  the  last 
week  of  May  to  late  November  1983. 
Traps  were  checked  1-2  times  each 
week.  Captures  were  usually  taken 
to  a  field  laboratory  for  identification 
and  measurements.  All  retained 
specimens  are  deposited  at  the  Na- 
tional Museum  of  Natural  History. 


RESULTS 

Tinne-Constrained  Searches  (TCS) 
Yield 

During  spring  TCS,  we  collected  258 
amphibians  and  4  reptiles  (table  1)  at 
the  18  Oregon  sites  (1.8  animals  per 
person-hr)  and  we  took  78  amphibi- 
ans and  4  reptiles  (table  2)  at  12 
Washington  sites  (0.85  per  person- 
hr).  For  summer  TCS,  all  Washington 
captures  included  only  4  lizards  from 
one  clearcut,  one  mature  (drier  as- 
pect) and  an  old-growth  dry  stand 
(0.25  animals  p>er  hr)  whereas  in  Ore- 
gon we  captured  13  salamanders  (no 
new  species)  and  2  lizards  from  4 
sites  (0.9  animals  per  hr). 

Although  we  report  the  abun- 
dance of  herpetofauna  collected  by 
TCS  (tables  1  and  2),  we  did  not  ana- 
lyze these  results  based  on  the  age 
and  moisture  gradients  because  such 
abundance  data  can  be  biased. 


Habitat  Use 

TCS  provided  useful  information  on 
the  exact  position  where  individuals 
were  found  (table  3).  Oregon  ensati- 


Table  2.— Numbers  of  amphibians  and  reptiles  captured  during  TCS  3-12 
May  at  the  Wind  River  Experimental  Forest  In  Washington.  Old-growth 
stands  are  arranged  in  order  of  increasing  dryness. 

Old  growth 


Wet    Moderate    Dry      Mature      Young  Clearcut 


Species  Stand  No.  14   12  21   20  31   41   42  50  60  61   70  71 

Olympic  Salamander  2 

Oregon  Ensatina  3     7    13     5     5     4      1      1      1  1 

Larch  Mountain  Salamander  14 
Western  Red-backed 

Salamander  6 
Rough-skinned  Nev4  3     2  1 

Red-legged  Frog  1 

Pacific  Tree  Frog  1 
Rubber  Boa  2  1 

Common  Garter  Snake  1 


13 


Table  3.— Number  of  salamanders  (Oregon  data  only)  captured  in  different 
microhabitats.  Percentages  are  In  parentheses. 

Oregon 


Oregon 

Clouded 

Slender 

Position 

Ensotina 

Salamander 

Salamander 

On/Under  Litter 

3  (2.4) 

0  (0) 

1  (1.6) 

On/Under  Rock 

3  (2.4) 

0  (0) 

1  (1.6) 

On/Under  Log 

14(11.5) 

8  (10.2) 

6  (6.8) 

Inside  Log 

52  (42.6) 

27  (34.2) 

38  (62.3) 

Under  Bark  on  Log 

12  (9.8) 

37  (46.8) 

7  (11.5) 

Under  Bark  on  Ground 

38  (31.1) 

7  (8.9) 

8  (13.1) 

nas  {Ensatim  eschscholtzi;  fig.  2)  oc- 
curred more  evenly  and  in  more  mi- 
crohabitats than  did  the  other  two 
species.  Clouded  salamanders 
(Aneides  ferreus)  were  mostly  under 
bark  on  logs  and,  secondarily,  often 
were  in  logs  (81%  of  the  sites  occu- 
pied were  related  to  logs).  The  Ore- 
gon slender  salamander  (Batrachoseps 
wrighti)  predominately  occurred  in 
logs  (62%)  and  then  under  bark  on 
ground  or  on  logs  (87%  in  or  near 
logs).  Most  bark  on  the  ground  oc- 
curred in  piles  sloughed  from  fallen 
trees  or  snags  and  is  essentially  an 
extension  of  the  log  environment. 

Terrestrial  salamanders  that  were 
captured  in  or  near  downed  wood 
markedly  differed  in  their  use  of  dif- 
ferent decay  classes  of  CWD  (fig.  3). 


We  did  not  include  decay  class  1 
logs,  because  few  of  these  were 
searched  and  none  had  salamanders. 
These  logs  are  intact  material  and 
offer  little  cover  for  salamanders. 

We  calculated  Chi-square  statistics 
for  three  species  in  Oregon.  The 
clouded  salamander  was  most  abun- 
dant in  younger  (class  2)  logs  (P 
<0.001),  while  Oregon  slender  sala- 
manders were  found  more  often  than 
expected  in  the  more  decayed  class  4 
and  5  logs  (P  <  0.05).  Numbers  of 
Oregon  ensatina  generally  followed 
the  pattern  of  log  abundance  (fig.  3), 
except  that  they  were  found  less  of- 
ten than  expected  in  class  3  logs  (P 
<0.05).  These  results  are  consistent 
with  microhabitats  where  the  sala- 
manders were  captured  (table  3). 


Pitfall  Trapping 
Total  Nunnbers 

Pitfall  arrays  at  18  Oregon  sites  pro- 
vided 1,028  captures  (table  4):  685 
salamanders,  252  frogs,  64  lizards 
and  27  snakes.  Pitfalls  at  12  Washing- 
ton sites  yielded  1,152  animals  (table 
5):  460  salamanders,  663  frogs  and  29 
snakes.  Two  Washington  sites  had 
exceptional  catches:  253  tailed  frogs 
(Ascaphus  truei)  at  #21  Old-growth 
Moderate  and  119  red-legged  frogs 
(Ram  aurora)  at  #42  Mature. 


HtL Alive  FREQUCNCV 


Figure  3.— Frequency  of  occurrence  of 
clouded  salannanders,  Oregon  slender 
salannanders,  and  Oregon  ensatinos  occu- 
pying downed  wood  in  decay  classes  2-5. 
Density  of  logs  in  each  decay  class  are 
provided.  Data  are  from  18  sites  at  the  H.  J. 
Andrews  Experimental  Forest,  Oregon. 


Yield 

Summer  operation  of  the  pitfall  ar- 
rays added  a  few  reptiles  but  the 
bulk  of  the  catch  was  amphibians  in 
the  fall  months  during  and  after 
heavy  seasonal  rains  (Bury  and  Corn 
1987).  There  was  a  low  catch  of  rep- 
tiles (Oregon,  mean  =  5  per  site; 
Washington,  mean  =  2.4). 

Species  richness  did  not  differ 
across  the  chronosequence  gradient 
(table  6,  fig.  4).  Moderate  and  dry 
old-growth  stands  had  the  highest 
mean  abundance  across  the  moisture 
gradient,  which  was  caused  by  the 
capture  of  large  numbers  of  several 
migratory  species. 


Figure  2.— Adult  ensatina  (Ensatina  eschscholtzi)  from  Douglas  Co.,  Oregon. 


14 


r 


Table  4.— Abundance  of  amphibians  and  reptiles  captured  by  pitfall  arrays  at  ttie  H.  J.  Andrews  Experimental  Forest  in 
Oregon.  Arrays  of  pitfall  traps  withi  drift  fences  were  operated  continuously  for  180  days  in  1983.  Old-growth  stands 
are  arrariged  in  order  of  increasing  dryness. 


Old  growth 


Wet 


Moderate 


Dry 


Mature 


Young 


Clearcut 


Species 


Stand  r4o.    15    03    24    02    17    33    25    29    11    35    42    39    47    48    75    55  291  391 


Northwestern  Salamander 
Pacific  Giant  Salamander 
Clouded  Salamander 
Oregon  Slender  Salamander  1 
Oregon  Ensatina  8 
Dunn's  Salamander 
Rough-skinned  Newt  21 
Tailed  Frog 
Red-legged  Frog 

Pacific  Tree  Frog  2 
Western  Skink 
Norhtern  Alligator  Lizard 
Western  Fence  Lizard 
Rubber  Boa 

Northwestern  Garter  Snake  1 
Common  Garter  Snake  1 


1 
3 
2 


28 


2 
1 

10 

3 
5 


18 

26 
3 


1 

22 

5 
4 


13 
1 

17 


26 


119 


3 
1 


27 
7 
4 

21 
1 

62 
46 
23 

3 
4 


1 

4 


28 
3 


11 
14 
5 


1 

2 


10 

16 

14 

20 

30 

12 

10 

1 

15 

13 

36 

5 

14 

16 

2 

6 

28 

30 

3 

2 

4 

1 

3 

3 

2 

5 

9 
8 
3 


1  11 


Table  5.— Abundance  of  amphibians  and  reptiles  captured  by  pitfall  arrays 
at  the  Wind  River  Experimental  Forest  in  Washington.  Arrays  of  pitfall  traps 
with  drift  fences  were  operated  continuously  for  180  days  in  1983.  Old- 
growth  stands  are  arranged  in  order  of  increasing  dryness. 


Old  growth 


Wet    Moderate    Dry      Mature      Young  Clearcut 


Species           Stand  No. 

14 

12 

21 

20 

31 

41 

42 

50 

60 

61 

70 

71 

Northwestern  Salamander 

2 

5 

15 

4 

1 

1 

1 

9 

10 

2 

Pacific  Giant  Salamander 

1 

Olympic  Salamander 

3 

1 

1 

Oregon  Ensatina 

7 

35 

29 

18 

39 

14 

13 

3 

24 

25 

0 

1 

Larch  Mountain 

Salamander 

10 

Western  Red-backed 

Salamander 

19 

Rough-skinned  Newt 

10 

4 

5 

40 

1 

10 

4 

7 

38 

37 

7 

4 

Tailed  Frog 

44 

22  253 

4 

27 

50 

4 

2 

1 

4 

Red-legged  Frog 

8 

1 

3 

15 

1 

19  119 

40 

5 

23 

6 

Pacific  Tree  Frog 

3 

9 

Northern  Alligator  Lizard 

1 

1 

12 

1 

Northwestern  Garter  Snake 

2 

1 

4 

Common  Garter  Snake 


Differences  in  Closed-Canopy 
Stands 

For  Oregon  and  Washington  data 
combined,  mean  abundance  of  com- 
mon species  (3  salamanders,  2  frogs) 
appeared  to  differ  across  either  forest 
development  (age)  or  moisture  gradi- 
ent (fig.  5).  However,  except  for  the 
Oregon  ensatina,  none  of  the  differ- 
ences were  statistically  significant 
(table  6).  High  numbers  of  individu- 
als at  a  few  stands  resulted  in  large 
variances  in  catch  at  stand  types. 

Large  numbers  of  both  the  rough- 
skinned  newt  (Taricha  granulosa)  and 
Northwestern  salamander 
(Ambystoma  gracile)  were  captured  in 
a  few  stands  (tables  4-5).  Most  of  the 
tailed  frogs  taken  were  juveniles  at 
one  old-growth  site  in  Washington 
(table  5),  and  these  were  apparently 
dispersing  away  from  a  nearby 
stream.  Similarly,  most  (78%)  of  the 
red-legged  frogs  were  taken  at  5  sites 
(tables  4-5);  the  largest  number  (n  = 


15 


119)  were  juveniles  captured  at  one 
mature  stand  in  Washington. 

The  only  species  showing  a  signifi- 
cant difference  (table  6)  across  the 
chronosequence  of  stands  was  the 
Oregon  ensatina.  Its  numbers  were 
lower  in  mature  stands  (fig.  5),  per- 
haps related  to  amounts  of  CWD  in 
different  age  classes  (fig.  6).  Abun- 
dance of  Oregon  ensatinas  was  most 
highly  correlated  with  the  number  of 
decay  class  4  and  5  logs  per  hectare 
(Pearson  r  =  0.48,  n  =  29,  P  <  0.01) 
and  the  mean  diameter  (d.b.h.)  of 
large-sized  canopy  trees  (r  =  0.51,  n  = 
29,  P  <  0.01).  A  discussion  of  the 
habitat  variables  used  here  is  pro- 
vided in  Corn  et  al.  (1988).  Mean 
abundance  of  Oregon  ensatina  also 
differed  across  the  moisture  gradient 
in  old-growth  stands  with  fewer 
present  in  wetter  sites  than  drier. 
Paradoxically,  most  OGW  stands 
have  large  amounts  of  CWD  (fig.  6). 
Oregon  ensatina  may  be  associated 
with  the  amount  of  CWD,  but  there 
are  other  components  of  the  habitat 
that  may  be  under  represented  in 
OGW  stands. 


Clearcut  Stands 

We  also  trapped  5  clearcut  sites  (all 
<10  years  old)  to  describe  herpe- 
tofauna  occurrence  in  managed 


stands.  The  relative  abundance  of  the 
herpetofauna  in  these  clearcuts 
markedly  differed  from  6  compa- 
rable young  stands  (fig.  7).  Reptiles 
predominate  in  clearcuts,  most  likely 
responding  to  increased  ambient 
temperature  in  such  areas.  The  Pa- 
cific treefrog  (Hyla  regilla)  also  was 
most  abundant  in  clearcuts. 


DISCUSSION 

Comparison  and  Improvements  in 
Tectiniques 

Time-constrained  searches  (TCS) 
provided  insufficent  animals  for 
quantitative  analyses  in  most  stands. 
The  technique  might  be  more  worth- 
while under  optimal  environmental 
conditions  (e.g.,  after  heavy  rains  for 
amphibians)  and  with  increased  ef- 
fort (16+  person-hr  per  site).  Summer 
searches  added  the  occurrence  of  liz- 
ards to  some  stands,  but  in  general 
the  effort  was  not  worth  the  time  in- 
vestment in  forested  stands  of  the 
Cascade  Mountains. 

However,  TCS  can  be  effective  to 
sample  terrestrial  species  of  salaman- 
ders. Our  pitfall  trapping  (180  days) 
caught  257  ensatina,  44  clouded  sala- 
manders, and  13  Oregon  slender 
salamanders,  whereas  TCS  yielded 
113  ensatina  (0.44  times  that  of  pit- 


r   '.  "  ^ 

Table  6.— Analysis  of  variance  of  species  richness  and  abundance  (log 
transformed)  categorized  by  age  (old  growth,  mature,  and  young)  and 
moisture  (wet,  moderate,  and  dry).  Wet  and  dry  old  growth  stands  were 
not  used  in  the  analysis  of  stand  age,  and  mature  and  young  stands  were 
not  used  in  the  analysis  of  stand  moisture. 


Age  (n  =  1 7)  Moisture  <n  =  1 3) 


F 

P 

F 

P 

Species  Richness 

2.02 

0,17 

0.30 

0.75 

Total  Abundance 

0.92 

0.42 

2.40 

0.14 

Northwest  Salamander 

0.38 

0.69 

1.90 

0.20 

Rough-skinned  Newt 

0.91 

0.43 

0.26 

0.78 

Oregon  Ensatina 

8.09 

0.005 

11.4 

0.003 

Tailed  Frog 

0.92 

0.42 

0.06 

0.94 

Red-legged  Frog 

0.65 

0.54 

0.12 

0.89 

falls),  76  clouded  salamanders  (1.7  X 
pitfalls),  and  57  slender  salamanders 
(4-4  X  pitfalls).  The  clouded  salaman- 
der is  a  common  denizen  of  Oregon 
forests  and  sometimes  the  most  fre- 
quently encountered  species,  but  pit- 
fall traps  caught  few.  This  species 
has  large  toes  and  is  adept  at  climb- 
ing, and  perhaps  escaped.  Or,  they 
rarely  free-fall  into  traps  on  the 
ground.  The  Oregon  slender  sala- 
mander seems  to  be  associated  with 

SPECIES  RICHNESS 

MEAN  I  OF  SPECIES 


ABUNDANCE 


MEAN  TOIAL  CAPTURES 


Figure  4.— Mean  species  richness  and 
mean  total  abundance  of  annpt»ibians  and 
reptiles  in  closed-canopy  forest  stands. 


16 


Ambystoma  gracila       Taricha  granulosa 


Ensatina  aschschaltzi    Rana  aurara 


downed  woody  debris  and  the  best- 
known  method  to  sample  such  mate- 
rial is  with  TCS,  area-constrained 
searches  (Bury  and  Raphael  1983, 
Raphael  and  Rosenberg  1983),  or 
hand-collecting  of  specific  amounts 
and  types  of  CWD. 

For  several  reasons,  we  refrained 
from  using  TCS  to  compare  differ- 
ences in  herpetofauna  across  stand 
ages  and  moisture  gradients.  In  1983, 
we  did  not  record  the  number  nor 
amount  of  litter  (CWD)  searched  in 
each  study  site,  which  could  have 
affected  the  results.  Unless  cover 
items  are  scarce,  TCS  will  result  in 
equivalent  numbers  of  cover  items 
searched,  e.g.,  20  logs  per  person-hr 
of  search.  However,  the  type,  num- 
ber and  biomass  of  logs  differs 
among  stands.  Thus,  the  number  of 
animals  collected  is  not  related  to  the 
availability  of  cover  (Corn  and  Bury 
unpublished  data). 

On  the  other  hand,  sites  with  large 
amounts  of  CWD  may  be  occupied 
by  many  individuals  yet  few  are  re- 
vealed because  they  are  dispersed. 
Douglas-fir  forests  can  have  over 
1600  mVha  of  CWD  (Spies  et  al.  in 
press).  Recently,  we  found  that  the 
density  of  salamanders  in  the  Oregon 
Coast  Range  (number/ m^  of  CWD) 
was  inversely  related  to  the  amount 
of  CWD  present  in  the  stand  (Corn 
and  Bury  unpublished  data).  TCS 
will  underestimate  abundance  in 
stands  with  large  amounts  of  CWD 
relative  to  stands  with  less  CWD. 
Underestimation  of  the  numbers  of 
amphibians  and  reptiles  in  ecosys- 
tems is  often  more  common  than 
overestimation.  Furthermore,  we  dis- 
covered that  some  collectors  tended 
to  focus  on  older  decay  classes  of 
CWD  (that  often  yield  the  highest 
catch)  rather  than  uniformly  search- 
ing all  objects. 

To  estimate  abundance  of  sala- 
manders, we  suggest  recording  the 
volume  of  CWD  searched,  control  for 
time  per  object  (e.g.,  15  minutes 
maximum),  balance  effort  (e.g., 
equivalent  search  between  different 
decay  classes  of  CWD),  and  relate 


17 


catch  per  volume  of  objects  to  sepa- 
rate estimates  of  the  total  CWD  per 
hectare.  These  changes  are  needed  to 
improve  the  value  of  TCS  techniques 
for  sampling  the  herpetofauna  of  for- 
est ecosystems. 

Pitfall  traps  catch  the  large  num- 
bers of  individuals  needed  for  quan- 
tified analyses  of  differences  between 
forest  stand  types.  They  proved  to  be 
particularly  important  for  sampling 
migratory  species  of  amphibians, 
which  we  found  to  be  common  in 
Cascade  forests.  Also,  our  recent  re- 
sults indicate  for  the  first  time  that 
tailed  frogs  occur  in  "upland"  for- 
ested habitats. 

Vogt  and  Hine  (1982)  pointed  out 
that  pitfall  traps  were  most  efficient 
during  periods  of  precipitation  or 
soon  thereafter.  Our  results  confirm 
these  observations  and,  lately,  we 
have  reduced  pitfall  operations  to  SO- 
SO  days  in  the  fall  only.  Also,  the 
triad  design  used  here  was  highly 
effective  but  required  great  effort 
(900  m  of  drift  fence  was  installed)  in 
Pacific  Northwest  forests,  which 
have  large  tree  roots  and  rocky  soils. 
Drift  fences  are  more  cost-effective  in 
sandy  areas  where  they  can  be  more 
readily  installed. 

We  caught  few  reptiles  in  the  Cas- 
cade Mountains  and  pitfall  traps 
were  ineffective  for  these  animals, 
even  in  the  warmer  summer  months. 
Reptiles  may  be  numerous  in  certain 
clearcuts  (e.g.,  tables  4-5),  in  drier 
regions  such  as  interior  areas  of 
northern  California  (e.g.,  Raphael 
and  Barrett  1984,  Raphael,  this  vol- 
ume) and,  based  on  our  prior  experi- 
ence, in  some  young  managed  stands 
(10-30  years  old).  When  present, 
these  would  be  worth  sampling  with 
pitfall  traps. 

Pitfall  traps  alone  are  adequate  to 
capture  most  amphibians  and  small 
mammals  (Bury  and  Corn  1987)  but 
overall  sample  size  can  be  improved 
by  increasing  the  number  of  traps  per 
site.  Thus,  we  have  more  recently 
employed  a  6  by  6  pitfall  grid  (36 
traps;  15-m  spacing)  and  the  catch  is 
large  enough  for  quantitative  analy- 


ses. These  adjustments  greatly  in- 
crease the  use  and  effectiveness  of 
pitfall  trapping  in  the  Pacific  North- 
west and,  likely,  in  other  forested 
habitats. 


Association  of  Herpetofauna  with 
Old-Growth  Forests 

TCS  revealed  microhabitat  differ- 
ences between  terrestrial  species  of 
salamanders,  confirming  general  ob- 
servations about  these  species  (e.g., 
see  Nussbaum  et  al.  1983,  Stebbins 
1985).  However,  the  habitat  require- 
ments of  these  forms  need  better  in- 
vestigation. 

The  Oregon  slender  salamander 
seems  to  be  associated  with  coarse 
woody  debris  in  older  decay  classes, 
which  is  a  characteristic  feature  of 
old-growth  forests.  This  species  is 
endemic  to  the  Oregon  Cascades,  oc- 
curring only  in  Douglas-fir  and  sub- 
alpine  forests.  Thus,  timber  harvest 
might  affect  populations  of  slender 


LU 
O 
Z 
< 
Q 
Z 

< 
LU 

:^ 

Lli 
> 

< 
_J 
Ol 


7 
6 
5 
4 
3 
2 
1 

1 

2 
3 
4 
5 
6 
7 


S 


IMATUREI    YOUNG  laEAR-l 
ao  GROWTH  ^ 

Figure  6.--Biomass  of  all  (top)  and  class  4 
and  5  (bottom)  downed  wood  at  18  stands 
at  the  H.  J.  Andrews  Experimental  Forest, 
Oregon. 


YOUNG 


I — I 


TAl£D     ENSATMA     ROUGH-  NORTHWEST    RED-  PACFIC 
FROG  8KJNNED      SALA-       UE.QGED  TREE 

KEWT       MANDER        FROG         FROG       SNAKES  UZARDS 


n 


CLEARCUT 


78 


Figure  7.— Relative  abundance  of  herpetofauna  in  young  stands  and  clearcuts.  Above  the 
horizontal:  species  more  abundant  in  young  stands.  Below:  species  more  abundant  in 
clearcuts.  Values  are  the  greater  mean  adundance  divided  by  the  lesser,  e.g.,  lizards  were 
78  times  more  abundant  in  clearcuts  than  in  young  forest  stands. 


18 


salamanders,  and  this  species  merits 
special  study. 

The  Olympic  salamander  (Rhyacot- 
riton  olympicus)  occurs  in  or  near 
small  streams,  which  can  be  dis- 
rupted by  timber  harvest  (Bury  1988, 
Bury  and  Corn  1988,  Welsh,  this  vol- 
ume). Our  techniques  sampled  ter- 
restrial habitats  and  we  found  few  of 
this  species  (pitfall  traps  took  only  4 
in  old-growth  and  1  in  mature 
stands).  Many  tailed  frogs  were  cap- 
tured in  pitfall  traps  in  closed-can- 
opy forests,  but  they  were  absent  or 
rare  in  clearcuts  (only  1%  of  the  total 
catch).  Both  the  Olympic  slamander 
and  the  tailed  frog  seem  to  be  sensi- 
tive to  timber  harvest,  and  the  sur- 
vival of  these  species  may  depend  on 
protection  of  cool,  flowing  streams 
(required  for  breeding  and  larval  de- 
velopment) as  well  as  adjacent  for- 
ested habitats  (for  shade  and  reten- 
tion of  stream  substrate  quality,  see 
Bury  and  Corn  1988).  There  is  a  need 
to  assess  the  effects  of  logging  in 
streamside  and  upland  forests,  which 
may  directly  or  indirectly  affect  am- 
phibians in  headwaters  and  small 
streams  (Cooper  et  al.  1988,  Bury  and 
Corn  1988). 

Adults  of  the  rough-skinned  newt 
and  Northwestern  salamander  mi- 
grate to  ponds  for  breeding  and, 
later,  the  adults  and  juveniles  move 
back  to  land,  which  obfuscates  their 
relation  to  forest  type.  The  red- 
legged  frog  breeds  in  slow-moving 
creeks  or  ponds,  and  the  proximity 
of  such  waters  may  have  influenced 
the  abundance  of  the  frog  in  adjacent 
stands. 

Tailed  frogs  breed  in  small 
streams  and  the  location  of  these  wa- 
ters can  greatly  influence  the  occur- 
rence of  the  species  in  nearby  forest 
stands.  Also,  we  captured  some  juve- 
nile and  adult  tailed  frogs  100  to 
>300  m  from  the  nearest  stream 
(Bury  1988).  Before  our  study,  tailed 
frogs  were  not  thought  to  move  far 
from  water  (Metter  1964,  Nussbaum 
et  al.  1983).  Proximity  of  aquatic 
breeding  sites  apparently  influenced 
the  capture  of  several  species  in  up- 


land habitat.  At  the  same  time, 
aquatic  and  semi-aquatic  species 
might  depend  on  the  forest  habitat 
for  part  of  their  life  history,  e.g.,  dis- 
persal. We  suggest  that  future  re- 
search emphasize  the  life  history  re- 
quirements and  movement  patterns 
of  amphibians,  which  might  help  to 
resolve  which  factors  are  most  im- 
portant to  their  continued  local  oc- 
currence and  abundance. 

Fewer  Oregon  ensatina  were  cap- 
tured in  mature  forests  than  either 
young  or  old-growth  stands,  and  this 
salmander  might  be  associated  with 
large  amounts  of  CWD  in  the  Oregon 
Cascades.  Mature  forests  lack  input 
from  large  trees  and  snags  (see  dis- 
cussions by  Franklin  et  al.  1981,  Har- 
mon et  al.  1986,  Spies  et  al.  in  press). 
Disturbance  (fire  or  blow-down)  cre- 
ates new  young  stands  with  appre- 
ciable amounts  of  CWD. 

Similar  to  our  results,  Raphael  and 
Barrett  (1984)  found  that  the  abun- 
dance of  Oregon  ensatina  in  northern 
California  was  correlated  to  density 
of  large  Douglas-fir  trees.  However, 
they  found  few  ensatina  in  the 
youngest  stands  (<150  years)  they 
studied,  and  they  included  ensatina 
with  species  associated  with  old- 
growth  stands.  In  the  Oregon  Cas- 
cades, ensatina  were  ubiquitous  and 
there  is  no  apparent  correlation  with 
old-growth  stands. 

Clouded  salamanders  were  most 
abundant  under  the  bark  of  relatively 
young  logs.  They  may  prefer  class  2 
and  3  logs,  particularly  occupying 
logs  with  loose  bark.  Also,  clouded 
salamanders  appear  to  be  common  in 
clearcuts  (table  1).  This  species  does 
not  appear  to  be  associated  with  old- 
growth  conditions. 

In  Washington,  we  only  found  the 
Larch  Mountain  salamander  (Pletho- 
don  larselli)  at  one  old-growth  stand 
(table  2).  This  species  may  be  associ- 
ated with  forested  stands  (Herring- 
ton  and  Larson  1985),  but  the  relation 
needs  further  inquiry  and  verifica- 
tion. 


Management  Considerations 

Current  evidence  suggests  that  rich, 
abundant  populations  of  herpe- 
tofauna  occur  in  naturally  regener- 
ated forests.  Within  these  stands, 
however,  we  found  few  differences 
in  amphibians  between  wet,  moder- 
ate, and  dry  old-growth  sites  and  be- 
tween young,  mature,  and  old- 
growth  stands.  These  results  might 
be  related  to  '"old-growth"  features 
occurring  in  many  or  all  of  these 
stands.  For  example,  young  and  ma- 
ture sites  retained  many  characteris- 
tics of  old-growth  forests:  complex 
structure,  snags,  and  large  amounts 
of  downed  woody  debris,  particu- 
larly in  older  decay  classes  (fig.  6). 
Such  material  is  the  result  of  wildfire 
that  burns  and  kills  larger  trees, 
which  later  fall  to  the  ground. 

Wildfire  often  burns  unevenly 
through  stands,  resulting  in  patches 
of  lightly  burned  or  unburned  vege- 
tation surrounded  by  areas  more  in- 
tensively affected  by  fire.  Some  large 
trees  might  not  be  killed  during  fires 
and  these  persist  into  the  regenerated 
stand.  Burned  trees  become  snags 
that  later  fall  to  the  forest  floor,  creat- 
ing huge  amounts  of  CWD.  This 
heterogeneity  and  large  amounts  of 
CWD  in  naturally  regenerated  forest 
likely  maintain  favorable  conditions 
for  many  species  of  the  herpetofauna. 

Managed  stands  (clearcuts)  had 
little  downed  CWD  in  older  decay 
clases  (fig.  6)  and,  generally,  no  snags 
nor  trees  (except  for  a  rare  spar  pole 
or  small  planted  trees).  Current  for- 
estry practices  usually  fell  all  trees 
and  snags  at  sites,  eliminating  vari- 
ability in  stand  age  and  structure. 
Logging  is  generally  followed  by  pre- 
scribed burning  of  slash  and  cull 
logs,  reducing  CWD  by  50%  or  more 
(Bartels  et  al.  1985,  Maser  et  al.  1979). 
The  large  amount  of  CWD  at  one  of 
our  Oregon  clearcuts  reflects  light 
burning  (fig.  6).  Also,  this  site  was 
surrounded  by  dense,  old-growth 
forest,  which  probably  contributed 
large  amounts  of  CWD  before  burn- 
ing. 


19 


Often,  the  result  of  current  timber 
harvest  is  even-aged  stands  with 
little  CWD,  especially  in  larger  sizes. 
Present  logging  differs  from  that  per- 
formed 30  or  more  years  ago,  when 
more  CWD  was  left  on  the  forest 
floor  and  smaller  trees  were  left  in- 
tact or  ignored.  Also,  earlier  practices 
tended  to  harvest  larger,  more  valu- 
able trees  with  little  or  no  site  prepa- 
ration (except  tree-planting),  particu- 
larly on  private  lands.  These  were 
economic  decisions,  but  the  resultant 
second-growth  stands  may  differ 
markedly  from  current  intensive 
management  of  forests. 

In  contrast  to  clearcuts,  young 
stands  (naturally  regenerated)  we 
studied  were  closed-canopy  and  had 
much  downed  woody  debris.  The 
predominant  species  were  the  tailed 
frog  and  ensatina,  and  young  stands 
had  more  newts.  Northwestern  sala- 
manders and  red-legged  frogs  than 
did  clearcuts  (fig.  7).  Thus,  there 
seem  to  be  major  differences  in  the 
herpetofaunas  of  pre-canopy 
clearcuts  and  naturally  regenerated 
stands  (young  to  old-growth). 

There  is  a  critical  need  to  compare 
differences  in  wildlife  in  intensively 
managed  stands  and  those  subjected 
to  other  treatments  (e.g.,  prior  log- 
ging practices,  select-cut).  At  this 
time,  there  is  a  lack  of  information  on 
herpetofaunas  or  other  wildlife  in 
managed  second-growth  forests. 
Managed  forests  soon  will  be  the 
predominate  forest  type  in  the  Pacific 
Northwest  and  the  bulk  of  our  wild- 
life probably  will  occur  in  these 
stands.  Wise  management  of  these 
forests  should  be  of  foremost  concern 
for  wildlife  managers,  and  done  in 
concert  with  protection  of  isolated 
habitat  patches  (old-growth  forest). 

ACKNOWLEDGMENTS 

We  thank  our  field  crew  for  their 
untiring  efforts:  S.  Boyle,  R.  Hayes,  L. 
Hanebury,  S.  Martin,  T.  Olson,  and  S. 
Woodis.  We  thank  J.  Dragavon,  L. 
Jones,  P.  Morrison,  R.  Pastor,  and  D. 


Smith  for  checking  traps.  A.  McKee 
and  J.  Moreau  at  the  H.  J.  Andrews 
Experimental  Forest,  Oregon,  and 
personnel  at  Wind  River  Experimen- 
tal Forest  and  the  Carson  National 
Fish  Hatchery,  Washington,  supplied 
logistical  support.  Data  in  figure  6 
was  provided  by  T.  Spies.  We  also 
thank  R.  E.  Beiswenger,  A.  B.  Carey, 
and  M.  G.  Raphael  for  review  com- 
ments. This  is  contribution  number 
67  of  the  USD  A  Forest  Service  Proj- 
ect, Wildlife  Habitat  Relationships  in 
Western  Washington  and  Oregon. 

LITERATURE  CITED 

Anthony,  R.  C,  E.  D.  Forsman,  G.  A. 
Green,  G.  Witmer,  and  S.  K.  Nel- 
son. 1987.  Small  mammal  popula- 
tions in  riparian  zones  of  different- 
aged  coniferous  forests.  Murrelet 
68:94-102. 

Bartels,  Ronald,  John  D.  Dell,  Rich- 
ard L.  Knight,  and  Gail  Schaefer. 
1985.  Dead  and  down  woody  ma- 
terial, p.  171-186.  In  E.  Reade 
Brown,  tech.  ed..  Management  of 
wildlife  and  fish  habitats  in  forests 
of  western  Oregon  and  Washing- 
ton. Part  1  -  chapter  narratives. 
USDA  Forest  Service,  Pacific 
Northwest  Publ.  No.  R6-F&WL- 
192-1985. 

Bennett,  Stephen  H.,  J.  Whitfield  Gib- 
bons, and  Jill  Glanville.  1980.  Ter- 
restrial activity,  abundance  and 
diversity  of  amphibians  in  differ- 
ently managed  forest  types. 
American  Midland  Naturalist 
103:412-416. 

Bruce,  Charles,  Dan  Edwards,  Kim 
Mellen,  Anita  McMillan,  Tom 
Owens,  and  Harold  Sturgis.  1985. 
Wildlife  relationships  to  plant 
communities  and  stand  condi- 
tions, p.  33-55.  In  E.  Reade  Brown, 
tech.  ed..  Management  of  wildlife 
and  fish  habitats  in  forests  of  west- 
ern Oregon  and  Washington.  Part 
1  -  chapter  narratives.  USDA  For- 
est Service,  Pacific  Northwest 
Publ.  No.  R6-F&WL-192-1985. 

Bury,  R.  Bruce.  1983.  Differences  in 


amphibian  populations  in  logged 
and  old  growth  redwood  forest. 
Northwest  Science  57:167-178. 

Bury,  R.  Bruce.  1988.  Habitat  rela- 
tionships and  ecological  impor- 
tance of  amphibians  and  reptiles. 
[In  press]  In  Kenneth  J.  Raedeke, 
ed.,  Streamside  management:  ri- 
parian wildlife  and  forestry  inter- 
actions. Proceedings  of  a  Sympo- 
sium [University  of  Washington, 
Seattle,  February  11-13, 1987]  Insti- 
tute of  Forest  Resourses,  Univer- 
sity of  Washington. 

Bury,  R.  Bruce,  and  Paul  Stephen 
Com.  1987.  Evaluation  of  pitfall 
trapping  in  northwestern  forests: 
trap  arrays  with  drift  fences.  Jour- 
nal of  Wildlife  Management 
51:112-119. 

Bury,  R.  Bruce,  and  Paul  Stephen 
Corn.  1988.  Responses  of  aquatic 
and  streamside  amphibians  to  tim- 
ber harvest.  [In  press]  In  Kenneth 
J.  Raedeke,  ed.,  Streamside  man- 
agement: riparian  wildlife  and  for- 
estry interactions.  Proceedings  of  a 
Symposium  [University  of  Wash- 
ington, Seattle,  February  11-13, 
1987]  Institute  of  Forest  Resourses, 
University  of  Washington. 

Bury,  R.  Bruce,  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles,  p. 
416-419.  In  John  F.  Bell  and  Toby 
Atterbury,  eds..  Renewable  re- 
source inventories  for  monitoring 
changes  and  trends:  Proceedings 
of  an  international  conference. 
[Corvallis,  Oregon,  August  15-19, 
1983]  Society  of  American  Forest- 
ers 83-14. 

Campbell,  Howard  W.,  and  Stephen 
P.  Christman.  1982.  Field  tech- 
niques for  herpetofaunal  commu- 
nity analysis,  p.  193-200.  In  Nor- 
man J.  Scott,  Jr.,  ed.,  Herpetologi- 
cal  Communities.  U.S.  Fish  and 
Wildlife  Service,  Wildlife  Research 
Report  No.  13. 

Carey,  Andrew  B.  1988.  The  influ- 
ence of  small  streams  on  the  com- 
position of  upland  bird  communi- 
ties. [In  press]  In  Kenneth  J. 
Raedeke,  ed.,  Streamside  manage- 


20 


ment:  riparian  wildlife  and  for- 
estry interactions.  Proceedings  of  a 
Symposium  [University  of  Wash- 
ington, Seattle,  February  11-13, 
1987]  Institute  of  Forest  Resourses, 
University  of  Washington. 

Cooper,  Edward  J.,  William  J.  Trush, 
and  Allen  W.  Knight.  1988.  Effects 
of  logging  on  Pacific  giant  sala- 
manders: influence  of  age-class 
composition  and  habitat  complex- 
ity. Bulletin  of  the  Ecological  Soci- 
ety of  America  69(suppl.):104-105. 

Enge,  Kevin  E.,  and  Wayne  R.  Mar- 
ion. 1986.  Effects  of  clearcutting 
and  site  preparation  on  herpe- 
tofauna  of  a  north  Rorida  flat- 
woods.  Forest  Ecology  and  Man- 
agement 14:177-192. 

Fosburg,  Whit.  1986.  Wildlife  issues 
in  the  National  Forest  system,  p. 
159-173.  In  Roger  L.  Di  Silvestro 
ed.,  Audubon  Wildlife  Report 
1986.  The  National  Audubon  Soci- 
ety, New  York.  1094  p. 

Franklin,  Jerry  F.,  and  Thomas  A. 
Spies.  1984.  Characteristics  of  old- 
growth  Douglas-fir  forests,  p.  328- 
334.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [Portland,  Oregon,  Octo- 
ber 16-20, 1983]  Society  of  Ameri- 
can Foresters,  Washington,  D.C. 

Franklin,  Jerry  F.,  Kermit  Cromack, 
Jr.,  William  Denison,  Arthur 
McKee,  Chris  Maser,  James  Sedell, 
Fred  Swanson,  and  Glen  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-fir  forests. 
U.S.  Forest  Service,  General  Tech- 
nical Report  PNW-118.  48  p. 

Friend,  Gordon  R.  1984.  Relative  effi- 
ciency of  two  pitfall-drift  fence 
systems  for  sampling  vertebrates. 
Austalian  Zoologist  21:423-433. 

Gibbons,  J.  Whitfield,  and  Raymond 
D.  Semlitsch.  1981.  Terrestrial  drift 
fences  with  pitfall  traps:  an  effec- 
tive technique  for  quantitative 
sampling  of  animal  populations. 
Brimleyana  7:1-16. 
Harmon,  M.  E.,  J.  F.  Franklin,  F.  J. 
Swanson,  P.  Sollins,  S.  V.  Gregory, 
J.  D.  Lattin,  N.  H.  Anderson,  S.  P. 
Cline,  N.  G.  Aumen,  J.  R.  Sedell, 


G.  W.  Lienkaemper,  K.  Cromack, 
Jr.,  and  K.  W.  Cummins.  1986. 
Ecology  of  coarse  woody  debris  in 
temperate  ecosystems.  Advances 
in  Ecological  Research  15:133-302. 

Harris,  Larry  D.  1984.  The  Frag- 
mented Forest:  Island  Biogeogra- 
phy  Theory  and  the  Preservation 
of  Biotic  Diversity.  University  of 
Chicago  Press.  211  p. 

Herrington,  Robert  E.,  and  John  H. 
Larson.  1985.  Current  status,  habi- 
tat requirements  and  management 
of  the  Larch  Mountain  salamander 
Plethodon  larselli  Bums.  Biological 
Conservation  34:169-179. 

Jones,  K.  Bruce.  1981.  Effects  of  graz- 
ing on  lizard  abundance  and  di- 
versity in  western  Arizona.  South- 
west Naturalist  26:107-115. 

Jones,  K.  Bruce.  1986.  Amphibians 
and  reptiles,  p.  267-290.  In  Allen  Y. 
Cooperrider,  Raymond  J.  Boyd, 
and  Hanson  R.  Stuart  eds..  Inven- 
tory and  monitoring  wildlife  habi- 
tat. U.S.  Dept.  Interior,  Bureau  of 
Land  Management,  Service  Cen- 
ter, Denver,  Colorado.  858  p. 

Kerrick,  Michael  A.,  Kathy  Johnson, 
and  Richard  J.  Pedersen.  1984. 
What  information  is  necessary  for 
planning  the  management  of  old- 
growth  forest  for  wildlife?  p.  383- 
386.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [Portland,  Oregon,  Octo- 
ber 16-20, 1983]  Society  of  Ameri- 
can Foresters,  Washington,  D.C. 

Manuwal,  David  A.,  and  Mark  H. 
Huff.  1987.  Spring  and  winter  bird 
populations  in  a  Douglas-fir  forest 
sere.  Journal  of  Wildlife  Manage- 
ment 51:586-595. 

Maser,  Chris,  Ralph  G.  Anderson, 
Kermit  Cromack,  Jr.,  Jerry  T.  Wil- 
liams, and  Robert  E.  Martin.  1979. 
Dead  and  down  woody  material, 
p.  78-95.  In  Jack  Ward  Thomas, 
tech.  ed..  Wildlife  habitats  in  man- 
aged forests:  the  Blue  Mountains 
of  Oregon  and  Washington.  Agri- 
cultural Handbook  553.  U.S.  Dept. 
Agriculture,  Washington,  D.C. 

Maser,  Chris,  and  James  M.  Trappe 
(tech.  eds.).  1984.  The  seen  and 


unseen  world  of  the  fallen  tree. 
USDA  Forest  Service  General 
Technical  Report  PNW-164.  56  p. 

Metter,  Dean  E.  1964.  A  morphologi- 
cal and  ecological  comparison  of 
two  populations  of  the  tailed  frog, 
Ascaphus  truei  Stejneger.  Copeia 
1964:181-195. 

Nussbaum,  Ronald  A.,  Edmund  D. 
Brodie,  Jr.,  and  Robert  M.  Storm. 
1983.  Amphibians  and  Reptiles  of 
the  Pacific  Northwest.  University 
Press  of  Idaho,  Moscow.  332  p. 

Rough,  F.  Harvey,  Ellen  M.  Smith, 
Donald  H.  Rhodes,  and  Andres 
Collazo.  1987.  The  abundance  of 
salamanders  in  forest  stands  with 
different  histories  of  disturbance. 
Forest  Ecology  and  Management 
20:1-9. 

Raphael,  Martin  G.  1984.  Wildlife 
populations  in  relation  to  stand 
age  and  area  in  Douglas-fir  forests 
of  northwestern  California,  p.  259- 
274  In  William  R.  Meehan,  Theo- 
dore R.  Merrell,  Jr.,  and  Thomas 
A.  Hanley,  eds..  Fish  and  wildlife 
relationships  in  old-growth  for- 
ests. Proceedings  of  a  symposium. 
[Juneau,  Alaska,  April  12-15,  1982] 
425  p.  American  Institute  of  Fish- 
ery Research  Biologists,  (available 
from  John  W.  Reintjes,  Rt.  4,  Box 
85,  Morehead  City,  North  Caro- 
lina 28557). 

Raphael,  Martin  G.,  and  Reginald  H. 
Barrett.  1984.  Diversity  and  abun- 
dance of  wildlife  in  late  succes- 
sional  Douglas-fir  forests,  p.  352- 
360.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [Portland,  Oregon,  Octo- 
ber 16-20, 1983]  Society  of  Ameri- 
can Foresters,  Washington,  D.C. 

Raphael,  Martin  G.,  and  Kenneth  V. 
Rosenberg.  1983.  An  integrated 
approach  to  wildlife  inventories  in 
forested  habitats,  p.  219-222.  In 
John  F.  Bell  and  Toby  Atterbury, 
eds..  Renewable  resource  invento- 
ries for  monitoring  changes  and 
trends:  Proceedings  of  an  interna- 
tional conference.  [Corvallis,  Ore- 
gon, August  15-19,  1983]  Society 
of  American  Foresters  83-14. 


21 


Rosenberg,  Kenneth  V.,  and  Martin 
G.  Raphael.  1986.  Effects  of  forest 
fragmentation  on  vertebrates  in 
Douglas-fir  forests,  p.  263-272.  In 
Jared  Verner,  Michael  L.  Morri- 
son, and  C.  John  Ralph,  eds.. 
Wildlife  2000.  University  of  Wis- 
consin Press.  470  p. 

Ruggiero,  Leonard  F.,  and  Andrew 
B.  Carey.  1984.  A  programmatic 
approach  to  the  study  of  old- 
growth  forest — wildlife  relation- 
ships, p.  340-345.  In  Proceedings, 
Society  of  American  Foresters  na- 
tional convention.  [Portland,  Ore- 
gon, October  16-20, 1983]  Society 
of  American  Foresters,  Washing- 
ton, D.C. 

Salwasser,  Hal.  1987.  Spotted  owls: 
turning  a  battleground  into  a  blue- 
print. Ecology  68:776-779. 

Scott,  Norman  J.,  Jr.  (Ed.).  1982.  Her- 
petological  Communities.  U.S. 
Fish  and  Wildlife  Service,  Wildlife 
Research  Report  13.  239  p. 

Spies,  Thomas  A.,  Jerry  F.  Franklin, 
and  Ted  B.  Thomas.  In  press. 
Coarse  woody  debris  in  Douglas- 
fir  forests  of  western  Oregon  and 
Washington.  Ecology. 

Stebbins,  Robert  C.  1985.  A  Field 
Guide  to  Western  Reptiles  and 
Amphibians.  Houghton  Mifflin 
Co.,  Boston.  336  p. 

Thomas,  Don.  In  press.  Forest  age  as- 
sociation of  bats  in  western  Ore- 
gon and  Washington.  Journal  of 
Wildlife  Management. 

Vogt,  Richard  C,  and  Ruth  L.  Hine. 
1982.  Evaluation  of  techniques  for 
assessment  of  amphibian  and  rep- 
tile populations  in  Wisconsin,  p. 
201-217.  In  Norman  J.  Scott,  Jr., 
ed.,  Herpetological  Communities. 
U.S.  Fish  and  Wildlife  Service, 
Wildlife  Research  Report  13. 

Webb,  G.  A.  1985.  Habitat  use  and 
activity  patterns  in  some  southeast 
Australian  skinks.  p.  23-30.  In  Gor- 
don Grigg,  Richard  Shine,  and 
Harry  Ehmann,  eds..  Biology  of 
Australasian  Frogs  and  Reptiles. 
Surrey  Beatty  &  Sons  Pty  Ltd, 
New  South  Wales,  Australia. 


West,  Stephen  D.  1985.  Differential 
capture  between  old  and  new 
models  of  the  Museum  Special 
snap-trap.  Journal  of  Mammalogy 
66:798-800. 

Wilcove,  David  S.  1987.  Public  lands 
management  and  the  fate  of  the 
spotted  owl.  American  Birds 
41:361-367. 


22 


Long-Term  Trends  in 
Abundance  of  Annphlbians, 
Reptiles,  and  Mammals  in 
Douglas- Fir  Forests  of 
Northwestern  California^ 

Martin  G.  RaphaeP 


Abstract.— Relative  abundance  of  55  species  of 
amphibians,  reptiles,  and  mammals  was  estimated 
at  166  sites  representing  early  clearcut  through  old- 
growth  Douglas-fir  forest  in  northwestern  California. 
Nine  species  were  strongly  associated  with  older 
stands  and  1 1  species  were  strongly  associated  with 
younger  stands.  The  remaining  species  were  either 
too  rare  to  analyze  statistically  (22  species)  or 
exhibited  no  clear  trends  of  abundance  in  relation 
to  stand  age  (13  species).  Estimates  of  relative 
abundance  of  each  species  in  each  stage,  coupled 
with  data  on  historical,  present,  and  future  acreage 
of  timber  in  each  serai  stage,  were  used  to 
approximate  the  long-term  impacts  of  timber 
harvest  on  the  fauna  of  the  Douglas-fir  region  in 
northwestern  California. 


Management  of  old-growth  Douglas- 
fir  (Pseudotsuga  menziesii)  forests  is 
controversial  in  the  Pacific  North- 
west, primarily  because  of  the  pos- 
sible value  of  old-growth  as  habitat 
for  certain  wildlife  species  versus  the 
revenues  represented  by  old-growth 
trees  (Meslow  et  al.  1981,  Harris  et  al. 
1982).  Management  to  provide  wild- 
life habitat  requires  an  inventory  of 
associated  wildlife  species  and  an 
assessment  of  their  old-growth  de- 
pendency. An  analysis  of  the  size 
and  distribution  of  habitat  patches 
necessary  to  support  viable  popula- 
tions of  those  species  is  also  critical 
(Burgess  and  Sharp  1981,  Rosenberg 
and  Raphael  1986,  Scott  et  al.  1987). 

This  study  describes  the  relative 
abundance  of  amphibians,  reptiles, 
and  mammals  in  six  serai  stages  rep- 
resenting clearcuts,  young  timber 
stands,  and  mature  forest  in  north- 
western California.  These  estimates 
of  relative  abundance  were  used  to 
project  probable  long-term  changes 
in  population  size  of  amphibians, 
reptiles,  and  mammals  as  each  serai 


^ Paper  presented  at  Symposium,  Man- 
agement of  Amphibians,  Reptiles  and  Small 
Mammals  in  North  America  (Flagstaff,  AZ, 
July  19-21,  1988). 

'Research  Ecologist,  Forestry  Sciences 
Laboratory,  USDA  Forest  Service,  Rocky 
Mountain  Forest  and  Range  Experiment 
Station,  222  South  22nd  Street,  Laramie, 
Wyoming  82070. 


Stage  responds  to  forest  management 
practices. 

METHODS 

Stand  Selection 

Study  stands  were  on  the  Six  Rivers, 
Klamath,  and  Shasta-Trinity  National 
Forests  within  a  50-km  radius  of  Wil- 
low Creek,  Calif.  Forest  cover  was 
dominated  by  Douglas-fir,  usually  in 
association  with  an  understory  of 
tanoak  (Lithocarpus  ensiflorus)  and  Pa- 
cific madrone  (Arbutus  menziesii).  Ele- 
vations varied  from  400  to  1300  m. 


Stage 

1 

2 
3 

4 

5 

6 


Raphael  and  Barrett  (1984)  describe 
methods  for  aging  these  stands. 
Ground  surveys  were  used  to  verify 
stand  conditions.  Forest  Service 
stand  designations  were  used  to 
guide  stand  selection,  but  the  final 
classification  of  each  stand  into  serai 
stages  was  based  on  measured  vege- 
tation characteristics. 


The  study  region  is  characterized  by 
warm,  dry  summers  and  cool,  wet 
winters;  total  precipitation  averages 
60-170  cm  per  year. 

After  selecting  potential  study 
stands  using  timber  maps  and  aerial 
photographs,  I  then  located  all  stands 
that  were  accessible  by  road,  were 
relatively  homogeneous  with  respect 
to  tree  cover,  included  no  large  clear- 
ings or  other  anomalous  features, 
and  were  free  from  scheduled  timber 
harvest  for  at  least  the  next  3  years. 

From  this  restricted  subset  of 
stands,  I  randomly  chose  10  to  15 
stands  representing  each  of  six  serai 
stages: 


Vegetation  Sampling 

The  structure  and  composition  of 
vegetation  on  each  stand  in  the  three 
older  serai  stages  was  measured  in 
three,  randomly  selected,  0.04-ha  cir- 
cular subplots  within  a  90-m  radius 
of  each  plot  center.  Within  each  sub- 
plot, observers  recorded  species. 


Serai  state 

Age  (yrs) 

Early 

<10 

Late 

10-20 

Pole 

20-50 

Sawtimber 

50-150 

Mature 

150-250 

Old-growth 

>250 

Classification 

Clearcut  (brush/sapling) 
Young  forest  (pole/sawtimber) 
Mature  forest 


23 


height,  diameter  at  breast  height 
(d.b.h.)  and  crown  dimensions  of 
each  tree  or  shrub  >2.0  m  tall.  In  ad- 
dition, all  trees  >90-cm  d.b.h.  were 
counted  on  one  0.50-ha  circular  sub- 
plot centered  on  the  plot.  This 
sample  permitted  a  better  estimate  of 
the  density  of  large-diameter  trees. 
Numbers  of  larger  (>8-cm  diameter) 
logs  and  volume  of  other  downed 
woody  debris  were  estimated  along  a 
30-m  transect  crossing  the  center  of 
each  0.04-ha  subplot  (Brown  1974). 
Marcot  (1984)  sampled  vegetation  in 
a  similar  manner  on  stands  in  the 
three  early-seral  stages. 

Vertebrate  Sampling 

All  field  data  were  collected  by  a 
team  of  three  to  six  biologists.  We 
used  a  variety  of  techniques  to 
sample  various  taxonomic  groups. 

Pitfall  Arrays 

We  used  pitfall  arrays  to  capture 
small  mammals  (especially  insecti- 
vores),  reptiles,  and  salamanders.  An 
array  was  composed  of  ten  2-gallon 
plastic  buckets  buried  flush  with  the 
ground  and  covered  with  plywood 
lids,  arranged  in  a  2  x  5  grid  with  20- 
m  spacing.  We  placed  one  array 
within  each  stand  center  and  checked 
traps  at  weekly  to  monthly  intervals 
from  December  1981  (sawtimber, 
mature,  old-growth;  n  =  27,  56,  and 
52  sites  in  each  stage,  respectively)  or 
August  1982  (early  shrub-sapling, 
late  shrub-sapling,  pole;  n  =  10  sites 
each)  until  October  1983.  All  live  ani- 
mals were  marked  and  released;  re- 
captures were  excluded  from  analy- 
ses. Dead  animals  were  collected  and 
prepared  for  permanent  deposit  in 
museum  collections.  Results  for  each 
species  were  expressed  as  captures 
per  1000  trapnights  on  each  stand. 
Raphael  and  Rosenberg  (1983)  dem- 
onstrated that  abundance  estimates 
(capture  rates)  had  stabilized  after  15 
months  of  continuous  trapping. 


Drift  Fence  Arrays 

To  better  sample  snakes,  we  installed 
a  drift  fence  array  (Campbell  and 
Christman  1982,  Vogt  and  Hine  1982) 
on  each  of  60  randomly  selected 
stands  (10  of  each  of  the  three  early 
stages  and  sawtimber,  8  mature,  and 
12  old-growth).  An  array  consisted  of 
two  5-gallon  buckets  placed  7.6  m 
apart  and  connected  by  an  aluminum 
fence  7.6  m  long  and  50  cm  tall  with 
two  20  X  76  cm  cylindrical  funnel 
traps,  one  on  each  side  of  the  center 
of  the  fence.  These  fences  were  oper- 
ated from  May  through  September 
1983.  All  captures  were  combined 
with  those  from  the  pitfall  arrays 
along  with  the  associated  trapnights 
from  each  stand. 


Track  Stations 

Tracks  of  squirrels  and  other  larger 
mammals  were  recorded  on  each  site 
on  a  smoked  aluminum  plate  baited 
with  tuna  pet  food  (Barrett  1983,  Ra- 
phael and  Barrett  1981,  Raphael  et  al. 
1986,  Taylor  and  Raphael  1988). 
Based  on  results  of  a  pilot  study  (Ra- 
phael and  Barrett  1981),  observers 
monitored  each  station  for  8  days  in 
August  or  September  in  1981-1983, 
sampling  20  stations  in  each  of  the 
three  early  stages  and  81, 168,  and 
157  stations  in  the  sawtimber,  ma- 
ture, and  old-growth  stages,  respec- 
tively. The  proportion  of  stations  in 
each  serai  stage  on  which  a  species 
occurred  was  as  an  index  of  that  spe- 
cies' abundance. 


Livetrap  Grids 

To  better  estimate  abundance  of 
small  mammals  that  were  liable  to 
escape  from  pitfalls,  we  established 
27  livetrap  grids  (3  in  each  of  the 
three  earliest  stages  and  5,  7,  and  6  in 
the  three  later  stages),  each  of  which 
usually  consisted  of  100  25-cm  Sher- 
man livetraps  arranged  in  a  10  x  10 
grid  with  20-m  spacing.  Other  grid 


sizes  or  shapes  were  used  when  the 
plot  configuration  would  not  contain 
the  standard  grid.  Traps  were 
checked  each  day  for  5  days  (based 
on  pilot  studies,  Raphael  and  Barrett 
1981)  during  July  in  1981  (late  stages 
only),  1982,  and  1983  (all  stages).  Re- 
sults for  each  species  were  expressed 
as  mean  number  of  captures  per  100 
trapnights. 

Surface  Searcti 

To  better  sample  certain  amphibian 
species,  we  conducted  time-  and 
area-constrained  searches  (Bury  and 
Raphael  1983,  Raphael  1984)  on  a 
subset  of  sites  in  1981  (late  stages), 
1982,  and  1983  (all  stages).  A  two- 
person  team  searched  under  all  mov- 
able objects  and  within  logs  on  three 
randomly  located  0.04-ha  circular 
subplots  (fall  1981, 1982)  or  within  a 
1-ha  area  for  4  working  hours  (spring 
1983).  We  conducted  20  surveys  in 
each  of  the  three  early  stages  and  29, 
39,  and  48  surveys  in  the  three  late 
stages. 

Opportunistic  Observations 

Observers  recorded  the  presence  of 
vertebrates  or  identifiable  vertebrate 
sign  incidental  to  the  above  proce- 
dures. We  tallied  observations  to  cal- 
culate frequency  of  occurrence  of 
rarer  species  within  each  stage. 

Forest  Area  Trends 

Estimates  of  historical,  current,  and 
future  acreage  in  each  serai  stage 
were  taken  from  Raphael  et  al.  (in 
press).  For  these  analyses,  I  com- 
bined similar  pairs  of  serai  stages 
into  three  generalized  stages  repre- 
senting brush/sapling,  pole/ sawtim- 
ber, and  mature  timber.  I  then  com- 
puted relative  abundance  of  each 
vertebrate  species  in  these  three 
stages  using  a  weighted  average 
(weights  based  on  sampling  effort)  of 


24 


estimates  from  each  of  the  two  stages 
forming  the  pair.  Population  esti- 
mates for  historical,  present,  and  fu- 
ture time  periods  were  computed 
using  the  formula: 

3 

where  P.^  was  the  relative  population 
size  of  the  zth  vertebrate  species  at 
time  t,D..  was  the  relative  abundance 
of  the  zth  vertebrate  in  the  /th  serai 
stage,  and  A  .^  was  the  total  area  of 
each  of  the  tkree  serai  stages  at  time 


RESULTS 

Vegetation  Structure 

Comparisons  of  vegetation  structure 
among  the  serai  stages  (table  1) 
showed  that  older  stands  had  greater 


canopy  volume,  basal  area,  litter 
depth,  and  density  of  Douglas-fir 
stems  >90  cm  d.b.h.  Downed  wood 
mass  differed  among  stages,  but  the 
greatest  volume  occurred  in  the 
youngest  stands,  probably  in  the 
form  of  logging  slash,  and  the  lowest 
volume  occurred  in  pole  and  sawtim- 
ber  stages.  Early-seral  stands  were 
higher  in  elevation  than  older  stands, 
probably  because  of  the  logistics  of 
timber  harvest  in  the  area  (most 
clearcuts  were  located  along  ridg- 
etops).  Stands  in  the  two  earliest 
serai  stages,  also  because  of  logging, 
were  smaller  in  area  than  stands  in 
the  four  older  stages. 


Vertebrate  Abundance  and 
Diversity 

Among  all  plots  and  years  of  study, 
we  recorded  9,928  captures  of  all 


Table  1  .—Comparisons  of  vegetation  characteristics  among  serai  stages  of 
Douglas-fir  forest,  northwestern  California,  1981-1983. 


Characteristic 


Early  Late 

brush/  brush/  Old- 
sapiing  sapling    Pole  SawtimberMature  growth 


Canopy  volume  (jTT'lrv?) 

10.77 

n.26 

^3.64 

7.15 

7,52 

7.47 

Live  stem  basal  area  (m^hc) 

^2.6 

^52.8 

50.5 

60.2 

65.6 

Snog  basal  area  (m^/ha) 

2_ 

4.7 

6.1 

5.3 

Downed  wood  moss  (metric  tons/ha) 

<8  cm  diameter 

^9.7 

V.9 

m.9 

12.9 

12.3 

11.5 

>8  cm  diameter 

^81.4 

V4.7 

^52.4 

32.3 

43.6 

67.3 

Utter  depth  (cm) 

^2.2 

M.8 

^6.0 

6.2 

5.1 

7.1 

Douglas-fir  >90  cm  d.b.h.  (n/ha)  — 

3.6 

19.3 

25.7 

Elevation  (m) 

1128 

1016 

972 

660 

832 

904 

Stand  area  (ha) 

12,3 

21.9 

41.2 

47.1 

62.0 

84.2 

Solar  radiation  index^ 

0.34 

0.41 

0.51 

0.49 

0.49 

0.43 

Slope  (%) 

48 

30 

31 

36 

41 

52 

Age  (years  since  clearcut. 

or  index) 

9 

14 

123 

206 

294 

'Dofo  are  from  Marcot(1984),  with  permission,  and  represent  a  larger  number  of 
sites  than  v/ere  sampled  in  tt)e  present  study. 

'Dasties  indicate  no  values  were  available. 

^ Index  of  total  yearly  solar  energy  flux  (Frank  and  lee  1966).  Larger  values  indicate 
warmer,  drier  sites. 


J 


Species  during  898,431  trapnights 
from  pitfalls  and  drift  fences;  1,636 
captures  of  amphibians  during  sur- 
face searches;  3,066  small  mammal 
captures  during  35,070  trapnights 
from  Hvetrap  grids;  and  510  detec- 
tions of  larger  mammals  from  track 
stations.  Relative  abundances  of  55 
species,  based  on  the  most  appropri- 
ate sampling  method  for  each  spe- 
cies, are  summarized  in  table  2.  Val- 
ues are  comparable  across  stages  but 
not  among  taxa  if  different  sampling 
methods  were  used.  Amphibians 
were  much  more  abundant  in  for- 
ested than  in  clearcut  stands, 
whereas  reptiles  were  more  abun- 
dant in  clearcuts.  None  of  the  am- 
phibians and  reptiles  [except  rarer 
species  such  as  northwestern  sala- 
mander (see  appendix  for  scientific 
names  of  vertebrates)]  was  absent 
from  any  stage. 

Mammals  exhibited  a  greater  vari- 
ety of  responses  to  serai  stage.  Some 
(e.g.,  Douglas'  squirrel,  western  red- 
backed  vole)  increased  in  abundance 
from  earliest  to  latest  serai  stages; 
others  (e.g.,  deer  mouse)  decreased 
along  this  gradient.  A  number  of  spe- 
cies (e.g.,  Allen's  chipmunk,  dusky- 
footed  woodrat,  pinyon  mouse,  Cali- 
fornia vole)  were  most  abundant 
both  in  late  shrub-sapling  and  ma- 
ture or  old-growth  stands. 

Mean  numbers  of  mammal  and 
reptile  species  recorded  per  stand 
differed  among  serai  stages,  but 
mean  numbers  of  amphibian  species 
did  not  differ  significantly  (fig.  1). 
Among  mammals,  mean  numbers  of 
species  were  greatest  in  mature  and 
old-growth  stages.  In  contrast,  mean 
numbers  of  reptile  species  were 
greatest  in  the  two  earliest  stages. 


Long-Term  Trends 

Estimates  of  land  area  in  each  serai 
stage  through  time  (table  3)  indicate 
more  area  is  occupied  by  early  serai 
stages  currently  than  during  historic 
or  future  times.  Mature  and  old- 
growth  stages  currently  occupy 


25 


Table  2.~Mean  relative  abundance  of  amphibians,  reptiles,  and  mammals  among  serai  stages  of  Douglas-fir  forest, 
nortliwestern  California,  1981 -1983. 


Species' 


Sampling  . 
metliod(s)2 


Total 
captures 


Relative  abundance 
among  serai  stages^ 


6  Significance* 


Salamanders 

Northwestern  salamander^ 
Pacific  giant  salamander 
Olympic  salamander^ 
Rough-skinned  newt 
Del  Norte  salamander 
Ensatina 

Black  salamander 
Clouded  salamander 

Frogs  and  toads 
Tailed  frog^ 
Western  toad 
Pacific  treefrog 
Foothill  yellow-legged  frog^ 
Bullfrog^ 

Turtles 

Western  pond  turtle^ 


PD.  TC^ 
PD 
TC 
TC^ 


PD,TC^ 

6 

0 

0 

0 

1 

3 

4 

PD 

28 

0 

0.05 

0 

0.01 

0.02 

0.04 

PD,TC6 

5 

0 

0 

0 

1 

1 

3 

PD 

68 

0.02 

0 

0 

0.05 

0.09 

0.04 

TC 

196 

0.70 

0.60 

0.05 

0.07 

1.92 

1.92 

TC 

1116 

2.40 

1.85 

8.10 

6.28 

8.15 

7.69 

TC 

32 

0.05 

0.05 

0.05 

0.03 

0.21 

0.42 

TC 

103 

0.35 

1.55 

0.50 

0.10 

0.31 

0.83 

0,114 

0.403 
0.035 
0.001 

o.on 

0.009 


3 

0 

0 

0 

0 

2 

0 

54 

0.18 

0.03 

0.02 

0.08 

0,06 

0.01 

0.035 

51 

0.60 

0.05 

0,10 

0.55 

0,03 

0,06 

0.000 

6 

1 

0 

0 

1 

0 

0 

3 

0 

0 

0 

1 

0 

0 

5 

0 

0 

0 

4 

0 

0 

Lizards 

Western  fence  lizard  PD 

Sagebrush  lizard  PD 

Western  skink  PD 

Southern  alligator  lizard  PD 

Northern  alligator  lizard  PD 

Snakes 

Rubber  boa^ 
Rtngneck  snake^ 
Sharp-tailed  snake^ 
Racer^ 

Gopher  snake-^ 
Common  kingsnake^ 
Common  gartersnake5 
Western  terrestrial  gartersnakeS 
Western  ratt-lesnake5 

Mammals 

Pacific  shrew  PD 

Trowbridge's  shrew  PD 

Shrev/-mole  PD 

Coast  mole^  PD 

Allen's  chipmunk  LT 

Weste rn  g ray  squi rrel  TP^ 

Douglas'  squirrel  TP* 

Northern  flying  squirrel  TP* 


523 

1.77 

2.38 

0.30 

0.94 

0.54 

0.11 

0.000 

196 

2.66 

0.76 

0.25 

0.09 

0.11 

0.01 

0.000 

584 

3.05 

3.47 

0.78 

0.73 

0,42 

0.13 

0,000 

41 

0.03 

0 

0 

0.11 

0,05 

0.03 

0.085 

586 

0.81 

1.03 

0.90 

0.97 

0.60 

0.44 

0,029 

CO,  PD^ 

7 

0 

20 

10 

0 

4 

0 

OCPD* 

6 

0 

0 

0 

0 

4 

0 

PD^ 

22 

10 

20 

30 

0 

5 

4 

OCPD^ 

8 

0 

20 

0 

4 

5 

4 

GO,  PD* 

0 

0 

10 

0 

2 

0 

00,PD* 

1 

0 

0 

0 

4 

0 

0 

GO,  PD6 

19 

20 

10 

20 

0 

5 

11 

400,PD 

11 

0 

20 

10 

7 

4 

6 

oo.  pd;tc6 

5 

0 

10 

0 

0 

2 

6 

89 

0.02 

0.08 

0 

0.07 

0.07 

0.17 

2384 

2.70 

4.01 

2.83 

3.04 

3.16 

3.80 

479 

0.04 

0.16 

0.25 

0.76 

0.55 

0.55 

15 

0 

0.03 

0 

0.02 

0.05 

0.06 

254 

16.7 

29.5 

0.8 

2.8 

5.2 

5.0 

48 

0 

0 

10 

12 

12 

9 

104 

0 

0 

20 

16 

22 

30 

43 

0 

0 

15 

9 

18 

13 

0.004 
0.215 
0.002 

0.003 
0.378 
0.001 
0.046 

(continued) 


26 


Table  2. —(continued). 


Species' 


Sampling 
method(s)2 


Total 
captures 


Relative  abundance 
among  serai  stages^ 


6  Significance* 


Deer  mouse 

PD 

1  Ml 

5.09 

0  C\~l 

o.U/ 

U.OV 

u.bo 

U.Vo 

1.28 

0.000 

Brush  mouse 

LT 

33 

0 

0.33 

3.67 

0.25 

0.25 

0 

0.216 

Pinyon  mouse 

LT 

222 

1.35 

10.34 

4.67 

10.63 

3.86 

2.76 

0.086 

Dusky-footed  woodrat 

LT 

115 

1.9 

3.5 

0.2 

1.2 

4.4 

3.4 

0.000 

Western  red-backed  vole 

PD 

669 

0,35 

0.36 

0.46 

0.45 

0.82 

0.97 

0.015 

Red  tree  vole 

PD 

19 

0 

0.10 

0 

0.07 

0.11 

0.15 

0.586 

California  vole 

PD 

106 

0.89 

1.70 

0.03 

0.02 

0.01 

0.01 

0.000 

Creeping  vole 

PD 

22 

0.09 

0.03 

0.05 

0.04 

0.01 

0.01 

0,038 

Western  jumping  mouse^ 

PD 

2 

0 

0 

0 

0 

0.04 

0.02 

Coyote^ 

ALL^ 

7 

10 

30 

0 

15 

9 

15 

Gray  fox 

TP^ 

63 

20 

15 

10 

30 

11 

8 

0.001 

Black  bear 

TP 

196 

20 

25 

5 

42 

45 

48 

0.028 

Ringtail 

TP 

25 

0 

0 

0 

10 

6 

4 

0.249 

Raccoon^ 

TP 

3 

0 

0 

0 

0 

1 

1 

Fisher 

TP 

58 

0 

5 

25 

6 

13 

15 

0.060 

Ermine^ 

PD 

2 

0 

0 

0 

0 

0,02 

0.02 

Western  spotted  skunk 

TP 

70 

10 

15 

5 

10 

18 

15 

0.426 

Striped  skunk^ 

TP 

17 

0 

0 

0 

7 

6 

1 

Bobcat^ 

TP 

3 

5 

5 

0 

1 

2 

0 

'A//  names  follow  Laudenslayer  and  Grenf ell  (1983). 

^PD  =  Pitfall  plus  drift  fence,  TC  =  Time-  and  area-constrained  searcti.  OO  =  Opportunistic  observations,  TP  =  Track  plots.  LT=  Live 
traps,  ALL  =  all  observation  methiods  combined. 

^Seral  stages  (and  numbers  of  stands  sampled)  are:  1  — early  brustt/sapling  (n=  10):  2— late  brusti/sapling  (n=  10):  3— pole  (n=  10):  4— 
sawtimber  (n=27):  mature  (r]=56);  5—old-growtht  (n=63). 

"Significance  from  analysis  of  variance  (means)  or  cN-square  analysis  (frequencies)  comparing  abundances  among  stages.  A  dashi 
indicates  th\at  no  test  was  performed. 

^Too  rare  for  subsequent  analyses. 

'^Abundance  values  based  on  percent  frequencies. 


about  half  of  historic  acreage,  and 
these  stages  will  probably  occupy 
only  about  30%  of  current  acreage 
under  the  most  likely  harvest  pat- 
terns of  the  future  (table  3). 

The  implications  of  these  changing 
distributions  of  serai  stages  for  am- 
phibians, reptiles,  and  mammals  are 
summarized  in  figure  2.  Nearly  equal 
numbers  of  species  are  likely  to  have 
increased  or  decreased  by  more  than 
25%  relative  to  historic  abundance  at 
present  and  in  the  future.  Three  of 
the  five  reptile  species  are  presently 
more  abundant  than  in  historic  times 
and  all  five  species  will  likely  be 
more  abundant  in  the  future.  Am- 
phibians showed  an  opposite  pattern. 


Four  of  the  eight  species  are  pres- 
ently less  abundant  and  five  of  the 
eight  may  be  less  abundant  in  the  fu- 
ture. Among  the  20  mammal  species, 
seven  are  presently  less  abundant 
than  in  historic  times  whereas  five 
are  more  abundant.  Eight  species 
will  probably  be  less  abundant  in  the 
future  and  six  more  abundant. 


DISCUSSION 

Abundance  in  Serai  Stages 

Results  suggest  late  brush/ sapling 
and  mature/old-growth  serai  stages 
provided  more  productive  wildlife 


habitat  than  early  brush/ sapling, 
pole,  and  sawtimber  stages.  Among 
amphibians,  only  ensatinas  were  cap- 
tured frequently  in  pole  sites. 
Clouded  salamanders  were  generally 
under  bark  or  inside  downed  logs 
and  persisted  in  clearcut  stands  as 
long  as  adequate  numbers  of  logs 
were  retained,  especially  in  late  sites 
(Raphael  1987,  Welsh,  this  volume). 

Lizards  were  more  abundant  in 
earlier  serai  stages  than  in  pole  and 
mature  stages.  Among  snakes,  only 
sharp-tailed  snakes  were  observed 
on  early  sites;  other  species  occurred 
on  later  sites.  However,  sampling 
was  not  sufficient  for  definitive  con- 
clusions. 


27 


With  the  exception  of  the  deer 
mouse,  small  mammals  were  more 
abundant  on  late  brush/ sapling  sites. 
Dusky-footed  woodrats  were  of  spe- 
cial interest  in  this  regard  as  we  ob- 
served many  woodrat  nests  built 
among  the  stems  of  tanoak  and  Pa- 
cific madrone  in  late  brush/ sapling 
sites.  The  combination  of  abundant 
mast,  good  nesting  substrate,  and 
protection  from  predation  (spotted 
owls  rarely  forage  in  old,  brush- 
dominated  clearcuts)  provided  by 
the  dense,  brushy  cover  were  proba- 
bly the  reasons  that  woodrats  and 
other  small  mammals  were  more 
numerous  in  late  clearcut  sites  (Ra- 
phael 1987). 

Tree  squirrels  were  most  abun- 
dant in  mature  forest  sites  and 
ground  squirrels  were  more  abun- 
dant in  early  clearcut  sites.  Chip- 
munks were  the  only  squirrel  that 
reached  peak  abundance  in  early 
serai  sites.  Their  abundance  was  cor- 
related with  the  cover  of  tanoak  in 
the  understory  (Raphael  1987).  Man- 
agement actions,  such  as  herbicide 
treatments,  that  shorten  or  delete  the 
late  brush/ sapling  stage  are  probably 
detrimental  to  chipmunks,  woodrats, 
and  certain  other  rodents. 

Several  carnivorous  mammals 
were  abundant  in  the  late  brush/sap- 
ling stage.  Greater  prey  density  in 
late  compared  to  early  and  pole  sites 
may  explain  this  higher  frequency  of 
carnivores  although  more  data  will 
be  necessary  to  confirm  this  observa- 
tion. 

Of  the  55  species  observed,  20 
were  strongly  associated  with  either 
older  (9  species)  or  younger  (11  spe- 
cies) stands  (table  4).  Three  salaman- 
ders and  six  mammals  were  associ- 
ated with  older  stands.  One  toad, 
one  frog,  five  lizards,  and  four  mam- 
mals were  associated  with  younger 
stands.  Five  species  associated  with 
old-growth  were  also  abundant  in 
late  (brushy)  clearcut  stages  (table  2). 
These  species  peak  in  abundance  in 
old  stands  and  late  clearcuts,  with 
low  abundance  in  intermediate  age 
classes. 


Table  3  — Approximate  area  (millions  of  ha)  of  serai  stages  in  Douglas-fir 
forests  of  northwestern  California  In  historic,  present,  and  future  time  peri- 
ods (after  Raphael  et  al.,  in  press). 


Serai  stage 

Historical 

Present 

Likely 
future 

Worst 
case 
future' 

Brush/sapling 

0.14 

0.49 

0.20 

0.24 

Young  forest 

0.14 

0.20 

0.77 

0.85 

Mature  forest 

0.81 

0.40 

0.12 

0.00 

'Assumes  that  all  mature  and  old-growth  stands  ore  harvested  and  all  lands  man- 
aged under  short  rotatiorts. 


CO 


PRESOff 


12  r 


3  4 

SERAL  STAGE 


Figure  1  .—Mean  numbers  of  amphibian, 
reptile,  and  mammal  species  observed  in 
serai  stages  of  Douglas-fir  forest,  northwest- 
ern California,  1981-1983.  Serai  stages  (and 
numbers  of  stands  sampled)  are:  1  -  early 
brush/sapling  (n  =  10);  2  -  late  brush/sap- 
ling (n  =  1 0);  3  -  pole  (n  =  1 0);  4  -  sawtimber 
(n  =  27);  5  -  mature  (n  =  56);  6  -  old-growth 
(n  =  S3).  Vertical  lines  indicate  95%  confi- 
dence intervals. 


UttLYFUIUt 


iMMi 

jlBTlB 


<-75 


-75       -50       -TS  • 

CHANGE  h  ASUNQANCE  (1 


Figure  2.— Percent  change  in  population 
size  of  amphibian,  reptile,  and  mammal 
species  at  present  and  in  the  future  relative 
to  estimated  historical  populations.  Histo- 
grams represent  the  numbers  of  species 
increasing  or  decreasing  by  specified  per- 
centages. 


28 


I  examined  habitat  associations 
among  each  of  the  above  9  species  by 
computing  correlations  of  their  abun- 
dance with  specific  habitat  compo- 
nents (table  5).  Density  of  large  trees 
and  hardwood  volume  were  corre- 
lated with  the  abundance  of  most 
species.  Moisture,  as  measured  by 
the  presence  of  surface  water,  mois- 
ture-loving tree  species,  or  north-fac- 
ing slopes,  was  important  for  most 
mammals  and  one  salamander  spe- 
cies. Four  mammal  S}:>ecies  were  sig- 
nificantly more  abundant  on  higher 
elevation  stands.  Downed  wood  vol- 
ume also  was  significantly  and  posi- 
tively correlated  with  abundance  of 
four  amphibian  and  mammal  spe- 
cies. The  abundance  of  hardwoods  in 
the  understory  was  important  for 
many  species  in  each  group.  In  con- 
trast, snag  density  was  not  positively 
correlated  with  the  abundance  of  any 
species. 


Long-Term  Trends 

The  list  of  sensitive  species  (table  4) 
is  tentative  pending  results  of  addi- 


r 


Table  5.— Habttat  components  that  were  correlated  with  relative  abun- 
dance of  amphibians  and  mammals  associated  with  late-seral  Douglas-fir 
forests  of  northwestem  California. 


Density  of 

Hardwood 

Downed 

conifers 

under- 

wood 

Standing 

Species                            >90-cm  d.b.h. 

story 

mass 

snags 

Moisture 

Elevation 

Del  Norte  salamander 

X 

Black  salamander 

X 

X 

Clouded  salamander  X 

X 

Pacific  shrew  X 

X 

X 

X 

Douglas'  squirrel  X 

^(X) 

X 

X 

X 

Northern  flying  squirrel 

X 

(X) 

Dusky  footed  woodrat  X 

X 

X 

X 

Western  red-backed  vole  X 

X 

X 

X 

Fisher  X 

X 

'Parentheses  indicate  negative  correlations. 


tional  surveys  and  more  intensive, 
species-specific  research.  The  projec- 
tions, although  based  on  an  intensive 
sampling  effort,  are  highly  specula- 
tive. Three  assumptions  must  be  rec- 
ognized to  interpret  these  results. 
First,  I  assumed  that  greater  relative 
abundance  in  a  serai  stage  indicates  a 
species'  preference  for  that  stage  and 
that  preferences  remain  constant 
with  shifting  distribution  of  acreage 


r 


Table  4.— Amphibian,  reptile,  and  mammal  species  most  strongly  affected 
by  future  harvest  of  old-growth  Douglas-fir  forest,  northwestern  California.' 


Decreasers— associated 
with  late-seral  forest 


Increasers— associated 
with  early-seral  forest 


Species 


decline^  Species 


%  increase^ 


Del  Norte  salamander 

75 

Western  toad 

45 

Black  salamander 

71 

Pacific  treefrog 

160 

Clouded  salamander 

29 

Western  fence  lizard 

60 

Pacific  shrew 

39 

Sagebrush  lizard 

44 

Douglas'  squirrel 

31 

Western  skink 

59 

Northern  flying  squirrel 

31 

Southern  alligator  lizard 

60 

Dusky-footed  woodrat 

55 

Northern  alligator  lizard 

43 

Western  red-backed  vole 

37 

Pinyon  mouse 

70 

Fisher 

26 

California  vole 

44 

Creeping  vole 

102 

Gray  fox 

78 

'Species  were  listed  if  thieir  estimated  future  abundance  differed  by  more  ftian  25% 
from  estimated  historical  abundance  and  if  mean  abundance  differed  significantly  (P 
<0.10)  among  serai  stages  (table  2). 

'Percent  increase  or  decrease  in  estimated  hjture  abundance  compared  with 
estimated  historic  abundance. 


in  each  stage.  Some  species  have  (or 
could)  adapt  to  new  stages  over  time. 

Second,  I  assumed  total  acreage  of 
each  serai  stage  can  be  used  to  esti- 
mate responses  of  vertebrates  with- 
out regard  to  size  and  juxtaposition 
of  stands  comprising  each  stage. 
However,  continued  fragmentation 
of  forest  habitats  may  result  in  dis- 
junct patches  so  small  they  cannot 
support  a  species  that  would  other- 
wise find  the  habitat  suitable.  Rosen- 
berg and  Raphael  (1986)  found  that 
at  least  eight  species  of  amphibians 
(2),  reptiles  (2),  and  mammals  (4) 
were  significantly  less  abundant  in 
stands  <10  ha  in  size  than  in  larger 
stands.  Some  of  these  (e.g.,  western 
gray  squirrel)  were  not  listed  in  this 
study  among  the  sensitive  species 
(table  4),  but  the  effects  of  habitat 
fragmentation  may  nonetheless  be 
cause  for  concern. 

A  third  assumption  is  that  young 
forested  stands  (pole,  sawtimber)  in 
this  study  represent  young  stands  of 
the  future.  Naturally  occurring  pole 
and  sawtimber  stands  contain  some 
large  Douglas-fir  stems  and  a  sub- 
stantial amount  of  standing  and 
downed  wood  (table  1).  If  future 
management  activities  result  in  fewer 
large  live  trees,  snags,  and  downed 
logs,  the  abundance  of  vertebrates 
associated  with  these  habitat  compo- 
nents may  also  decline.  In  this  case. 


29 


responses  of  vertebrates  to  forest 
management  may  be  more  extreme 
than  those  projected. 

The  overall  trend  is  for  increased 
abundance  among  species  of  south- 
ern affinity  that  are  associated  with 
open,  drier  habitats  in  other  parts  of 
their  ranges,  and  decreased  abun- 
dance among  species  of  boreal  affin- 
ity that  are  primarily  associated  with 
moist  coniferous  forest  throughout 
their  ranges.  Furthermore,  most  of 
the  increasers  are  widespread  species 
with  large  distributions  that  include 
many  nonthreatened  habitats.  In  con- 
trast, the  decreasers  are  almost  all 
species  with  rather  restricted  total 
ranges,  most  of  which  are  in  threat- 
ened habitats.  Therefore,  even 
though  total  numbers  of  increasers 
and  decreasers  are  nearly  equal,  the 
effects  of  old-growth  reduction 
should  not  be  viewed  as  neutral. 

Because  many  of  the  decreasers 
are  affected  by  soil  moisture  and 
other  microclimatic  conditions,  man- 
agement to  protect  stream  edges, 
moist  ravines,  and  other  moist  sites 
may  provide  refuges  for  species  that 
can  later  recolonize  maturing  stands. 
Management  efforts  to  retain  (or  rec- 
reate) natural  components  of  regen- 
erating stands,  such  as  hardwood 
understory,  snags,  and  logs,  may 
help  mitigate  against  wildlife  losses 
in  future  forests.  It  is  not  stand  age, 
per  se,  but  the  structural  characteris- 
tics of  forests  of  various  ages  that  are 
important  to  survival  of  most  spe- 
cies. 

Finally,  results  of  this  study  ad- 
dress another  important  forest  man- 
agement issue  in  the  northwest; 
What  should  managers  use  as  a 
baseline  for  evaluation  of  impacts: 
historic  or  present  conditions?  It  is 
apparent  that  many  species  are  pres- 
ently much  less  abundant  compared 
with  historic  numbers  (fig.  2).  Addi- 
tional reductions  because  of  contin- 
ued timber  harvest  will  cause  further 
declines  in  some  species  but  most 
major  declines  have  already  oc- 
curred. Therefore,  I  believe  that  esti- 
mates of  historic  populations  should 


be  used  as  baselines  for  monitoring 
biological  diversity,  rather  than  pre- 
sent populations. 

ACKNOWLEDGMENTS 

Field  studies  were  funded  by  the  Pa- 
cific Southwest  Region  and  the  Pa- 
cific Southwest  Forest  and  Range 
Experiment  Station  of  the  USDA  For- 
est Service  and  by  the  University  of 
California,  Agricultural  Experiment 
Station  3501  MS.  I  especially  thank 
my  field  assistants  (Paul  Barrett,  John 
Brack,  Cathy  Brown,  Christopher 
Canaday,  Lawrence  Jones,  Ronald 
LaValley,  Kenneth  Rosenberg,  and 
Cathy  Taylor)  for  their  dedication 
and  blisters;  R.  H.  Barrett,  C.  J. 
Ralph,  and  J.  Verner  for  their  sup- 
port; Bruce  G.  Marcot  for  freely  shar- 
ing information  from  his  studies  and 
for  valuable  discussions;  and  Ken- 
neth V.  Rosenberg,  Fred  B.  Samson, 
and  Hobart  M.  Smith  for  their  com- 
ments on  an  earlier  draft  of  this 
manuscript. 

LITERATURE  CITED 

Barrett,  Reginald  H.  1983.  Smoked 
aluminum  track  plots  for  deter- 
mining furbearer  distribution  and 
abundance.  California  Fish  and 
Game  69:188-190. 

Burgess,  Robert  L.,  and  David  M. 
Sharpe.  1981.  Forest  island  dy- 
namics in  man-dominated  land- 
scpates.  Springer- Verlag,  New 
York.  310  p. 

Brown,  James  K.  1974.  Handbook  for 
inventorying  downed  woody  ma- 
terial. USDA  Forest  Service  Gen- 
eral Technical  Report  INT-16.  24  p. 

Bury,  R.  Bruce,  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles,  p. 
426-419.  In  J.  F.  Bell  and  T.  Atter- 
bury  (eds.).  Renewable  Resource 
Inventories  for  Monitoring 
Changes  and  Trends.  College  of 
Forestry,  Oregon  State  University, 
Corvallis,  Oregon. 


Campbell,  H.  W.,  and  S.  P.  Christ- 
man.  1982.  Field  techniques  for 
herptofaunal  community  analysis, 
p.  193-200.  In  N.  J.  Scott  (ed.).  Her- 
petological  Communities.  USDI 
Fish  and  Wildlife  Service  Wildlife 
Research  Paper  13.  239  p. 

Frank  Ernest  C,  and  Richard  Lee. 
1966.  Potential  solar  beam  irradia- 
tion on  slopes.  USDA  Forest  Serv- 
ice Research  Paper  RM-18. 

Harris,  Larry  D.,  Chris  Maser,  and 
Arthur  McKee.  1982.  Patterns  of 
old  growth  harvest  and  implica- 
tions for  Cascades  wildlife.  Trans- 
actions of  North  American  Wild- 
life and  Natural  Resource  Confer- 
ence 47:374-392. 

Laudenslayer,  William  F.,  Jr.,  and 
William  E.  Grenfell,  Jr.  1983.  A  list 
of  amphibians,  reptiles,  birds  and 
mammals  of  California.  Outdoor 
California  44:5-14. 

Marcot,  Bruce  G.  1984.  Habitat  rela- 
tionships of  birds  and  young- 
growth  Douglas-fir  in  northwest- 
ern California.  Corvallis,  OR:  Ore- 
gon State  University;  282  p.  Ph.D. 
dissertation. 

Meslow,  E.  Charles,  Chris  Maser, 
and  Jared  Verner.  1981.  Old- 
growth  forests  as  wildlife  habitat. 
Transactions  of  North  American 
Wildlife  and  Natural  Resource 
Conference  46:329-344. 

Raphael,  Martin  G.  1984.  Wildlife  di- 
versity and  abundance  in  relation 
to  stand  age  and  area  in  Douglas- 
fir  forests  of  northwestern  Califor- 
nia, p.  259-274.  In  Meehan,  W.  R., 
T.  T.  Merrell,  Jr.,  and  T.  A.  Hanley 
(tech.  eds.).  Fish  and  Wildlife  Rela- 
tionships in  Old-growth  Forests: 
proceedings  of  a  symposium  (Jun- 
eau, Alaska,  12-17  April  1982). 
Bookmasters,  Ashland,  Ohio. 

Raphael,  Martin  G.  1987.  Wildlife 
tanoak  associations  in  Douglas-fir 
forests  of  northwestern  California, 
p.  183-189.  In  Plumb,  T.  R.,  N.  H. 
Pillsbury,  (tech.  coord.).  Proceed- 
ings of  the  Symposium  on  Mul- 
tiple-Use Management  in  Califor- 
nia's Hardwood  Resources;  No- 
vember 12-14, 1986,  San  Luis 


30 


Obispo,  CA.  General  Technical 
Report  PSW-100.  Berkeley,  CA: 
Pacific  Southwest  Forest  and 
Range  Experiment  Station,  Forest 
Service,  U.S.  Department  of  Agri- 
culture, 462  p. 
Raphael,  Martin  G.,  and  Reginald  H. 
Barrett.  1981.  Methodologies  for  a 
comprehensive  wildlife  survey 
and  habitat  analysis  in  old-growth 
Douglas-fir  forests.  Gal-Neva 
Wildlife  1981:106-121. 
Raphael,  Martin  G.,  and  Reginald  H. 
Barrett.  1984.  Diversity  and  abun- 
dance of  wildlife  in  late  succes- 
sional  Douglas-fir  forests,  p.  352- 
360.  In  New  Forests  for  a  Chang- 
ing World.  Proceedings  1983  Con- 
vention of  the  Society  of  American 
Foresters.  650  p. 
Raphael,  Martin  G.,  and  Kenneth  V. 
Rosenberg.  1983.  An  integrated 
approach  to  inventories  of  wildlife 


in  forested  habitats,  p.  219-222.  In 
J.  F.  Bell  and  T.  Atterbury  (eds.). 
Proceedings,  conference  on  renew- 
able resource  inventories  for 
monitoring  changes  in  trends. 
Corvallis,  Oregon,  1983. 

Raphael,  Martin  G.,  Kenneth  V. 
Rosenberg,  and  Bruce  G.  Marcot. 
In  press.  Large-scale  changes  in 
bird  populations  of  Douglas-fir 
forests,  northwestern  California. 
Bird  Conservation  3. 

Raphael,  Martin  G.,  Cathy  A.  Taylor, 
and  Reginald  H.  Barrett.  1986. 
Sooted  aluminum  track  stations 
record  flying  squirrel  occurrence. 
Pacific  Southwest  Forest  and 
Range  Experiment  Station  Re- 
search Note  PSW-384. 

Rosenberg,  Kenneth  V.,  and  Martin 
G.  Raphael.  1986.  Effects  of  forest 
fragmentation  on  wildlife  commu- 
nities of  Douglas-fir.  p.  263-272.  In 

Appendix 


Verner,  J.,  M.  L.  Morrison,  and  C. 
J.  Ralph  (eds.).  Modeling  habitat 
relationships  of  terrestrial  verte- 
brates. University  of  Wisconsin 
Press,  Madison,  WI. 

Scott,  J.  Michael,  Blair  Csuti,  James 
D.  Jacobi,  and  John  E.  Estes.  1987. 
Species  richness — a  geographic  ap- 
proach to  protecting  future  bio- 
logical diversity.  Bioscience 
37:782-788. 

Taylor,  Cathy  A.,  and  Martin  G.  Ra- 
phael. 1988.  Identification  of  mam- 
mal tracks  from  sooted  track  sta- 
tions in  the  Pacific  Northwest. 
California  Fish  and  Game  74:4-11. 

Vogt,  R.  C,  and  R.  L.  Hine.  1982. 
Evaluation  of  techniques  for  as- 
sessment of  amphibian  and  reptile 
populations  in  Wisconsin,  p.  201- 
217.  In  N.  J.  Scott,  (ed.).  Herpeto- 
logical  Communities.  USDI  Fish 
and  Wildlife  Service  Research  Re- 
port 13,  239  p. 


Common  and  scientific  names  of  vertebrates  mentioned  in  text  (nomenclature  follows  Laudenslayer  and 

Grenfell  (1983)). 


Salamanders 

Northwestern  salamander   Amhystoma  gracile 

Pacific  giant  salamander    Dicamptodon  ensatus 

Olympic  salamander  Rhyacotriton  olympicus 

Rough-skinned  newt   Taricha  granulosa 

Del  Norte  salamander  Plethodon  elongatus 

Ensatina   Ensatina  eschscholtzi 

Black  salamander   Aneides  flavipunctatus 

Qouded  salamander  Aneides  ferreus 

Frogs  and  toads 

TaUed  frog  Ascaphus  truei 

Western  toad   Bufo  boreas 

Pacific  treefrog   Hyla  regilla 

Foothill  yellow-legged  frog   Rana  boylei 

Bullfrog   Rana  catesbeiana 

Turtles 

Western  pond  turtle    Clemmys  marmorala 

Lizards 

Western  fence  lizard   Sceloporus  occidentalis 

Sagebrush  lizard    Sceloporus  graciosus 

Western  skink   Eumeces  skiltonianus 

Southern  alligator  lizard    Gerrhonotus  multicarinatus 

Northern  alligator  lizard    Gerrhonotus  coeruleus 

Snakes 

Rubber  boa    Charina  botlae 

Ringneck  snake   Diadophis  punctatus 

Sharp-tailed  snake   Phyllorhynchus  decurtatus 

Racer   Coluber  constrictor 

Gopher  snake    Pituophis  melanoleucus 

Common  kingsnake   Lampropeltis  zonula 


Common  gartersnake  Thamnophis  sirtalis 

Western  terrestrial  gartersnake  Thamnophis  elegans 

Western  rattlesnake   Crotalis  viridis 

Mammals 

Pacific  shrew   Sorex  pacificus 

Trowbridge's  shrew  Sorex  trowbridgii 

Shrew-mole   Neurotrichus  gibbsii 

Coast  mole   Scapanus  orarius 

AUen's  chipmunk  Tamias  senex 

Western  gray  squirrel  Sciurus  griseus 

Douglas'  squirrel  Tamiasciurus  douglasii 

Northern  flying  squirrel   Glaucomys  sabrinus 

Deer  mouse   Peromyscus  maniculatus 

Brush  mouse  Peromyscus  boylii 

Pin  yon  mouse   Peromyscus  truei 

Dusky-footed  woodrat  Neotoma  fuscipes 

Western  red-backed  vole  Clethrionomys  californicus 

Red  tree  vole   Arborimus  longicaudus 

California  vole  Microtus  californicus 

Creeping  vole  Microtus  oregoni 

Western  jumping  mouse  Zapus  princeps 

Coyote  Canis  latrans 

Gray  fox  Urocyon  cinereoargenteus 

Black  bear  Ursus  americanus 

Ringtail   Bassanscus  astutus 

Raccoon  Procyon  lotor 

Fisher  Martes  pennanti 

Ermine  Mustela  erminea 

Western  spotted  skunk   Spologale  gracilis 

Striped  skunk  Mephitis  mephitis 

Bobcat   Lynx  rufus 


31 


Use  of  Woody  Debris  by 
Plethodontid  Salamanders  In 
Douglas-Fir  Forests  In 
Washington 

Keith  B.  Aubry,^  Lawrence  L  C.  Jones,^  and 
Patricia  A.  Hali^ 


Abstract.— Ensaf/no  eschscholfzii \f^os  found  most 
often  under  pieces  of  bark,  whereas  Plefhodon 
vehiculum  occurred  primarily  under  logs.  Captures 
of  both  species  were  highest  in  young  stands,  but 
occurred  in  all  age  classes.  Our  results  suggest  that 
the  retention  of  coarse  woody  debris  in  managed 
forests  would  provide  for  the  habitat  needs  of  these 
species. 


The  harvesting  of  old-growth 
Douglas-fir  (Pseudotsuga  menziesii) 
forests  in  the  Pacific  Northwest,  and 
its  potential  effects  on  wildlife  spe- 
cies has  been  the  focus  of  much  con- 
cern in  recent  years  (e.g..  Lumen  and 
Nietro  1980,  Franklin  et  al.  1981, 
Meslow  et  al.  1981,  Meehan  et  al. 
1984,  Gutierrez  and  Carey  1985). 
Most  of  this  attention  has  been  di- 
rected towards  birds  and  mammals 
such  as  the  spotted  owl  (Strix  occiden- 
talis),  Vaux's  swift  (Chaetura  vauxi), 
northern  flying  squirrel  (Glaucomys 
sahrinus),  and  red  ti'ee  vole  (Ar- 
borimus  longicaudus);  little  concern 
has  been  expressed  about  amphibi- 
ans and  reptiles.  These  groups  have 
not  been  studied  extensively  in  the 
Pacific  Northwest.  Only  recently  has 
research  been  conducted  on  habitat 
associations  among  different  forest 
age  classes  (Raphael  1984,  Raphael 
and  Barrett  1984,  Ruggiero  and 
Carey  1984). 

From  1983  to  1986,  the  USDA  For- 
est Service  and  USDI  Bureau  of  Land 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ,  July  19-21,  1988). 

^Researcti  Wildlife  Biologist,  USDA  Forest 
Service,  Pacific  Nortt)west  Research)  Station, 
3625  93rd  Ave.  SW,  Olympia,  WA  98502. 

^Biological  Technician,  USDA  Forest  Serv- 
ice, Pacific  Northwest  Research  Station, 
3625  93rd  Ave.  SW,  Olympia,  WA  98502. 

"Wildlife  Biologist,  USDA  Forest  Service, 
Pacific  Northwest  Research  Station,  3625 
93rd  Ave.  SW,  Olympia,  WA  98502. 


Management  funded  a  major  re- 
search effort  aimed  at  identifying 
wildlife  species  that  occur  in  highest 
abundances  in  old-growth  Douglas- 
fir  forests  and  investigating  the  eco- 
logical basis  of  observed  patterns  of 
association 

Amphibian  communities  were 
sampled  using  pitfall  traps,  stream 
surveys,  and  time-constrained 
searches  (Standard  Sampling  Proto- 
cols on  file  at  the  Forestry  Sciences 
Laboratory,  Olympia,  WA).  Some  of 
the  results  of  these  studies  are  re- 
ported elsewhere  in  this  volume 
(Bury  and  Corn  1988,  Welsh  1988). 
Here,  we  report  the  results  of  time- 
constrained  searches  conducted  in 
southern  Washington  in  1984.  Our 
objectives  are  to  (1)  identify  potential 
habitat  associations,  (2)  examine  pat- 
terns of  cover  object  use,  and  (3) 
evaluate  the  efficacy  of  this  technique 
for  studying  amphibians  in  this  re- 
gion. 

Study  Area 

Forty-five  forest  stands  were 
sampled  in  the  southern  portion  of 
the  Cascade  Range  in  Washington 
(fig.  1).  Stands  ranged  in  age  from  55 
to  730  yr  and  were  at  least  20  ha  in 
size.  All  stands  were  located  within 
the  western  hemlock  (Tsuga  heter- 
phylla)  zone  and  lower  elevations  of 
the  Pacific  silver  fir  (Abies  amabilis) 
zone  (Franklin  and  Dyrness  1973), 
which  are  characterized  by  a  wet  and 


Figure  1  .—Location  of  study  stands  by  age 
class  in  the  southern  Washington  Cascade 
Range. 

mild  maritime  climate.  Snow  rarely 
accumulates  at  our  sites. 

Old-grovv^th  stands  (210-730  yr) 
typically  contained  high  proportions 
of  Douglas-fir  and  western  hemlock 
and,  in  wet  sites,  western  redcedar 
(Thuja  plicata).  Mature  (95-190  yr) 
and  young  (55-80  yr)  stands  were 


32 


dominated  by  Douglas-fir.  In  all  age 
classes,  other  species  such  as  red 
alder  (Alnus  rubra),  vine  maple  (Acer 
circinatum),  bigleaf  maple  (A.  macro- 
phyllum),  Pacific  silver  fir,  and  west- 
ern hemlock  occurred  in  lesser 
amounts. 

Average  age  of  each  stand  was 
determined  through  growth  ring 
counts,  either  by  increment  coring  or 
examination  of  cut  stumps  in  nearby 
stands.  Old-growth  stands  were  clas- 
sified into  wet,  moderate,  and  dry 
moisture  classes  on  the  basis  of  flo- 
ristic  and  physiographic  characteris- 
tics; all  young  and  mature  stands 
were  in  the  moderate  moisture  class 
(T.  A.  Spies,  unpubl.  data).  All  stands 
had  resulted  from  natural  regenera- 
tion following  fires;  none  had  under- 
gone silvicultural  treatments. 


Methods 

Surveys  for  terrestrial  amphibians 
were  conducted  from  16  April  to  12 
June  1984;  all  but  four  high-elevation 
stands  were  sampled  by  4  May.  A 
time-  constrained  search  method  was 
used  (Campbell  and  Christman 
1982).  A  crew  of  two  to  four  persons 
actively  searched  each  stand  for  am- 
phibians for  a  total  of  4  person- 
hours.  An  initial  search  area  was  se- 
lected at  least  50  m  within  the  stand 
to  avoid  edge  effects. 

In  general,  woody  debris  such  as 
logs,  snags,  and  pieces  of  bark  was 
abundant  in  each  stand  and  consti- 
tuted virtually  all  potential  cover  ob- 
jects. An  area  was  searched  for  0.5 
person-hours,  after  which  we  moved 
a  minimum  of  25  m  to  search  another 
suitable  area;  sampling  areas  were 
not  spatially  constrained.  This  was 
repeated  until  the  sampling  period 
was  over.  All  potential  cover  objects 
were  searched  by  hand  or  with  po- 
tato rakes,  but  no  single  object  was 
searched  for  more  than  20  min.  Logs 
of  all  sizes  in  advanced  stages  of  de- 
composition were  pulled  apart  with 
potato  rakes.  Areas  beneath  large 
undecomposed  logs  could  not  be 


Table  1.— Amphibian  species  captured  during  time-constrained  searches 
in  the  southern  Washington  Cascade  Range  by  stand  type.' 


Species 


Mean  Captures  ±  Standard  Error 

(N=9)  <N=9)  (N=6)  (N=17)  (N=4) 
YNG       MAT      OGW     OGM  OGD 


Caudata 

Plethodontidae 

Ensafina  eschscholfzh 

Plefhodon  vehiculum 
Ambystomatidae 

Ambysfoma  gracile 

A.  macrodacfylum 
Salamadridae 

Taricha  granulosa 
Dicamptodontidae 

Rhyacofrifon  olympicus 
Anura 

Leiopelmatidae 

Ascaphus  fruei 
Ranidae 

Rana  aurora 

R.  cascadae 


5.9±1.8  2,6±0.8   1.2±1.2  2.8±0.9  2.5+1.0 

3.1±2,2  Q.4±0.3  0.2±0.2  0.5+0.3  2.3±2.3 

0.3±0.1 
0.2+0.2  0.1+0.1 


0.1+0.0 


0.2+0,1 


0.1+0.1 


0.1±0.1  0.5+0.5 
0.1+0.1 

0.2+0.1 

0.1+0.1 


'YNG^Young.  MAT=Mafure.  OGW=Wef  Old  Growth.  OGtv1=Moderafe  Old  Growth. 
OGD=Dry  Od  Growth 


searched  in  most  cases.  Little  effort 
was  expended  searching  leaf  litter,  as 
this  has  been  shown  to  be  relatively 
ineffective  when  sampling  amphibi- 
ans in  Douglas-fir  forests  (Bury  and 
Raphael  1983).  Areas  near  seeps, 
streams,  ponds,  rock  outcrops,  and 
other  areas  not  representative  of  the 
stand  were  avoided. 

Modifications  of  methods  devel- 
oped by  Raphael  (1984)  were  used  to 
describe  capture  sites.  The  following 
information  was  recorded  for  each 
individual  captured:  species,  vertical 
position  in  relation  to  cover  object, 
snag  or  log  decay  class,  length  and 
width  of  cover  object,  and  slope  and 
aspect  of  capture  site.  All  amphibians 
were  collected,  measured,  and  pre- 
served, usually  on  the  same  day. 
Snout- vent  length  (to  anterior  margin 
of  vent),  total  length,  and  weight 
were  recorded.  Specimens  were  de- 
posited in  the  Museum  of  Vertebrate 
Zoology,  University  of  California, 
Berkeley. 


Results 


Captures 

A  total  of  214  amphibians,  including 
6  species  of  salamanders  and  3  spe- 
cies of  frogs,  were  captured;  no  rep- 
tiles were  encountered  (table  1).  Only 
two  species  of  plethodontid  salaman- 
ders, the  ensatina  (Ensatina  es- 
chscholtzii)  and  western  redback 
salamander  (Plethodon  vehiculum), 
were  captured  in  sufficient  numbers 
(141  and  50,  respectively)  to  permit 
comparisons  of  abundance  among 
stand  types  or  to  conduct  analyses  of 
cover  object  use. 


Habitat  Occupancy 

Ensatinas  and  redback  salamanders 
occurred  in  all  forest  age  and  mois- 
ture classes.  Although  both  species 


33 


were  most  abundant  in  young  forests 
(table  1),  a  one-way  AN  OVA  re- 
vealed no  significant  differences 
among  stand  types.  Mean  captures 
for  both  species  were  lowest  in  wet 
old-growth  stands.  The  proportion  of 
stands  containing  ensatinas  was  also 
relatively  low  in  wet  old  growth: 
fewer  than  20%  of  wet  old -growth 
stands  sampled  contained  ensatinas, 
whereas  all  other  stand  types  had  a 
frequency  of  occurrence  of  65%  or 
greater  (fig.  2).  The  proportion  of 
stands  containing  redback  salaman- 
ders was  generally  low  in  all  stand 
types  (fig.  2),  suggesting  that  at  the 
time  of  our  sampling,  redback  sala- 
manders were  less  abundant  or  more 
clumped  in  distribution  than  ensati- 
nas. We  found  no  amphibians  in  67% 
of  old-growth  wet  stands,  11%  of 
young  and  mature  stands,  12%  of 
moderate  old-growth  stands,  and  0% 
of  dry  old-growth  stands. 


\  \  \  \ 


\  \  \  \  \  \ 
\  \  \ 


Use  of  Woody  Debris 

Cover  object  selection  varied  be- 
tween the  two  species.  Ensatinas 


O.Bt 


OGW 
(N=6) 

STAND  TYPE 

Figure  2.— Proportion  of  stands  in  each  stand  type  with  captures  of  Ensatina  eschscholtzii 
(ENES)  and  Plethodon  vehiculum  (PLVE)  in  the  southern  Washington  Cascades  Range.  Stand 
type  YNG= Young.  MAT=Mature,  OGW=Wet  Old  Growth,  OGM=Moderate  Old  Growth, 
OGD=Dry  Old  Growth. 


Figure  3.— Use  of  cover  objects  by  Ensatina  eschscholtzii  (ENES)  and  Plethodon  vehiculum 
(PLVE)  in  the  southern  Washington  Cascade  Range. 


were  most  often  found  under  pieces 
of  bark  (generally  within  1  m  of  a 
snag  or  log)  and  secondarily  under 
logs  (fig.  3).  The  pattern  was  re- 
versed for  redback  salamanders.  Nei- 
ther species  was  found  under  bark 
on  snags.  When  found  under  pieces 
of  bark,  ensatinas  most  often  oc- 
curred in  bark  piles  at  the  base  of 
moderately  decayed  snags  (see  Tho- 
mas et  al.  1979,  p.  64).  Seventy-four 
percent  of  these  captures  occurred 
next  to  snags  in  which  the  top  had 
broken  off,  the  wood  was  soft,  and 
most  or  all  of  the  bark  had  sloughed 
onto  the  ground.  Logs  where  ensati- 
nas and  redback  salamanders  were 
captured  were  most  often  10-  30  cm 
in  diameter  (fig.  4).  Both  species  were 
captured  in  low  numbers  in  associa- 
tion with  very  large  logs  (diameter 
>30  cm),  but  our  inability  to  ade- 
quately search  this  cover  type  may 
account  for  these  results.  Virtually  all 
logs  where  ensatinas  and  redback 


34 


salamanders  were  found  were  in  in- 
termediate stages  of  decay  (fig.  5) 
(see  Maser  et  al.  1979,  p.  80).  Only  a 
few  captures  of  either  species  oc- 
curred in  association  with  intact  or 
extensively  decomposed  logs.  Nei- 
ther species  was  commonly  found 
under  rocks,  but  this  cover  type  is 
relatively  rare  in  Douglas-fir  forests. 
No  correlations  between  slope  or  as- 
pect and  amphibian  capture  sites 
could  be  detected. 


Discussion 

Old-growth  forests  do  not  appear  to 
provide  unique  habitat  for  either  en- 
satinas  or  western  redback  salaman- 
ders; both  species  were  well-repre- 
sented in  all  age  classes.  Our  results 
suggest  that  abundance  levels  of 
these  salamanders  are  more  likely  a 
function  of  the  availability  of  woody 
debris  for  cover  than  age  of  the  over- 
story.  Wet  old-growth  stands  in 
southern  Washington,  however,  ap- 
parently provide  low  quality  habitat 
for  these  plethodontids,  especially 
ensatinas  (table  1,  fig.  2).  Soils  in 
these  stands  were  often  saturated 


INTACT 


MODERATELY  DECOMPOSED 
DECAY  CLASS 


DECOI^OSED 


Figure  5.— Use  of  logs  by  Ensatina  esc/)scho//z//(E NES)  and  Plethodon  vehiculum  (PLVE)  by 
decay  class  In  the  southern  Washington  Cascade  Range. 


with  water,  and  such  conditions  may 
reduce  the  availability  of  microenvi- 
ronments  suitable  for  cover,  mainte- 
nance of  water  balance,  and  success- 
ful reproduction.  In  addition,  these 


<  10  CM 
FINE  WOODY  DEBRIS 


10 -30  CM  >  30  CM 

COARSE  WOODY  DEBRIS 


Figure  4.— Use  of  logs  by  Ensatina  eschscholtzii  i£N£S)  and  Plethodon  vehiculum  (PLVE)  by 
dianneter  class  in  the  southern  Washington  Cascade  Range. 


Stands  were  located  in  topographi- 
cally low  sites  where  cold  air  accu- 
mulates, which  may  create  unfavor- 
able microclimatic  conditions  for  ple- 
thodonhd  salamanders.  Our  results 
also  suggest  that  plethodontid  sala- 
manders may  prefer  certain  types  of 
woody  debris  as  cover,  especially 
those  associated  with  large,  moder- 
ately to  well-decomposed  snags  and 
logs.  Captures  of  ensatinas  were 
most  common  under  pieces  of  bark, 
especially  in  bark  piles  at  the  base  of 
well-  decayed  snags  (fig.  3).  Snags  in 
the  early  stages  of  decomposition 
with  shallow  or  no  bark  piles  at  their 
bases  provide  few  suitable  mi- 
crohabitats  for  salamanders.  Depth 
of  these  bark  piles  increases  as 
sloughing  continues  until  all  bark  has 
fallen  off.  Later  stages  of  snag  de- 
composition provide  no  additional 
bark  to  the  pile  and  habitable  spaces 
become  compressed  as  the  lower  lay- 
ers of  bark  decay  and  mix  with  the 
underlying  substrate. 

Bark  microhabitats  formed  by  the 
deterioration  of  snags  differ  in  struc- 
ture from  those  formed  by  the  de- 


35 


composition  of  logs.  As  logs  decay,  a 
single  layer  of  bark  is  deposited  on 
the  forest  floor,  whereas  bark  slough- 
ing from  snags  forms  multilayered, 
structurally  complex  cover.  Such 
bark  piles  could  provide  microcli- 
matic conditions  more  resistant  to 
fluctuations  in  temperature  and 
moisture  than  those  found  under 
bark  on  the  ground.  Additional  for- 
aging habitat  may  also  be  available. 

Redback  salamanders,  on  the 
other  hand,  were  most  often  found 
under  moderately  decayed  logs  10-30 
cm  in  diameter  (figs.  3-5).  In  the  early 
stages  of  decay,  bark  has  not  begun 
to  slough  and  branches  suspend  the 
log  above  the  ground.  As  the  bark 
begins  to  slough  and  branches  dete- 
riorate, increased  cover  and  moisture 
are  provided  along  the  length  of  the 
bole  where  it  comes  in  contact  with 
the  forest  floor  (Maser  and  Trappe 
1984).  The  quality  of  this  environ- 
ment for  salamanders  continues  to 
improve  with  further  decay  until  the 
organic  matter  becomes  incorporated 
into  the  underlying  substrate  and 
habitable  interstices  become  com- 
pressed in  the  advanced  stages  of 
decomposition. 

All  known  nest  sites  of  ensatinas 
in  the  Pacific  Northwest  have  been 
found  in  association  with  large,  mod- 
erately decayed  logs  (Norman  and 
Norman  1980,  Maser  and  Trappe 
1984,  Jones  and  Aubry  1985,  Norman 
1986,  L.  L.  C.  Jones  unpubl.  data). 
This  habitat  feature  may  be  impor- 
tant for  the  persistence  of  ensatinas 
in  these  forests.  We  do  not  know  to 
what  extent  coarse  woody  debris 
may  be  important  for  reproduction 
of  redback  salamanders  in  Douglas- 
fir  forests;  only  one  nest  site  has  been 
found,  and  this  was  in  moist  talus  in 
the  Oregon  Coast  Range  (Hanlin  et 
al.  1978). 

In  Douglas-fir  stands  of  the  Cas- 
cade Range  that  have  regenerated 
after  catastrophic  fires,  levels  of 
coarse  woody  debris  (CWD)  (logs 
and  snags  >  10  cm  in  diameter)  are 
moderate  in  young  stands,  lowest  in 
mature  stands,  and  highest  in  old- 


growth  stands  (Spies  et  al.  in  press). 
In  general,  this  is  due  to  the  inheri- 
tance of  high  levels  of  CWD  in  young 
stands  from  the  preceding  old- 
growth  stands,  a  low  accumulation 
of  CWD  in  mature  stands  as  CWD 
decays  but  inputs  are  low,  and  high 
inputs  of  CWD  in  older  stands  as  the 
large  Douglas-firs  die  and  accumu- 
late as  snags  and  logs.  Intensive  for- 
est management  results  in  levels  of 
CWD  substantially  lower  than  that 
encountered  in  unmanaged  forests 
(Spies  and  Cline  in  press).  This  is  be- 
cause plantations  inherit  little  CWD 
from  the  preceding  stand  when  it  is 
clearcut  and  existing  CWD  is  re- 
moved and  fragmented.  In  addition, 
thinning  operations  reduce  the  input 
of  CWD  from  suppression  mortality 
and  short  rotations  prevent  the  accu- 
mulation of  CWD.  Maintaining  even 
moderate  amounts  of  CWD  in  man- 
aged forests  will  require  modifica- 
tions of  current  harvesting  and 
silvicultural  practices  (Harmon  et  al. 
1986,  Spies  et  al.  in  press). 

Virtually  all  available  cover  ob- 
jects we  encountered  were  woody 
debris,  and  both  species  were  found 
most  often  in  association  with  large, 
moderately  decayed  logs  and  snags. 
Our  results  suggest  that  the  availabil- 
ity of  coarse  woody  debris  may  be 
important  for  maintaining  salaman- 
der populations  in  Douglas-fir  for- 
ests. Additional  studies  of  terrestrial 
salamanders  in  managed  vs.  unman- 
aged forests  are  necessary  to  deter- 
mine the  extent  to  which  they  may  be 
affected  by  intensive  forest  manage- 
ment. 

In  general,  our  study  yielded  a 
relatively  low  number  of  captures. 
Only  two  common  species  (Nuss- 
baum  et  al.  1983)  were  captured  in 
high  enough  numbers  to  permit 
analyses  of  the  data;  captures  of  all 
other  species  were  incidental.  The 
total  number  of  species  detected  was 
also  low  in  relation  to  known  species 
richness:  pitfall  trapping  for  approxi- 
mately 1000  trap  nights  in  each  of  the 
same  study  sites  in  the  fall  of  1984 
yielded  916  captures  of  13  species  (K. 


B.  Aubry  unpubl.  data).  Research  us- 
ing time-constrained  searches  to 
study  all  but  the  most  common  spe- 
cies in  this  region  would  require  sub- 
stantially more  search  time.  Sam- 
pling should  also  be  conducted  dur- 
ing all  seasons  of  the  year  to  detect 
seasonal  shifts  in  habitat  selection  or 
cover  object  use,  and  to  sample  spe- 
cies that  are  active  at  other  times  of 
the  year. 

Acknowledgements 

We  thank  R.  W.  Lundquist,  J.  B. 
Buchanan,  B.  A.  Schrader,  A.  B. 
Humphrey,  M.  Q.  Affolter,  M.  J. 
Reed,  B.  F.  Aubry,  and  M.  J.  Crites 
for  assistance.  T.  A.  Spies  at  the  For- 
estry Sciences  Laboratory,  Corvallis, 
OR  provided  data  on  stand  charac- 
teristics. R.  W.  Lundquist  provided 
the  map  used  in  figure  1 .  This  study 
was  funded  under  USDA  Forest 
Service  Cooperative  Agreement 
PNW-83-  219  to  S.  D.  West  and  D.  A. 
Manuwal  at  the  Univ.  of  Washing- 
ton. We  thank  personnel  of  the  Gif- 
ford  Pinchot  National  Forest  and 
Mount  Rainier  National  Park  for 
their  cooperation  and  support.  M.  G. 
Raphael,  K.  E.  Severson,  T.  A.  Spies, 
and  A.  B.  Carey  provided  construc- 
tive comments  on  a  previous  draft  of 
the  manuscript.  This  paper  is  Contri- 
bution No.  65  of  the  Old-growth  For- 
est Wildlife  Habitat  Project,  USDA 
Forest  Service,  Pacific  Northwest  Re- 
search Station,  Olympia,  WA. 

Literature  Cited 

Bury,  R.  B.  and  P.  S.  Corn.  1988. 
Douglas-fir  forests  in  the  Oregon 
and  Washington  Cascades:  rela- 
tionship of  herpetofauna  to  stand 
age  and  moisture.  This  volume. 

Bury,  R.  B.  and  M.  G.  Raphael.  1983. 
Inventory  methods  for  amphibians 
and  reptiles,  p.  416-419.  In].  F.  Bell 
and  T.  Atterbury,  eds.  Renewable 
resource  inventories  for  monitor- 
ing changes  and  trends:  Proceed- 


36 


ings  of  an  international  conference 
(Corvallis,  Oregon,  August  15-19, 
1983)  Society  of  American  Forest- 
ers 83-14).  737  p. 
Campbell,  H.  W.  and  S.  P.  Christ- 
man.  1982.  Field  techniques  for 
herpetofaunal  community  analy- 
sis, p.  193-200.  In  N.  J.  Scott,  Jr., 
ed.  Herpetological  communities. 
USDI  Fish  and  Wildlife  Service 
Wildlife  Research  Report  13. 
239  p. 

Franklin,  J.  F.,  K.  Cromack,  Jr.,  W. 
Denison,  A.  McKee,  C.  Maser,  J. 
Sedell,  F.  Swanson,  and  G.  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-fir  forests. 
USDA  Forest  Service  General 
Technical  Report  PNW-118, 48  p. 

Franklin,  J.  F.  and  C.  T.  Dyrness. 
1973.  Natural  vegetation  of  Ore- 
gon and  Washington.  USDA  For- 
est Service  General  Technical  Re- 
port PNW-8, 417  p. 

Gutierrez,  R.  J  and  A.  B.  Carey,  eds. 
1985.  Ecology  and  management  of 
the  spotted  owl  in  the  Pacific 
Northwest.  USDA  Forest  Service 
General  Technical  Report  PNW- 
184. 119  p. 

Hanlin,  H.  G.,  J.  J.  Beatty,  and  S.  W. 
Hanlin.  1978.  A  nest  site  of  the 
western  red-backed  salamander 
Plethodon  vehiculum.  Journal  of 
Herpetology  13(2):212-214. 

Harmon,  M.  E.,  J.  F.  Franklin,  F.  J. 
Swanson,  P.  Sollins,  S.  V.  Gregory, 
J.  D.  Lattin,  N.  H.  Anderson,  S.  P. 
Cline,  N.  G.  Aumen,  J.  R.  Sedell, 
G.  W.  Lienkaemper,  K.  Cromack, 
Jr.,  and  K.  W.  Cummins.  1986. 
Ecology  of  coarse  woody  debris  in 
temperate  ecosystems.  In  A.  Mac- 
Fadyen  and  E.  D.  Ford,  eds.  Ad- 
vances in  Ecological  Research 
15:133-302. 

Jones,  L.  L.  C.  and  K.  B.  Aubry.  1985. 
Ensatina  eschscholtzii  (Oregon  en- 
satina).  Reproduction.  Herpeto- 
logical Review  16(1):26. 

Lumen,  I.  D.  and  W.  A.  Neitro.  1980. 
Preservation  of  mature  forest  serai 
stages  to  provide  wildlife  habitat 
diversity.  Transactions  of  the 
North  American  Wildlife  and 


Natural  Resources  Conference 
43:78-88. 

Maser,  C,  R.  G.  Anderson,  K. 

Cromack,  Jr.,  J.  T.  Williams,  and  R. 
E.  Martin.  1979.  Dead  and  down 
woody  material,  p.  78-95.  In  J.  W. 
Thomas,  ed.  Wildlife  habitats  in 
managed  forests-  the  Blue  Moun- 
tains of  Oregon  and  Washington. 
USDA  Forest  Service  Agricultural 
Handbook  553,  512  p. 

Maser,  C.  and  J.  M.  Trappe.  1984. 
The  seen  and  unseen  world  of  the 
fallen  tree.  USDA  Forest  Service 
General  Technical  Report  PNW- 
164, 56  p. 

Meehan,  W.  R.,  G.  R.  Merrell,  Jr.,  and 
T.  A.  Hanley,  eds.  1984.  Fish  and 
wildlife  relationships  in  old- 
growth  forests:  Proceedings  of  a 
symposium.  (American  Institute 
of  Fishery  Research  Biologists, 
Juneau,  Alaska,  April  1982).  425  p. 

Meslow,  E.  C,  C.  Maser,  and  J. 
Vemer.  1981.  Old-growth  forests 
as  wildlife  habitat.  Transactions  of 
the  North  American  Wildlife  and 
Natural  Resources  Conference 
46:329-335. 

Norman,  B.  R.  1986.  Ensatina  es- 
chscholtzii oregonensis  (Oregon  en- 
satina). Reproduction.  Herpeto- 
logical Review  17(4):89. 

Norman,  C.  E.  and  B.  R.  Norman. 
1980.  Notes  on  the  egg  clusters 
and  hatchlings  of  Ensatina  es- 
chscholtzi  oregonensis.  Bulletin  of 
the  Chicago  Herpetological  Soci- 
ety 15(4):99-100. 

Nussbaum,  R.  A.,  E.  D.  Brodie,  Jr., 
and  R.  M.  Storm.  1983.  Amphibi- 
ans and  reptiles  of  the  Pacific 
Northwest.  332  p.  University  Press 
of  Idaho,  Moscow. 

Raphael,  M.  G.  1984.  Wildlife  diver- 
sity and  abundance  in  relation  to 
stand  age  and  area  in  Douglas-fir 
forests  of  northwestern  California, 
p.  259-274.  In  W.  R.  Meehan,  T.  R. 
Merrell,  Jr.,  and  T.  A.  Hanley,  eds. 
Fish  and  wildlife  relationships  in 
old-growth  forests:  Proceedings  of 
a  symposium.  (American  Institute 
of  Fishery  Research  Biologists, 
Juneau,  Alaska,  April  1982).  419  p. 


Raphael,  M.  G.  and  R.  H.  Barrett. 
1984.  Diversity  and  abundance  of 
wildlife  in  late  successional 
Douglas-fir  forests,  p.  352-360.  In 
New  Forests  for  a  changing  world. 
(Society  of  American  Foresters 
Convention,  Portland,  Oregon, 
1983). 

Ruggiero,  L.  F.  and  A.  B.  Carey.  1984. 
A  programmatic  approach  to  the 
study  of  old-growth  forest  -  wild- 
life relationships,  p.  340-345.  In 
New  forests  for  a  changing  world. 
(Society  of  American  Foresters 
Convention,  Portland,  Oregon, 
1983). 

Spies,  T.  A.  and  S.  P.  Cline.  In  Press. 
Coarse  woody  debris  in  manipu- 
lated and  unmanipulated  coastal 
Oregon  forests.  In  C.  Maser,  R.  F. 
Tarrant,  J.  M.  Trappe,  and  J.  F. 
Franklin,  eds.  From  the  forest  to 
the  ocean — a  story  of  fallen  trees. 
USDA  Forest  Service  General 
Technical  Rep>ort. 

Spies,  T.  A.,  J.  F.  Franklin,  and  T.  B. 
Thomas.  In  Press.  Coarse  woody 
debris  in  Douglas-fir  forests  of 
western  Oregon  and  Washington. 
Ecology.  Thomas,  J.  W.,  R.  G.  An- 
derson, C.  Maser,  and  E.  L.  Bull. 
1979.  Snags.  In  J.  W.  Thomas,  ed. 
Wildlife  habitats  in  managed  for- 
ests -  the  Blue  Mountains  of  Ore- 
gon and  Washington.  USDA  For- 
est Service  Agricultural  Handbook 
553.  512  p. 

Welsh,  H.  H.,  Jr.  1988.  Old  growth 
forests  and  the  distribution  of  the 
terrestrial  herpetofauna.  This  vol- 
ume. 


37 


Forestry  Operations  and 
Terrestrial  Salamanders: 
Techniques  in  a  Study  of  the 
Cow  Knob  Salamander, 
Plethodon  punctatus^ 

Kurt  A.  Buhlmann,2  Christopher  A.  Pague,^ 
Joseph  C.  Mitchell,^  and  Robert  B.  Glasgow^ 


Abstract.— The  status  and  ecology  of  Plethodon 
punctaf us  \^/as  investigated  in  George  Washington 
National  Forest,  Virginia  to  determine  potential 
effects  of  logging.  Pitfall  traps  and  mark-recapture 
supplemented  searching  by  hand.  Elevation, 
aspect,  soil  characteristics,  and  number  of  cover 
objects  (rocks)  ore  the  most  important  features  that 
identify  P.  panc^ofus  habitat.  Intensive  logging 
operations  appear  to  be  detrimental  to  this  species. 


Increasing  emphasis  is  being  placed 
on  conservation  and  preservation  of 
biological  diversity  worldwide 
(Norse  et  al.,  1986;  Wilson,  1988). 
U.S.  federal  and  state  agencies  have 
become  concerned  about  the  bio- 
diversity of  their  managed  lands  and 
are  directing  efforts  towards  preserv- 
ing natural  biota.  From  a  manage- 
ment perspective,  research  on  am- 
phibians and  reptiles  lags  behind  that 
devoted  to  game  animals,  such  as 
some  mammals,  birds,  and  fish  (Bury 
et  al.,  1980).  This  is  partly  due  to  a 
previous  lack  of  interest  in  these 
groups,  but  also  because  some  spe- 
cies can  be  more  difficult  to  observe 
or  investigate. 

The  Cow  Knob  salamander,  Ple- 
thodon punctatus,  is  a  dark,  moder- 
ately large  (to  74  mm  snout-vent 
length),  woodland,  fossorial  amphib- 
ian (Martof  et  al.,  1982)  found  only 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ,  July  19-21,  1988). 

^ Kurt  A.  Buhlmann  is  a  consulting  biolo- 
gist for  tlie  U.S.  Forest  Service.  Buhlmann's 
current  address  is  2001  N.  Main  St., 
Blacksburg,  VA  24060. 

^Christophier  A.  Pague  is  a  doctoral  can- 
didate in  ttie  Ecological  Sciences  Program, 
Department  of  Biological  Sciences,  Old 
Dominion  University,  Norfolk,  VA  23508. 

^Josept)  C.  Mitchiell  is  a  Research)  Biolo- 
gist, Department  of  Biology,  University  of 
Richmond,  Richmond,  VA  23173. 

^Robert  B.  Glasgow  is  the  Wildlife  Biolo- 
gist for  the  George  Washington  National 
Forest,  Harrison  Plaza,  P.O.  Box  233,  Harri- 
sonburg, VA  22801. 


on  Shenandoah  and  North  Mountain 
of  western  Virginia  and  eastern  West 
Virginia  (Highton,  1972,  Tobey, 
1985).  Most  of  the  known  range  of 
this  recently  described  species 
(Highton,  1972)  is  in  the  George 
Washington  National  Forest.  Fraser 
(1976)  compared  some  aspects  of  the 
ecology  of  this  species  with  a  sympa- 
tric  congener,  Plethodon  hoffmani. 
Little  else  is  known  of  the  ecology  of 
this  salamander.  Because  of  its  rela- 
tively small  range  and  unknown 
status,  P.  punctatus  was  added  to  the 
U.S.  Fish  and  Wildhfe  Service's  Cate- 
gory 2  list  (U.S.  Fish  and  Wildlife 
Service,  1985).  Potential  timber  har- 
vesting within  the  range  of  this  spe- 
cies (USD A  Forest  Service,  1986) 
prompted  us  to  examine  its  status  in 
forest  stands  of  various  ages.  In  this 
paper  we  report  the  following  as- 
pects of  this  study:  techniques  of  cap- 
ture and  data  collection,  salamander 
habitat  characteristics,  and  potential 
effects  of  logging  operations.  Our  ob- 
jective in  this  paper  is  to  make  other 
researchers  aware  of  the  techniques 
we  used  and  the  problems  we  en- 
countered in  developing  useful  man- 
agement recommendations  for  the 
protection  of  an  apparently  rare  ter- 
restrial salamander. 


Materials  and  Methiods 

We  conducted  this  study  on  Shenan- 
doah Mountain,  Augusta  and  Rock- 
ingham Counties,  George  Washing- 


ton National  Forest,  Virginia.  Before 
its  purchase,  between  1911  and  1940, 
by  the  U.S.  government,  this  area 
was  repeatedly  logged  and  burned 
(Leichter,  1987;  original  land  deed 
documents).  Few  virgin  stands  of 
forest  remain,  and  regrowth  and  log- 
ging operations  has  resulted  in  a  mo- 
saic of  mixed  hardwoods  of  various 
ages. 

We  selected  five  sites  of  different 
aged  forest  to  determine  the  relative 
abundance  of  Plethodon  punctatus  (fig. 
1)  to  see  if  its  presence  was  affected 
by  logging.  All  sites  selected  had 
similar  aspects  (S-SE)  and  elevation 
(914- 1127  m)  (table  1).  We  used 
USDA  Forest  Service  compartment 
descriptions  and  maps  to  aid  in  se- 
lection of  sites  and  to  obtain  informa- 
tion on  the  history,  physical  and  bio- 
logical descriptions,  and  future  man- 
agement goals  for  each  site.  A  com- 
partment is  divided  into  a  series  of 
stands,  each  of  which  defines  a  for- 
ested area  of  similar  tree  species  by 
composition,  age,  and  stand  condi- 
tion. Stand  age  is  defined  by  the  age 
of  dominant  canopy  trees.  Final 
choices  of  sites  were  made  only  after 
each  was  checked  in  the  field  and 
tree  age  was  verified  by  tree  ring 
analysis. 

In  each  site  we  erected  drift  fence 
arrays  (Campbell  and  Christman, 
1982)  consisting  of  four  60  cm  x  7.5  m 
secHons  of  aluminum  flashing  ar- 
ranged in  a  cross  pattern.  Opposite 
arms  of  the  cross  were  separated  by 
15  m  and  all  sections  were  sunk  in 


38 


the  ground  approximately  10  cm.  A 
5-gal  plastic  bucket  was  placed  in  the 
center  of  each  arm  and  #10  cans  were 
placed  in  the  ground  on  either  side  of 
the  ends  of  each  arm  so  that  the  tops 
of  the  pitfalls  were  flush  with  the 
ground  surface.  Sites  A  and  C  con- 
tained two  drift  fence  arrays,  and  the 
remaining  three  sites  had  one  array 
each.  In  each  pitfall  we  put  4-10  cm 
of  10%  formal-dehyde  solution  to  in- 
sure adequate  preservation  of  the 
salamanders.  We  selected  this 
method  to  obtain  samples  of  all  the 
terrestrial  fauna  for  a  range  of  stud- 


ies on  reproductive  cycles  and  ecol- 
ogy. Pitfalls  were  checked  and  all 
captures  (including  other  vertebrates 
and  all  invertebrates)  were  collected 
weekly  May  5  -  June  18, 1987,  bi- 
weekly July  7  -  November  22,  1987 
and  monthly  December  1987  and 
January  1988.  Samples  were  sorted  in 
the  laboratory  and  the  vertebrates 
stored  in  10%  neutral  buffered  for- 
maldehyde. Invertebrates  were 
stored  in  70%  isopropanol. 

Hand-collecting  supplemented 
drift  fence  collection  and  was  used  to 
determine  the  elevational  range  of  P. 


Figure  1  .—Plethodon  punctatus,  the  Cow  Knob  salamander,  from  Shenandoah  Mountain, 
Augusta  County,  Virginia.  Photograph  by  Kurt  A.  Buhlmann. 


punctatus  and  to  obtain  information 
on  range  and  habitat  characteristics. 
Results  from  timed  collecting  periods 
allowed  comparison  among  sites  and 
dates  of  collection.  Between  April  20 
and  June  2,  1987,  we  collected  data 
on  eleven  microhabitat  variables  at 
67  sites  to  evaluate  those  most  im- 
portant in  predicting  the  presence  of 
this  salamander.  These  variables 
were  elevation,  aspect,  slope,  soil 
temperature  under  cover  object,  soil 
moisture,  soil  pH,  soil  description, 
canopy  cover,  number  of  cover  ob- 
jects available  within  a  2  m  circle, 
type  of  cover  object  (e.g.,  rock,  log), 
and  forest  type. 

One  site  >1  km  away  from  any  of 
the  collection  sites  was  selected  for 
estimation  of  population  size  and 
data  on  individual  movements.  We 
searched  for  salamanders  in  daylight 
by  turning  and  replacing  all  surface 
objects  and  at  night  while  they  were 
active  on  the  surface  (i.e.,  during 
conditions  of  near  100%  relative  hu- 
midity [sensu  Heatwole,  1960;  Jaeger 
1978]).  Each  individual  was  meas- 
ured (snout-vent  length,  tail  length  to 
nearest  mm),  weighed  (nearest  0.1  g), 
the  sex  determined,  assigned  to  adult 
or  juvenile  age-classes,  marked  by 
toe-clipping,  and  released  at  its  cap- 
ture site.  We  marked  each  capture 
site  with  survey  flags  on  which  the 
salamander's  number  and  capture 


Table  1.— Descriptions  of  drift  fence  study  sites  for  Plethodon  punctatus  on  ShenarKloah  Mountain,  Virginia.  Slope 
angle  is  in  degrees  and  site  age  is  in  years  since  last  logging  activity. 


Site 

Timber  descrip. 

Slope 

Manag.  type 

Site  age 

Stand  condition 

Past  logging  history 

A 

1  yr  old  white  pine 
several  hardwood 
seed  trees 

30 

white  pine 

2 

seedling/ 
sapling 

90%  clearcut 
few  hardwood  trees 

B 

white  oak/ 
red  oak/hickory 

.  45 

oak/hickory 

8 

sparse 
saw  timber 

thinned  due  to  ice 
damage,  1979 

C 

white  pine/mixed 

25 

white  pine 

30 

immature 

cut  in  1956,  planted  in 

hardwoods 

pole  timber 

white  pine,  some  hardwood 

seed  trees 

D 

white  oak/hickory 

30 

oak/hickory 

60-100 

mature  saw 
timber 

no  recent  management 

E 

white  oak/ 
red  oak/  hickory 

5 

none 

virgin? 

low  quality 
saw  timber 

none  known 

39 


date  were  written.  We  noted  all  re- 
captures and  measured  movements 
in  linear  fashion  (0.1  m)  between  cap- 
ture points. 

Results  and  Discussion 

Capture  Tectiniques 

Nineteen  P.  punctatus  were  caught  in 
the  pitfall  traps,  2.0%  of  the  total 
number  of  salamanders.  Of  the  17 
recorded,  12  were  caught  in  5-gal. 
buckets  and  5  in  #10  cans.  Eleven  P. 
punctatus  were  caught  in  Site  E,  six  in 
Site  B,  and  two  Site  D.  None  were 
caught  in  Site  A  or  the  Site  C.  In  con- 
trast, by  hand  collecting  in  areas  ad- 
jacent to  Site  E,  we  found  38  P.  punc- 
tatus in  7.7  man  hours  of  searching. 
The  drift  fence  method  appears  only 
moderately  effective  in  sampling  this 
salamander.  It  is  feasible  that  P.  punc- 
tatus is  less  likely  to  fall  into  the  pit- 
falls than  other  salamander  species. 
We  observed  several  individuals 
climbing  rocks  and  tree  trunks  dur- 
ing nocturnal  surface  activity.  This 
suggests  that  this  salamander  is  able 
to  detect  precipices  and  avoid  falling 
into  pitfalls.  Also,  this  species  may  be 
active  on  the  horizontal  surface  only 
for  limited  periods  of  time  and  under 
specific  environmental  conditions. 
Thus,  the  drift  fence  technique, 
which  depends  on  horizontal  activ- 
ity, may  not  be  an  effective  sampling 
method  for  this  salamander  (R.D. 
Semlitsch,  pers.  comm.). 

Data  from  pitfall  traps,  combined 
with  data  from  hand  collecting,  can 
provide  information  for  management 
decisions.  For  instance,  seasonal 
trends  in  surface  activity  were  simi- 
larly indicated  by  both  drift  fence  re- 
sults (fig.  2)  and  captures  based  on 
hand  collection  (fig.  3).  Comparison 
of  P.  punctatus  with  that  of  its  sympa- 
tric  congener  P.  cinereus  (fig.  3)  re- 
veals concordance  in  seasonal  activ- 
ity and  suggests  similar  responses  to 
surface  environmental  conditions. 
This  information  could  be  used  to 
determine  the  times  logging  opera- 


MAY    JUN     JUL    AUG    SEP    OCT    NOV    DEC  JAN 


Figure  2. —Seasonality  of  drift  fence  captures  of  Plethodon  cinereus  and  P.  punctatus  at  Site  E 
(Tomat>awl<  Mountain),  George  Washington  National  Forest.  Adults  and  juveniles  are  in- 
cluded, but  not  tiatctilings.  Sampling  period  is  5  May  1987  to  24  January  1988. 


20 


APR  22  MAY  5  MAY  24  JUL  6  JUL  21  AUG  31  SEP  28  OCT  12  NOV  22 


Figure  3.— Seasonality  of  captures  per  man  hiour  of  Plett)odon  cinereus  and  P.  punctatus  on 
Tomatiawk  Mountain,  George  Wastiington  National  Forest.  Black  bars  represent  P.  punctatus 
and  bars  witli  diagonal  lines  represent  P.  cinereus.  Sampling  dates  are  22  April  to  22  Novem- 
ber 1987. 

40 


tions  would  cause  the  least  impact  on 
salamanders  at  or  near  the  surface. 

The  benefits  of  the  drift  fence  tech- 
nique outweighed  the  low  numbers 
of  captures  of  P.  punctatus.  We 
probably  would  not  have  otherwise 
found  this  species  in  Site  D  because 
there  were  few  surface  rocks  to  turn 
over.  Although  the  individuals 
caught  may  have  been  transients,  this 
species  does  occasionally  occur  at 
this  site.  This  result  would  not  have 
been  obtained  by  hand-collecting 
alone. 

The  drift  fence  method  also  pro- 
vided estimates  of  the  relative  abun- 
dance of  the  salamander  fauna  and 
other  species  in  the  community.  The 
relative  numbers  of  these  species  and 
species  groups  can  generate  addi- 
tional information  on  the  structure  of 
the  community  in  which  the  focal 
species  lives.  Drift  fence  techniques 
have  been  used  for  a  variety  of  eco- 
logical studies  (e.g..  Gibbons,  1970; 
Gill,  1978;  Pechmann  and  Semlitsch, 
1986)  but  only  recently  to  answer 
questions  about  vertebrate  communi- 
ties in  relation  to  forest  management 
(e.g.,  Bennett  et  al.,  1980;  Gibbons 
and  Semlitsch,  1981;  Enge  and  Mar- 


Table  2.— Seasonal  differences  in 
surface  abundance  of  Plethodon 
punctatus  at  Flagpole  Knob  and 
Skidmore  Tract,  Shenandoah 
f^ountain,  George  Washington  For- 
est. These  sites  are  <1  km  apart. 
Flagpole  Knob  is  a  rocky,  grassy 
ridge  habitat  containing  young 
oak  (Quercus  sp.)  and  maple 
(Acer  sp.)  pole  timber,  and  Skid- 
more  is  a  virgin  hemlock  (Tsuga 
canadensis)/ye\\ow  birch  (Betula 
lutea)  forest.  Numbers  of  salaman- 
ders are  followed  by  number  of 
man  hours  In  parentheses.  All  data 
are  based  on  hand-collecting  re- 
sults. 


Date         Flagpole  Skidmore 

June  2         11  (0,5)  10  (1.7) 

June  8          0  (0.5)  2  (2.0) 

Sept.  28        0  (1.0)  3  (3.0) 

V   J 


ion,  1986;  Bury  and  Com,  1987).  Our 
results  indicate  this  technique  can  be 
effective  in  mountainous  terrain  and 
can  be  used  to  gain  information  on 
apparently  rare  terrestrial  salaman- 
ders. 

If  an  endangered  or  otherwise 
protected  species  is  the  focus  of 
study  and  cannot  be  collected,  then 
slight  modifications  of  the  drift 
fence-pitfall  design  must  be  made. 
Traps  would  need  to  be  checked  on  a 
daily  basis,  or  nearly  so,  in  order  to 
release  the  animals  unharmed  (Gib- 
bons and  Semlitsch,  1981).  Water  or 
wet  leaves  can  be  placed  in  the  pit- 
falls for  cover  and  moisture.  Poten- 
tial problems  include  killing  of  the 
salamanders  in  the  pitfalls  by  small 
mammals,  especially  shrews,  and 
desiccation.  The  loss  of  animals  by 
shrew  or  raccoon  predation  in  pitfall 
containers  affects  the  samples  and 
may  prevent  quantitative  compari- 
sons among  sites.  Data  obtained 
from  visitation  frequencies  of  every 
three  days  (Bury  and  Corn,  1987)  to 
once  a  week  (Enge  and  Marion,  1986) 
probably  underestimate  actual  cap- 
tures. 

The  detection  of  P.  punctatus  at  a 
particular  site  depends  on  the  time  of 
year,  substrate  type,  soil  depth,  soil 
moisture,  soil  temperature,  and 
weather  conditions  (see  Habitat  Re- 
quirements). A  simple  survey  of  sites 
by  hand  searching  and  rock  turning 
in  daylight  hours  without  attention 
to  weather  and  seasonality  will  un- 
derestimate actual  abundance  and 
fail  to  detect  presence  of  a  species. 
Table  2  contains  comparative  data 
for  two  sites  searched  the  same  day 
at  different  times  of  the  year  and 
demonstrates  a  strong  seasonal  ef- 
fect. In  order  to  construct  effective 
management  plans,  the  range  and 
abundance  of  a  terrestrial  salaman- 
der must  be  known.  Therefore,  re- 
searchers conducting  distributional 
surveys  must  take  seasonal  and  diet 
changes  in  surface  activity  into  con- 
sideration. 

Results  of  our  1987  mark-recap- 
ture efforts  are  preliminary;  only 


four  recaptures  were  made.  One  P. 
punctatus  captured  28  May  was  re- 
captured on  15  October.  It  had 
moved  17.4  m.  Three  salamanders 
were  recaptured  within  ten  days  of 
original  capture  and  had  moved  <  2 
m.  Knowledge  of  movement  capabili- 
ties by  P.  punctatus  is  an  important 
part  of  evaluating  the  consequences 
of  population  fragmentation  through 
logging  operations.  Are  salamanders 
able  to  move  out  of  a  logged  area  or 
repopulate  it  when  suitable  habitat 
conditions  return?  We  believe  mark- 
recapture  studies  can  provide  useful 
information  on  rare  terrestrial  sala- 
manders, but  realize  that  data  may 
need  to  be  collected  over  several 
years  and  under  standardized  condi- 
tions in  order  to  provide  direct  an- 
swers. 


Habitat  Characteristics 

Preliminary  evaluation  of  microhabi- 
tat  data  indicate  that  four  site  charac- 
teristics are  most  important  in  deter- 
mining the  presence  of  P.  punctatus. 
We  found  P.  punctatus  at  elevations 
between  732  m  and  1317  m  (fig.  4). 
Most  sites  (87%)  with  this  species  oc- 
curred above  960  m.  Plethodon  punc- 
tatus occurred  on  all  slopes  but  were 
more  common  on  north-facing  as- 
pects (87%  of  11  sites)  than  east  (38% 
of  13),  south  (36%  of  8),  or  west  as- 
pects (40%  of  7).  Most  of  the  captures 
(67%  of  21)  were  on  slopes  of  20-45°. 
Seven  sites  were  on  slopes  less  than 
20°  and  between  46°  and  60°.  Sites 
without  this  salamander  were  on  a 
similar  range  of  slopes  (<  20°,  28.6%; 
20-45°,  57.1%;  >  45°,  14.3%). 

Soil  temperatures  under  cover  ob- 
jects at  sites  with  P.  punctatus  (x  = 
12.3  C,  9.4-16.1,  n  =  36)  were  nearly 
identical  to  temperatures  at  sites 
without  this  species  (x  =  12.8  C,  9.4- 
15.8,  n  =  15).  Soil  pH  under  cover  ob- 
jects were  also  similar  (with  P.  punc- 
tatus: X  =  6.3,  5.4-6.8;  without  P.  punc- 
tatus: X  =  6.4,  5.8-6.8).  Average  soil 
moisture  at  sites  with  P.  punctatus 
was  37.1%  (12-  70%)  and  42.8%  (24- 


41 


80%)  at  sites  without  this  species. 
Soils  in  which  P.  punctatus  were 
found  are  characterized  by  shallow 
black  humus  intermixed  with  rocks 
(72%  of  39  sites).  One  site  where 
eleven  salamanders  were  captured 
consisted  of  brown  humus  and  ex- 
tensive log  cover,  but  few  rocks. 
Cover  objects  under  which  this  sala- 
mander was  found  were  rocks  <  645 
cm2  (13.6%),  rocks  645-1290  cm^ 
(40.0%),  rocks  >  1290  cm^  (34.8%), 
and  logs  (10.6%).  Over  89%  of  the 
captures  were  found  under  rock 
cover.  Number  of  cover  objects 
within  a  2  m  circle  of  the  captured 
salamander  averaged  15.1  (1-45). 
Sites  without  P.  punctatus  ranged 
from  100%  rock  cover  to  0%  rock 
cover.  Sites  with  canopy  cover  equal 
to  or  greater  than  50%  accounted  for 
88.2%  of  the  captures  (n  =  52). 

We  found  P.  punctatus  in  the  fol- 
lowing forest  types:  mature  oak/ 
hickory  (38.5%  of  13),  oak/maple/ 
birch  (62.5%  of  8),  oak/pine  (33.3%  of 
3),  young  oak/ maple/  hemlock  (50% 
of  8),  virgin  hemlock /yellow  birch 
(100%  of  2),  hemlock/maple/bass- 
wood  (62.5%  of  8),  white  pine  (0%  of 
2),  and  grassy  balds  (20%  of  5).  Of 
the  site  characteristics  we  examined, 
the  following  appear  to  be  most  im- 
portant in  identifying  P.  punctatus 
habitat:  elevation,  aspect,  soil  charac- 
teristics, and  number  of  cover  objects 
(rocks). 

Habitats  of  terrestrial  salamanders 
differ  among  species  and,  in  some 
cases,  among  geographic  areas 
within  species  (e.g.,  Semlitsch,  1980; 
Tilley,  1973).  Data  derived  from  the 
literature  for  management  studies 
and  plans  must  be  used  with  caution. 
Baseline  habitat  and  life  history  stud- 
ies should  be  conducted  on  the  focal 
species  at  the  location  in  question  be- 
fore developing  management  plans. 

Effects  of  Logging 

Tree  removal  effects  the  terrestrial 
salamander  community  in  several 
ways.  Removal  of  canopy  cover 


eliminates  the  moisture-retaining  po- 
tenUal  of  the  soil  and  leaf  litter,  al- 
lows an  increase  in  insolation  (with  a 
concomitant  increase  in  soil  tempera- 
tures), and  increases  soil  erosion 
(Bury,  1983). 

The  use  of  heavy  machinery  com- 
pacts soil  and  destroys  leaf  litter. 
Enge  and  Marion  (1986)  found  that 
machine  site  preparation  and 
clearcutting  had  little  effect  on  am- 
phibian species  richness  in  a  Florida 
slash  pine  forest.  However,  of  the  15 
amphibian  species  they  recorded, 
none  was  a  terrestrial  salamander. 
On  Shenandoah  Mountain,  where 
most  of  the  terrestrial  amphibian 
community  is  comprised  of  terres- 
trial salamanders,  logging  and 
clearcutting  are  likely  to  have  detri- 
mental effects.  Salamander  abun- 
dance in  a  60-100  yr-old  deciduous 
forest  in  another  Virginia  site  was 
more  than  four  times  that  in  2  yr-old 
and  6-7  yr-old  clearcuts  (Blymer  and 
McGinnes,  1977).  Bury  (1983)  found 
that  terrestrial  salamanders  were 
more  abundant  in  old  growth  com- 
pared to  logged  redwood  forest 


habitats.  Plethodon  cinereus  was  sig- 
nificantly less  abundant  in  a  clearcut 
site  compared  to  an  old-growth  site 
in  a  deciduous  forest  in  New  York 
(Pough  et  al.,  1987). 

Populations  of  Plethodon  punctatus 
inhabiting  rocky  substrates  with  a 
thin  soil  cover  may  be  able  to  with- 
stand some  logging  operations.  Our 
Site  B  was  logged  in  a  salvage  opera- 
tion after  ice  storm  damage.  Not  all 
trees  were  removed  and  the  sub- 
strate was  not  as  damaged  as  that  in 
Site  A,  which  was  clearcut.  These  fac- 
tors, combined  with  the  presence  of  a 
seep  near  the  drift  fence  array, 
probably  explain  the  high  numbers  of 
P.  punctatus  found  at  Site  B  com- 
pared to  other  logged  sites. 

We  found  no  P.  punctatus  on  Sites 
A  and  C  for  apparently  different  rea- 
sons. Site  A  was  clearcut,  the  sub- 
strate was  greatly  disturbed,  and  the 
lack  of  canopy  cover  prevented  mois- 
ture retention.  The  fact  that  P.  punc- 
tatus occurred  on  the  same  ridge  in  a 
nearby  hardwood  stand  suggests  this 
salamander  may  have  occurred  on 
Site  A  prior  to  logging.  Site  C  was 


4(37-610  611-762 


763-914    915-1067   1068-1219     1220  +■ 

Elevation  (m) 


Figure  4.— Elevational  distribution  of  Pleftiodon  punctatus  on  Stienandoati  Mountain,  George 
Washiington  National  Forest.  Solid  bars  represent  sites  wtiere  P.  punctatus  was  not  found  and 
bars  witti  diagonal  lines  represent  sites  whiere  thiis  species  was  found. 


42 


logged  30  years  ago  but  was  re- 
planted with  white  pine  (table  1).  The 
logging  operation  and  change  in 
vegetation  type  may  have  affected 
the  salamander  populations  previ- 
ously present.  However,  because  of 
the  lack  of  rocky  substrate,  we  can- 
not disprove  the  hypothesis  that  P. 
punctatus  may  not  have  occurred 
there  historically. 

Plethodon  punctatus  appears  to  oc- 
cur in  greatest  abundance  on  rocky 
sites  that  contain  virgin  hardwoods 
(Site  E)  and  sites  that  are  not  heavily 
disturbed  by  logging  operations  (Site 
B).  Clearcutting  and  associated  dis- 
turbance does  appear  to  eliminate 
populations  of  this  salamander.  Sala- 
mander mortality  can  be  minimized 
if  logging  operations  are  conducted 
outside  the  seasonal  activity  period. 
If  size  of  the  area  logged  is  small,  or 
if  the  area  is  logged  in  a  mosaic,  or  if 
corridors  are  allowed  to  remain,  rein- 
vasion  may  eventually  be  possible 
from  peripheral  populations  when 
suitable  conditions  return.  Fragmen- 
tation of  the  limited  range  of  P.  punc- 
tatus by  a  patchwork  of  clearcuts 
could  seriously  affect  its  long-term 
survival. 


Conclusions 

Because  of  budget  and  time  con- 
straints, our  study  attempted  to  ob- 
tain baseline  data  and  evaluate  the 
effects  of  logging  simultaneously.  We 
offer  the  following  conclusions  to  re- 
searchers and  managers  who  must 
study  a  salamander  whose  ecology  is 
little  known. 

1 .  Multiple  capture  techniques 
should  be  used  when  study- 
ing an  apparently  rare  terres- 
trial salamander. 

2.  The  life  history  and  basic 
ecology  of  the  study  species 
needs  to  be  understood  be- 
fore the  project^ s  experimen- 
tal design  can  be  erected  to 
evaluate  logging  effects. 


3.  Seasonal  and  daily  activity 
patterns  of  salamander  activ- 
ity must  be  taken  into  con- 
sideration when  surveys  are 
conducted  to  determine 
range  and  population  abun- 
dance. 

4.  Project  proposals  to  federal 
and  state  agencies  should 
contain  a  two  step  process,  a 
field  survey  phase  to  obtain 
baseline  data  on  ecology  and 
life  history  and  an  experi- 
mental phase  in  which  log- 
ging or  other  concerns  are 
evaluated.  The  design  of  the 
experimental  phase  should 
be  based  on  the  results  of  the 
field  survey. 

Acknowledgments 

We  are  grateful  to  the  following 
people  for  field  assistance:  Christian 
A.  Buhlmann,  Kara  S.  Buhlmann, 
Lana  C.  Buhlmann,  Susan  J.  Fortuna, 
Joshua  C.  Mitchell,  and  Scott  M. 
Smith.  David  A.  Young  deserves  spe- 
cial thanks  for  helping  install  the 
drift  fence  arrays.  The  rangers  and 
foresters  at  the  Dry  River  District 
provided  logistical  support,  equip- 
ment, and  help  with  site  selection. 
This  study  was  supported  by  a  Chal- 
lenge Cost  Share  from  the  U.S.  Forest 
Service.  Additional  support  was  pro- 
vided by  the  Nongame  Wildlife  and 
Endangered  Species  Program  of  the 
Virginia  Department  of  Game  and 
Inland  Fisheries. 


Literature  Cited 

Bennett,  Stephen  H.,  J.  Whitfield  Gib- 
bons, and  Jill  Glanville.  1980.  Ter- 
restrial activity,  abundance  and 
diversity  of  amphibians  in  differ- 
ently managed  forest  types. 
American  Midland  Naturalist 
103:412-416. 

Blymer,  Michael  J.,  and  Burd  S. 
McGinnes.  1977.  Observations  on 


possible  detrimental  effects  of 
clearcutting  on  terrestrial  amphibi- 
ans. Bulletin  of  the  Maryland  Her- 
petological  Society  13:79-83. 

Bury,  R.  Bruce.  1983.  Differences  in 
amphibian  populahons  in  logged 
and  old  growth  redwood  forest. 
Northwest  Science  57:167-178. 

Bury,  R.  Bruce,  Howard  W. 

Campbell,  and  Norman  J.  Scott,  Jr. 
1980.  Role  and  importance  of 
nongame  wildlife.  Transactions  of 
the  45th  North  American  Wildlife 
and  Natural  Resources  Confer- 
ence, 45:197-  207. 

Bury,  R.  Bruce,  and  Paul  Stephen 
Corn.  1987.  Evaluation  of  pitfall 
trapping  in  northwestern  forests: 
trap  arrays  with  drift  fences.  Jour- 
nal of  Wildlife  Management 
51:112-119. 

Campbell,  Howard  W.,  and  Stephen 
P.  Christman.  1982.  Field  tech- 
niques for  herpetofaunal  commu- 
nity analysis,  p.193-200.  In  Herpe- 
tological  Communities.  N.  J.  Scott, 
Jr.,  editor.  U.S.  Fish  and  Wildlife 
Service,  Wildlife  Research  Report 
13. 

Enge,  Kevin  M.,  and  Wayne  R.  Mar- 
ion. 1986.  Effects  of  clearcutting 
and  site  preparation  on  herpe- 
tofauna  of  a  north  Rorida  flat- 
woods.  Forest  Ecology  and  Man- 
agement 14:177- 192. 

Eraser,  Douglas  F.  1976.  Coexistence 
of  salamanders  in  the  genus  Ple- 
thodon: A  variation  on  the  Santa 
Rosalia  theme.  Ecology  57:238-251. 

Gibbons,  J.  Whitfield.  1970.  Terres- 
trial activity  and  the  population 
dynamics  of  aquatic  turtles. 
American  Midland  Naturalist 
83:404-414. 

Gibbons,  J.  Whitfield,  and  Raymond 
D.  Semlitsch.  1981.  Terrestrial  drift 
fences  with  pitfall  traps:  an  effec- 
tive technique  for  quantitative 
sampling  of  animal  populations. 
Brimleyana  7:1-16. 

Gill,  Douglas  E.  1978.  The  metapopu- 
lation  ecology  of  the  red-spotted 
newt,  Notophthalmus  viridescens 
(Rafinesque).  Ecological  Mono- 
graphs 48:145-166. 


43 


Heatwole,  Harold.  1960.  Burrowing 
ability  and  behavioral  responses  to 
desiccation  of  the  salamander,  Ple- 
thodon  cinereus.  Ecology  41:661- 
668. 

Highton,  Richard.  1972.  Distribu- 
tional interactions  among  eastern 
North  American  salamanders  of 
the  genus  Plethodon.  p.  139- 188.  In 
The  Distributional  History  of  the 
Biota  of  the  Southern  Appalachian 
Mountains.  Part  III:  Vertebrates. 
Virginia  Polytechnic  Institute  and 
State  University,  Research  Divi- 
sion Monographs  4. 

Jaeger,  Robert  G.  1978.  Plant  climb- 
ing ability  by  salamanders:  peri- 
odic availability  of  plant-  dwelling 
prey.  Copeia  1978:686-  691. 

Leichter,  Bill.  1987.  The  land  nobody 
wanted.  Highlands  of  the  Vir- 
ginia's Magazine,  Autumn  1987/ 
Winter  1988,  pp.  3-7. 

Martof,  Bernard  S.,  William  M. 
Palmer,  Joseph  R.  Bailey,  and  Jul- 
ian R.  Harrison,  III.  1980.  Am- 
phibians and  Reptiles  of  the  Caro- 
linas  and  Virginia.  University  of 
North  Carolina  Press,  Chapel  Hill, 
264  p. 

Norse,  Elliott  A.,  Kenneth  L.  Rosen- 
baum,  David  S.  Wilcove,  Bruce  A. 
Wilcox,  William  H.  Romme, 
David  W.  Johnston,  and  Martha  L. 
Stout.  1986.  Conserving  Biological 
Diversity  in  our  National  Forests. 
The  Wilderness  Society,  Washing- 
ton, D.C.,  116  pp. 

Pechmann,  Joseph  H.  K.,  and  Ray- 
mond D.  Semlitsch.  1986.  Diel  ac- 
tivity patterns  in  the  breeding  mi- 
grations of  winter-breeding  anu- 
rans.  Canadian  Journal  of  Zoology 
64:1116-1120. 

Pough,  P.  Harvey,  Ellen  M.  Smith, 
Donald  H.  Rhodes,  and  Andres 
Collazo.  1987.  The  abundance  of 
salamanders  in  forest  stands  with 
different  histories  of  disturbance. 
Forest  Ecology  and  Management 
20:1-9. 


Semlitsch,  Raymond  D.  1980.  Geo- 
graphic and  local  variation  in 
population  parameters  of  the 
slimy  salamander  Plethodon  gluti- 
nosus.  Herpetologica  36:6-16. 

Tilley,  Stephen  G.  1973.  Life  histories 
and  natural  selection  in  popula- 
tions of  the  salamander 
Desmognathus  ochrophaeus.  Ecology 
54:3-17. 

Tobey,  Franklin  J.  1985.  Virginia's 
amphibians  and  reptiles,  a  dis- 
tributional survey.  Privately  pub- 
lished, Virginia  Herpetological  So- 
ciety, 114  p. 

USDA  Forest  Service.  1986.  Final 
land  and  resource  management 
plan  for  the  George  Washington 
National  Forest.  U.S.  Department 
of  Agriculture. 

U.S.  Fish  and  Wildlife  Service.  1985. 
Endangered  and  threatened  wild- 
life and  plants:  review  of  verte- 
brate wildlife.  Federal  Register 
50:37958-37967. 

Wilson,  Edward  0. 1988.  Biodiver- 
sity. National  Academy  Press. 
Washington,  D.C. 


44 


Conserving  Genetically 
Distinctive  Populations:  The 
Case  of  the  Huachuca  Tiger 
Salamander  (Ambystoma 
tigrinum  stebbinsi  Lowey 

James  P.  Collins,^  Thomas  R.  Jones,^  and 
Howard  J.  Berna^ 


Abstract.— Huachuca  tiger  salamanders  are  a 
genetically  distinctive  race  of  Ambystoma  tigrinum 
found  only  in  1 7  localities  in  the  San  Rafael  Valley 
(SRV)  in  southeastern  AZ.  Populations  of  SRV 
salamanders  are  threatened  by  introduction  of 
exotic  fishes  and  disease.  Salamanders  were  largely 
eliminated  from  four  habitats  after  introduction  of 
sunfish  and/or  catfish.  An  unknown  fatal  disease 
killed  all  aquatic  morphs  in  two  other  habitats.  An 
additional  threat  includes  possible  hybridization  and 
introgression  of  SRV  populations  resulting  from 
introduction  of  exotic  salamanders.  Introduced 
bullfrogs  may  also  prey  on  salamanders,  or  act  as 
vectors  for  disease. 


Technological  advances  in  genetics 
now  enable  characterization  of  vari- 
ation within  a  species  at  increasingly 
finer  levels  of  description.  These  de- 
velopments are  allowing  us  to  begin 
the  difficult  task  of  identifying  which 
gene  pools  should  be  protected  to 
preserve  genetic  attributes  significant 
for  conserving  present  and  future 
generations  of  a  species  (Echelle 
1988,  Meffe  and  Vrijenhoek  1988, 
Ryman  and  Utter  1987).  Rather  than 
considering  simply  which  species  to 
conserve,  we  can  now  ask  whether  a 
conservation  effort  should  be  di- 
rected at  the  species,  subspecies,  or 
population  levels  (Allendorf  and 
Leary  1988,  Behnke  1972,  Ryder 
1986). 

Tiger  salamanders,  Ambystoma  ti- 
grinum Green,  range  throughout 
much  of  North  America  from  south- 
ern Canada  to  the  central  Mexican 
Plateau,  and  from  the  east  coast  of 
the  United  States  to  California 
(Gehlbach  1967).  This  complex  spe- 
cies is  divided  into  eight  subspecies 
(Collins  et  al.  1980,  Gehlbach  1967, 

' Paper  presented  of  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff, AZ,  July  19-21,  1988). 

'James  P.  Collins  is  Associate  Professor  of 
Zoology.  Department  of  Zoology,  Arizona 
State  University,  Tempe,  AZ  85287-1501 . 

^Ihomas  R.  Jones  and  Howard  J.  Berna 
are  Graduate  Students,  Department  of  Zo- 
ology, Arizona  State  University,  Tempe,  AZ 
85287-1501. 


Jones  et  al.  1988),  several  of  which 
(tigrinum,  mavortium,  nebulosum,  and 
melanostictum)  are  widespread  geo- 
graphically, locally  abundant,  and 
apparently  not  in  need  of  protection 
at  this  time.  More  information  is 
needed  on  the  Mexican  subspecies, 
velasci,  and  a  north-central  USA  race, 
diaboli,  before  conservation  needs  can 
be  confidently  assessed.  Two  races 
need  consideration  now. 

A.  t.  californiense  occurs  only  in  the 
Central  Valley  and  adjacent  oak 
woodlands  of  California,  placing  it 
among  the  more  geographically  re- 
stricted tiger  salamanders.  Further, 
A.  t.  californiense  appears  to  have 
been  isolated  from  the  other  races  of 
A.  tigrinum  for  several  million  years, 
and  has  a  level  of  genetic  divergence 
equalling  species-level  differences 
among  many  ambystomatid  taxa 
(Jones  1988).  Two  factors  suggest  this 
taxon  warrants  special  conservation 
efforts.  First,  California  populations 
are  as  distinct  genetically  from  other 
races  of  tiger  salamanders,  as  other 
species  of  Ambystoma  are  from  each 
other.  Second,  the  geographic  isola- 
tion and  apparent  spatial  subdivision 
of  A.  t.  californiense  populations 
(Gehlbach  1967)  likely  increases  their 
probability  of  extinction  (Soule  1987). 

A.  t.  stebbinsi  has  properties  like  A. 
t.  californiense,  suggesting  it  too  needs 
special  conservation  efforts  despite 
being  classified  as  only  part  of  a  very 
wide-ranging  species.  Populations  of 
A.  t.  stebbinsi  occur  only  in  the  San 


Rafael  Valley  (SRV)  in  the  border- 
lands between  Arizona  and  Sonora, 
Mexico.  In  addition  to  being  geo- 
graphically restricted,  the  race  is  also 
genetically  distinctive.  Average 
heterozygosity  among  SRV  popula- 
tions is  the  lowest  in  Ambystoma 
(Jones  et  al.  1988).  Electrophoretic 
analysis,  as  well  as  variation  in  exter- 
nal morphology,  indicates  A.  t.  steb- 
binsi is  phylogenetically  most  closely 
related  to  A.  t.  mavortium.  In  contrast, 
analysis  of  the  mitochondrial  DNA 
(mtDNA)  in  these  populations  indi- 
cates there  is  a  single  mitochondrial 
clone  in  the  San  Rafael  Valley.  This 
clone  is  derived  from  A.  t.  nebulosum, 
not  A.  t.  mavortium,  suggesting  A.  t. 
stebbinsi  actually  arose  from  hybridi- 
zation between  A.  t.  nebulosum  and  A. 
t.  mavortium  (Collins  1988). 

A  recent  paper  describes  patterns 
of  variation  in  external  morphology, 
allozymes,  and  geographic  isolation 
that  suggest  A.  t.  stebbinsi  is  a  distinc- 
tive race  within  the  A.  tigrinum  com- 
plex (Jones  et  al.  1988).  In  a  future 
paper  we  will  describe  mitochondrial 
DNA  variation  in  these  populations 
(Collins  et  al.,  in  prep.).  Our  present 
goal  is  to  summarize  several  aspects 
of  the  population  biology  of  A.  t.  steb- 
binsi. In  addition  to  being  restricted 
geographically,  our  research  indi- 
cates salamander  populations  in  SRV 
are  threatened  by  several  factors  in- 
cluding disease,  and  factors  sur- 
rounding the  introduction  of  exotic 
fishes  and  salamanders. 


45 


Materials  and  Methods 

SRV  is  a  Plains  grassland-Madrean 
evergreen  woodland  habitat  extend- 
ing from  southeastern  AZ  into  north- 
eastern Sonora  (Brown  1982).  A  sur- 
vey of  aquatic  habitats  in  southern 
AZ  and  northern  Sonora  and  Chi- 
huahua indicated  salamanders  refer- 
able to  A.  i.  stebbinsi  occurred  only  in 
SRV  (Jones  et  al.  1988). 

From  June  1979  to  February  1988, 
we  sampled  seven  natural  and  23 
man-made  or  man-altered  aquatic 
habitats  in  SRV  and  adjacent  slopes 
of  the  Patagonia  and  Huachuca 
mountains.  Altered  habitats  were 
primarily  livestock  watering  tanks 
constructed  where  natural  water  for- 
merly existed.  Bog  Hole  tank  is  a 
large,  impounded  cienega  (sensu 
Hendrickson  and  Minckley  1985), 
and  another  may  be  an  impounded 
spring.  Salamanders  occurred  in  only 
17  of  the  30  habitats  sampled  in  SRV 
(appendix  1,  fig.  1).  We  report  life 
history  variation  in  A.  t.  stebbinsi,  and 
the  influence  of  disease  and  intro- 
duced exotic  animals  on  this  taxon. 
For  describing  life  history  variation 
we  emphasize  four  tanks  (Parker 
Canyon  #1,  Huachuca,  Upper  13,  and 
Bodie  Canyon)  sampled  routinely. 
We  also  present  additional  informa- 
tion from  irregular  collections  at  all 
other  SRV  tanks  with  salamanders. 

We  usually  collected  specimens 
using  seines  and  dipnets,  but  occa- 
sionally used  gill  nets.  Depending  on 
our  plans  for  using  a  particular  col- 
lection, we  either  marked  and  re- 
leased salamanders,  returned  them 
to  the  laboratory  alive,  or  preserved 
them  in  the  field  for  later  analysis. 
All  preserved  specimens  are  in  the 
Lower  Vertebrate  Collections  at  Ari- 
zona State  University. 

To  summarize  life  history  vari- 
ation in  A.  t.  stebbinsi,  we  classified 
salamanders  by  life  history  stage  and 
morphology  using  internal  and  exter- 
nal characters  (table  1).  Stages  1  or  2 
were  immature,  and  3-5  were  ma- 
ture. Metamorphosed  salamanders 
lack  gills  and  a  caudal  fin,  while  lar- 


vae and  mature  branchiate  salaman- 
ders have  those  structures.  All  meas- 
urements are  in  mm;  snout-vent 
length  (SVL)  is  the  distance  from 
snout  to  posterior  margin  of  the  vent. 

Results 

Life  History  Variation 

Ambystoma  tigrinum  has  the  most 
complicated  pattern  of  morphologi- 
cal and  life  history  variation  known 
in  salamanders.  After  an  egg  hatches 
a  larva  begins  growing  in  an  aquatic 
habitat.  At  about  30  mm  SVL,  larvae 
of  A.  t.  nebulosum,  A.  t.  mavortium,  or 
A.  t.  tigrinum  can  continue  develop- 
ment as  a  typical  larva,  or  develop  as 
a  cannibalistic  larval  morph.  This 
dimorphism  is  unknown  in  the  other 
subspecies  (Collins  et  al.  1980).  At 
about  70  mm  SVL,  larvae  of  all  sub- 


species except  A.  t.  californiense  con- 
tinue developing  in  one  of  two  ways. 
They  may  metamorphose,  often  leave 
the  aquatic  habitat,  and  must  eventu- 
ally return  to  freshwater  to  breed. 
Alternatively,  a  larva  continues 
growing  beyond  70  mm  SVL,  ma- 
tures, and  breeds  as  a  larval-like 
form,  or  paedomorph  (Gould  1977). 
Thus,  depending  on  the  subspecies,  a 
single  population  might  have  two 
juvenile  morphs,  typical  or  cannibal, 
and  four  adult  morphs,  typical  and 
cannibal,  mature,  branchiate  morphs 
or  metamorphosed  morphs  of  either 
type.  Relative  frequency  of  each 
morph  varies  among  populations  in  a 
subspecies  (Collins  1981,  Rose  and 
Armentrout  1976). 

In  SRV,  most  populations  have 
mature,  typical,  branchiate  morphs 
as  well  as  mature,  typical,  metamor- 
phosed morphs.  Judy  Tank  is  one 
population  in  which  we  have  col- 


Figure  1  .—Map  of  the  San  Rafael  Valley,  Arizona.  Symbols  indicate  sampling  sites  (see  ap- 
pendix 1).  Electrophoretic  samples  were  from  sites  1-8;  mtDNA  samples  were  from  sites 
1,2,5,6,9;  F=sites  with  fish;  D=sites  with  diseased  salamanders;  arrow=J.F.  Jones  Ranch,  type 
locality  for  stebbinsi. 


46 


lected  no  mature,  branchiate  morphs 
thus  far.  Mature,  typical,  branchiate 
morphs  dominated  the  SRV  popula- 
tions. From  July  1979  to  August  1985, 
we  collected  more  than  1200  mature, 
branchiate  salamanders  and  only  64 
mature,  metamorphosed  animals. 
We  conservatively  estimated  popula- 
tion sizes  of  mature,  branchiate 
morphs  as  varying  from  50  (Upper 
13  Reservoir,  1984)  to  several 
hundred  (Huachuca  Tank,  1983, 
1984).  No  population  in  the  SRV  had 
cannibalistic  morphs.  Absence  of  the 
cannibal  morph  is  a  distinctive  fea- 
ture of  these  populations,  since  the 
morph  can  be  common  in  A.  t.  nebu- 


losum  and  A.  t.  mavortium,  the  nearest 
relatives  of  A.  t.  stehbinsi. 

Salamanders  in  SRV  bred  as  early 
as  mid-February  and  as  late  as  early 
May.  Most  egg  laying  occurred  from 
mid-March  to  late  April.  Animals 
hatched  within  several  weeks  and 
grew  rapidly,  so  that  larvae  <40  mm 
SVL  were  often  abundant  by  late 
spring  (tables  2-5).  By  mid-July,  lar- 
vae were  usually  about  60  mm  SVL, 
and  those  that  metamorphosed  gen- 
erally did  so  from  late  July  to  early 
September.  A  relatively  small  per- 
centage of  larvae  metamorphose  an- 
nually—about 17%  to  40%  based  on 
estimates  from  Bodie  Canyon  Tank. 


r 


Table  1  .—Criteria  used  to  classify  salamanders  into  stages  of  breeding 
readiness.  Numerals  In  parentheses  refer  to  diameter  in  mm  (after  Collins 
1981). 


Oviduct,  ovary,  peritoneum,  and 
cloacal  characters 


Wolffian  duct,  testes,  peritoneum, 
and  cloacal  characters 


1 .  Gonadal  tissue  primarily  white  and  flacid;  Wolffian  ducts  or  oviducts 
narrow,  with  few  folds;  cloacal  margins  not  swollen;  peritoneum  largely 
unpigmented. 


2.  Oviducts  enlarged  (0.5-1),  white, 
weal<ly  convoluted;  ova  small  (<1), 
mostly  white-cream  colored;  dor- 
sal third  of  peritoneum  light  grey; 
cloacal  margins  not  swollen. 


3.  Oviducts  large  (3-4),  convo- 
luted, white;  ova  small  and  white, 
medium  (1-  1 .5)  and  cream  or 
cream-tan  or  black,  with  some 
perhaps  large  (1 .5-2)  and  bipolar 
cream  and  tan;  at  least  dorsal 
two-thirds  of  peritoneum  grey  to 
black;  cloacal  margins  swollen, 
bulbous  with  interior  margins  light 
grey  to  black  and  rugose. 

4.  Oviducts  large,  convoluted, 
white,  distended  in  coils;  ova  small 
and  white  or  large  and  bipolar 
cream  and  tan;  peritoneum  and 
cloaca  as  in  3. 

5.  All  characters  as  in  3  except 
most  ova  small  and  white  with  a 
few  darkly  pigmented. 


2.  Duct  enlarged  (0.5-1),  convo- 
luted, but  not  distended  in  coils; 
testes  small,  flacid;  peritoneum 
black;  cloacal  margins  swollen, 
with  grey  to  grey-black  borders, 
especially  posterior. 

3.  Duct  large(>l), 
convoluted, cream  colored  with 
localized  black  pigment;  testes 
turgid;  cloaca!  margins  swollen , 
grey  to  grey-black,  rugose  borders, 
especially  posterior;  peritoneum 
black,  especially  densely  pig- 
mented dorsally. 


4.  Duct  large,  convoluted,  cream 
colored  with  scattered  black  pig- 
ment spots,  distended  in  coils;  tes- 
tes turgid,  enlarged;  cloaca  and 
peritoneum  as  in  3. 


By  early  autumn,  first  year  ani- 
mals that  did  not  metamorphose  be- 
gan to  mature  (tables  2-5).  From  late 
autumn  through  winter  most  SRV 
branchiate  salamanders  were  >100 
mm  SVL  (tables  2-5),  and  ready  to 
breed  (figs.  2-3).  These  data  indicate 
branchiate  salamanders  in  SRV  breed 
for  the  first  time  when  one  year  old. 


Disease 

During  July  and  August  1985,  all 
branchiate  salamanders  in  Inez, 
Huachuca,  and  Parker  Canyon  #1 
Tanks  were  killed  by  an  undiagnosed 
disease  (fig  1).  Salamanders  in  the 
field  and  laboratory  showed  little  re- 
sistance to  the  disease  which  was 
100%  fatal  within  a  few  days  of  the 
appearance  of  symptoms.  Attempts 
to  culture  the  pathogen(s)  were  in- 
conclusive, but  many  symptoms  re- 
sembled those  characteristic  of  Aero- 
monas  infection  [red  legl  (Fowler 
1978),  including  lethargy,  loss  of  ap- 
petite, and  the  epidermis  can  become 
red  from  infusion  of  blood.  This  type 
of  epidemic  disease  in  the  aquatic 
environment  is  particularly  devastat- 
ing in  A.  t.  stebbinsi,  because  popula- 
tion structure  in  SRV  is  strongly 
skewed  toward  larvae  and  mature 
branchiate  animals.  In  addition  to 
death  of  larvae,  therefore,  most 
adults  may  have  been  killed  in  highly 
infected  populations. 

Parker  C!anyon  #1  and  Inez  were 
recolonized  by  metamorphosed  sala- 
manders that  presumably  escaped 
the  disease  while  in  terrestrial  sites. 
We  collected  two  metamorphosed 
adults  (male  and  female)  and  one 
larva  in  Inez  Tank  in  April  1986  and 
collected  eggs  in  April  1987.  We  also 
collected  eggs  in  Parker  Canyon  #1  in 
April  1987,  and  five  mature  branchi- 
ate morphs  (3  males,  2  females)  in 
January  1988.  Since  all  branchiate 
morphs  in  Parker  Canyon  were  killed 
in  1985,  and  none  was  collected  in 
1986,  these  five  animals  also  sup- 
ported our  conclusion  that  in  SRV 
branchiate  salamanders  can  reach 


47 


sexual  maturity  when  a  year  old.  We 
collected  no  salamanders  in 
Huachuca  Tank  as  late  as  spring  1988 
(see  below). 

Introduction  of  Exotic  Animals 

Fishes. — A  few  exceptional  species  of 
salamanders  can  coexist  with  fishes, 
but  most  cannot.  In  SRV  exotic 
fishes,  especially  centrarchids  and 
ictalurids,  invariably  eliminate  sala- 
manders. We  do  not  know  the  effect 
of  native  fish  on  A.  t.  stebbinsi,  but  no 
salamanders  occur  in  four  natural 
SRV  habitats  (Heron  Spring,  Sheehy 
Spring,  Sharp  Spring,  Santa  Cruz 
River  and  tributaries)  that  have  na- 
tive fishes  (Gila  topminnow,  Poecili- 
opsis  0.  occidentalis  and  Gila  chub, 
Gila  intermedia).  We  base  our  general 
conclusions  concerning  exotic  fishes 
and  salamanders  in  SRV  on  the  fol- 
lowing observations  (fig.l). 

J.F.  Jones  Ranch  Tank. — This  is 
the  type  locality  for  A.  t.  stebbinsi. 
Largemouth  bass  (Micropterus  salmoi- 
des)  and  bluegill  (Lepomis  macrochirus) 
were  introduced  in  the  1950s,  and 
salamanders  no  longer  occur  here 
(see  photograph  of  this  site  in  Lowe 
1964:106).  It  is  apparently  a  popular 
local  fishing  spot. 

FS  58  Tank.— We  first  collected 
mature,  branchiate  and  larval  sala- 
manders here  in  July  1979.  There 
were  only  yellow  bullheads 
(Ameiurus  natalis)  in  June  1980.  In 
August  1984,  we  collected  19  mature, 
branchiate  salamanders,  no  catfish, 
and  hundreds  of  sunfish  (Lepomis 
sp.). 

Huachuca  Tank. — First  sampled  in 
May  1982,  this  tank  was  a  reliable 
source  of  salamanders  and  natural- 
history  information  for  the  next  two 
years.  On  22  August  1984  we  found 
one  yellow  bullhead,  plus  many  lar- 
val and  mature  branchiate  salaman- 
ders. On  5  July  1985  we  netted  >100 
salamanders  in  each  of  several  seine 
hauls.  Routine  sampling  on  24  Au- 
gust 1985  yielded  several  thousand 
fingerling  catfish  and  no  salaman- 


 > 

Table  2.— Seasonal  variation  in  number  and  size  (SVL)  of  salamanders  in 
eacti  breeding  stage  collected  from  Parker  Canyon  Tank. 


Snout/Vent  Length  (mm) 


Date 

Stage 

0-19 

20-39 

40-59 

60-79 

80-89 

100-119 

120-140 

8  Jan 

4 

5 

28-29  Mar 

1 

13 

4 

7 

2 

22-28  Apr 

1 

4 

6 

4 

39 

8 

5 

4 

13-25  Jun 

1 

3 

36 

2 

1 

3 

6 

9 

1 

4 

1 

19 

10 

5 

4 

4 

14 

5-10  Jul 

1 

1 

18 

4 

7 

1 

5 

1 

22  Aug 

1 

7 

2 

2 

3 

7 

3 

4 

9 

1 

2  Dec 

4 

10 

18 

Table  3.— Seasonal  variation  in  number  and  size  (SVl^  of  salamanders  In 
each  breeding  stage  collected  from  Upper  13  Resen/oir. 


Snout/Vent  Length  (mm) 


Date 

Stage 

0-19 

20-39 

40-59 

60-79 

80-89 

100-119 

8  Jan 

4 

2 

17Mar 

4 

5 

7  May 

1 

n 

43 

24  Jun- 

1 

20 

79 

24 

8 

9  Jul 

2 

13 

4 

1 

7 

5 

2 

5 

23-28  Aug 

1 

1 

1 

2 

3 

4 

n 

4 

3 

9  Oct 

1 

2 

2 

4 

2 

3 

1 

1 

2 

4 

4 

10  Nov 

1 

1 

2 

3 

1 

2  Dec 

4 

5 

48 


ders.  We  resampled  this  site  several 
times  through  February  1988.  Each 
time  we  caught  only  catfish,  although 
salamanders  were  abundant  in 
nearby  tanks. 

In  this  instance  disease  as  well  as 
predation  may  have  contributed  to 
decline  of  the  salamander  popula- 
tion. On  24  August  1985  we  found 


three  dead  mature,  branchiate 
morphs  in  the  tank.  We  also  ob- 
served a  significant  decline  in  sala- 
mander p)opulations  on  this  date  at 
two  other  tanks  with  diseased  sala- 
manders. Yellow  bullheads  are 
highly  carnivorous  (Minckley  1973), 
and  we  do  not  expect  salamanders  to 
successfully  recruit  at  Huachuca 


Table  4.— Seasonal  variation  In  number  and  size  (SVL)  of  salamanders  In 
each  breeding  stage  collected  from  Huachuca  Tank. 


Snout/Vent  Length  (mm) 


Date 

Stage 

29  Mar 

1 

4/5 

21  Apr 

1 

4/5 

25  Jun 

1 

4/5 

5  Jul 

1 

3 

4 

5 

22-28  Aug 

1 

2 

3 

4 

5 

2  Dec 

1 

4 

Stage  0-19    20-39  40-59   60-79    80-89    100-119  120-140 


13 
46 


78 


21 


24 


18 
1 


9 
9 
1 
2 
5 
13 

1 

2 
14 


2 

16 

20 
1 

12 
17 


8 
7 

11 


1 
1 

4 


Table  5.— Seasonal  variation  In  number  and  size  (SVL)  of  salamanders  in 
each  breeding  stage  collected  from  Bodle  Canyon  Tank. 


Snout/Vent  Length  (mm) 


Date 


Stage  0-19    20-39  40-59   60-79    80-89    100-119  120-140 


28  Mar  1 
3 

25  Aug  1 
2 
3 
4 

26-27  Sep 
3 
4 

11  Nov  3 
4 


10 


74 


21 


18 
3 


21 
18 


8 

2 
2 


10 


2 
3 
2 
6 
3 
1 


2 
3 
1 

4 
1 


Tank  as  long  as  the  catfish  popula- 
tion remains  high.  Catfish  will  pre- 
sumably eat  eggs,  larvae,  and  all  but 
the  largest  salamanders.  We  know  of 
no  experiments  demonstrating  the 
minimum  number  of  catfish  that  will 
prohibit  salamander  reproduction. 

Bog  Hole  Tank. — We  collected 
salamanders  here  in  1979,  1980,  and 
one  larva  in  1982.  Native  fishes  com- 
prised longfin  dace  (Agosia  chrysogas- 
ter)  and  Gila  topminnow.  Since  the 
1970s,  several  exotic  fishes  including 
Gambusia  affinis,  Cyprinodon  macular- 
ius  eremus,  Lepomis  spp.,  and  Microp- 
terus  salmoides  (W.L.  Jvlinckley,  pers. 
comm.;  Minckley  and  Brooks  1985) 
have  become  established.  Disappear- 
ance of  A.  t.  stebbinsi,  and  the  two  na- 
tive fish  species,  correlates  with  es- 
tablishment of  non-native  fish  popu- 
lations. 

Frogs. — During  the  last  decade 
bullfrogs  (Ram  catesbeiana)  were  in- 
troduced in  SRV.  Their  introduction 
correlates  with  reduction  in  native 
frog  populations  in  the  valley,  but 
the  impact  of  bullfrogs  on  A.  t.  steb- 
binsi is  unknown.  Bullfrog  larvae 
may  eat  salamander  eggs,  while 
adults  may  prey  on  larval  salaman- 
ders. Bullfrogs  may  also  act  as  vec- 
tors for  disease,  since  in  the  three 
tanks  where  salamanders  were  heav- 
ily affected  by  disease,  bullfrog 
populations  were  large  and  appar- 
ently unaffected.  Frogs  may  be  a 
natural  reservoir  for  disease,  and  suf- 
fer few  negative  effects  from  the 
pathogen(s).  Since  they  disperse 
readily  to  colonize  surrounding  habi- 
tats, they  may  also  help  spread  dis- 
ease among  amphibian  populations. 

S  alam  anders. — Commercial 
baitdealers  (waterdoggers),  fisher- 
men, and  private  landowners  intro- 
duce native  and  exotic  salamanders 
into  aquatic  habitats  in  Arizona 
(Collins  1981).  Salamanders  are  used 
commonly  as  bait  by  fishermen  in  the 
American  Southwest  (table  6),  and 
Lowe  (1955)  first  noted  that  salaman- 
ders were  being  introduced  into  Ari- 
zona for  this  purpose.  SRV  is  closed 
to  "waterdog"  collecting  under  Ari- 


49 


zona  Game  and  Fish  Commission 
order  #R1 2-4-311.  Enforcement  is  dif- 
ficult, because  SRV  is  large  and 
sparsely  settled.  It  would  be  easy  to 
introduce  exotic  A.  Hgrinum  into  this 
valley.  Pre-mating  and  post-mating 
isolating  mechanisms  in  the  A.  H- 
grinum  species  group  within  Amby- 
stoma  are  weak  (Brandon  1972,  Nel- 
son and  Humphrey  1972). 
Introduced  A.  Hgrinum  would  be  ex- 
pected, therefore,  to  easily  interbreed 
with  native  tiger  salamanders. 

Discussion 

In  theory,  average  heterozygosity  or 
gene  diversity  of  organisms  in  an 
area  can  be  decomposed  into  gene 
diversities  within  and  between  any 
subpopulations  comprising  the  total 
number  of  organisms  in  the  popula- 
tion (Nei  1987).  If  all  organisms  in  a 
population  are  a  panmictic  aggre- 
gate, then  the  component  describing 
variation  between  subpopulations  is 
zero.  We  have  no  information  on  dis- 
persal between  tanks  in  SRV,  so  for 
this  discussion  we  arbitrarily  con- 
sider each  tank  a  subpopulation  and 
together  all  tanks  comprise  the  total 
population  of  SRV  salamanders. 
Within  this  context  our  results  high- 
light several  factors  to  consider  in 
trying  to  understand  the  evolution- 
ary genetics  of  SRV  tiger  salaman- 
ders. 

Mean  heterozygosity  (.0015)  for  A. 
t.  stebbinsi  is  the  lowest  reported  for 
any  salamander  (Jones  et  al.  1988). 
Salamanders  in  SRV  went  through 
one  or  more  bottlenecks  at  some 
point  in  their  history,  but  cause(s) 
and  time  of  reduction  in  numbers 
and  associated  genetic  diversity  are 
unknown.  The  effect  of  a  one-time 
bottleneck  is  a  drastic  decrease  in  ex- 
pected heterozygosity  of  a  popula- 
tion, and  in  theory,  repeated  bottle- 
necks could  reduce  gene  diversity 
even  more  (Motro  and  Thomson 
1982). 

Current  factors  affecting  changes 
in  SRV  salamander  numbers  may 


provide  some  insight  into  the  origin 
and /or  perhaps  maintenance  of  low 
gene  diversity  in  SRV.  Increased 
heterozygosity  generally  correlates 
positively  with  traits  associated  with 
high  individual  vigor  and  fitness, 
plus  population  stability  (Mitton  and 
Grant  1984).  Susceptibility  to  disease 
or  apparent  reduced  ability  to  over- 
come infection  may  thus  be  conse- 
quences of  reduced  genetic  variation 
in  SRV  salamanders.  A  historical  bot- 
tleneck in  population  size  with  asso- 
ciated loss  of  gene  diversity  in  SRV 
salamanders,  therefore,  could  have 
resulted  in  populations  more  suscep- 
tible to  disease.  This  susceptibility,  as 
seen  in  contemporary  stocktanks, 
could  easily  cause  severe  reductions 
in  numbers  of  salamanders  and  re- 
tard any  expected  increase  in  gene 
diversity.  O'Brien  et  al.  (1985)  pro- 
vide a  related  example.  They  de- 
scribe how  extremely  low  genetic 
variation  in  the  South  African  chee- 
tah may  derive  from  a  population 


bottleneck.  Low  genetic  variation 
seen  in  structural  loci  also  extends  to 
the  major  histocompatibility  com- 
plex. This  extreme  monomorphism 
correlates  with  a  hypersensitivity  in 
cheetahs  to  some  viral  pathogens, 
and  they  feel  the  sensitivity  of  this 
genetically  uniform  species  to  patho- 
gens provides  an  example  of  the  pro- 
tection against  disease  genetic  vari- 
ation provides  to  species.  The  mecha- 
nism connecting  low  genetic  vari- 
ation revealed  by  electrophoresis  and 
susceptibility  to  disease  is  unclear. 
Hence,  for  both  cheetahs  and  SRV 
salamanders  it  is  uncertain  if  reduc- 
tion of  population  size  and  loss  of 
genetic  variation  increased  suscepti- 
bility to  disease,  or  alternatively,  sus- 
ceptibility increased  for  some  other 
reason,  and  this  lead  to  reductions  in 
population  numbers. 

Two  additional  factors,  again 
found  in  present  stocktanks,  would 
reinforce  this  pattern  of  change  in 
numbers  of  salamanders  and  reduc- 


> 


120" 


100" 


80" 


60" 


40" 


20 


2         3  4 
Breeding  Stage 


Figure  2.— Variation  in  SVL  with  breeding  stage  for  aninnals  from  four  SRV  populations:  soiid 
llne=Upper  13  Reservoir,  dotted  line=Parl<er  Canyon  Tank  #1,  dots+dasties=Huactiuca  Tank, 
daslies=Bodie  Canyon  Tank.  (Circles=mean,  vertical  line=lSE,  perpendicular  tiorizontal 
line=linnits.) 


50 


tion  in  heterozygosity.  First,  in  SRV, 
most  salamanders  occur  in  aquatic 
habitats  and  most,  if  not  all,  salaman- 
ders in  the  water  at  the  time  of  an 
epidemic  are  apparently  killed.  Since 
most  SRV  salamanders  are  adult, 
branchiate  animals,  aquatic  disease 
dramatically  reduces  effective  popu- 
lation size.  Furthermore,  future 
population  recruitment  is  reduced 


since  a  larval  year  class  is  also  lost. 
Thus,  the  preponderance  of  branchi- 
ate morphs  in  SRV  subpopulations 
exacerbates  any  negative  effects  of 
aquatic  diseases  on  population  size 
and  heterozygosity.  If  disease  is  a 
predictable  selection  pressure,  how- 
ever, it  is  not  obvious  why  relative 
frequency  of  adult  morphs  in  a  sub- 
population  has  not  shifted  from 


2         3  4 
Breeding  Stage 


Figure  3.— Variation  in  SVL  with  breeding  stage  for  ail  SRV  populations.  Synnbols  as  In  figure  2. 


Table  6.— Total  bait  sales  in  the  Lower  Colorado  River  basin  (modified  after 
Espinoza  et  al.  1970). 

Area  Value  of  sales  ($)  Volume  of  sales 


Salamanders 

Minnows 

1 .  Las  Vegas-Lake  Mead 

190,000 

1,250,000 

750,000 

2.  Mid-river 

110,000 

570,000 

325,000 

3.  Parker  Dam 

80,000 

400,000 

185,000 

4,  Yuma 

53,000 

190,000 

290,000 

5.  BIythe-Palo  Verde 

24,000 

30,000 

230,000 

Total  in  1968 

457,000 

2,440,000 

1,780,000 

51 


branchiate  to  metamorphosed 
morphs.  Since  disease  appears  to 
equally  affect  metamorphosed  and 
branchiate  morphs,  this  may  indicate 
there  is  little  or  negligible  difference 
in  heritable  variation  for  disease  re- 
sistance betv/een  morphs.  Selection, 
therefore,  would  have  little  or  no  ef- 
fect on  relative  morph  frequencies. 
Likewise,  the  genetic  basis  of  paedo- 
morphosis  versus  metamorphosis  is 
poorly  understood.  It  may  be  that 
genetic  differences  between  morphs 
are  slight,  with  environmental  condi- 
tions largely  determining  relative  fre- 
quency of  each  adult  morph  in  a  sub- 
population. 

Second,  exotic  predaceous  fishes, 
like  an  aquahc-borne  disease,  will 
quickly  reduce  adult  and  larval  sala- 
mander numbers,  and  coincidently 
genetic  diversity,  in  any  stocktank  in 
which  they  are  introduced.  Haphaz- 
ard introduction  of  fishes  in  SRV 
habitats  may  help  maintain  low  lev- 
els of  genetic  diversity. 

Other  than  by  mutation,  heterozy- 
gosity in  SRV  could  be  increased  by 
the  introduction  of  exotic  A.  tigrinum, 
and  their  interbreeding  with  native 
SRV  salamanders.  The  only  report  on 
salamander  introductions  in  AZ  is  20 
yrs  old,  summarizes  use  of  salaman- 
ders as  bait  in  only  the  extreme  west- 
ern part  of  AZ,  and  provides  no  in- 
formation on  relative  numbers  im- 
ported into  AZ,  as  opposed  to  sala- 
manders moved  within  AZ  (Espi- 
noza et  al.  1970).  Nonetheless,  in 
1968,  about  2.5  million  salamanders 
in  western  AZ  were  available  for  po- 
tential introduction  into  aquatic  habi- 
tats. The  increased  number  of  people 
living  in  AZ  means  these  numbers 
are  probably  higher  now.  Further- 
more, salamanders  are  regularly  sold 
for  bait  in  all  major  population  cen- 
ters in  AZ,  not  just  along  the  Colo- 
rado River.  Salamanders  sold  in  AZ 
come  from  three  primary  sources:  (1) 
seined  from  AZ  populations;  (2)  col- 
lected and  imported  from  popula- 
tions in  at  least  NM,  OK,  CO,  TX, 
and  NE;  and  (3)  adults  and/or  larvae 
collected  in  AZ  or  other  states,  intro- 


duced  into  AZ  habitats  as  "brood 
stock,"  and  larvae  from  these  ani- 
mals collected  in  subsequent  years 
and  sold  as  bait. 

We  know  from  discussions  with 
residents  that  salamanders  are  at 
least  occasionally  moved  between 
tanks  in  SRV.  We  have  no  evidence 
salamanders  are  introduced  into  SRV 
from  elsewhere,  and  two  facts  sug- 
gest such  events  are  rare  or  non-exis- 
tent. First,  our  electrophoretic  data 
show  heterozygosity  is  uniformly 
low  for  SRV  animals  from  eight  sub- 
populations  separated  by  as  much  as 
25  km  (fig.  1)  (Jones  et  al.  1988).  Alle- 
lic diversity  should  be  higher  if  sala- 
manders are  regularly  being  intro- 
duced into  SRV.  Second,  there  is  only 
one  mitochondrial  DNA  clone  in 
SRV.  Again,  regular  introductions 
would  be  expected  to  result  in  more 
than  one  mtDNA  haplotype  in  SRV. 
Nonetheless,  continued  active  use  of 
salamanders  for  bait  in  AZ  means 
there  is  always  the  possibility  exotic 
animals  might  be  introduced.  This 
could  lead  to  introgressive  hybridiza- 
tion between  species  or  subspecies, 
or  perhaps  interbreeding  between 
genetically  distinctive  populations  of 
the  same  species.  Furthermore,  we 
cannot  completely  exclude  the  possi- 
bility that  A.  t.  nebulosum  and/or  A.  t. 
mavortium  was  deliberately  or  ac- 
cidently  introduced  into  SRV,  thus 
creating  the  opportunity  for  hybridi- 
zation between  these  races.  How- 
ever, several  arguments  suggest  sala- 
manders were  native  in  SRV  (Jones  et 
al.  1988). 

Among  tiger  salamanders  in  SRV, 
color  pattern  of  metamorphosed  ani- 
mals, relative  frequency  of  typical 
and  cannibal  morphs,  nuclear  gene 
frequencies  derived  from  electro- 
phoresis, and  mitochondrial  DNA 
genotype  each  show  distinctive  vari- 
ation relative  to  the  entire  A.  tigrinum 
complex.  We  conclude,  therefore, 
that  SRV  tiger  salamander  popula- 
tions are  sufficiently  different  to  war- 
rant at  least  subspecific  status  as  A.  t. 
stebbinsi  (Collins  1988,  Jones  et  al. 
1988).  Likewise,  the  small  number 


and  restricted  geographic  range  of 
SRV  populations  increases  their  like- 
lihood of  extinction.  These  facts 
coupled  with  our  information  con- 
cerning life  history,  incidence  of  dis- 
ease, and  potential  negative  effects  of 
exotic  animals  in  SRV,  argue  that 
conservation  efforts  and  careful 
management  of  A.  t.  stebbinsi  is 
needed.  Although  A.  tigrinum  has  a 
wide  distribution,  in  some  races  spe- 
cial effort  needs  to  be  directed  at  pro- 
tecting locally  adapted  populations 
to  conserve  the  diversity  of  genetic 
and  life  history  traits  characteristic  of 
this  polytypic  species. 

Acknowledgments 

We  thank  the  following  for  help  in 
the  field:  T.  Corbin,  B.D.  DeMarais, 
P.J.  Fernandez,  W.C.  Hunter,  A.S.O. 
Jones,  L.F.  Elliott,  P.  Rosen,  C.A. 
Schmidt,  and  C.W.  Seyle.  We  appre- 
ciate criticism  of  any  earlier  draft  of 
this  paper  provided  by  L.  Allison,  D. 
Begun,  W.L.  Minckley,  and  C.A. 
Schmidt.  G.  Goodwin  and  F.  Sharp 
generously  provided  access  to  their 
land  in  SRV.  Arizona  Game  and  Fish 
Department  provided  collecting  per- 
mits. This  research  was  supported  by 
funds  from  U.S.  Fish  and  Wildlife 
Service,  Office  of  Endangered  Spe- 
cies (order  #20181-0746-83)  awarded 
to  JPC,  a  National  Science  Founda- 
tion grant  to  JPC  (BSR-8407930),  and 
a  Tucson  Audubon  Society  grant  to 
TRJ. 

Literature  Cited 

Allendorf,  Fred  W.  and  Robb  F. 
Leary.  1988.  Conservation  and  dis- 
tribution of  genetic  variation  in  a 
polytypic  species,  the  cutthroat 
trout.  Conservation  Biology  2:170- 
184. 

Behnke,  R.J.  1972.  The  systematics  of 
salmonid  fishes  of  recentiy  glaci- 
ated lakes.  Journal  of  the  Fisheries 
Research  Board  of  Canada  29:639- 
671. 


Brandon,  Ronald  A.  1972.  Hybridiza- 
tion between  Mexican 
salamanders  Ambystoma  dumerilii 
and  Ambystoma  mexicanum  under 
laboratory  conditions.  Herpetol- 
ogica  28:199-207. 

Brown,  D.E.,  ed.  1982.  Biotic  commu- 
nities of  the  American  Southwest- 
United  States  and  Mexico.  Desert 
Plants  4:1-342. 

Collins,  James  P.  1981.  Distribution, 
habitats,  and  life  history  variation 
in  the  tiger  salamander,  Amby- 
stoma tigrinum,  in  east-central  and 
southeast  Arizona.  Copeia 
1981(3):666-675. 

Collins,  James  P.  1988.  Evolutionary 
relationships  of  Ambystoma  ti- 
grinum stebbinsi  to  other  South- 
western A.  tigrinum  based  on  mi- 
tochondrial DNA.  Final  Report, 
Nongame  Branch,  Arizona  Game 
and  Fish  Department:  28  p. 

Collins,  James  P.,  Jeffry  B.  Mitton, 
and  Benjamin  A.  Pierce.  1980.  Am- 
bystoma tigrinum:  a  multispecies 
conglomerate?  Copeia  1980(4):938- 
941 

Echelle,  Anthony  A.  1988,  in  press. 
Conservation  genetics  and  genie 
diversity  in  threatened  fishes  of 
western  North  America.  In  Battie 
against  extinction:  Twenty  years  of 
native  fish  management  in  the 
desert  West.  W.  L.  Minckley  and 
James  E.  Deacon,  eds.  University 
of  Arizona  Press,  Tucson. 

Espinoza,  F.A.,  James  E.  Deacon,  and 
Andre  Simmons.  1970.  An  eco- 
nomic and  biostatistical  analysis  of 
the  bait  fish  industry  in  the  lower 
Colorado  River.  Special  Publica- 
tion of  the  University  of  Nevada, 
Las  Vegas:  1-87. 

Fowler,  Murray,  E.,ed.  1978.  Zoo  and 
wild  animal  medicine.  W.B.  Saun- 
ders, Philadelphia,  PA. 

Gehlbach,  F.R.  1967.  Ambystoma  ti- 
grinum (Green).  Catalogue  of 
American  Amphibians  and  Rep- 
tiles: 52.1-52.4. 

Gould,  Stephen  J.  1977.  Ontogeny 
and  phylogeny.  Harvard  Univer- 
sity Press,  Cambridge,  MA:  501  p. 


52 


Hendrickson,  Dean  A.  and  W.L. 
Minckley.  1985.  Cienegas- Vanish- 
ing climax  communities  of  the 
American  Southwest.  Desert 
Plants  6:131-175. 

Jones,  Thomas  R.  1988.  Macrogeogra- 
phic  and  microgeographic  pat- 
terns of  evolutionary  differentia- 
tion in  tiger  salamanders  (Amby- 
stoma  tigrinum)  in  the  Southwest- 
ern U.S.A.,  Ph.  D.  Dissertation, 
Arizona  State  University,  Tempe, 
AZ. 

Jones,  Thomas  R.,  James  P.  Collins, 
Thomas  D.  Kocher,  and  Jeffry  B. 
Mitton.  1988.  Systematic  status 
and  distribution  of  Ambystoma  ti- 
grinum stebbinsi  Lowe 
(Amphibia:Caudata).  Copeia 
1988(3):621-635. 

Lowe,  Charles  H.  1955.  The  salaman- 
ders of  Arizona.  Transactions  of 
the  Kansas  Academy  of  Sciences 
58:237-251. 

Lowe,  Charles  H.,  ed.  1964.  The  ver- 
tebrates of  Arizona.  University  of 
Arizona  Press,  Tucson,  Arizona: 
270  p. 


Meffe,  Gary  K.  and  Robert  C.  Vrijen- 
hoek.  1988.  Conservation  genetics 
in  the  management  of  desert 
fishes.  Conservation  Biology 
2:157-169. 

Minckley,  W.L.  1973.  Fishes  of  Ari- 
zona. Arizona  Game  and  Fish  De- 
partment, Phoenix,  Arizona:  293  p. 

Minckley,  W.L.  and  James  E.  Brooks. 
1985.  Transplantations  of  native 
Arizona  fishes:  records  through 
1980.  Journal  of  the  Arizona-Ne- 
vada Academy  of  Sciences  20:73- 
89. 

Mitton,  J.B.  and  M.C.  Grant.  1984. 
Associations  among  protein 
heterozygosity,  growth  rate,  and 
developmental  homeostasis.  An- 
nual Review  of  Ecology  and  Sys- 
tematics  15:479-499. 

Motro,  Uzi  and  Glenys  Thomson. 
1982.  On  heterozygosity  and  the 
effective  size  of  populations  sub- 
ject to  size  changes.  Evolution 
36:1059-1066. 

Nei,  Masatoshi.  1987.  Molecular  evo- 
lutionary genetics.  Columbia  Uni- 
versity Press,  New  York,  N.Y.: 
512  p. 

Appendix 


Nelson,  Craig  E.  and  R.R. 

Humphrey.  1972.  Artificial  inter- 
specific hybridization  among  Am- 
bystoma. Herpetologica  28:27-32. 

O'Brien,  S.J.,  M.E.  Roelke,  L.  Marker, 
A.  Newman,  C.  A.  Winkler,  D. 
Meltzer,  L.  Colly,  J.  F.  Evermann, 
M.  Bush,  and  D.  E.  Wildt.  1985. 
Genetic  basis  for  species  vulnera- 
bility in  the  cheetah.  Science 
227:1428-1434. 

Rose,  Francis  R.  and  D.  Armentrout. 
1976.  Adaptive  strategies  of  Amby- 
stoma tigrinum  (Green)  inhabiting 
the  Llano  Estacado  of  West  Texas. 
Journal  of  Animal  Ecology  45:713- 
729. 

Ryder,  Oliver  A.  1986.  Species  con- 
servation and  systematics:  the  di- 
lemma of  subspecies.  Trends  in 
Ecology  and  Evolution  1:  9-10. 

Ryman,  Nils  and  Fred  Utter,  eds. 
Population  genetics  and  fishery 
management.  University  of  Wash- 
ington Press,  Seattle,  WA:  420  p. 

Soule,  Michael  A.,ed.  1987.  Viable 
populations  for  conservation. 
Cambridge  University  Press,  Cam- 
bridge, England:  189  p. 


Locality  data  for  all  populations  of  A.  t.  stebbinsi,  taken  from  ttie  following  U.S.G.S.  7V2  nnin.  quadrangles: 
Cannpini  Mesa,  Canelo  Pass,  Duquesne,  Harstiaw,  Huactiuca  Peak,  Lochiel. 


Site  Locality  Map  codes        Site  Locality  Map  codes 


Bodie  Canyon  Tank: 
Bog  Hole  Tank: 
Campini  Mesa  Tank  #1: 
FS  58  Tank: 
FS  799  Tank: 
Grennan  Tank: 
Heron  Springs  Tank: 
Huachuca  Tank: 


NW»  SE«  sec.2,  T.24S,  R.18E  8 

31*22'  30"N,  no  •28'  45"W 

NW»  SE«  sec33,  T.22S,  R17E  F 

31*  28'  36'H  110°  37"  06"W 

SW»  E»  sec.l9,  T.24S,  R.19E 

31*21' 00",  110*26'  45"W 

NE»  NE»  sec6,  T.23S,  RITE  F 

31*27' 03"N,  110*38'  49"W 

SW-  NE»  sec36,  T22S,  R.17E  7 

31*28'  48"N,  110*34'  09"W 

S  center  sec.  14,  T.23S,  R.16E  6 

31*25'  29'H  110°40'47"W 

SW-  NE-  sec.l4,  T24S,  R17E 

31*20'  39"N,  110*34'  54"W 

NE-  NW»  sec.l5,  T24S,  R18E  3,F,D 

31*21'  12"N,  110°30'15"W 


Inez  Tank: 
Judy  Tank: 

Ki-He-Kah  Ranch  Tank: 

Lower  13  Reservoir: 

Meadow  Valley  Flat 
Tank  #1 

Parker  Canyon  Tank  #1 : 
School  Canyon  Tank  #1 : 
School  Canyon  Tank  #2: 
Upper  13  Reservoir: 


SW-  NW-  sec.2,  T.24S,  R18E  D 

31*22'  30"N,  110*29'30"W 

SE»  SE»  sec35,  T23S,  R.18E  9 

31*23'  04"N,  110*29' 19"W 

SW»  SW-  sec.l,  T.23S,  R17E 

31*26'  26"N,  110*35'  22"W 

SW»  NE»  sec.l8,T.24S,  R17E 

31*20'  49"N,  110*39'  05"W 

SW  •  NE«  sec.6,  T.22S,  R.17E  1 

31*27'  49"N,  110*38'  47"W 

NE-  NE»  sec.l9,  T.24S,  R.18E  2,D 

31*20'  16"N,  110*32' 42"W 

NE»  SE-  sec9,  T.24S,  R.18E  4 

31*21' 28"N,  110*24'  04"W 

NE»  SE»  sec.l7,  T.24S,  R19E 

31*21'  14"N,  110*24'  24"W 

S  center  sec.7,  T.24S,  R.17E.  5 

31*21' 18"N,  110*39' 16"W 


53 


Habitat  Requirements  of  New 
IVIexico's  Endangered 
Salannanders^ 

Cynthia  A.  Ramotnik^  and  Nornnan  J. 
Scott,  Jr.^ 


Abstract.— We  measured  habitat  components  for 
two  state-listed  endangered  salamanders  in  New 
Mexico  in  1986  and  1987.  Both  species  ore  restricted 
to  mesic  environments  within  high-elevation,  mixed 
coniferous  forests.  Steep  slope  and  high  elevation 
were  the  most  useful  variables  for  predicting  the 
occurrence  of  Jemez  Mountains  salamanders  and 
Sacramento  Mountain  salamanders,  respectively. 
Although  the  discriminant  models  show  some 
predictive  value  in  detecting  salamanders  based  on 
habitat  variables,  we  believe  that  the  best  survey 
technique  is  ground-truth  surveys  in  wet  weather.  A 
better  fit  of  the  discriminant  models  might  be 
obtained  by  including  variables  not  measured  e.g., 
fire  and  logging  history,  and  soil  characteristics.  We 
offer  interim  management  guidelines  as  a  result  of 
our  analysis. 


Two  of  the  three  species  of  salaman- 
ders that  occur  in  New  Mexico  are 
restricted  to  coniferous  forests  at 
high  elevations.  The  Jemez  Moun- 
tains salamander  (Plethodon  neomexi- 
canus)  (fig.  1)  is  known  only  from 
north-central  New  Mexico  at  the 
southern  terminus  of  the  Rocky 
Mountains  (Reagan  1972).  The  Sacra- 
mento Mountain  salamander  (Aneides 
hardii)  (fig.  2)  occurs  in  the  Capitan 
and  Sacramento  Mountains  in  south- 
central  New  Mexico  (Williams  1976). 
These  lungless  salamanders,  with 
small  body  sizes  and  terrestrial  juve- 
nile development,  are  restricted  to 
mesic  environments.  Lowe  (1950) 
suggested  that  both  species  are  rel- 
icts of  the  mid-Tertiary  Rocky  Moun- 
tain fauna. 

In  1975,  both  species  were  listed 
by  the  state  of  New  Mexico  as  endan- 
gered due  to  their  restricted  distribu- 
tion (Hubbard  et  al.  1979).  Since 
1980,  increases  in  timber  harvest  by 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff, AZ.  July  19-21  1988). 

^Museum  Specialist,  U.S.  F/sh  &  Wildlife 
Service,  National  Ecology  Researchi  Center. 
1300  Blue  Spruce  Drive.  Fort  Collins,  CO 
80524. 

^Zoologist,  U.S.  Fisfi  &  Wildlife  Service. 
National  Ecology  Research)  Center,  Mu- 
seum of  Southwestern  Biology,  University  of 
New  Mexico,  Albuquerque.  NM  87131. 


the  U.S.  Forest  Service  (USPS)  and 
changes  in  timber  practices  have 
prompted  concern  about  the  effect  of 
logging  on  these  salamanders  (Scott 
et  al.  1987,  U.S.  Fish  &  Wildlife  Serv- 
ice 1986).  Most  of  the  range  of  each 
species  occurs  on  National  Forest 
(NF)  lands,  and  the  close  association 
of  these  salamanders  with  mixed  co- 
niferous forests  may  make  them  vul- 
nerable to  some  forest-management 
practices.  In  1985,  both  species  were 
placed  under  review  as  potentially 
threatened  or  endangered  species 
under  the  Federal  Endangered  Spe- 
cies Act  (Ramotnik  1986,  Staub  1986). 
As  a  result,  an  interagency  commit- 
tee was  established  to  identify  data 
and  management  needs  and  develop 
strategies  to  address  these  needs. 


Figure  1  .—Jemez  Mountain  salannander 
(Plethodon  neomexicanus).  Photo  by 
Stephen  Corn. 


Figure  2.— Sacramento  Mountain 
salamander  (Aneides  hardii).  Photo  by 
Stephen  Corn. 

In  1986,  the  U.S.  Fish  &  Wildlife 
Service  (USFWS)  contracted  with  the 
USFS  to  study  these  species  on  NF 
lands.  The  primary  objectives  were 
to  survey  for  salamanders  in  plan- 
ning units  under  consideration  for 
future  logging  operations  and  to 
characterize  salamander  habitats  us- 
ing habitat  components  that  are 
meaningful  and  useful  to  USFS  biolo- 
gists and  land  managers.  This  infor- 
mation would  be  used  to  assess  po- 
tential salamander  habitat  from  maps 
or  aerial  photos,  thereby  reducing 
the  need  to  inventory  areas  by 
ground-truth  assessment. 

In  this  paper,  we  characterize 
habitats  of  Jemez  Mountains  sala- 
manders and  Sacramento  Mountain 
salamanders  based  on  general  site 
characteristics  and  surface  cover 


54 


items  that  could  serve  as  refugia  for 
salamanders.  We  use  a  multivariate 
analysis  of  habitat  characteristics  that 
describes  areas  with  and  without 
salamanders,  and  present  manage- 
ment guidelines  as  a  result  of  this 
analysis. 

Study  Areas 

We  studied  the  Jemez  Mountains 
salamander  within  the  Santa  Fe  NF 
in  the  Jemez  Mountains  (Los  Alamos, 
Rio  Arriba,  and  Sandoval  Counties, 
New  Mexico),  which  are  located  ap- 
proximately 100  km  north  of  Al- 
buquerque (fig.  3).  The  Jemez  Moun- 
tains are  volcanic  in  origin  and  are 
underlain  by  volcanic  rock,  ash,  and 
pumice.  The  predominant  feature  in 
the  area  is  the  volcanic  caldera,  the 
Valle  Grande,  around  which  the 
mountains  lie.  Fieldwork  on  the  Sac- 
ramento Mountain  salamander  was 
conducted  in  the  Sacramento  Moun- 
tains, within  the  Lincoln  NF,  Otero 
County,  New  Mexico  fig.  3).  Volcanic 
intrusions  occur  within  the  Paleozoic 
strata  of  the  Sacramento  Mountains. 
Elevations  in  the  Jemez  Mountains 
range  from  2130-3410  m,  and  from 
2290-3600  m  in  the  Sacramento 
Mountains. 

Habitat  types  within  these  eleva- 
tional  ranges  occur  within  the  Rocky 
Mountain  upper  montane  (2290-2900 
m)  and  subalpine  (2900-3660  m)  for- 
est association  (Castetter  1956).  The 
upper  montane  forest  association 
(Shelf ord  1963)  is  characterized  by 
mixed  coniferous  forests  dominated 
by  white  fir  (Abies  concolor),  Douglas- 
fir  (Pseudotsuga  menziesii),  Engelmann 
spruce  (Picea  engelmannii),  and  blue 
spruce  (Picea  pungens).  Deciduous 
components  include  quaking  aspen 
(Populus  tremuloides),  Rocky  Moun- 
tain maple  (Acer  glabrum),  oak  {Quer- 
cus  spp.),  New  Mexico  locust  (Robinia 
neomexicam),  and  oceanspray  (Holo- 
discus  dumosus).  Ponderosa  pine 
(Pinus  ponderosa)  stands  predominate 
at  the  lower  elevations,  particularly 
on  south-facing  slopes.  Within  the 


subalpine  forest  association,  Engel- 
mann spruce,  Douglas-fir,  and  white 
fir  are  the  most  common  trees.  Aspen 
and  Rocky  Mountain  maple  are 
found  to  a  lesser  extent.  Aspen 
groves,  talus  fields,  and  open  mead- 
ows are  present  at  higher  elevations. 
Annual  precipitation  in  the  Jemez 
Mountains  ranges  from  400-550  mm 
(Castetter  1956)  and  is  slightly  higher 
in  the  Sacramento  Mountains.  Much 
of  the  precipitation  falls  between  July 
and  September  (Kunkel  1984). 

Methods 

We  conducted  fieldwork  in  the  sum- 
mers of  1986  and  1987  (Jemez  Moun- 
tains: 28  July-14  August  1986,  29 
June-11  July  1987,  24  August-5  Sep- 
tember 1987;  Sacramento  Mountains: 
22  August-10  September  1986, 8-20 
June  1987;  20  July-1  August  1987). 
These  dates  included  the  surface  ac- 
tivity periods  of  Jemez  Mountains 
salamanders  (Reagan  1972)  and  Sac- 
ramento Mountain  salamanders 
(Williams  1976). 

Transects  were  established  in  for- 
ested areas;  most  were  located  in 
planning  units  selected  by  USPS  per- 
sonnel. Within  these  areas,  locations 
of  transects  were  selected  from  topo- 
graphic maps  to  sample  a  variety  of 
topographic  aspects.  South-facing 
slopes  were  not  searched  in  the 
Jemez  Mountains  due  to  the  diffi- 
culty in  locating  salamanders  on 
these  slopes  (Ramotnik  1988).  To  en- 
sure having  sites  occupied  by  sala- 
manders, we  visited  known  localities 
or  areas  where  salamanders  had  re- 
cently been  found.  A  small  number 
of  sites  outside  planning  units  were 
chosen  from  topographic  maps. 

We  established  100-m^  transects  (2 
m  X  50  m)  oriented  uphill  from  near 
the  bottoms  of  slopes.  Our  transect  is 
modified  from  area-constrained 
searches,  a  technique  developed  by 
others,  e.g..  Bury  (1983),  Bury  and 
Corn  (this  volume).  Bury  and  Ra- 
phael (1983),  Campbell  and  Christ- 
man  (1982),  Raphael  (this  volume). 


and  Raphael  and  Rosenberg  (1983). 
The  areas  of  four  classes  of  cover 
items  (rock,  bark,  fine  woody  debris, 
and  coarse  woody  debris)  were  esti- 
mated visually.  We  further  divided 
coarse  woody  debris  (CWD)  into 
three  decay  classes,  adapted  from  a 
five-class  scheme  for  rating  decom- 
position of  Douglas-fir  logs  (Franklin 
et  al.  1981).  To  emphasize  differences 
between  decay  classes,  we  combined 
classes  1  and  2  (CWDl),  and  classes  3 
and  4  (CWD3),  and  placed  the  most 
decayed  logs,  class  5,  in  a  third  cate- 
gory (CWD5). 

Aspect  was  taken  with  a  magnetic 
compass  at  10,  30,  and  50  m.  Com- 
pass readings  were  assigned  to  one 
of  four  aspect  classes  where  316-45°  = 
north-facing;  46-135°  =  east-facing; 
136-225°  =  south-facing;  and  226-315° 
=  west-facing.  Percent  slope  was  de- 
termined with  a  clinometer,  and  per- 
cent canopy  cover  was  estimated 
with  a  spherical  densiometer 
(Lemmon  1956).  Both  measurements 
were  recorded  at  10-m  intervals.  All 
readings  were  made  along  the 
transect  and  averaged  for  the 


Figure  3.— Distribution  of  Jemez  Mountairw 
salamanders  (Plethodon  neomexicanus) 
and  Sacramento  Mountain  salamanders 
(Aneides  hardii)  in  New  Mexico. 


55 


transect.  Numbers  of  white  fir  and 
Douglas-fir  were  pooled  in  a  single 
class  (TFIR),  as  were  Engelmann  and 
blue  spruce  (TSPRUCE),  and  Pinus 
spp.  (TPINE).  Numbers  of  trees 
within  tree  classes  were  counted  in  a 
20-m  X  50-m  plot  centered  over  the 
transect.  Twenty-three  measured  and 
derived  variables  were  used  in  the 
analyses  (table  1). 

We  determined  numbers  of  sala- 
manders on  transects  by  searching  all 
cover  items  manually  or  with  potato 
rakes.  The  locations  of  salamanders 
in  other  than  the  four  classes  of  cover 
items  also  were  recorded.  When  a 
salamander  was  found,  we  recorded 
snout-vent  length  (distance  from  tip 
of  snout  to  anterior  edge  of  vent), 
sex,  and  dimensions  and  type  of 
cover  item.  For  coniferous  logs,  we 
also  recorded  salamander  position 
relative  to  the  log  (in,  under,  or  un- 
der bark)  and  decay  class  (modified 
from  Com  and  Bury,  in  press,  Ra- 
phael and  Rosenberg  1983).  These 
data  were  used  to  calculate  densities 
of  salamanders  on  transects  and  to 
determine  cover  item  use  by  sala- 
manders. We  acquired  additional 
data  on  cover  item  use  by  salaman- 
ders by  locating  salamanders  in  areas 
on  both  sides  of  the  transects. 


Statistical  Analysis 

Data  for  transects  with  and  without 
salamanders  were  pooled  separately. 
We  calculated  descriptive  statistics 
(mean,  standard  error,  range)  for 
habitat  variables  in  the  two  groups 
and  used  a  one-way  analysis  of  vari- 
ance to  compare  transformed  vari- 
ables between  groups.  Size  classes  of 
fir  and  spruce  were  compared  be- 
tween the  two  groups  with  a  t-test. 

The  following  transformations 
were  applied  to  stabilize  the  variance 
of  the  habitat  variables  (Snedecor 
and  Cochran  1967)  and  to  increase 
the  probability  of  a  normal  distribu- 
tion: arcsine  (SLOPE  CANOPY); 
square  root  +  0.5  (tree  densities);  and 
log  +  0.5  (cover  items).  Elevahon  was 


not  transformed  because  values  were 
distributed  normally. 

A  stepwise  variable  entry  proce- 
dure (STEPDISC)  selected  the  "best 
set"  of  habitat  variables  to  discrimi- 
nate between  groups  and  reduced 
the  corriplexity  of  the  original  vari- 
able set.  Because  the  models  selected 
by  STEPDISC  are  not  necessarily  the 
best  possible  models  (SAS  Institute 
Inc  1982),  cross-validation  was  ac- 
complished by  using  canonical  analy- 
sis (CANDISC)  or  descriptive  dis- 
criminant analysis  (DDA)  (Williams 
1983).  DDA  attempts  to  establish  op- 


timal separation  between  groups  us- 
ing linear  transformations  of  the  in- 
dependent variables  based  on  vari- 
ables selected  by  the  stepwise  proce- 
dure. The  Mahalanobis  distance  be- 
tween group  means  was  tested  using 
an  F-statistic. 

Predictive  discriminant  analysis 
(PDA)  (Williams  1983)  (DISCRIM) 
was  used  to  test  the  discriminatory 
power  of  the  variables  selected  by 
DDA.  We  used  chi-square  analysis  to 
compare  cover  item  use  (of  the  four 
classes)  to  availability  and  to  com- 
pare aspects  of  transects  with  and 


Table  1.— Description  of  measured  and  derived  habitat  variables  used  In 
habitat  selection  analysis  of  two  species  of  New  Mexico  salamanders. 


Sampling  unit 
mnemonic 


50-m  x2-m  transect 

BARK 
CANOPY 

CWDl 

CWD3 

CWD5 

CWD 
ELEV 

FWD 

ROCK 
SLOPE 


Description 


50-m  X  20-m  plot 

SFIR 

Number  of 

MFIR 

Number  of 

LFIR 

Number  of 

TFIR 

SFIR  +  MFIR 

SSPRUCE 

Number  of 

MSPRUCE 

Number  of 

LSPRUCE 

Number  of 

TSPRUCE 

SSPRUCE  + 

TASPEN 

Number  of 

TNOD 

Number  of 

TOAK 

Number  of 

TPINE 

Number  of 

TSNAGS 

Number  of 

Estimate  of  amount  of  bark  on  ground  (m^) 
Average  percent  canopy  cover  recorded  with 
G  spherical  der^iometer 
Estimate  of  amount  of  poorly  decayed  coarse 
woody  debris  (m^) 

Estimate  of  amount  of  moderately  decayed 
coarse  woody  debris  (m^) 
Estimate  of  amount  of  well-decayed  coarse 
woody  debris  (m^) 
CWD1  +CWD3  +  CWD5 
Estimated  from  a  U.S.  Geological  Survey  topo- 
graphic map  (m) 

Estimate  of  amount  of  fine  woody  debris 
(sticks)  (m^ 

Estimate  of  amount  of  surface  rock  (m^) 
Average  percent  slope  measured  with  a  cli- 
nometer 


small  fir  (<20  cm  dbh) 
medium  fir  (20-50  cm  dbh) 
large  fir  (>50  cm  dbh) 
+  LFIR 

small  spruce  (<20  cm  dbh) 

medium  spruce  (20-50  cm  dbh) 

large  spruce  (>50  cm  dbh) 

MSPRUCE  +  LSPRUCE 

aspen  (all  sizes) 

non-oak  deciduous  (all  sizes) 

oak  (all  sizes) 

pine  (all  sizes) 

snags  (all  sizes) 


56 


without  salamanders.  The  Statistical 
Analysis  System  computer  package 
(SAS,  Version  5)  was  used  for  all 
analyses  (SAS  Institute  Inc  1982).  Sig- 
nificance levels  were  set  at  P  <  0.05 
unless  otherwise  indicated. 


Results 

Jemez  Mountains  Salanaander 

Salamanders  (N  =  28)  were  present 
on  10  of  43  transects  (23%)  with  a 
mean  density  of  3/100  m^  in  occu- 
pied areas.  One  hundred  twenty 
salamanders  were  found  in  areas  off 
the  transects.  Transects  with  sala- 
manders occurred  on  significantly 
steeper  slopes  and  at  lower  eleva- 
tions than  transects  without  salaman- 
ders (table  2).  Analysis  of  size  classes 
of  fir  and  spruce  showed  no  signifi- 
cant differences  between  transects 
with  and  without  salamanders.  Pro- 
portions of  decay  classes  of  CWD 


also  did  not  differ  significantly  be- 
tween the  two  groups  of  transects  (X^ 
=  0.28,  df  =  2,  P  >  0.90).  The  amount 
of  CWDl  was  similar  between 
groups  but  amounts  of  CWD3  and 
CWD5  were  higher  on  transects  with 
salamanders.  Although  no  south-fac- 
ing slopes  were  searched,  propor- 
tions of  other  aspects  occupied  by 
salamanders  were  not  different  from 
the  proportions  of  total  aspects 
searched  (X^  =  1.3,  df  =  2,  P  >  0.50). 

Three  of  the  original  20  variables 
were  selected  by  the  stepwise  vari- 
able entry  procedure  for  inclusion  in 
the  descriptive  discriminant  model: 
SLOPE,  TPINE,  and  LSPRUCE  (table 
3).  Subsequent  analysis  by  DDA  re- 
tained these  variables.  The  resultant 
discriminant  function  explained  38% 
of  the  between-group  variance;  how- 
ever, it  did  not  have  significant 
power  in  discriminating  between 
groups  (F  =  2.34,  P  =  0.09).  This  func- 
tion describes  a  multivariate  gradient 
that  ranges  from  steep  slopes  with 


Table  2,— Comparison  of  habitat  variables  measured  on  transects  with  and 
without  Jemez  Mountains  salamanders,  Santa  Fe  National  Forest,  1986- 
1987.  Significance  is  based  on  one-way  analysis  of  variance.  Mnemonic 
codes  for  habitat  variables  are  explained  in  table  1 . 


Transects  (N  =  10) 
with  salamanders 


Transects  (N  =  33) 
without  salamanders 


Mnemonic 

X  ±  se  (range) 

X  ±  se  (range)  Significance 

ELEV 

2526  +35.8 

(2359-2621) 

2635  +22.0 

(2332-2886) 

• 

SLOPE 

66  +  2.5 

(55-84) 

44+  2.8 

(0-82) 

CANOPY 

62  +  1,8 

(56-65)^ 

64+  2.1 

(21-82)2 

NS 

TFIR 

72  ±10,4 

(29-156) 

95  +  10.3 

(22-292) 

NS 

TSPRUCE 

17  +  6,6 

(0-59) 

20+  5,9 

(0-163) 

NS 

TPINE 

25  ±  7,8 

(0-63) 

9  +  2.1 

(0-56) 

NS 

TASPEN 

20+  8.8 

(1-96) 

17  +  2.5 

(0-60) 

NS 

TOAK 

10+  6,6 

(0-59) 

7±  2.4 

(0-50) 

NS 

TSNAGS 

33  +  6.1 

(5-64) 

27  +  3.3 

(3-82) 

NS 

TNOD 

29  +  10.4 

(0-103) 

8+  2.0 

(0-51) 

NS 

ROCK 

11  +  2,6 

(3-26) 

7  +  1.6 

(0-37) 

NS 

FWD 

4+  1.1 

(2-12) 

4+  0,5 

(0-15) 

NS 

BARK 

1+1.0 

(0-10) 

1  ±  0.1 

(0-3) 

NS 

CWD 

10+  1.9 

(1-20) 

9±  0.8 

(1-26) 

NS 

'P<0.05 
"P<  0.005 

'Data  are  available  for  5  frar\secfs. 
'Data  are  available  for  29  transects. 

many  pine  and  large  spruce  trees 
containing  salamanders,  to  shallow 
slopes  with  few  pine  or  large  spruce 
trees  without  salamanders.  SLOPE 
had  the  highest  discriminating  power 
(r^  =  0.73).  PDA  correctly  classified 
91%  of  the  33  transects  without  sala- 
manders and  80%  of  the  10  transects 
with  salamanders. 

The  10  transects  and  additional 
searches  produced  148  Jemez  Moun- 
tains salamanders;  the  type  of  cover 
item  was  known  for  all  but  one  sala- 
mander. Ninety-six  percent  (141/ 
147)  of  salamanders  were  distributed 
among  the  four  major  cover  classes 
as  follows:  CWD,  100  (68%);  ROCK, 
40  (27%);  FWD,  1  (1%).  No  salaman- 
ders were  found  under  BARK.  Three 
salamanders  (2%)  were  found  on 
transects  under  surface  litter  and 
three  salamanders  (2%)  were  found 
under  aspen  logs.  The  frequency  of 
salamanders  associated  with  CWD 
by  decay  class  was  CWDl — 4%; 
CWD3— 66%;  CWD5— 30%.  Of  28 
salamanders  found  on  transects,  24 
salamanders  were  associated  with 
one  of  the  four  classes  of  cover  items. 
Because  of  the  small  sample  size,  we 
were  unable  to  determine  a  correla- 
tion between  cover  item  availability 
and  use. 


Sacramento  Mountain 
Salanaander 

Salamanders  (N  =  233)  were  present 
on  26  of  80  transects  (33%)  with  a 
mean  density  of  6/100  m^  in  occu- 
pied areas.  We  located  387  salaman- 
ders in  areas  off  the  transects. 
Transects  with  and  without  salaman- 
ders differed  in  several  respects: 
transects  with  salamanders  occurred 
at  significantly  higher  elevations,  on 
shallower  slopes,  and  had  higher 
numbers  of  spruce  and  lower  num- 
bers of  pine  than  transects  without 
salamanders  (table  4).  Analysis  of 
size  classes  of  fir  and  spruce  revealed 
that  densities  of  large  fir  and  all  size 
classes  of  spruce  were  significantly 
higher  on  transects  with  salamanders 


57 


(LFIR:  t  =  3.38,  P  =  0.001;  SSPRUCE:  t 
=  2.85,  P  =  0.008;  MSPRUCE:  t  =  2.56, 
P  =  0.016;  LSPRUCE:  t  =  3.04,  P  = 
0.003)  (fig.  4).  Although  the  total 
amount  of  CWD  on  transects  with 
and  without  salamanders  was  not 
significantly  different,  there  was  sig- 
nificantly more  CWD5  on  transects 
with  salamanders  (X^  =  6.93,  df  =  2,  P 
>  0.05).  The  proportions  of  transects 
by  aspect  did  not  differ  between  the 
two  groups  (X2  =  3.83,  df  =  3,  P  > 
0.10). 

Because  numbers  of  the  three  size 
classes  of  spruce  were  significantly 
higher  on  transects  with  salaman- 
ders, we  substituted  TSPRUCE  for 
SSPRUCE,  MSPRUCE,  and 
LSPRUCE  in  subsequent  analyses.  A 
stepwise  variable  entry  procedure  se- 
lected eight  of  the  original  20  vari- 
ables for  inclusion  in  the  descriptive 
discriminant  model  (table  5).  Subse- 
quent DDA  kept  all  but  three 
(SLOPE,  CWDl,  and  TAPSEN)  in  the 
model.  The  resultant  discriminant 
function  explained  49%  of  the  be- 
tween-group  variance  and  had  sig- 
nificant power  in  discriminating  be- 
tween groups  (F  =  6.87,  P  <  0.0001). 
This  function  can  be  interpreted  ecol- 
ogically to  describe  a  gradient  that 
ranges  from  low  elevations  with 
many  pine,  few  spruce  and  large  fir, 
and  infrequent  CWD5  without  sala- 
manders, to  higher  elevations,  few 
pine,  many  spruce  and  large  fir,  and 
abundant  CWD5  that  contain  sala- 
manders. ELEV  had  the  highest  dis- 
criminating power  (r^  =  0.64).  PDA 
correctly  classified  96%  of  the  54 
transects  without  salamanders  and 
58%  of  the  26  transects  with  salaman- 
ders. 

The  26  occupied  transects  and  ad- 
ditional searches  produced  620  Sac- 
ramento Mountain  salamanders. 
Ninety-five  percent  (589)  were  dis- 
tributed among  the  four  major  cover 
classes  as  follows:  CWD,  377  (64%); 
ROCK,  127  (22%);  BARK,  58  (10%); 
and  FWD,  27  (4%).  Fourteen  sala- 
manders (2%)  were  found  under  as- 
pen logs  and  17  salamanders  (3%) 
were  above  or  below  surface  litter. 


The  frequency  of  salamanders  associ- 
ated with  CWD  in  the  three  decay 
classes  was  CWDl— 13%;  CWD3— 
62%;  CWD5— 25%.  Of  233  salaman- 
ders found  on  transects,  209  sala- 
manders were  associated  with  one  of 
the  four  classes  of  cover  items.  Ex- 
amination of  cover  item  availability 
and  use  for  these  salamanders  re- 
vealed that  salamanders  are  associ- 
ated with  some  cover  items  dispro- 
portionate to  their  availability  (X^  = 
59.9,  df  =  3,  P  <  0.001).  In  particular, 
Aneides  was  found  in  association 
with  FWD  proportionately  less  fre- 
quent than  expected,  and  used  well- 
decayed  and  moderately  decayed 
logs  to  a  greater  extent  than  expected 
(X2  =  62.1,  df  =2,  P<  0.001). 


Discussion 

Jemez  Mountains  Salamander 

While  canonical  analysis  did  not  dis- 
criminate between  transects  with  and 


without  salamanders,  it  did  identify 
steep  slopes  as  the  most  useful  vari- 
able in  determining  the  occurrence  of 
Jemez  Mountains  salamanders.  It  is 
possible  that  steep  slopes  contain 
more  interstitial  spaces  in  the  soil 
than  do  shallower  slopes.  The  soils  of 
steep  slopes  may  be  less  compacted 
than  those  of  more  gentle  slopes  due 
to  the  combined  effects  of  gravity, 
and  movement  of  water  and  soil.  As 
a  consequence  of  steep  slope  and  the 
presence  of  underlying  volcanic  rock 
characteristic  of  the  Jemez  Mountains 
(Burton  1982),  spaces  within  this  ma- 


Tobie  3.— Correlations  of  habitat 
variables  with  discriminant  scores 
for  transects  with  and  without 
Jemez  Mountains  salamanders. 


Mnemonic 


DFl 


SLOPE 

TPINE 

LSPRUCE 


0.73 
0.52 
0.35 


Table  4.--Comparlson  of  habitat  variables  measured  on  transects  with  and 
without  Sacramento  Mountain  salamanders,  Lincoln  National  Forest,  1986- 
1987.  Significance  Is  based  on  one-way  analysis  of  variance.  Mnemonic 
codes  for  habitat  variables  are  explained  In  Table  1 . 


Transects  (N  =  26) 
with  salamanders 


Transects  (N  =  54) 
without  salamanders 


Mnemonic 

x  ±  se  (Range) 

X  ±  se  (Range) 

Significa 

ELEV 

2779 

+  17.6 

(2618-2890) 

2682 

+ 

8.7 

(2450-2792)  " 

SLOPE 

39 

+  2.7 

(21-65) 

41 

+ 

1.6 

(17-70) 

♦* 

CANOPY 

72 

+  1.3 

(59-88) 

71 

+ 

1.3 

(53-90) 

NS 

TFIR 

67 

+  6.3 

(8-122) 

64 

+ 

4.0 

(14-144) 

NS 

TSPRUCE 

17 

+  7.6 

(0-186) 

1 

+ 

0.6 

(0-30) 

*• 

TPINE 

7 

+  2.1 

(0-50) 

22 

+ 

2.3 

(0-71) 

♦ 

TASPEN 

14 

+  4.1 

(0-74) 

17 

+ 

3.3 

(0-107) 

NS 

TOAK 

5 

+  2.4 

(0-59) 

18 

3.8 

(0-104) 

NS 

TSNAGS 

24 

+  2.8 

(6-56) 

25 

+ 

2.5 

(1-106) 

NS 

TNOD 

33 

+  7.7 

(4-180) 

34 

+ 

5.6 

(0-222) 

NS 

ROCK 

7 

+  1.7 

(0-33) 

7 

+ 

0.9 

(0-29) 

NS 

FWD 

6 

+  0.6 

(2-13) 

5 

+ 

0.5 

(0-14) 

NS 

BARK 

1 

+  0.3 

(0-6) 

1 

+ 

0.2 

(0-10) 

NS 

CWD 

12 

+  1.2 

(4-24) 

8 

+ 

0.8 

(0-26) 

NS 

"P<  0.005 
'P  <  0.05 


58 


130 


SPRUCE 


50  -I 

40 

30 

20  - 
10  - 
0 

130 


WITH  SALAMANDERS 
WITHOUT  SALAMANDERS 


120  A 
80 


FIR 


50  4-/-^ 


40 


30 


20  - 


10  - 


<20CM  20-50  CM  >  50  CM 


D.B.H. 


Figure  4.— Comparisons  of  average  size  classes  (d.b.h.)  of  spruce  and  fir  on  transects  witti 
and  without  Sacramento  Mountain  salamanders.  Boxes  indicate  95%  confidence  Inten/als 
for  ttie  mean.  Levels  of  significance  indicated  by  asterisks  are  0.05  (*)  and  0.005  (**). 


trix  of  rocky  soil  may  provide  refugia 
for  salamanders  during  inhospitable 
times  and,  thus,  may  provide  a  clue 
to  the  survival  of  this  salamander  in 
the  harsh  environment  of  the  Rocky 
Mountains.  The  largest  concentra- 
tions of  P.  neomexicanus  have  been 
found  in  association  with  talus  slopes 
(Whitford  and  Ludwig  1975,  Clyde 
Jones  pers.  comm.),  which  are  also 
important  to  many  other  western  Ple- 
thodon  (Brodie  1970).  Other  pletho- 
dontids  are  virtually  restricted  to  ar- 
eas with  a  loose  rocky  soil  (Aubry  et 
al.  1987,  French  and  Mount  1978, 
Herrington  and  Larsen  1985,  Jaeger 
1971). 

The  variables  selected  by  canoni- 
cal analysis  showed  some  predictive 
value.  Although  three  transects  with- 
out salamanders  were  misclassified 
by  PDA  as  transects  with 
salamanders,  Plethodon  was  found  in 
areas  adjacent  to  the  transects.  The 
two  transects  misclassified  as 
transects  without  salamanders  had 
values  for  TPINE  and  LSPRUCE 
closer  to  values  usually  associated 
with  transects  without  salamanders. 
Because  a  larger  percentage  of 
transects  without  salamaders  were 
correctly  classified  by  PDA,  these 
three  variables  may  better  describe 
the  conditions  under  which  salaman- 
ders are  absent  from  an  area,  rather 
than  describing  favorable  conditions 
under  which  they  would  occur. 

The  limited  discriminatory  and 
predictive  power  of  the  variables  se- 


Table  5.— Correlations  of  habitat 
variables  v/ith  discriminant  scores 
for  transects  with  and  v^lthout  Sac- 
ramento Mountain  salamanders. 


Mnemonic 


DFl 


ELEV 

TSPRUCE 

TPINE 

CWD5 

LFIR 

CWDl 

SLOPE 

TASPEN 


0.55 
0.42 
-0.47 
0.44 
0.34 
-0.05 
-0.06 
-0.02 


59 


lected  by  multivariate  techniques 
may  reflect  our  inability  to  reliably 
and  consistently  detect  the  presence 
of  Plethodon  at  a  site.  We  believe  that 
our  ability  to  detect  salamanders  is 
fairly  good  and  repeatable,  but  we 
realize  that  environmental  factors 
can  influence  the  relative  numbers  of 
salamanders.  During  repeated  visits 
to  the  same  sites,  Plethodon  was  more 
abundant  when  we  searched  under 
wet  conditions,  and  other  studies 
have  reported  a  significant  correla- 
tion between  movement  and  activity 
of  salamanders,  and  precipitation 
(Barbour  et  al.  1969,  Kleeberger  and 
Werner  1982,  MacCullough  and 
Bider  1975).  Low  densities  and 
patchiness  of  P.  neomexicanus  popula- 
tions also  can  hinder  detection  of  the 
animal.  In  comparison  with  densities 
of  red-backed  salamanders,  P.  cin- 
ereus,  (0.9-2.2  individuals/ m^;  Heat- 
wole  1962,  Jaeger  1980),  our  density 
estimates  for  Jemez  Mountains  sala- 
manders are  extremely  low  (0.03  in- 
dividuals/m^).  Although  Williams 
(1972)  reported  estimates  of  Jemez 
Mountains  salamanders  ten  times 
greater  than  ours,  he  noted  that  their 
distribution  was  spotty. 

A  better  fit  to  a  discriminant 
model  might  be  obtained  by  includ- 
ing variables  that  we  did  not  meas- 
ure, e.g.,  fire  and  logging  history  and 
soil  characteristics  (moisture,  pH, 
and  compaction).  Williams  (1976) 
suggested  that  logging  may  have 
eliminated  Jemez  Mountains  sala- 
manders from  part  of  Peralta  Canyon 
due  to  dry  conditions  resulting  from 
removal  of  most  of  the  canopy.  How- 
ever, there  was  no  documentation 
that  salamanders  occurred  at  the  site 
prior  to  logging.  Soil  characteristics, 
which  can  be  affected  by  fire  and  log- 
ging practices  (Childs  and  Hint  1987, 
DeByle  1981,  Krag  et  al.  1986),  also 
can  influence  the  distribution  of  ple- 
thodontid  salamanders,  that  occupy 
the  soil-litter  interface.  Plethodon  cin- 
ereus  was  excluded  from  27%  of  for- 
est habitat  in  eastern  deciduous  for- 
ests because  of  low  soil  pH  (Wyman 
and  Hawksley-Lescault  1987),  while 


the  distributions  of  up  to  10  amphibi- 
ans in  southeastern  New  York  were 
significantly  influenced  by  soil  pH 
and  moisture  (Wyman  1988). 

Salamanders  also  may  be  absent 
from  a  given  site  for  reasons  other 
than  unsuitability  of  habitat.  For  ex- 
ample, access  to  a  particular  area  by 
salamanders  may  be  impossible  due 
to  the  unsuitability  of  the  area  that 
surrounds  it,  e.g.,  dry,  open  field.  Or, 
a  climatic  event  may  have  eliminated 
salamanders  from  a  given  area  with- 
out sufficient  time  occurring  for  them 
to  recolonize  the  site. 


Sacramento  Mountain 
Salamander 

The  variables  selected  by  canonical 
analysis  were  able  to  discriminate  be- 
tween transects  with  and  without 
salamanders.  However,  these  vari- 
ables had  limited  predictive  value. 
Although  a  larger  percentage  of 
transects  without  salamanders  were 
correctly  classified  by  PDA,  there  is 
still  a  one-in-five  chance  of  being 
wrong  in  predicting  that  salaman- 
ders are  absent  from  a  site.  For  most 
management  decisions,  this  level  of 
uncertainty  will  not  be  acceptable, 
and  ground-truth  searches  will  have 
to  be  made. 

High  elevation  was  the  best  pre- 
dictor of  the  presence  of  Sacramento 
Mountain  salamanders  (table  5). 
Weigmann  et  al.  (1980)  also  found 
significantly  more  Sacramento 
Mountain  salamanders  on  transects 
at  higher  elevations.  The  higher  ele- 
vations of  the  Sacramento  Mountains 
experience  greater  rainfall,  cooler 
temperatures,  and  lower 
evapotranspiration  rates  than  the 
lower  elevations  and  therefore  may 
be  more  hospitable  to  plethodontid 
salamanders.  The  low  critical  ther- 
mal maximum  of  Aneides  probably 
reflects  adaptations  to  the  low  tem- 
peratures characteristic  of  their  mi- 
crohabitat  (Whitford  1968)  and  may 
restrict  salamanders  to  high  eleva- 
tions. 


Aneides  is  often  present  where  the 
best  habitat  predictors  indicate  they 
should  not  occur.  While  high-eleva- 
tion, wet,  north-facing  slopes  with  a 
mature  mixed-conifer  forest  do  har- 
bor Aneides,  salamanders  are  also 
found  less  predictably  in  areas  that 
may  be  drier  and  more  exposed  than 
the  model  would  indicate.  With  the 
exception  of  elevation,  the  ranges  of 
habitat  variables  on  transects  occu- 
pied by  salamanders  are  not  strik- 
ingly different  from  those  on  plots 
without  salamanders  (table  4).  This 
overlap  may  be  due  to  factors  not 
measured,  e.g.,  fire  and  logging  his- 
tory, and  it  may  show  an  ability  of 
salamanders  to  persist  after  habitats 
have  been  altered. 


Management  Guidelines 

Our  data  show  that,  despite  some 
predictive  power  of  the  habitat  vari- 
ables, the  level  of  uncertainty  in  pre- 
dicting salamander  occurrence  may 
preclude  their  use  by  the  USFS.  At 
this  time,  we  feel  the  best  survey 
technique  for  salamanders  is  ground- 
truth  surveys  in  wet  weather  during 
the  activity  season  of  each  species. 
Under  proper  conditions,  both  spe- 
cies are  easy  to  find  and  relatively 
unskilled  persons  can  be  quickly 
trained  to  survey  habitats.  Our  im- 
pression was  that  Plethodon  was 
more  difficult  to  survey,  because  it 
tended  to  retreat  underground  dur- 
ing dry  periods.  Aneides,  however, 
can  usually  be  found  even  during  ex- 
tended dry  periods. 

Our  attempts  to  explain  the  ab- 
sence of  salamanders  from  a  given 
area,  i.e.,  potential  difficulty  of  de- 
tecting all  salamanders  present,  and 
low  density  or  patchy  distribution  of 
populations,  may  overlook  the  possi- 
bility that  absence  is  not  solely  due  to 
unsuitable  habitat.  Absence  does  not 
necessarily  mean  avoidance,  but  may 
be  due  to  insufficient  time  for  the 
animal  to  recolonize  an  area,  or  inac- 
cessibility of  a  suitable  area  due  to 
unsuitable  habitat  surrounding  it. 


60 


In  lieu  of  specific  recommenda- 
tions, the  USPS  needs  interim  man- 
agement guidelines  to  protect  the 
salamanders  from  population  de- 
clines. We  suggest  the  following 
steps: 

1.  Salamander  surveys  should 
be  made  on  specific  sale  ar- 
eas as  early  in  the  planning 
process  as  possible.  The 
USPS  could  maintain  a  team 
of  seasonal  employees  for 
such  surveys  and  for  other 
activities  related  to  endan- 
gered species. 

2.  To  the  extent  possible,  inten- 
sive logging  operations  (i.e., 
clearcuts,  seed-tree  cuts,  trac- 
tor logging)  should  not  be 
conducted  in  areas  occupied 
by  salamanders.  Cable  log- 
ging in  winter,  when  the 
ground  is  frozen  and  the 
salamanders  are  under- 
ground, is  probably  the  least 
damaging  activity.  In  com- 
parison, tractor  logging  on 
wet  soils  can  compact  the 
soil  to  such  a  degree  that 
salamanders  cannot  use  it. 

3.  Modifications  of  current 
practices,  such  as  leaving 
slash  where  it  falls  or  leaving 
as  much  canopy  as  possible, 
help  prevent  the  soil  surface 
from  drying  out  and  will 
probably  benefit  salaman- 
ders. 

4.  Because  current  timber  har- 
vest schedules  will  inevitably 
lead  to  younger-aged  stands 
with  few  or  only  small 
downed  logs,  a  mix  of  young 
and  old  logs  should  be  main- 
tained to  ensure  short-term 
and  long-term  habitat  com- 
ponents. Old  logs  provide 
cover  to  Aneides  and  Pletho- 
don,  while  younger  logs  are 
potential  sources  of  cover  in 
future  years. 


Other  studies  provide  some  evi- 
dence for  negative  effects  of  logging 
on  amphibian  populations  (Bennet  et 
al.  1980,  Blymer  and  McGinnes  1977, 
Bury  1983,  Gordon  et  al.  1962,  Her- 
rington  and  Larsen  1985,  Pough  et  al. 
1987,  Ramotnik  1988,  Staub  1986,  and 
Williams  1976)  and  we  suspect  that 
intensive  logging,  slash  removal,  and 
burning  will  reduce  or  eliminate 
populations  of  Plethodon  neomexica- 
nus  and  Aneides  hardii.  Only  intensive 
observations  of  salamander  popula- 
tions throughout  the  logging  cycle 
will  provide  the  information  needed 
to  make  management  recommenda- 
tions. These  studies  are  in  progress, 
but  may  require  years  before  defini- 
tive results  are  available  to  assess  the 
effects  of  logging  on  Plethodon  and 
Aneides. 


Acknowledgments 

We  thank  the  following  U.S.  Porest 
Service  personnel:  Santa  Pe  National 
Porest — R.  Alvarado,  D.  Delorenzo, 
and  M.  Morrison;  Lincoln  National 
Porest — R.  Dancker,  D.  Edwards,  S. 
Lucas,  J.  Peterson,  and  D.  Zaborske; 
and  L.  Pisher,  Regional  Office.  Much 
of  the  funding  was  provided  by  the 
U.S.  Porest  Service  (Southwestern 
Region). 

Pield  personnel  included  M.  J.  Al- 
tenbach,  R.  R.  Beatson,  A.  Bridegam, 
R.  B.  Bury,  C.  Campbell,  S.  Com,  T. 
H.  Pritts,  B.  E.  Smith,  and  M.  C. 
Tremble.  S.  Stefferud  (Endangered 
Species,  U.S.  Pish  &  Wildlife  Service) 
and  C.  Painter  (Endangered  Species 
Program,  New  Mexico  Department 
of  Game  &  Pish)  were  welcome  field 
companions.  S.  Corn  provided  pho- 
tographs. K.  Aubry,  K.  Buhlmann,  S. 
Corn,  C.  K.  Dodd,  and  C.  Painter 
provided  helpful  criticism  of  earlier 
drafts. 


Literature  Cited 

Aubry,  Keith  B.,  Clyde  M.  Senger, 
and  Rodney  L.  Crawford.  1987. 


Discovery  of  Larch  Mountain  sala- 
manders Plethodon  larselli  in  the 
central  Cascade  Range  of  Wash- 
ington. Biological  Conservation 
42:147-152. 

Barbour,  Roger  W.,  James  W.  Har- 
din, James  P.  Schafer,  and  Michael 
J.  Harvey.  1969.  Home  range, 
movements,  and  activity  of  the 
dusky  salamander,  Desmognathus 
fuscus.  Copeia  1969:293-297. 

Bennet,  Stephen  H.,  J.  Whitfield  Gib- 
bons, and  Jill  Glanville.  1980.  Ter- 
restrial activity,  abundance  and 
diversity  of  amphibians  in  differ- 
ently managed  forest  types. 
American  Midland  Naturalist 
103:412-416. 

Blymer,  Michael  J.  and  Burd  S. 
McGinnes.  1977.  Observations  on 
possible  detrimental  effects  of 
clearcutting  on  terrestrial  amphibi- 
ans. Bulletin  of  the  Maryland  Her- 
petological  Society  13:79-83. 

Brodie,  Edmund  D.,  Jr.  1970.  West- 
ern salamanders  of  the  genus  Ple- 
thodon: systematics  and  geo- 
graphic variation.  Herpetologica 
26:468-516. 

Burton,  Barry  W.  1982.  Geologic  evo- 
lution of  the  Jemez  Mountains  and 
their  potential  for  future  volcanic 
activity.  Los  Alamos  National 
Laboratory,  LA-8795-GEOL.  Los 
Alamos,  New  Mexico.  31  p. 

Bury,  R.  Bruce.  1983.  Differences  in 
amphibian  populations  in  logged 
and  old  growth  redwood  forest. 
Northwest  Science  57:167-178. 

Bury,  R.  Bruce  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles,  p. 
416-419.  In  Renewable  resource 
inventories  for  monitoring 
changes  and  trends.  J.  P.  Bell  and 
T.  Atterbury,  eds.  Oregon  State 
University,  Corvallis. 

Campbell,  Howard  W.  and  Stephen 
P.  Christman.  1982.  Pield  tech- 
niques for  herpetofaunal  commu- 
nity analysis,  p.  193-200.  In  Herpe- 
tological  communities.  Norman  J. 
Scott,  ed.  U.S  Pish  &  Wildlife 
Service,  Wildlife  Research  Rep.  13. 

Castetter,  Edward,  P.  1956.  The  vege- 


61 


tation  of  New  Mexico.  New  Mex- 
ico Quarterly  26:256-285. 

Childs,  S.  W.  and  L.  E.  Hint.  1987.  Ef- 
fect of  shadecards,  shelterwoods, 
and  clearcuts  on  temperature  and 
moisture  environments.  Forest 
Ecol.  and  Manage.  18:205-217. 

Corn,  P.  Steven  and  R.  Bruce  Bury.  In 
press.  Sampling  terrestrial  am- 
phibians and  reptiles.  In  Popula- 
tion Monitoring  Techniques  for 
Wildlife  in  Pacific  Northwest  For- 
ests. A.  B.  Carey  and  L.  F.  Ruggi- 
ero,  eds.  USDA  Forest  Service 
General  Technical  Report.  Pacific 
Northwest  Research  Station,  Port- 
land, Oregon. 

DeByle,  Norbert  V.  1981.  Clearcut- 
ting  and  fire  in  the  larch/ Douglas- 
fir  forests  of  western  Montana-a 
multifaceted  research  summary. 
General  Technical  Report  INT-99. 
USDA  Forest  Service,  Intermoun- 
tain  Forest  and  Range  Experiment 
Station.  73  p. 

Franklin,  J.  F.,  K.  Cromack,  Jr.,  W. 
Denison,  A.  McKee,  C.  Maser,  J. 
Sedell,  F.  Swanson,  and  G.  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-fir  forests. 
General  Technical  Report  PNW- 
118.  USDA  Forest  Service,  Port- 
land, Oregon.  48  p. 

French,  Thomas  W.  and  Robert  H. 
Mount.  1978.  Current  status  of  the 
Red  Hills  salamander, 
Phaeognathus  huhrichti  Highton, 
and  factors  affecting  its  distribu- 
tion. Journal  of  the  Alabama 
Academy  of  Science  49:172-179. 

Gordon,  Robert  E.,  James  A.  MacMa- 
hon,  and  David  B.  Wake.  1962. 
Relative  abundance,  microhabitat 
and  behavior  of  some  Southern 
Appalachian  salamanders.  Zool- 
ogica  47:9-14. 

Heatwole,  Harold.  1962.  Environ- 
mental factors  influencing  local 
distribution  and  activity  of  the 
salamander,  Plethodon  cinereus. 
Ecology  43:460-472. 

Herrington,  Robert  E.  and  John  H. 
Larsen.  1985.  Current  status,  habi- 
tat requirements  and  management 
of  the  Larch  Mountain  salamander 


Plethodon  larselli  Bums.  Biological 
Conservation  34:169-179. 

Hubbard,  John  P.,  Marshall  C.  Con- 
way, Howard  Campbell,  Gregory 
Schmitt,  and  Michael  D.  Hatch. 
1979.  Handbook  of  species  endan- 
gered in  New  Mexico.  New  Mex- 
ico Department  of  Game  &  Fish. 

Jaeger,  Robert  G.  1971.  Moisture  as  a 
factor  influencing  the  distributions 
of  two  species  of  terrestrial  sala- 
manders. Oecologia  6:191-207. 

Jaeger,  Robert  G.  1980.  Microhabitats 
of  a  terrestrial  forest  salamander. 
Copeia  1980:265-268. 

Kleeberger,  Steven  R.  and  J.  Kirwin 
Werner.  1982.  Home  range  and 
homing  behavior  of  Plethodon  cin- 
ereus in  northern  Michigan.  Copeia 
1982:409-415. 

Krag,  R.,  K.  Higginbotham,  and  R. 
Rothwell.  1986.  Logging  and  soil 
disturbance  in  southeast  British 
Columbia.  Canadian  Journal  of 
Forest  Research  16:1345-1354. 

Kunkel,  Kenneth  E.  1984.  Tempera- 
ture and  precipitation  summaries 
for  selected  New  Mexico  locations. 
Climate  report.  Office  of  State  Cli- 
matologist.  New  Mexico  Depart- 
ment of  Agriculture,  Las  Cruces, 
New  Mexico. 

Lemmon,  Paul  E.  1956.  A  spherical 
densiometer  for  estimating  forest 
overstory  density.  Forest  Science 
2:314-320. 

Lowe,  Charles  H.,  Jr.  1950.  The  sys- 
tematic status  of  the  salamander 
Plethodon  hardii,  with  a  discussion 
of  biogeographic  problems  in 
Aneides.  Copeia  1950:92-99. 

MacCullough,  Ross  D.  and  J.  R. 
Bider.  1975.  Phenology,  migra- 
tions, circadian  rhythm  and  the  ef- 
fect of  precipitation  on  the  activity 
of  Eurycea  b.  bislineata  in  Quebec. 
Herpetologica  31:433-439. 

Rough,  F.  Harvey,  Ellen  M.  Smith, 
Donald  H.  Rhodes,  and  Andres 
Collazo.  1987.  The  abundance  of 
salamanders  in  forest  stands  with 
different  histories  of  disturbance. 
Forest  Ecol.  and  Manage.  20:1-9. 

Ramotnik,  Cynthia  A.  1986.  Status 
report  Plethodon  neomexicanus 


Stebbins  and  Riemer,  Jemez 
Mountains  salamander.  Submitted 
to  the  U.S.  Fish  &  Wildlife  Service, 
Region  2,  Office  of  Endangered 
Species,  Albuquerque,  NM.  55  p. 

Ramotnik,  Cynthia  A.  1988.  Habitat 
requirements  and  movements  of 
Jemez  Mountains  salamanders, 
Plethodon  neomexicanus.  M.S.  the- 
sis, Colorado  State  University, 
Fort  Collins.  84  p. 

Raphael,  Martin  G.  and  K.  V.  Rosen- 
berg. 1983.  An  integrated  ap- 
proach to  wildlife  inventories  in 
forested  habitats,  p.  219-222.  In 
Renewable  resource  inventories 
for  monitoring  changes  and 
trends.  J.  F.  Bell  and  T.  Atterbury, 
eds.  Oregon  State  University, 
Corvallis. 

Reagan,  Douglas  P.  1972.  Ecology 
and  distribution  of  the  Jemez 
Mountains  salamander,  Plethodon 
neomexicanus.  Copeia  1972:486-492. 

SAS  Institute  Inc.  1982.  SAS  user's 
guide:  statistics,  1982  ed.  Gary, 
North  Carolina:  SAS  Institute  Inc. 
584  p. 

Scott,  Norman  J.  Jr.,  Cynthia  A.  Ra- 
motnik, Marilyn  J.  Altenbach,  and 
Brian  E.  Smith.  1987.  Distribution 
and  ecological  requirements  of 
endemic  salamanders  in  relation 
to  forestry  management.  Summary 
of  1987  activities.  Part  2:  Santa  Fe 
National  Forest.  Final  report  sub- 
mitted to  the  USDA  Forest  Serv- 
ice, Albuquerque,  New  Mexico. 
33  p. 

Shelford,  Victor  E.  1963.  The  ecology 
of  North  America.  University  of  Il- 
linois Press,  Urbana.  610  p. 

Snedecor,  George  W.  and  William  G. 
Cochran.  1967.  Statistical  methods. 
6th  ed.  The  Iowa  State  University 
Press,  Ames,  Iowa.  593  p. 

Staub,  Nancy.  1986.  A  status  survey 
of  the  Sacramento  Mountain 
salamander,  Aneides  hardii,  with  an 
assessment  of  the  impact  of  log- 
ging on  salamander  abundance. 
Final  report  submitted  to  the  U.S. 
Fish  and  Wildlife  Service,  Region 
2,  Endangered  Species  Program, 
Albuquerque,  New  Mexico.  42  p. 


62 


U.S.  Fish  &  Wildlife  Service.  1986. 
Distribution  and  ecological  re- 
quirements of  endemic  salaman- 
ders in  relation  to  forestry  man- 
agement. Part  2:  Santa  Fe  National 
Forest.  Report  submitted  to  the 
U.S.  Forest  Service,  Albuquerque, 
New  Mexico.  40  p. 

Weigmann,  Diana  L.,  Mark  Hakkila, 
Karl  Whitmore,  and  Richard  A. 
Cole.  1980.  Survey  of  Sacramento 
Mountain  salamander  habitat  of 
the  Cloudcroft  and  Mayhill  dis- 
tricts of  the  Lincoln  National  For- 
est. Final  report  submitted  to  the 
U.S.  Forest  Service,  Lincoln  Na- 
tional Forest.  45  p. 

Whitford,  Walter  G.  1968.  Physiologi- 
cal responses  to  temperature  and 
desiccation  in  the  endemic  New 
Mexico  plethodontids,  Plethodon 
neomexicanus  and  Aneides  hardii. 
Copeia  1968:247-251. 

Whitford,  Walter  G.  and  John 
Ludwig.  1975.  The  biota  of  the 
Baca  geothermal  site.  Report  for 
Union  Oil  Co. -Geothermal  Divi- 
sion. 140  p. 

Williams,  Byron  K.  1983.  Some  obser- 
vations on  the  use  of  discriminant 
analysis  in  ecology.  Ecology 
64:1283-1291. 

Williams,  Stephen  R.  1972.  Repro- 
duction and  ecology  of  the  Jemez 
Mountains  salamander,  Plethodon 
neomexicanus.  M.  S.  thesis,  Univ.  of 
New  Mexico,  Albuquerque.  98  p. 

Williams,  Stephen  R.  1976.  Compara- 
tive ecology  and  reproduction  of 
the  endemic  New  Mexico  pletho- 
dontid  salamanders,  Plethodon  ne- 
omexicanus and  Aneides  hardii.  Ph. 
D.  thesis.  University  of  New  Mex- 
ico, Albuquerque.  152  p. 

Wyman,  Richard  L.  1988.  Soil  acidity 
and  moisture  and  the  distribution 
of  amphibians  in  five  forests  of 
southcentral  New  York.  Copeia 
1988:394-399. 

Wyman,  Richard  L.  and  Dianne  S. 
Hawksley-Lescault.  1987.  Soil 
acidity  affects  distribution,  behav- 
ior, and  physiology  of  the 
salamander  Plethodon  cinereus. 
Ecology  68:1819-1827. 


utilization  Of  Abandoned 
Mine  Drifts  and  Fracture 
Caves  By  Bats  and 
Salamanders:  Unique 
Subterranean  Habitat  In  The 
Ouachita  Mountains^ 


Abstract.— Twenty-seven  abandoned  mine  drifts 
and  four  fracture  caves  constitute  one  of  the  most 
unique  habitats  in  and  adjacent  to  the  Ouachita 
National  Forest,  an  area  devoid  of  solutional  caves. 
Six  species  of  salamanders  and  nine  species  of  bats 
were  found  to  utilize  these  areas. 


David  A.  Saugey,^  Gary  A.  Heidt,^  Darrell  R. 
Heathi^ 


Caves  and  mines  play  an  important 
role  in  the  ecology  of  many  species, 
serving  as  permanent  or  temporary 
habitats.  Culver  (1986)  stated,  "the 
variety  of  species  that  depends  on 
caves  during  some  critical  time  in 
their  life  cycle,  such  as  hibernation  in 
bats,  is  impressive  and  usually 
underestimated."  To  this  statement, 
we  add  mines. 

Bear  Den  Caves  are  located  in 
Winding  Stair  Mountain,  LeFlore 
County,  in  southeastern  Oklahoma. 
These  four  caves  occur  in  an  outcrop 
belt  of  a  massive  sandstone  unit  and 
were  formed  by  a  number  of  factors, 
the  most  important  being  gravita- 
tional sliding  and  slumpage  of  sand- 
stone. These  four  caves  have  more 
than  365  meters  of  mapped  passage- 
way and  represent  the  only  known 
caves  in  the  Ouachita  National  Forest 
(Puckette  1974-75). 

Additional  subterranean  habitat 
was  formed  from  1870  to  1890,  when 
the  area  extending  west  from  Hot 
Springs  to  Mena,  Arkansas  was  the 
scene  of  a  gold,  lead,  silver  and  zinc 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff ,  AZ,  July  19-2h  1988). 

^David  A.  Saugey  is  a  Wildlife  Biologist, 
U.S.  Forest  Sen/ice,  Ouachita  National  For- 
est, Hot  Springs,  AR.  71902. 

^GaryA.  Heidt  is  Professor  of  Biology, 
University  of  Arkansas  at  Little  Rock,  AR. 
72204. 

"Darrell  R.  Heath  is  an  Undergraduate 
Student,  University  of  Arkansas  at  Little 
Rock,  AR.  72204. 


rush.  During  the  period  of  greatest 
activity,  1885  to  1888,  over  a  dozen 
gold  mines  were  in  operation,  rang- 
ing from  shallow  test  holes  to  exten- 
sive linear  and  L-shaped  drifts  ex- 
tending up  to  150  meters  into  the 
surrounding  mountains  (Harrington 
1986,  Hudgins  1971,  U.S.  Army 
Corps  of  Engineers  1980).  The  "gold 
and  silver  boom"  effectively  ended 
with  the  issuance  of  a  report  which 
in  effect  stated  there  were  no  pre- 
cious metals  in  paying  quantities  to 
be  found  in  the  area  (Branner  1888). 
Soon  thereafter,  many  mines  were 
abandoned  as  prospectors  moved 
West  (Harrington  1986,  Hudgins 
1971).  Through  the  years,  other  min- 
erals, such  as  manganese  and  mer- 
cury, have  been  mined  from  the  Ou- 
achitas  resulting  in  the  excavation  of 
numerous  additional  drifts;  but  for  a 
variety  of  reasons,  most  have  been 
abandoned  (Clardy  and  Bush  1976, 
Stone  and  Bush  1984).  The  legacy  of 
these  mining  activities  has  not  been 
riches  and  new-found  wealth,  but  the 
creation  of  unusual  and  unique  wild- 
life habitat. 

The  objectives  of  this  study  were 
to  review,  compile,  and  consolidate 
existing  literature  concerning  utiliza- 
tion of  caves  and  mine  drifts  by  bats 
and  salamanders  in  the  Ouachita 
Mountains.  In  addition,  we  provide 
new  data  and  propose  recommenda- 
tions concerning  management  of 
caves  and  mines  in  the  Ouachita  Na- 
tional Forest  and  on  other  public  and 
private  lands. 


METHODS 

During  the  past  six  years,  27  aban- 
doned mines  in  Garland  (8), 
Montgomery  (3),  Pike  (4)  and  Polk 
(12)  counties,  Arkansas  (fig.  1)  were 
located  and  visited  a  minimum  of 
eight  times  (at  least  once  each  sea- 
son). In  several  cases,  where  endemic 
or  Category  II  (U.S.  Federal  Register 
1985)  species  occurred  or  breeding 
populations  were  found,  mines  were 
visited  much  more  often.  Mist  net- 
ting of  entrances  for  bats  was  con- 
ducted in  spring,  summer,  and  fall. 
Bear  Den  Caves  came  to  our  atten- 
tion during  1987  and  were  visited 
several  times.  Collections  were  mini- 
mal (mines  only)  and  voucher  speci- 


OKLA 


Figure  1  .—Location  of  Ouachita  National 
Forest  (backslastied  area)  and  study  area 
(crosshiatchied  area). 


64 


mens  are  located  in  the  Vertebrate 
Collections  at  the  University  of  Ar- 
kansas at  Little  Rock  and  Arkansas 
State  University. 

Following  McDaniel  and  Smith 
(1976),  we  include  the  probable  eco- 
logical position  of  the  species  in  the 
cave  and  mine  environments.  This  is 
followed  by  comments  concerning 
the  status  or  life  history  of  each  spe- 
cies. Following  Barr  (1963)  and 
McDaniel  and  Smith  (1976)  the  terms 
"troglophile"  (commonly  found  in 
caves),  "trogloxene"  (may  be  com- 
mon in  caves  but  must  leave  to  com- 
plete their  life  history),  and  "acciden- 
tal" (unable  to  survive  long  in  the 
cave  environment)  have  been  em- 
ployed in  the  species  accounts. 

RESULTS 

Nine  species  of  bats  and  six  species 
of  salamanders  were  found  to  utilize 
caves  and  abandoned  mine  drifts 
during  some  portion  of  their  annual 
cycles. 

Annotated  List  of  Bats  and 
Salamanders  Utilizing  Caves  and 
Abandoned  Mine  Drifts 

CLASS  AMPHIBIA 

Order  Urodela 

Family  Plethodontidae 

Desmognafhus  brimleyorum 
(Stefneger).  Troglophile. 

Means  (1974)  stated  the  Ouachita 
dusky  salamander  was  confined  to 
rocky,  gravelly,  streams  in  the  Ou- 
achita Mountains.  Rock  falls  along 
the  upper  portions  of  streams  repre- 
sented particularly  good  adult  habi- 
tat. This  species  was  most  abundant 
where  water  percolated  through 
rocky  substrate  in  streambeds  and 
along  stream  sides.  Description  of 
egg  clutch  characteristics  and 
stream/streamside  deposition  were 


given  by  Means  (1974)  and  Trauth 
(1988)  provided  descriptions  of 
deposition  sites  in  seepage  areas  dur- 
ing the  severe  summer  drought  in 
1980.  Heath  et  al.  (1986)  reported  the 
occurrence  of  this  endemic  salaman- 
der in  four  drifts,  with  egg  clutches 
deposited  on  the  underside  of  rocks 
in  one  mine  and  the  presence  of  lar- 
vae in  two  others.  In  those  mines 
with  larvae,  pools  contained  abun- 
dant leaf  litter  and  isopods.  On  one 
occasion,  larvae  were  observed  feed- 
ing on  isopods.  Since  these  observa- 
tions were  made,  numerous  addi- 
tional visits  to  these  four  mines  re- 
vealed the  presence  of  D^smognathus 
when  epigean  conditions  would  be 
considered  ideal.  The  pools  within 
these  and  other  drifts  are  the  result 
of  seepage  through  walls  which,  in 
some  instances,  provided  sufficient 
volumes  of  water  to  have  small 
streams  flowing  from  their  entrances. 
However,  unlike  the  preferred, 
gravel-bottomed  stream  habitat, 
pools  typically  exhibited  silted  sub- 
strates with  very  little  rubble  and 
few  rocks  large  enough  for  egg  at- 
tachment. 


Eurycea  multiplicata  (Cope). 
Trogloptiile. 

The  many-ribbed  salamander  is  pri- 
marily an  aquatic  species  endemic  to 
the  Interior  Highland  region  and  ad- 
jacent areas  that  contain  suitable 
habitat.  It  may  be  found  under 
stones,  logs,  and  other  debris  in 
clear,  rock  or  gravel-bottomed 
streams  (Bishop  1943,  Ireland  1971, 
Reagan  1974).  It  inhabits  essentially 
the  same  habitat  as  Desmognathus 
brimleyorum  (Strecker  1908).  Hurter 
and  Strecker  (1909)  noted 
Desmognathus  eating  Eurycea  indi- 
viduals with  which  they  were  con- 
fined. Heath  et  al.  (1986)  reported 
both  larvae  and  adults  in  two  mines 
and  in  one,  larvae  shared  the  same 
pools  with  Desmognathus  larvae.  Both 
mines  contained  shallow  streams 
with  a  gravel  substrate.  One  addi- 


tional mine  contained  larvae  of  this 
species.  A  seepage  stream  in  this 
mine  was  approximately  five  centi- 
meters wide,  one  centimeter  deep, 
and  extended  a  distance  of  sixty 
centimeters  before  dropping  into  a 
large  pool  at  the  entrance.  The  pool 
connected  directly  to  an  epigean 
stream. 


Plethodon  caddoensis  Pope  and 
Pope.  Trogloptiile. 

Large  aggregations  of  the  endemic 
Caddo  Mountain  salamander  using 
drifts  as  refugia  to  escape  heat  and 
dryness  during  summer  and  fall 
were  first  reported  by  Saugey  et  al. 
(1985).  Over  100  individuals  were 
discovered  in  each  of  two  drifts, 
from  June  through  September  1983. 
Subsequent  visits  to  these  and  other 
drifts  revealed  limited  use  of  three 
additional  drifts  and  use  of  one  of 
the  original  aggregation  sites  for  egg 
deposition  and  breeding  (Heath  et  al. 
1986).  Since  these  observations  were 
made,  summer  aggregations  of  this 
salamander  have  numbered  as  high 
as  383  individuals  and  additional  egg 
clutches  have  been  observed  and 
monitored.  Known  only  from  the 
Novaculite  Uplift  area  of  the  Ou- 
achita Mountains  in  Howard, 
Montgomery,  and  Polk  counties  in 
Arkansas  (Blair  and  Lindsey  1965, 
Robison  and  Smith  1982),  this  sala- 
mander and  its  habitat  are  of  special 
concern  to  the  Arkansas  Natural 
Heritage  Commission  (ANHC) 
(Smith  1984).  In  1985,  the  U.S.  Fish 
and  Wildlife  Service  (USFWS)  desig- 
nated it  a  Category  II  species.  In 
1986,  the  U.S.  Forest  Service  (Ou- 
achita National  Forest)  began  infor- 
mal consultation  with  the  USFWS 
(Jackson,  Mississippi,  Endangered 
Species  Field  Station)  and  requested 
field  assistance  from  the  ANHC  con- 
cerning preservation  of  critical  mine 
aggregation  sites  and  protection  of 
their  vulnerable  populations.  Place- 
ment of  a  gate  at  one  sensitive  site  is 
planned  in  1988  (fig.  2). 


65 


Plethodon  glutinosus  glutinosus 
(Green).  Troglophile. 

The  slimy  salamander,  a  woodland 
species,  is  widely  distributed,  ex- 
ploiting virtually  every  available  ter- 
restrial habitat.  This  species  is  com- 
monly found  under  rocks,  in  and 
under  well  rotted  logs  and  stumps, 
and  buried  deep  in  moist  layers  of 
leaf  litter.  During  hotter  and  drier 
portions  of  the  year,  they  usually  re- 
treat deeper  into  the  substrate.  Al- 
though primarily  epigean,  this  sala- 
mander has  been  reported  to  use 
caves  for  aggregation  sites,  egg  depo- 
sition and  brooding,  and  escape  from 
inhospitable  surface  environmental 
conditions  (Barnett  1970,  Noble  and 
Marshall  1929).  Heath  et  al.  (1986) 
reported  this  salamander  from  five 
mines;  two  contained  breeding  popu- 
lations and  brooding  behavior  has 
been  observed  several  times.  Subse- 
quent observations  have  confirmed 
another  of  the  five  mines  as  an  egg 
deposition  and  brooding  site.  One  of 
the  mines  reported  with  a  breeding 
population  (Heath  et  al.  1986)  is  the 
site  of  an  annual  aggregation  of  slimy 
salamanders  exceeding  600  individu- 
als. A  gate  (fig.  2)  has  been  con- 
structed by  the  U.S.  Army  Corps  of 
Engineers  to  protect  this  population. 
Continuing  studies  to  determine  the 
effect  of  gating  will  allow  compari- 
son of  pre-  and  post-gating  data. 

Plethodon  ouachitae  Dunn  and 
Heinze.  Troglophile. 

Endemic  to  the  Ouachita  Mountains 
of  Arkansas  and  Oklahoma,  the  Rich 
Mountain  salamander  may  be  found 
living  beneath  rotting  logs  and 
stumps.  However,  it  lives  primarily 
under  pieces  of  sandstone  on  heavily 
overgrown  talus  north  slopes  (Black 
1974,  Dunn  and  Heinze  1933,  Pope 
and  Pope  1951,  Sievert  1986).  Reagan 
(1974)  listed  this  species  as  "endan- 
gered and  vulnerable"  in  Arkansas. 
Ashton  (1976)  and  Black  (1980)  both 
considered  this  salamander  "threat- 


ened" in  Oklahoma.  Sievert  (1986) 
proposed  it  as  a  species  of  "special 
concern,"  conditional  on  his  recom- 
mendations concerning  silvicultural 
practices  on  National  Forest  lands. 
Black  (1974)  reported  this  salaman- 
der in  Bear  Den  Caves  where  they 
were  found  throughout,  but  most 
commonly  within  the  first  19  meters 
or  twilight  zone.  A  small  juvenile 
with  a  snout-vent  length  (SVL)  of  <  7 
mm  was  found  in  an  entrance  and 
the  presence  of  numerous  juveniles 
with  SVLs  of  >  30mm  may  indicate 
egg  deposition  and  brooding  activi- 
ties. One  of  the  authors  (DAS)  visited 
these  caves  in  December,  1987  and 
observed  one  adult  Rich  Mountain 
salamander  near  the  entrance  of  one 
cave.  An  additional  visit  in  June  1988 
resulted  in  the  observation  of  30+ 
salamanders  of  various  size  classes. 
Considerable  human  refuse  and  a 
well  worn  path  indicated  substantial 
numbers  of  visitors.  Considering  the 
uniqueness  of  this  area  and  the  Cate- 
gory II  status  of  this  salamander, 
steps  are  being  taken  to  exclude  ex- 
cessive visitation  and  protect  this 
population  from  vandalism  and 
overcollection.  These  caves  are  util- 
ized by  the  small-footed  bat,  Myotis 
leibii,  (Caire  1985)  also  a  Category  II 
species. 


Plethodon  serratus  Grobman. 
Troglophile. 

The  endemic  Ouachita  Red-backed 
salamander  is  commonly  found  be- 
neath rocks,  logs,  and  in  leaf  litter  at 
all  elevations  throughout  the  Ou- 
achita Mountains.  This  species  has 
been  observed  in  one  mine  on  two 
separate  occasions.  In  both  cases,  it 
has  been  in  association  with  large 
aggregations  of  the  Caddo  Mountain 
salamander  during  extremely  dry 
epigean  conditions.  Reagan  (1974) 
frequently  found  this  species  in  asso- 
ciation with  the  Caddo  Mountain 
and  Rich  Mountain  salamanders. 


CLASS  MAMMALIA 
Order  Chiroptera 

Family  Vespertilionidae 

Myotis  austroriparius  (Rhoads). 
Trogloxene. 

The  first  Arkansas  specimens  of  the 
southeastern  bat  were  collected  from 
one  of  several  drifts  located  12  miles 
northwest  of  Hot  Springs,  Garland 
County,  Arkansas  (Davis  et  al.  1955). 


66 


At  the  time  of  collection  (November 
1952)  and  during  a  subsequent  visit, 
this  species  was  found  in  association 
with  the  little  brown  bat,  Myotis  luci- 
fugus,  and  Keen's  bat,  Myotis  keenii. 
This  particular  drift  was  inundated 
by  the  filling  of  Lake  Ouachita  in 
1955  and,  since  that  time,  no  addi- 
tional specimens  have  been  observed 
in  nearby  drifts.  The  second  occur- 
rence of  this  species  in  the  Ouachita 
Mountain  area  was  from  abandoned 
Cinnabar  mines  located  on  an  penin- 
sula in  Lake  Greeson,  Pike  County, 
Arkansas  (Heath  et  al.  1986).  During 
a  winter  visit  (January  1984)  over  150 
individuals  of  both  red  and  gray 
color  phases  were  observed  in  deep 
torpor.  A  subsequent  early  spring 
visit  (March  1986),  revealed  15  indi- 
viduals. During  December,  1986, 
only  a  few  scattered  individuals  were 
found.  According  to  personnel  famil- 
iar with  the  drift,  considerable  hu- 
man visitation  and  disturbance  may 
have  been  the  cause  of  sharp  decline 
in  use  of  this  excavation.  Mumford 
and  Whitaker  (1982)  suggested  the 
southeastern  bat  does  not  tolerate 
disturbance  and  is  likely  to  change  its 
roosting  and  hibernation  sites  quite 
readily.  Caire  (1985)  did  not  report 
this  species,  but  records  exist  for  the 
Little  River  drainage  in  southeastern 
Oklahoma  (Glass  and  Ward  1959). 
The  southeastern  bat  is  listed  as  a 
Category  II  species  in  the  U.S.  Fed- 
eral Register  (1985). 

Myotis  keenii  (Merriam). 
Trogloxene. 

Utilization  of  caves  and  mines  by 
Keen's  bat  has  been  well  documented 
(Barbour  and  Davis  1969,  Heath  et  al. 
1986,  McDaniel  and  Gardner  1977). 
Sealander  and  Young  (1955)  first  re- 
ported the  occurrence  of  Keen's  bat 
from  the  Ouachita  Mountain  area 
when  three  specimens  were  collected 
from  the  drift  located  12  miles  north- 
west of  Hot  Springs.  Caire  (1985) 
mist-netted  a  number  of  specimens  at 
Bear  Den  Caves;  the  majority  were 


males  with  a  few  postlactating  fe- 
males. Heath  et  al.  (1986)  found  this 
bat  in  12  drifts.  The  largest  hibernat- 
ing aggregation  consisted  of  12  bats, 
including  both  males  and  females. 
Normally,  from  one  to  three  indi- 
viduals (usually  males)  were  found 
hibernating  in  small  cracks  and  crev- 
ices near  entrances.  On  occasion,  two 
have  been  found  together  in  drill 
holes  in  ceilings  and  walls  and,  less 
frequently,  individuals  were  ob- 
served hanging  in  the  open.  The  larg- 
est non-hibernating  cluster  was  57 
females  found  in  the  spring  of  1985. 
Three  were  collected  and  found  to  be 
pregnant  (drifts  were  not  used  as 
maternity  roosts).  Although  utilized 
more  frequently  during  winter 
months,  these  drifts  contained  from 
one  to  several  Keen's  bats  through- 
out most  of  the  year. 

Myotis  ieibii  (Audubon  and 
Bachman).  Trogloxene. 

The  small-footed  bat  is  very  common 
and  widespread  in  the  western 
United  States  where  it  readily  uses 
caves  and  mines  for  hibernation.  In 
the  eastern  United  States  it  is  consid- 
ered to  be  rare  (Barbour  and  Davis 
1969,  Smith  1984).  Caire  (1985)  re- 
ported mist-netting  four  males,  three 
adults  and  one  subadult,  at  Bear  Den 
Caves.  Specimens  collected  in  Sep- 
tember had  descended  testes.  Heath 
et  al.  (1986)  did  not  record  this  bat 
from  drifts  in  Arkansas.  According 
to  Barbour  and  Davis  (1969),  the  only 
known  winter  habitats  for  this  spe- 
cies are  caves  and  mines.  Preferred 
hibernation  sites  are  near  entrances 
where  temperatures  drop  below 
freezing  and  humidity  is  relatively 
low.  Abandoned  drifts  in  the  Ou- 
achitas  generally  have  one,  small, 
partially  collapsed  entrance  which 
ensures  relatively  warm  interiors  (18 
C)  with  high  humidities,  which  is  un- 
suitable hibernating  habitat.  Mist- 
netting  of  creeks  and  drift  entrances 
and  subsequent  winter  visits  to  drifts 
have  been  unsuccessful  in  locating 


this  bat.  Caire  (1985)  indicated  this 
species  is  probably  restricted  to  cave 
areas.  Thus,  the  few  caves  in  south- 
eastern Oklahoma  are  critical  to  the 
species  survival  and  are  in  need  of 
protection.  The  small-footed  bat  is  a 
Category  II  species  (U.S.  Federal 
Register  1985). 

Myotis  lucifugus  (LeConte). 
Trogloxene. 

The  little  brown  bat  appears  to  be 
extremely  rare  in  the  Ouachita 
Mountains.  It  had  been  reported 
from  one  drift  by  Sealander  and 
Young  (1955),  but  an  additional 
specimen  was  reported  by  Heath  et 
al.  (1986)  from  a  drift  in  Arkansas.  In 
Oklahoma,  the  little  brown  bat  has 
been  collected  only  from  Beavers 
Bend  State  Park  in  the  southeastern 
part  of  the  state  (Glass  and  Ward 
1959). 


Myotis  sodalis  Miller  and  Allen. 
Trogloxene. 

Sealander  and  Young  (1955)  reported 
a  misidentified  Indiana  bat  from  a 
now  inundated  drift  northwest  of 
Hot  Springs.  There  is  a  confirmed 
record  of  the  species  from  a  south- 
eastern Oklahoma  cave  (Glass  and 
Ward  1959).  Neither  Caire  (1985)  nor 
Heath  et  al.  (1986)  found  this  species 
inhabiting  mines  or  caves  in  the  Ou- 
achitas. 


Pipistrellus  subflavus  (F.  Cuvier). 
Trogloxene. 

The  eastern  pipistrelle  was  described 
as  fairly  abundant  in  southeastern 
Oklahoma  (Caire  1985)  and  as  wide- 
spread and  abundant  in  the  Arkansas 
portion  of  the  Ouachitas  (Heath  et  al. 
1986).  Barbour  and  Davis  (1969)  de- 
scribed it  as  the  most  abundant  bat 
over  much  of  the  eastern  United 
States.  Caves  and  mines  appear  to  be 
important  habitats  for  winter  hiber- 


67 


nation  sites  and  for  summer  night 
roosts  (Barbour  and  Davis  1969, 
McDaniel  and  Gardner  1977).  Caire 
(1985)  reported  capturing  many  indi- 
viduals at  Bear  Den  Caves  during 
summer  months.  Heath  et  al.  (1986) 
reported  this  species  had  been  ob- 
served in  every  drift  at  all  times  of 
the  year  and  that,  over  a  three  year 
period,  one  drift  had  an  annual 
population  of  between  600-800  hiber- 
nating individuals.  Visits  to  this 
hibernaculum  over  the  past  three 
years  have  revealed  the  number  of 
individuals  to  be  fairly  constant.  Pre- 
liminary observations  of  a  drift  that 
has  had  a  gate  in  its  entrance  for  two 
years  have  indicated  an  increase  in 
numbers  of  hibernating  pipistrelles. 

Epfesicus  fuscus  (Palisot  de 
Beauvois).  Trogloxene. 

Heath  et  al.  (1986)  reported  that,  al- 
though common  in  the  Ouachita 
Mountain  area,  the  big  brown  bat 
was  rarely  found  hibernating  in 
drifts.  The  four  drifts  used  during 
hibernation  had  larger,  less  re- 
stricted, openings  that  created  a  vari- 
able temperature  zone.  Rarely  were 
more  than  two  or  three  observed  in 
any  drift.  This  species  characteristi- 
cally chose  hibernating  sites  near  the 
entrance  where  temperature  and 
humidity  levels  were  lower.  Similar 
hibernating  behavior  has  been  docu- 
mented in  other  caves  and  mines 
(Barbour  and  Davis  1969,  Lacki  and 
Bookhout  1983).  Caire  (1985)  re- 
ported this  species  from  Bear  Den 
Caves. 


Lasionycteris  noctivagans 
(LeConte).  Trogloxene. 

Typically  considered  a  tree  bat,  the 
silver-haired  bat  has  been  found  in 
numerous  caves  and  mines  (Barbour 
and  Davis  1969,  Saugey  et  al.  1978, 
Whitaker  and  Winter  1977).  Heath  et, 
al.  (1986)  discovered  a  single  speci- 
men hibernating  in  a  breeze  way  of  a 


drift  near  Lake  Greeson;  the  ambient 
temperature  was  2  C. 

The  three  following  species  of  La- 
siurus,  normally  considered  tree  bats, 
have  been  captured  during  swarm- 
ing activities  at  the  entrances  of,  but 
not  inside  drifts  (Heath  et  al.  1983, 
1986).  Similar  behavior  in  tree  bats 
has  been  observed  at  caves  (Barbour 
and  Davis  1969,  Harvey  et  al.  1981). 

Lasiurus  borealis  (Muller). 
Accidental. 

The  red  bat  was  captured  at  the  en- 
trances of  three  drifts.  Caire  (1985) 
reported  capturing  this  species  at 
Bear  Den  Caves.  Red  bats  were  re- 
ported from  inside  two  Ozark  caves 
by  McDaniel  and  Gardner  (1977). 
Saugey  et  al.  (1978)  discovered  the 
remains  of  140  red  bats  in  one  Ozark 
cave. 


Lasiurus  seminolus  (Rhoads). 
Accidental. 

Heath  et  al.  (1983)  reported  the  cap- 
ture of  a  female  Seminole  bat  at  the 
entrance  to  a  drift  in  Polk  County, 
Arkansas,  during  September. 

Lasiurus  cinereus  (Palisot  de 
Beauvois).  Accidental. 

Previously  unreported,  a  male  hoary 
bat  was  captured  simultaneously 
with  the  above  mentioned  Seminole 
bat.  The  occurrence  of  this  species  in 
mines  and  caves  has  been  well  docu- 
mented (Barbour  and  Davis  1969, 
Saugey  et  al.  1978). 

DISCUSSION 

Caves  are  common  and  widely  dis- 
tributed in  the  United  States.  Caves 
are  known  in  every  state  and,  in 
some,  are  very  common.  It  has  been 
found  that  most  caves  contain  a  bio- 
logically interesting  fauna  (Culver 


1986).  Where  caves  are  scarse,  aban- 
doned mineshaf  ts  occasionally  pro- 
vide the  same  specialized  habitat  as 
do  natural  caves  (Barbour  and  Davis 
1969). 

Abandoned  mine  drifts  and  frac- 
ture caves  represent  important  habi- 
tat features  in  the  Ouachita  Moun- 
tains. Six  species  of  salamanders  and 
nine  species  of  bats  utilize  these 
structures  for  some  purpose.  In  addi- 
tion, four  of  the  six  salamanders  are 
endemic  to  the  Ouachita  Mountains, 
and  a  fifth  is  endemic  to  the  Interior 
Highlands.  Two  of  these 
salamanders,  Plethodon  caddoensis 
and  P.  ouachitae,  are  Category  II  spe- 
cies. For  all  of  these  salamanders, 
caves  and  mines  may  only  represent 
larger  versions  of  existing  subterra- 
nean microhabitats,  complimenting 
existing  situations  and  not  replacing 
them.  However,  caves  and  mines  do 
provide  ''natural  laboratories''  where 
insights  into  life  histories  and  species 
interactions,  otherwise  unobservable, 
may  be  studied  with  the  knowledge 
gained  applied  to  management  of 
surface  populations. 

Six  of  the  nine  species  of  bats 
regularly  frequent  caves  or  mines 
during  some  portion  of  their  annual 
cycles  and  two  of  these  are  listed  as 
Category  II  species  {Myotis  austrori- 
parius  and  M.  leibii).  Mines  provide  a 
key  habitat  component  for  bats 
where  natural  subterranean  hiber- 
nacula  are  scarce.  Hibernacula  can  be 
viewed  as  islands  of  different  sizes 
and  complexities  in  an  ocean  of  habi- 
tat inhospitable  for  hibernation 
(Gates  et  al.  1984).  Most  caves  and 
mines  in  the  Ouachitas  are  small  and 
marginal  as  hibernacula  when  com- 
pared with  extensive  and  complex 
cave  systems  of  other  regions.  How- 
ever, minor  hibernacula  may  become 
major  ones  (depending  on  their  size, 
configuration,  and  microclimate),  if 
the  latter  are  destroyed.  Further,  they 
may  function  to  promote  range  ex- 
pansions (Gates  et  al.  1984).  In  addi- 
tion, small  populations  become  in- 
creasingly important  in  species  man- 
agement when  large  populations  are 


68 


continually  threatened  (Humphrey 
1978). 

Fifty- three  vertebrate  taxa  use 
Ozark  caves  (McDaniel  and  Gardner 
1977).  Heath  et  al.  (1986)  reported  the 
occurrence  of  27  vertebrate  taxa  util- 
izing abandoned  mine  drifts  in  the 
Ouachita  Mountains.  Caire  (1985) 
and  Black  (1974)  reported  two  spe- 
cies from  Bear  Den  Caves.  We  report 
two  additional  species  from  aban- 
doned mines  (Lasiurus  cinereus  and 
Plethodon  serratus).  Of  the  31  re- 
corded species  that  use  caves  and 
mines  in  the  Ouachita  Mountains,  22 
are  common  to  both  the  Ouachitas 
and  Ozarks. 

These  data  further  support  Maser 
et  al  (1979)  when  they  stated, 
"Unique  habitats  occupy  a  very 
small  percent  of  the  total  forest  land 
base,  yet  they  are  disproportionately 
important  as  wildlife  habitats."  From 
our  measurement,  the  total  area  of  all 
known  and  inventoried  caves  and 
drifts  in  the  Ouachita  Mountains  is 
approximately  one  acre  in  a  forest 
with  nearly  1 .6  million  surface  acres. 
For  these  reasons,  resource  managers 
should  not  overlook  opportunities  to 
protect  and  conserve  what  may  ap- 
pear to  be  marginal  sites,  especially 
in  areas  where  these  unique  habitats 
may  be  a  limiting  factor. 

MANAGEMENT 
RECOMMENDATIONS 

While  the  National  Forest  Manage- 
ment Act  (1976)  and  Endangered 
Species  Act  (1973)  specify  objectives 
and  set  policy,  the  Forest  Service 
Manual  provides  guidance  and  di- 
rection to  realize  these  objectives  re- 
lating to  species  of  special  concern 
and  their  habitats.  These  documents 
mandate  consideration  of  these 
unique  and  valuable  resources  in  all 
phases  of  planning  and  project  im- 
plementation. 

Nieland  and  Thornton  (1985),  Nie- 
land  (1985),  Hathom  and  Thornton 
(1986),  and  Chaney  (1984)  provide 
additional  information,  guidance  and 


considerations  concerning  manage- 
ment, inventory  and  evaluation  of 
caves.  Caire  (1985)  made  recommen- 
dations about  habitat  management 
for  bats,  including  Bear  Den  Caves  in 
southeastern  Oklahoma,  and  Sievert 
(1986)  proposed  guidelines  for  pres- 
ervation of  habitat  for  the  endemic 
Rich  Mountain  salamander. 

Because  management  of  cave  re- 
sources are  adequately  addressed  in 
these  references,  the  following  rec- 
ommendations address  issues  con- 
cerning needed  management  of  aban- 
doned mine  drifts  whose  importance 
to  bats  and  other  vertebrates  has 
been  demonstrated  by  Heath  et  al. 
(1986),  Lacki  and  Bookhout  (1983), 
Saugey  et  al.  (1985),  Whitaker  and 
Winter  (1977)  and  this  study. 

In  line  with  these  studies,  we  rec- 
ommend the  following  actions  be 
taken  on  National  Forests,  other  pub- 
lic lands,  and  private  lands: 

1.  Address  abandoned  mine 
drifts  and  shafts  as  "unique 
subterranean  habitat"  in  the 
Cave  Management  section  of 
the  Forest  Service  Manual. 
Most  of  the  language  in  this 
chapter  is  directly  applicable 
to  these  excavations. 

2.  Incorporate  management 
prescriptions  for  abandoned 
mine  drifts  into  Forest  Land 
Management  Plans  and  other 
resource  management  plan- 
ning documents,  where  ap- 
plicable. 

3.  Develop  specific  supple- 
ments, for  individual  Na- 
tional Forests,  to  the  Forest 
Service  Manual  concerning 
the  inventory,  evaluation, 
and  management  of  these 
excavations. 

4.  Prepare  a  chapter  in  the  Ou- 
achita National  Forest  Wild- 
life Handbook  providing  di- 
rection and  guidance  con- 
cerning management  of 


abandoned  mine  drifts  and 
coordination  with  other  re- 
sources. 

5.  Use  full  seasonal  or  partial 
closures  to  protect  species  of 
special  concern  during  criti- 
cal periods  of  the  year. 

6.  Acquire  lands  within  agency 
administrative  authority  that 
contain  caves  and  aban- 
doned mine  drifts. 

7.  Prohibit  extraction  of  miner- 
als and  other  materials  from 
abandoned  mine  drifts. 

8.  Identify  and  designate  aban- 
doned mine  drifts,  caves, 
and  associated  above  ground 
habitat  as  "key  areas"  for 
wildlife  during  the  silvicultu- 
ral  prescription  process. 

9.  Set  aside  and  preserve  travel 
corridors  to  prevent  isolation 
and  loss  of  use  by  terrestrial 
vertebrates. 

10.  Establish  monitoring  activi- 
ties to  assess  changes  in  the 
drift  environment  and  asso- 
ciated wildlife  utilization. 

11.  Continue  inventory  of  spe- 
cies utilizing  drifts  and  de- 
termine how  and  what  they 
are  using  them  for. 

12.  Cooperate,  consult,  and  coor- 
dinate with  state  and  federal 
resource  management  agen- 
cies, universities  and  col- 
leges, public  and  private  con- 
servation organizations,  and 
other  interested  publics  to 
promote  conservation,  edu- 
cation, and  research. 

"Ultimately,  the  survival  of  most 
animal  species  depends  more  on 
habitat  protection  than  on  direct 
shielding  of  the  creatures  them- 
selves" (Smith  1984). 


69 


ACKNOWLEDGMENTS 

We  thank  District  Rangers  John  M. 
Archer  and  Rex  B.  Mann,  Resource 
Assistant  Clifford  F.  Hunt,  and  Wild- 
life Staff  Officer  Dr.  David  F.  Urb- 
ston,  all  of  the  Ouachita  National 
Forest,  for  their  support  and  encour- 
agement during  this  study.  Special 
appreciation  is  extended  to  Clark 
Efaw,  Belinda  Jonak,  Stan  Neal,  Di- 
anne  Saugey,  and  Derrick  Sugg  for 
valuable  assistance  in  the  field.  Le- 
onard Aleshire  and  David  Heath 
were  most  helpful  in  locating  aban- 
doned mines  in  the  Polk  County 
area.  The  Arkansas  Geological  Com- 
mission provided  useful  information 
concerning  the  location  of  mines. 
This  study  was  supported,  in  part,  by 
the  U.S.  Forest  Service  (Ouachita  Na- 
tional Forest),  a  University  of  Arkan- 
sas Faculty  Research  Grant  and  the 
University  of  Arkansas  at  Little  Rock 
College  of  Science's  Office  of  Re- 
search, Science  and  Technology. 

LITERATURE  CITED 

Ashton,  Ray  E.  1976.  Endangered 
and  threatened  amphibians  and 
reptiles  in  the  United  States.  Soci- 
ety Study  Amphibians  Reptiles, 
Herpetological  Circular  5, 65  pp. 

Barbour,  Roger  W.,  and  Wayne  H. 
Davis.  1969.  Bats  of  America.  Uni- 
versity Press  of  Kentucky,  Lexing- 
ton. 286  pp. 

Barnett,  Douglas  Eldon.  1970.  An 
ecological  investigation  of  cavemi- 
cole  populations  in  Mansell  Cave, 
Randolph  County,  Arkansas.  M.S. 
Thesis,  Northwestern  State  Uni- 
versity, Natchitoches,  LA.  51  pp. 

Barr,  Thomas  C.  1963.  Ecological 
classification  of  cavernicoles.  Cave 
Notes  5:9-12. 

Bishop,  Sherman  C.  1943.  Handbook 
of  Salamanders.  Comstock  Pub- 
lishing Company,  Ithaca.  555  pp. 

Black,  Jeffery  H.  1974.  Notes  on  Ple- 
thodon  ouachitae  in  Oklahoma.  Pro- 
ceedings Oklahoma  Academy  Sci- 
ence 54:88-89. 


Black,  Jeffery  H.  1980.  Amphibians  of 
Oklahoma — a  checklist.  Bulletin 
Oklahoma  Herpetological  Society 
4:78-80. 

Blair,  Albert  P.,  and  Hague  L.  Lind- 
say, Jr.  1965.  Color  pattern  vari- 
ation and  distribution  of  two  large 
Plethodon  salamanders  endemic  to 
the  Ouachita  Mountains  of  Okla- 
homa and  Arkansas.  Copeia  3:331- 
335. 

Branner,  John  C.  1888.  Annual  report 
of  the  Geological  Survey  of  Arkan- 
sas. Volume  1.  Little  Rock.  103pp. 

Caire,  William.  1985.  Summer  ecol- 
ogy of  the  bats  in  southeastern 
Oklahoma.  Final  Report.  Okla- 
homa Department  Wildlife  Con- 
servation, Oklahoma  City.  22  pp. 

Chaney,  Steve  W.  1984.  Cave  Man- 
agement Plan.  USDI,  National 
Park  Service,  Buffalo  National 
River,  Arkansas.  24  pp. 

Clardy,  Benjamin  F.,  and  William  V. 
Bush.  1976.  Mercury  District  of 
southwest  Arkansas.  Information 
Circular  23.  Arkansas  Geological 
Commission,  57  pp. 

Culver,  David  C.  1986.  Cave  Faunas. 
Pp.  427-443.  In  Conservation  Biol- 
ogy: The  science  of  scarcity  and 
diversity.  Michael  E.  Soule,  editor. 
Sinaeur  Assoc.  Inc.,  Publishers, 
Sunderland,  MA.  584  pp. 

Davis,  Wayne  H.,  William  Z.  Lid- 
icker,  Jr.,  and  John  A.  Sealander. 
1955.  Myotis  austroriparius  in  Ar- 
kansas. Journal  of  Mammalogy 
36:288. 

Dunn,  Emmett  R.,  and  Albert  A. 
Heinze.  1933.  A  new  salamander 
from  the  Ouachita  Mountains. 
Copeia  3:121-122. 

Gates,  J.  Edward,  George  A.  Feld- 
hamer,  Lizabeth  A.  Griffith,  and 
Richard  L.  Raesly.  1984.  Status  of 
cave-dwelling  bats  in  Maryland: 
Importance  of  marginal  habitats. 
Wildlife  Society  Bulletin  12:162- 
169. 

Glass,  Byron  P.,  and  Claud  M.  Ward. 
1959.  Bats  of  the  genus  Myotis 
from  Oklahoma.  Journal  of  Mam- 
malogy 40:194-201. 

Harrington,  Donald.  1986.  Bear  City, 


Arkansas,  p.  381-427.  In:  Let  us 
build  us  a  city.  Harcourt  Brace 
Jovanovich,  Publishers,  New  York. 

Harvey,  Michael  J.,  John  J.  Cassidy, 
and  Gary  G.  O'Hagan.  1981.  En- 
dangered bats  of  Arkansas:  Distri- 
bution, status,  ecology  and  man- 
agement. Final  Report  to  Arkansas 
Game  and  Fish  Commission,  U.S. 
Forest  Service  (Ozark  National 
Forest)  and  National  Park  Service 
(Buffalo  National  River).  Memphis 
State  University,  80  pp. 

Hathom,  Jim,  and  Jer  Thornton. 
1986.  The  common  sense  guide  to 
cave  gates.  American  Cave  Con- 
servation Association  1:23-44. 

Heath,  Darrell  R.,  Gary  A.  Heidt, 
David  A.  Saugey  and  V.  Rick 
McDaniel.  1983.  Arkansas  range 
extensions  of  the  seminole  bat  (La- 
siurus  seminolus)  and  eastern  big- 
eared  bat  (Plecotus  rafinesquii)  and 
additional  county  records  for  the 
hoary  bat  (Lasiurus  cinereus),  sil- 
ver-haired bat  (Lasionycteris  nocti- 
vagans)  and  evening  bat  (Nycticeius 
humeralis).  Proceedings  Arkansas 
Academy  Science  37:90-91 . 

Heath,  Darrell  R.,  David  A.  Saugey, 
and  Gary  A.  Heidt.  1986.  Aban- 
doned mine  fauna  of  the  Ouachita 
Mountains,  Arkansas:  Vertebrate 
taxa.  Proceedings  Arkansas  Acad- 
emy Science  40(in  press). 

Hudgins,  Mary  D.  1971.  Gold  and 
silver  boom  town  of  Bear  thrived 
for  a  decade.  The  Record:  Hot 
Springs-Garland  County  Histori- 
cal Society  12:98-102. 

Humphrey,  Stephen  R.  1978.  Status, 
winter  habitat,  and  management 
of  the  endangered  Indiana  bat, 
Myotis  sodalis.  Florida  Science 
41:65-76. 

Hurter,  J.  and  J.K.  Strecker,  Jr.  1909. 
Amphibians  and  reptiles  of  Ar- 
kansas. Transactions  Academy 
Science  St.  Louis  18:11-27. 

Ireland,  Patrick  H.  1971.  Systematics, 
reproduction  and  demography  of 
the  salamander  Eurycea  multipli- 
cata.  Ph.D.  Dissertation.  Univer- 
sity of  Arkansas,  Fayetteville.  127 
pp. 


70 


Lacki,  Michael  ].,  and  Theodore  A. 
Bookhout.  1983.  A  survey  of  bats 
in  Wayne  National  Forest,  Ohio. 
Ohio  Journal  Science  83:45-50. 

Maser,  Chris,  Jon  E.  Rodiek,  and  Jack 
Ward  Thomas.  1979.  Cliffs,  talus 
and  caves.  Pp.  96-103.  In  Wildlife 
habitats  in  managed  forests:  The 
Blue  Mountains  of  Oregon  and 
Washington.  Jack  Ward  Thomas, 
editor.  Agriculture  Handbook 
No.553,  USDA  Forest  Service, 
Washington  D.C.  512  pp. 

McDaniel,  V.  Rick,  and  Kenneth  L. 
Smith.  1976.  Cave  fauna  of  Arkan- 
sas: selected  invertebrate  taxa. 
Proceedings  Arkansas  Academy 
Science  30:57-60. 

McDaniel,  V.  Rick,  and  James  E. 
Gardner.  1977.  Cave  fauna  of  Ar- 
kansas: vertebrate  taxa.  Proceed- 
ings Arkansas  Academy  Science 
31:68-71. 

Means,  D.  Bruce.  1974.  The  Status  of 
Desmogmthus  brimleyorum 
Stejneger  and  an  analysis  of  the 
genus  Desmognathus 
(Amphibia:Urodela)  in  Florida. 
Bulletin  Rorida  State  Museum, 
Biological  Sciences  18. 100  pp. 
Mumford, 

Russell  E.  and  John  O.  Whitaker. 
1982.  Mammals  of  Indiana.  Indi- 
ana University  Press,  Blooming- 
ton.  537  pp. 

Nieland,  Jim.  1985.  Cave  manage- 
ment prescriptions.  American 
Cave  Conservation  Association 
1:36-41. 

Nieland,  Jim,  and  Jer  Thornton.  1985. 
Spelean  inventory  and  evaluation 
guide.  American  Cave  Conserva- 
tion Association  1:1-35. 

Noble,  G.K.,  and  Byron  C.  Marshall. 
1929.  The  breeding  habits  of  two 
salamanders.  American  Museum 
Novitates  347:2-12. 

Pope,  Clifford  H.,  and  Sarah  H. 
Pope.  1951.  A  study  of  the 
salamander,  Plethodon  ouachitae, 
and  the  description  of  an  allied 
form.  Bulletin  Chicago  Academy 
Science  9:129-152. 

Puckette,  Bill.  1974-75.  Bear  Caves. 
Oklahoma  Underground  7:15-17. 


Reagan,  Douglas  P.  1974.  Threatened 
native  amphibians  of  Arkansas. 
Pp.  93-99.  In  Arkansas  natural 
area  plan.  Arkansas  Department 
of  Planning.  Little  Rock,  AR.  248 
pp. 

Robison,  Henry  W.,  and  Kenneth  L. 
Smith.  1982.  The  endemic  flora 
and  fauna  of  Arkansas.  Proceed- 
ings Arkansas  Academy  Science 
36:52-57. 

Saugey,  David  A.,  Robert  H.  Baber, 
and  V.  Rick  McDaniel.  1978.  An 
unusual  accumulation  of  bat  re- 
mains from  an  Ozark  cave.  Pro- 
ceedings Arkansas  Academy  Sci- 
ence 32:92-93. 

Saugey,  David  A.,  Darrell  R.  Heath, 
and  Gary  A.  Heidt.  1985.  Summer 
use  of  abandoned  mines  by  the 
Caddo  Mountain  salamander,  Ple- 
thodon caddoensis  (Plethodontidae) 
in  Arkansas.  Southwestern  Natu- 
ralist 30:318-319. 

Sealander,  John  A.,  and  Howard 
Young.  1955.  Preliminary  observa- 
tions on  the  cave  bats  of  Arkansas. 
Proceedings  Arkansas  Academy 
Science  7:21-31. 

Sievert,  Gregory.  1986.  An  investiga- 
tion of  the  distribution  and  popu- 
lation status  of  the  Rich  Mountain 
Salamander  (Plethodon  ouachitae)  in 
Oklahoma.  Final  Report.  Okla- 
homa Department  Wildlife  Con- 
servation, Oklahoma  City.  37  pp. 

Smith,  Kenneth  L.  1984.  Animal  spe- 
cies. Pp.  48-70.  In  Arkansas's  natu- 
ral heritage.  Bill  Shepherd,  editor. 
August  House  Publishers,  Little 
Rock.  116  pp. 

Stone,  Charles  G.,  and  William  V. 
Bush.  1984.  General  geology  and 
mineral  resources  of  the  Caddo 
River  Watershed.  Information  Cir- 
cular 29.  Arkansas  Geological 
Commission,  Little  Rock.  32  pp. 

Strecker,  John  K.  Jr.  1908.  Notes  on 
the  habits  of  two  Arkansas  sala- 
manders and  a  list  of  batrachians 
and  reptiles  collected  at  Hot 
Springs.  Proceedings  Biological 
Society  Washington  21:85-89. 

Trauth,  Stanley  E.  1988.  Egg  clutches 
of  the  Ouachita  dusky 


salamander,  Desmognathus  brimley- 
orum (Caudata:  Plethodontidae), 
collected  in  Arkansas  during  a 
summer  drought.  Southwestern 
Naturalist  (in  press). 

U.S.  Army  Corps  of  Engineers.  1980. 
Lake  Ouachita  design  memoran- 
dum No.  2A.  Re-evaluated  and 
Updated  Master  Plan  for  Develop- 
ment and  Management  of  Lake 
Ouachita  and  Blakely  Mountain 
Dam.  Vicksburg,  MS.  254  pp. 

U.S.  Federal  Register.  1985.  Endan- 
gered and  threatened  wildlife  and 
plants;  review  of  vertebrate  wild- 
life; notice  of  review.  Volume  50, 
No.  181.  September  18. 10  pp. 

Wells,  Kentwood  D.,  and  Roger  A. 
Wells.  1976.  Patterns  of  movement 
in  a  population  of  the  slimy 
salamander,  Plethodon  glutinosus, 
with  observations  on  aggrega- 
tions. Herpetologica  32:165-162. 

Whitaker,  John  O.,  and  Francis  A. 
Winter.  1977.  Bats  of  the  caves  and 
mines  of  the  Shawnee  National 
Forest,  southern  Illinois.  Transac- 
tions Illinois  State  Academy  Sci- 
ence. 70:301-313. 


71 


The  Herpetofauna  of  Long 
Pine  Key,  Everglades 
National  Park,  in  Relation  to 
Vegetation  and  Hydrology' 

George  H.  Dalrymple^ 


Abstract.— The  amphibians  and  reptiles  of  the 
Long  Pine  Key  region.  Everglades  National  Park, 
were  surveyed  between  1984  and  1986.  This 
herpetofauna,  with  51  species,  is  well  represented 
by  habitat  generalists  and  Prairie  species,  but  the 
compliment  of  Upland  species,  primarily  Pineland 
species,  is  low  due  to  the  lack  of  natural  soil 
development  and  the  isolation  of  the  area. 


Many  authors  have  noted  a  general 
reduction  in  species  diversity  among 
animal  groups  as  latitude  decreases 
in  peninsular  Florida  (Dinnen  1984, 
Loftus  and  Kushlan  1987,  for  fishes; 
Duellman  and  Schwartz  1958,  Kiester 
1971,  for  amphibians  and  reptiles; 
Cook  1969,  Robertson  and  Kushlan 
1984,  for  birds;  Simpson  1964,  Layne 
1984,  for  manmials).  Simpson  (1964) 
considered  such  a  "peninsular  ef- 
fect to  be  due  to  a  greater  rate  of 
extinction  and,  or  a  lower  rate  of 
immigration  along  peninsulas  in 
comparison  to  the  mainland. 

Species  area  curves  (Preston  1962, 
Mac  Arthur  and  Wilson  1967)  for  liz- 
ards and  snakes  evaluated  by  Busack 
and  Hedges  (1984)  showed  that  there 
was  no  significant  peninsular  effect 
in  Florida.  There  was,  however,  a 
general  trend  for  reduced  species 
numbers  as  one  proceeds  down  the 
peninsula  of  Rorida,  most  likely 
caused  by  a  reduction  in  habitat 
quality.  Moreover,  Robertson's 
(1955)  study  of  breeding  land  birds 
of  the  Long  Pine  Key  region  of  Ever- 
glades National  Park,  the  southern 
most  Upland  region  on  the  mainland, 
revealed  both  lower  species  richness 
and  lower  densities  within  species 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ,  July  19-21,  1988.) 

'George  H.  Dalrymple  is  Associate  Pro- 
fessor, Department  of  Biological  Sciences, 
Florida  International  University,  Miami,  FL 
33199. 


than  in  other  areas.  This  reduced 
abundance  of  animals  agrees  with 
the  general  belief  that  productivity  is 
low  in  southern  Florida  Pinelands 
(oligotrophic,  Snyder  1986).  When 
Duellman  and  Schwartz  (1958)  de- 
scribed the  southern  Florida  herpe- 
tofauna as  "depauperate... for  a 
warm  lowland  area"  they  were  refer- 
ring to  the  lower  number  of  species 
(table  1).  It  has  remained  unclear 
whether  characterization  of  the  her- 
petofauna as  depauperate  applies  to 
all  habitat  types  in  the  region,  in- 


cludes both  low  species  and  popula- 
tion numbers  and  applies  to  all  taxa. 

The  main  objectives  of  this  study 
are  to: 

1.  develop  a  species  list  of  am- 
phibians and  reptiles  in  Long 
Pine  Key-Paradise  Key  area 
(abbreviated  LPK  herein), 

2.  describe  species  associations 
with  vegetation  characteris- 
tics. 


Table  1.— The  number  of  species  of  amphibians  and  reptiles  found  In  Flor- 
ida, southern  Florida  and  In  Long  Pine  Key' 


Taxa 

Florida 

Southern  Florida 
#  (%) 

Long  Pine  K 
#  (%) 

# 

Salamanders 

24 

4  (17) 

3 

(13) 

Frogs  and  toads 

29 

16  (55) 

12 

(41) 

Amphibian 

Subtotal 

53 

20  (38) 

15 

(28) 

Turtles 

20 

11  (55) 

8 

(40) 

Crocodilians 

2 

2(100) 

1 

(50) 

Lizards 

16 

11  (69) 

6 

(38) 

Snakes 

41 

28  (68) 

21 

(52) 

Reptile 

Subtotal 

79 

52  (66) 

36 

(46) 

Totals 

132 

72  (55) 

51 

(39) 

'The  data  for  Florida  and  souttiern  Florida  are  based  upon  current  species  lists 
(Wilson  arud  Porras,  1983;  Auffenberg,  1982).  The  numt>ers  for  Long  Pine  Key  are  for  the 
current  study  (see  text).  Since  Long  Pine  Key  column  includes  the  exotic  species 
Eleutherodactylus  planirostris.  Osteopilusseptentrionalis  onaf  Anolissagrei  they  have 
been  included  in  the  counts  for  the  first  two  columns  also.  (Salamander  list  includes 
Stereochilus  marginatus;  frog  list  includes  the  new  species  Rana  okaloosae  (Moler, 
1986). 


72 


3.  evaluate  correlations  be- 
tween species'  phenologies 
and  rainfall  patterns  in  the 
area, 

4.  estimate  abundances  of  spe- 
cies and  compare  them  to 
other  areas  in  North  Amer- 
ica. 


Study  Area 

The  Long  Pine  Key  (LPK)  region  was 
chosen  for  study  because  this  8000  ha 
area  is  the  principal  remaining  natu- 
ral upland  region  of  the  original  Mi- 
I    ami  (or  Atlantic)  Rock  Ridge  physi- 
ographic province  (Davis  1943)  and 
jj    as  part  of  Everglades  National  Park  it 
'    has  been  protected  from  human 
interference  for  nearly  40  years.  The 
region  includes  about  4650  ha  of  Pi- 
nelands  (Snyder  1986)  with  a  series 
of  ''transverse  or  finger  glades,"  or 


seasonally  flooded  Prairies,  inter- 
spersed throughout  the  Pinelands 
(fig.  1).  Within  the  Pinelands  there  is 
a  series  of  at  least  120  tropical  hard- 
wood Hammocks  (Olmsted  et  al. 
1983,  fig.  2)  varying  in  size  from  .1  ha 
to  91  ha  (Olmsted,  Loope  and  Hilsen- 
beck  1980).  Most  Hammocks  are 
completely  surrounded  by  Pineland 
and  are  kept  rather  small  due  to  the 
frequent  fires  (prescribed  burns  and 
natural  fires  from  lightning)  in  the 
region.  The  largest  Hammock,  Royal 
Palm,  is  surrounded  by  seasonally 
flooded  Prairies  and  has  almost  com- 
pletely overgrown  the  limestone  ele- 
vation known  as  Paradise  Key  (these 
names  are  sometimes  used  inter- 
changeably). Because  Paradise  Key 
figured  importantly  in  the  study  of 
Duellman  and  Schwartz  (1958),  I 
have  included  it  in  the  present  study 
as  part  of  the  general  area  described 
herein  as  LPK. 

On  the  southern  border  of  LPK 


about  3600  ha  of  land  were  farmed 
until  1975  (abandonment  was  an  at- 
tenuated process  from  the  1960's  to 
1975),  when  this  agricultural  area, 
known  as  the  "hole-in-the  donut," 
was  purchased  by  the  Park  Service. 
Early  farming  was  limited  to  areas 
with  deeper  soil,  and  involved  little 
alteration  of  the  underlying  bedrock. 
Starting  in  1954  (W.B.  Robertson,  Jr. 
pers.  comm.)  rock-plowing  of  the 
upper  20  cm  of  the  ground  surface 
created  an  artificial  soil:  "deeper, 
better  drained,  better  aerated,  and 
possibly  more  nutrient-rich  than  the 
pre-farming  soil"  on  1600  of  the  3600 
ha  (Ewel  et  al.  1982:1-2).  The  sub- 
strate alteration  proved  conducive  to 
the  establishment  of  exotic  vegeta- 
tion, especially  Brazilian  Pepper  (Sch- 
inus  terebinthifolius)  after  the  farm- 
land was  abandoned  (Ewel  et  al. 
1982). 

Existing  detailed  surveys  of  the 
region's  vegetation  in  relation  to  ele- 
vation, fire  and  hydrology  (e.g.  Olm- 
sted et  al  1980;  Olmsted  et  al.  1983; 
Olmsted  and  Loope  1984;  Taylor  and 
Herndon  1981)  as  well  as  an  ex- 
tremely detailed  vegetation  map  of 
the  area  (Johnson  et  al.  1983)  have 
made  it  much  easier  to  plan  the  cur- 
rent project.  Historical  surveys  of  the 
literature  in  the  above  cited  refer- 
ences, among  many  others,  make  it 
clear  that  the  LPK  region  has  not 
been  completely  free  from  distur- 
bances: logging  of  the  Pinelands  dur- 
ing the  1930's  and  1940's;  farming,  as 
described  above;  invasion  by  exotic 
vegetation;  development  of  elevated 
roadways  with  marl  dug  from  local 
pits  and  their  resulting  small  canals, 
culverts  and  ponds  bordering  the 
former  farmlands  (all  of  which  dis- 
tort the  original  associations  of  eleva- 
tion, soil,  vegetation  and  surface  wa- 
ter); fire  roads,  to  help  control  pre- 
scribed bums;  and  the  inevitable 
presence  of  humans  and  their  build- 
ings (both  those  for  visitors  and  the 
complex  of  staff  facilities).  All  of 
these  factors  play  a  role  in  determin- 
ing the  present  herpetofauna.  Cur- 
rent park  management  fosters  a  de- 


Figure  1.— Aerial  photograph  of  Pineland  and  Prairie  of  Long  Pine  Key. 


73 


large  enough  to  ensure  lasting  pres- 
ervation of  this  unique  ecosystem 
type. 

Materials  and  Methods 

General  Collecting  and  Road 
Cruising 

For  the  3  years  of  the  study  reported 
on  herein  many  hours  were  spent 
surveying  and  trapping  in  areas  for 
evidence  of  amphibians  and  reptiles. 
Each  time  the  traps  were  checked,  a 
50  km  section  of  unimproved  dirt 
roads  was  driven  over  by  van,  and 
an  additional  15  km  paved  road  was 
systematically  covered  by  van  for  a 
total  of  8  to  16  hours  per  week,  dur- 
ing which  all  animals  were  captured 
and  identified.  Searches  on  foot,  by 
teams  of  two  to  four  people,  were 
conducted  in  all  of  the  major  habitats 
each  week,  during  which  animals 
were  searched  for  at  the  surface  and 
under  rocks  and  logs.  The  time  spent 
collecting  and  road  cruising  was  di- 
vided between  day  and  night  to  en- 
sure that  all  species  in  LPK  might  be 
found. 


Trapping 

I  used  a  system  of  funnel  traps  at- 
tached to  drift  fences  and  transects 
(referred  to  throughout  as  '"arrays"). 
Many  researchers  have  used  arrays 
to  study  amphibians  and  reptiles 
(Campbell  and  Christman  1982b, 
Clawson  and  Baskett  1982,  Vogt  and 
Hine  1982,  Gibbons  and  Semlitsch 
1981,  Clark  1970),  however  they  all 
employed  arrays  that  included  both 
funnel  traps  and  pit  traps.  Usually 
the  pit  traps  are  placed  at  regular  in- 
tervals by  digging  holes  in  the 
ground.  However,  the  lack  of  well 
developed  soils  coupled  with  an  ir- 
regular limestone  surface  made  the 
use  of  pit  traps  impractical  to  use  in 
the  everglades. 

Each  array  was  constructed  of 
four  fifteen  meter  long  sheets  of 
shade  cloth  (one  meter  tall)  that 


intersected  in  the  middle  to  form  an 
"x."  The  shade  cloth  was  kept  up- 
right by  tieing  it  to  iron  rebars  that 
were  hammered  into  the  limestone. 
Traps  were  made  of  cylinders  of  one- 
eighth  inch  hardware  cloth  approxi- 
mately 1  m  in  length  and  30  cm  in 
diameter.  Each  trap  was  fitted  with 
two  funnels  (one  funnel  on  each  side 
of  the  shade  cloth  fencing)  made  of 
the  same  material.  Funnels  were  at- 
tached to  the  free  ends  of  the  four 
arms  of  the  array.  Shade  cloth  had 
12-cm  flaps  sewn  onto  the  bottom 
edge  to  conform  to  the  irregular  sur- 
faces of  the  everglades  terrain.  Flaps 
were  covered  with  natural  soils  and 
or  leaf  litter  so  that  animals  would 
not  crawl  under  them  (figs.  3  and  4). 
The  square  area  encompassed  by 
each  array  was  .10  ha. 

Arrays  were  placed  in  each  of  four 
main  habitat  types:  seasonally 
flooded  Prairies,  Pinelands,  tropical 


hardwood  Hammocks,  and  in  the 
area  of  secondary  succession  from 
former  farming,  the  "hole-in-the-do- 
nut."  The  latter  area  is  referred  to 
throughout  as  "Disturbed."  Thirteen 
arrays  were  maintained  starting  in 
May,  1984,  and  the  arrays  are  still 
checked  to  the  current  date.  Three 
arrays  were  placed  in  each  habitat 
type  within  Long  Pine  Key  and  one 
extra  hammock  array  was  main- 
tained in  Royal  Palm  Hammock  on 
Paradise  Key  (fig.  5).  Arrays  were 
temporarily  taken  down  during  park 
service  prescribed  burns  and  re- 
placed after  the  bums.  Because  ar- 
rays were  in  place  for  different  dura- 
tions, I  assessed  yield  in  terms  of  rate 
of  capture,  rather  than  absolute  cap- 
ture yield,  and  capture  rate  was  as- 
sessed separately  for  wet  and  dry 
seasons.  At  each  array  we  main- 
tained two  1-m^  pieces  of  tar-paper, 
under  which  we  commonly  collected 


seasons.  At  each  array  we  main- 
tained two        pieces  of  tar-paper, 
under  which  we  commonly  collected 
animals.  All  animals  caught  along  the 
fences  or  under  the  tar  paper  at  an 
array  were  counted  as  part  of  the 
capture  rate  at  the  array  in  question. 


Figure  3.— Aerial  photograph  of  locality  known  as  New  Wave  Prairie  in  Long  Pine  Key  with 
"x"-shaped  trapping  array  visible  at  left  (each  of  the  four  arnr>s  of  the  array  is  15  m  long). 


Symbolic  Star  Plot  Analysis 

Symbolic  Star  Plot  Analysis  (Cham- 
bers at  al.  1983)  was  chosen  as  a  use- 
ful multivariate  method  for  graphi- 
cally depicting  the  rates  of  capture  of 
species  in  the  major  habitats.  Only 
species  for  which  there  were  at  least 
ten  captures  were  chosen,  and  the 
analyses  were  based  on  the  number 
of  animals  trapped  per  1000  array 
days  because  the  raw  data  does  not 
reflect  the  fact  that  arrays  were  op- 
erational for  varying  time  periods. 
The  data  values  are  used  as  the 
lengths  of  the  rays  of  the  stars  for 
each  habitat.  All  data  values  were 
rescaled  to  range  from  1  to  c,  where  c 
is  the  length  of  the  smallest  ray  (set 
to  0.1  for  these  analyses).  According 
to  Chambers  et  al.  (1983:158):  "If  x..  is 
the    measurement  of  the  i""  variable 

then  the  scaled  variable  [x*.  ]  is 

') 

x*j  =  (1  -  c)(X|j  -  min,X|P 

/  (maX|X|j  -  min|X,j)  +  c." 

The  scaled  variables  are  arranged 
around  a  circle  at  equal  angles,  the 
number  of  angles  determined  by  the 
number  of  variables,  and  the  actual 
rays  are  drawn  by  connecting  points 
trigonometrically  calculated  for  an 
arbitrarily  chosen  maximum  radius 
for  the  circle. 

The  lengths  of  the  rays  (not  the 
area  adjoining  the  rays)  in  the  four 
habitat  stars  for  a  given  species  rep- 
resent the  proportion  of  all  captures 
for  that  species  in  each  habitat.  The 
result  is  intended  to  form  a  simple 
yet  "dramatic  and  memorable"  im- 
pression of  the  relationships  within 
species  and  between  habitat  types, 
for  further  details  see  Chambers  et  al. 
(1983:158-163). 


Figure  4.— Ground  level  view  of  trapping  array  fencing  in  Pineland, 


75 


Population  Abundance  Estimates 

For  most  species  the  actual  numbers 
presented  are  actual  numbers  of  indi- 
viduals captured.  All  snakes  and 
turtles  were  individually  marked. 
The  anurans  and  lizards  were 
marked  only  during  1984,  but  due  to 
the  lack  of  recaptures  I  stopped 
marking  in  1985.  The  marking 
method  used  for  snakes  was  that  of 
Brown  and  Parker  (1976),  and  even 
though  snakes  were  marked  for  four 
consecutive  years  (1984-1987)  the  re- 
capture rate  remained  very  low 
(<0.05,  Dalrymple,  in  prep.). 

Concentrations  of  amphibians  and 
reptiles  around  one  or  more  re- 
sources, such  as  water  (ponds  or 
lakes.  Carpenter  1952,  Reichenbach 
and  Dalrymple  1986),  hibemacula 
(caves,  pits  and  dens,  Woodbury 
1951,  Brown  and  Parker  1982a, 
Aleksiuk  and  Gregory  1974)  breeding 
sites  (Crump  1982,  Brown  and  Parker 
1982b,  Wiest  1982)  and  or  food  (Ha- 
milton 1951)  lead  to  recaptures  that 
allow  for  density  estimates  with  con- 
fidence limits  (cf.  Turner  1971).  These 
estimates  are  dependent  on  seasonal 
fluctuations,  and  may  differ  greatly 
from  estimates  of  crude  density. 
However,  few  concentrations  were 
found  on  LPK  particularly  because 
water  was  readily  available  in  nu- 
merous solution  holes  in  every  habi- 
tat. Moreover,  mild  winters  allowed 
most  species  to  be  active  throughout 
the  year,  and  the  ability  of  animals  to 
readily  go  underground  through  the 
porous  limestone  and  plentiful  solu- 
tion holes  found  in  all  habitats  re- 
sulted in  the  absence  of  group  hiber- 
nacula.  Further  complicating  density 
estimation  were  widespread  move- 
ments in  search  of  mates,  and  the  fact 
that  major  food  sources  were  not 
clumped. 

All  these  factors  lead  to  a  wide 
spread  distribution  of  most  species  in 
the  region  and  most  were  not  habitat 
specialists,  at  least  at  the  major  vege- 
tation type  level.  The  lack  of  concen- 
trations and  the  limited  number  of 
recaptures  permit  only  the  presenta- 


tion of  total  numbers  of  captures  and 
not  accurate  density  estimates  at  this 
time. 


Results 


Species  List 

Starting  in  January,  1984, 51  species 
of  amphibians  and  reptiles  were  ob- 
served or  collected  in  LPK  (table  2). 
Some  species  were  rare  because  they 
are  most  commonly  associated  with 
more  permanently  aquatic  habitats, 
such  as  the  Sloughs  (e.g.  Acris  gryllus, 
Ram  grylio,  Trionyx  ferox,  Farancia 
abacura,  Nerodia  cyclopion,  Nerodia  tax- 
ispilota,  Regim  alleni).  A  few  species 


that  have  been  recorded  in  the  larger 
geographic  region  were  not  found  in 
LPK  during  this  study  (Scaphiopus 
holbrooki,  Pseudobranchus  striatus, 
Semimtrix  pygaea,  Masticophis  flagel- 
lum,  Heterodon  platyrhinos,  Ophisaurus 
ventralis,  Sternotherus  odoratus). 


Trapping  Results 

Between  May,  1984  and  December, 
1986, 1709  amphibians  and  reptiles 
were  collected  either  in  the  traps, 
under  associated  tar  paper,  or  along 
array  fences  (table  3).  These  animals 
represent  37  of  the  51  species  (73%) 
known  from  our  overall  surveys.  I 
compared  the  four  habitats  by  re- 
cording the  number  of  animals  per 


Figure  5.— Map  of  the  Long  Pine  Key-Paradise  Key  region  of  Everglades  National  Park.  Array 
locations  are  nunnbered  and  referred  to  in  \he  text  as  follows:  1 .  Pine  Block  B,  2.  New  Wave 
Prairie,  3.  Pine  Block  E,  4.  Junk  Hammock,  5.  Serenoa  Prairie,  6.  Wrigfit  Hammock,  7.  Mud 
Prairie,  8.  Pine  Block  H,  9.  Palma  Vista  I  Hammock,  10.  Royal  Palm  Hammock,  11.  Burnout  Dis- 
turbed, 12.  Sctiinus  Disturbed,  13.  Grass  Disturbed. 


76 


Table  2.— List  of  species  of  amphibians  and  reptiles  observed  In  the  Long  Pine  Key  -  Paradise  Key  region  of  Ever- 
glades National  Park  during  present  study,  between  January,  1984  and  December,  1986.  The  reglonwide  natural 
habitat  associations  of  Duellman  and  Schwartz  (1958),  as  they  apply  In  the  study  area,  are  given  after  the  scientific 
name  for  each  species.  Pr  =  Prairie,  Pi  =  Pine,  H  =  Hammock,  A=  Permanently  Aquatic,  I.e.  Slough,  Canals. 


Scientific  name 


Common  name 


Scientific  name 


Common  name 


Urodela 

Amphiumo  means  -Pr  A 

Siren  lacerfina  -Pr  A 
Nofophfhalmus  viridescens  -Pr  A 

Anura 

Acris  gryllus  -Pr 

Bufo  quercicus -Pr  ?\  H 

Bufo  ferresfris  -Pr  Pi  H 

Eleufherodacfylus  planirosfris  -Pi 

Gasfrophryne  carolinensis  -Pr  Pi  H 

Hyla  cinerea  -Pr  Pi  H 

Hyla  squirella  -Pr  Pi  H 

Limnaoedus  ocularis  -Pr 

Osfeopilus  sepfenfrionalis  -H  * 

Pseudacris  nigrifa  -Pr  Pi 

Rana  grylio  -Pr  A 

Rana  splienocephala  -Pr  A 


Testudines 

Chelydra  serpentina  -Pr  A 
Clirysemys  floridana  -Pr  A 
Chrysemys  nelsoni  -Pr  A 
Deirocheiys  reficularia  -A 
Gopherus  polyphemus  -Pi 
Kinosfernon  bauri  -Pr  A 
Terrapene  Carolina  -Pr  Pi  H 
Trionyx  ferox -A 

Crocodylia 

Alligator  mississippiensis  -Pr  A 


two-toed  ampl-ii- 
uma 

greater  siren 
peninsula  newt 


Florida  cricket  frog 
oak  toad 
southern  toad 
greenhouse  frog 
eastern  narrow- 
mouthed  toad 
green  treefrog 
squirrel  treefrog 
little  grass  frog 
Cuban  treefrog 
Florida  chorus  frog 
pig  frog 

southern  leopard 
frog 


snapping  turtle 
peninsula  cooter 
red-bellied  turtle 
chicken  turtle 
gopher  tortoise 
striped  mud  turtle 
box  turtle 
Florida  soft-shelled 
turtle 


American  alligator 


Squamatalacertilia 

Anolis  carolinensis  -Pr  Pi  H 

Anolis  sagrei -P\ 

Eumeces  inexpectatus  -Pr  Pi  H 

Ophisaurus  compressus  -Pr  Pi 
Scincella  laterale  -Pi 
Sptiaerodactylus  notatus  -Pi 

Squamata  ,Serpentes 

Agkistrodon  piscivorus  -Pr  A 
Cemopt)ora  coccinea  -Pi 
Coluber  cor^strlctor -Pr  Pi  H 
Crotalus  adamanteus  -Pi 

Diadophis  punctatus  -Pr  Pi  H 
Drymarct)on  corais -Pr  Pi  H 
Eiaphe  guttata  -Pr  Pi  H 
Elaphe  obsoleta  -Pr  Pi  H 
Farancia  abacura  -Pr  A 
Lampropeltis  getulus  -Pr  Pi  H 
Lampropeltis  triangulum  -Pi 
Micrurus  fulvius  -Pi 
Nerodia  fasciata  -Pr  A 

Nerodia  cyclopion  -Pr  A 
Nerodia  taxispilota-  A 
Opheodrys  aestivus  -Pr  Pi  H 
Regina  alleni  -Pr 

Sistrurus  miliarius  -Pr  Pi 
Storeria  dekayi  -Pr  Pi  H 
Thamnophis  sauritus  -Pr  Pi  H 
Ttiamnophis  sirtalis  -Pr  Pi  H 


green  anole 
brown  anole 
southeastern  five- 
lined  skink 
island  glass  lizard 
ground  skink 
reef  gecko 


cottonmouth 
scarlet  snake 
black  racer 
eastern  diamond- 
back 

ringnecked  snake 
indigo  snake 
corn  snake 
yellow  rat  snake 
mud  snake 
kingsnake 
scarlet  kingsnake 
coral  snake 
banded  water 
snake 

green  water  snake 
brown  water  snake 
rough  green  snake 
striped  crayfish 
snake 

pigmy  rattlesnake 
brown  snake 
ribbon  snake 
garter  snake 


array  day.  The  highest  capture  rates 
were  in  seasonally  flooded  Prairie, 
which  had  both  the  most  individuals 
and  the  most  species  collected,  fol- 
lowed by  Disturbed  areas.  Hammock 
and  Pineland  (table  3). 

Monthly  total  rainfall  for  LPK  and 
maximum  water  level  from  well  sta- 
tion NP-72  in  the  same  area  for  data 
from  1984-1986  were  provided  from 
hydrological  stations  maintained  by 
the  South  Florida  Research  Center, 


Everglades  National  Park.  These  data 
were  correlated  with  the  monthly 
values  for  animals  trapped  per  check 
day.  There  were  significant  correla- 
tions between  number  of  animals 
caught  per  check  day  and  both 
monthly  rainfall  (r  =  0.55,  p  =  .001), 
and  monthly  maximum  water  levels 
(r  =  0.50,  p  =  .004)  for  the  three  year 
period  (fig.  6).  Rates  of  capture  were 
significantly  greater  during  the  wet 
season  than  the  dry  season  (table  4; 


Wilcoxin  matched  pairs  test,  T  =  3.0, 
p  <  .005).  Differences  in  overall  cap- 
ture rates  between  the  dry  and  wet 
seasons  is  greater  in  Hammock  and 
Disturbed  areas  than  in  the  Pinelands 
and  Prairie. 


Relative  Abundance 

Although  37  species  were  found  at 
arrays  they  were  not  all  equally  com- 


77 


Table  3.— Total  numbers  of  amphibians  and  reptiles  trapped,  May  1984-Dec  1986.  "Ctieck  days"  are  number  of  days 
on  which  traps  were  checked.  "Array  days"  are  number  of  total  days  arrays  were  standing.  Numbers  in  parentheses 
are  animals  per  1000  array  days.  Acronyms  at  right  of  table  are  for  species  used  in  figures  7-9. 


Taxa 

Prairie 

Pineland 

Hammock 

Disturbed 

Total 

A.  means 

9 

(3.5) 

0 

(0) 

0 

(0) 

0 

(0) 

9 

A.  gryllus 

1 

(0.4) 

0 

(0) 

0 

CO) 

0 

(0) 

\^y 

1 

B.  quercicus 

95 

(37.2) 

7 

(2,8) 

3 

(0  9) 

9 

(6.2) 

114 

Bq 

B.  ferresfris 

45 

(17.6) 

24 

(9.4) 

50 

(15.5) 

31 

(21.3) 

150 

Bt 

E.  plonirosfris 

15 

(5.9) 

17 

(6.7) 

50 

(15.5) 

6 

(4.1) 

88 

EP 

G.  carol inensis 

10 

(3.9) 

1 

(0.4) 

21 

(6  5) 

33 

(22.6) 

65 

Go 

H.  cinerea 

on 

1 
1 

7 

u-2; 

31 

He 

H.  squkella 

32 

(12.5) 

3 

(1.2) 

6 

d  9) 

4 

(2.7) 

45 

Hs 

OiSepfenfrionalis 

2 

(0.8) 

1 

(0.4) 

3 

CO  9) 

6 

(4.1) 

12 

Os 

P.nigrifa 

5 

(2.0) 

8 

(3.1) 

0 

CO) 

0 

(0) 

\^y 

13 

Pn 

R.  grylio 

5 

(2,0) 

V,*-  >^ J 

0 

(0) 

0 

CO) 

0 

(0) 

\^y 

5 

R.  sphenocephala 

135 

(52.8) 

10 

(3.9) 

106 

(32  8) 

20 

(13.7) 

271 

Rs 

A.  carolinensis 

170 

(66,5) 

136 

(52.3) 

19 

C5  9) 

19 

(13.0) 

344 

Ac 

A.  sagrei 

0 

(0) 

0 

(0) 

50 

(15.5) 

103 

(70.7) 

153 

As 

E.  inexpecfatus 

23 

(9.0) 

21 

(8.2) 

42 

CI  3.0) 

3 

(2.1) 

89 

El 

O.  compressus 

1 

(0.4) 

1 

(0.4) 

0 

(0) 

\^y 

1 

(0,4) 

3 

S.  late  rale 

30 

(11.7) 

9 

(3.5) 

3 

CO  9) 

0 

(0) 

\^y 

42 

SI 

S.  not  a  f us 

0 

(0) 

0 

(0) 

29 

(9.0) 

0 

(0) 

29 

Sn 

K.  bauri 

12 

(4.7) 

2 

(0.8) 

1 

(0.3) 

\  V  •  ^y 

1 

(0,7) 

16 

Kb 

T.  Carolina 

n 

(4.3) 

1 

(0.4) 

2 

(0.6) 

3 

(2.1) 

17 

To 

A.  piscivorus 

1 

(0.4) 

0 

(0) 

0 

(0) 

2 

(1.4) 

3 

C.  coccinea 

2 

(0.8) 

0 

(0) 

0 

(0) 

0 

(0) 

\^y 

2 

C.  constrictor 

8 

(3.1) 

30 

(1 1.8) 

14 

C4  3) 

14 

(9.6) 

v »  •  ^y 

66 

Cc 

C.  adamanteus 

0 

(0) 

0 

(0) 

0 

(0) 

\^y 

1 

(0,7) 

1 

D.  punctatus 

3 

(1.0) 

3 

(1.0) 

13 

(4  0) 

0 

(0) 

\^y 

19 

Dp 

D.  corals 

1 

(0.4) 

2 

(0.8) 

2 

(0,6) 

0 

(0) 

5 

E.  guttata 

0 

(0) 

1 

(0.4) 

0 

(0) 

0 

(0) 

1 

E.  obsoleta 

0 

(0) 

0 

(0) 

4 

(1.2) 

1 

(0.7) 

5 

L.getulus 

0 

(0) 

0 

(0) 

0 

(0) 

1 

(0.7) 

1 

L  triangulum 

1 

(0.4) 

0 

(0) 

0 

(0) 

0 

(0) 

1 

M.  fulvius 

0 

(0) 

0 

(0) 

4 

(1.2) 

0 

(0) 

4 

N.  fasciata 

3 

(1.2) 

0 

(0) 

0 

(0) 

0 

(0) 

3 

R.  alleni 

1 

(0.4) 

0 

(0) 

0 

(0) 

0 

(0) 

1 

S.  millarius 

14 

(5.5) 

8 

(3.1) 

3 

(0.9) 

6 

(4.1) 

31 

Sm 

S,  dekayi 

2 

(0.8) 

0 

(0) 

4 

(1.2) 

0 

(0) 

6 

T.  sauritus 

8 

(3.1) 

1 

(0.4) 

10 

(3.1) 

0 

(0) 

19 

Tsa 

T.  sirtalis 

30 

(11.7) 

5 

(2.0) 

2 

(0.6) 

7 

(4.8) 

44 

Tsi 

Totals 

695 

292 

448 

274 

1709 

No.  Check  days 

669 

663 

789 

361 

2482 

Anis/Check  day 

1.04 

0.44 

0.57 

0.76 

0.70 

No.  Species 

30 

22 

24 

21 

37 

No.  Array  days 

2555 

2550 

3229 

1458 

9792 

Anis/ Array  day 

0.27 

0.12 

0.14 

0.19 

0.18 

men.  The  most  common  species  were 
anurans  and  lizards  (table  3):  Ram 
sphenocephala,  Bufo  terrestris,  and  Ano- 
lis  carolinensis.  Of  the  20  species  of 
snakes  collected  during  the  study,  17 
were  trapped  but  only  five  were  cap- 


tured in  high  enough  frequency  to 
allow  for  more  detailed  study  (Col- 
uber constrictor,  Thamnophis  sirtalis, 
Sistrurus  miliarius,  Diadophis  punc- 
tatus, and  Thamnophis  sauritus).  As  a 
preliminary  method,  abundance  can 


be  minimally  estimated  as  the  actual 
counts  from  the  'Total"  column  of 
table  3  as  the  number  per  hectare  (12 
arrays,  each  one  covering  approxi- 
mately one-tenth  of  a  hectare  makes 
this  a  conservative  estimate). 


78 


Habitat  Use  And  Preference 

A  species'  likelihood  of  being 
trapped  is  more  a  funcrion  of  the 
number  of  individuals  in  the  vicinity 
of  an  array  than  a  result  of  any  dif- 


Figure  6.— Comparison  of  number  of  ani- 
mals trapped  per  check  day  per  month 
with  monthly  rainfall  and  water  table  values 
from  study  area  between  May  1984  and 
December  1986. 


ference  in  trap  functioning  between 
habitats.  For  species  with  high  cap- 
ture rates,  there  were  significant  dif- 
ferences in  habitat  use  for:  Coluber 
constrictor,  more  common  in  Pine- 
lands  (chi  square  =  14.59,  p  =  ,0007); 
Thamnophis  sirtalis,  Sistrurus  miliarius, 
Scincella  laterale  and  Bufo  quercicus  all 
more  common  in  Prairie  (chi  squares 
of  42.9,  9.6,  26.4,  71.8  respectively,  all 
with  p's  <  .01);  while  Bufo  terrestris  is 
equally  common  in  all  habitats  (chi 
square  =  2.36,  p  =  .51).  In  most  cases, 
species  were  found  in  more  than  one 
and  usually  three  habitats  (cf.  Duell- 
man  and  Schwartz  1958).  Among 
trapped  species,  41%  were  found  in 
all  four  habitat  types,  27%  in  two  or 
three,  and  32%  in  only  one  habitat 
type.  Seven  of  the  13  species  from 
only  1  habitat  type  were  from  Prairie. 


Table  4.— Results  of  1985  trapping  of  all  individuals  of  amphibians  and  rep- 
tiles at  13  array  sites  organized  by  vegetation  type,  and  season  (dry  =  No- 
vember-April; wet  =  May-October).  "Check-days"  are  the  number  of  days 
on  which  an  array  was  checked  for  animals.  Note  that  there  is  no  data  for 
the  wet  season  for  "Grass"  array  (see  Materials  and  Methods).  Variation 
within  habitat  types  is  as  great  as  between  habitat  types. 


Habitat/array 


No.  No.  No.  Animals 

Individuals       species       check-days  per  check  day 


Season: 

Dry 

Wet 

Dry 

Wet 

Dry 

Wet 

Dry 

Wet 

Prairie 

New  Wave 

65 

118 

12 

20 

54 

54 

1,2 

2.2 

Mud 

38 

64 

10 

18 

56 

54 

0.7 

1.2 

Serer^oa 

12 

20 

2 

10 

50 

51 

0,2 

0.4 

Pineiand 

Pine  Block  B 

26 

23 

8 

8 

50 

53 

0,5 

0.4 

Pine  Block  H 

25 

39 

8 

11 

56 

51 

0,5 

0,8 

Pine  Block  E 

16 

17 

4 

8 

51 

52 

0,3 

0,3 

Hammocks 

Royal  Palm 

18 

110 

7 

17 

56 

28 

0.3 

3.9 

Pclma  Vista  1 

11 

50 

6 

12 

56 

33 

0,2 

1.5 

Wright 

15 

21 

6 

8 

52 

53 

0.3 

0.4 

Junk 

17 

23 

7 

7 

53 

52 

0,3 

0.4 

Disturbed 

Schinus 

11 

76 

6 

12 

55 

33 

0.2 

2.3 

Burnout 

11 

16 

6 

7 

45 

17 

0.2 

0.9 

Grass 

14 

4 

18 

0.7 

Symbolic  star  plot  analyses 
(Chambers  et  al.  1983)  were  applied 
to  the  1984-1986  trap  data  for  the 
number  of  animals  per  1000  array 
days  as  the  data  set  (table  3),  for  the 
anurans  (fig.  7),  lizards  and  turtles 
(fig.  8),  and  snakes  (fig.  9).  Since  the 
qualitative  general  habitat  associa- 
tions of  Duellman  and  Schwartz 
(1958)  were  corroborated  in  this 
study,  I  restricted  this  quantitative 
analysis  to  those  species  for  which 
there  were  at  least  10  captures. 

It  is  obvious  from  the  anuran  plot 
that  the  majority  of  individuals  and 
species  are  most  prevalent  in  Prairie. 
Pseudacris  nigrita  is  strongly  repre- 
sented in  Pineiand,  as  was  noted  by 
Duellman  and  Schwartz  1958).  In 
Hammocks,  Eleutherodactylus 
planirostris,  Bufo  terrestris,  Gastro- 
phryne  carolinensis,  and  Hyla  cinerea 
were  dominant.  Rana  sphenocephala 
was  most  common  in  Prairie  but  was 
very  abundant  in  two  Hammocks 
that  are  adjacent  to  wet  Prairie  and 
that  retained  water  in  solution  holes 
throughout  most  of  the  year  (Royal 
Palm  and  Palma  Vista  I).  Bufo 
terrestris,  G.  carolinensis  and  the  exotic 
Cuban  tree  frog,  Osteopilus  septentri- 
onalis,  were  dominant  in  Disturbed 
habitat  (fig.  7). 

For  the  trap  data  for  turtles,  Kinos- 
ternon  bauri  and  Terrapene  Carolina, 
and  the  lizards.  Prairie  again  had  the 
greatest  abundance;  but  T.  Carolina 
was  commonly  found  in  the  Dis- 
turbed habitat.  Anolis  carolinensis  was 
well  represented  in  Pineiand  and 
Prairie,  as  were  the  skinks,  Eumeces 
inexpectatus  and  Scincella  laterale.  Ano- 
lis sagrei  was  restricted  to  Disturbed 
sites  and  Hammocks,  especially 
those  close  to  roads  and  parking  lots. 
Sphaerodactylus  notatus  is  most  often 
found  in  leaf  litter  of  Hammocks, 
and  E.  inexpectatus  is  also  well  repre- 
sented in  Hammocks  (fig.  8). 

For  snakes,  the  star  diagram 
analysis  was  restricted  to  the  five 
most  common  species;  again  the 
greatest  diversity  and  abundance  is 
found  in  Prairie.  Coluber  constrictor 
was  clearly  the  dominant  snake  in 


79 


prairie  pineland 


Pn  Rg 


Figure  7.— Star  plot  diagrams  of  anuran 
data  from  table  3,  comparing  th»e  frequen- 
cies of  trapping  (anurans  per  1000  array 
days)  of  ttie  species  in  \he  four  tiabitat 
types.  Genus  and  species  names  abbrevi- 
ated on  key  at  bottom  of  figure  correspond 
to  acronyms  given  in  table  3. 

Pineland.  Sistrurus  miliarius  was  well 
represented  in  all  habitats,  but  is 
least  common  in  Hammocks.  Tham- 
nophis  sirtalis  was  most  abundant  in 
Prairie,  while  T.  sauritus  was  most 
common  in  Prairie  and  Hammocks. 
Diadophis  punctatus  is  the  snake  spe- 
cies most  difficult  to  keep  in  traps 
(because  of  their  small  size  they 
could  more  readily  escape)  but  cur- 
rent data  indicate  that  they  are  most 
common  in  the  leaf  litter  environ- 
ment of  Hammocks  (fig.  9). 

The  most  similar  habitats  with  re- 
gard to  trap  data  were  Prairie  and 
Pineland,  the  least  similar  were  Pine- 
land and  Hammock  (table  5).  Table  5 
includes  the  only  data  from  the  ar- 
rays and  therefore  some  species  are 
excluded  from  the  similarity  index 
(because  the  index  used,  Morisita's 
index  (Horn  1966;  Brower  and  Zar 
1984)  requires  data  on  both  the  num- 
ber of  species  and  the  number  of  in- 
dividuals per  species  in  the  estima- 
tion of  degree  of  similarity). 


hammock  disturbed 


prairie  pineland 


Figure  8.— Star  plot  diagrams  of  lizard  and 
turtle  data  from  table  3,  comparing  fre- 
quencies of  trapping  (lizards  or  turtles  per 
1000  array  days)  in  the  four  tiabitat  types. 
Genus  and  species  names  abbreviated  on 
key  at  bottom  of  figure  correspond  to  acro- 
nyms given  in  table  3. 

Discussion 

Species  List 

Duellman  and  Schwartz  (1958)  gave 
a  complete  list  of  the  localities  from 
which  they  examined  specimens  but, 
unfortunately  this  list  does  not  serve 
as  an  effective  species  list  for  this 
study.  Since  the  intention  of  their 
study  was  a  survey  of  all  of  southern 
Florida,  they  did  not  collect  as  exten- 


Pralrie 


Prairie   30 

Pine  .736 

Hammocks   .608 

Disturbed  .314 

V   


hammock  disturbed 


prairie  pineland 


Figure  9.— Star  plot  diagranr>s  of  snake  data 
from  table  3,  comparing  frequencies  of 
trapping  (snakes  per  1000  array  days)  in 
ttie  four  tiabitat  types.  Genus  and  species 
names  abbreviated  on  key  at  bottom  of 
figure  correspond  to  acronynns  given  in 
table  3. 

sively  in  one  area  as  we  have  been 
able  to.  Nevertheless,  the  descrip- 
tions of  habitat  preferences  they  gave 
make  it  clear  that  a  few  more  species 
might  be  found  in  the  Long  Pine  Key 
region  if  I  continue  the  study.  There 
are  some  noticeable  absences  from 
their  list  for  the  Long  Pine  Key  and 
Paradise  Key  areas  however:  Storerk 
dekayi  and  Diadophis  punctatus.  It  is 
possible  that  these  species  were 
merely  overlooked  in  their  surveys 


Pine 

Hammocks 

Disturbed 

21 

20 

17 

22 

19 

16 

,303 

24 

17 

.253 

.589 

21 

Table  5.— Measures  of  similarity  among  arrays  grouped  by  vegetation  type 
based  on  data  from  table  3  (1984-1986,  above).  Numbers  above  the  di- 
agonal are  the  numbers  of  species  shared  between  habitats;  numbers 
along  the  diagonal,  boldfaced,  are  numbers  of  species  occurring  in  each 
habitat.  Numbers  below  the  diagonal,  underlined,  are  Morisita's  indices. 


80 


and  it  is  extremely  unlikely  that  these 
species  were  not  present  in  the  local 
area  thirty  years  ago  (Duellman  and 
Schwartz,  personal  communications). 

Salamanders  were  the  taxon  most 
poorly  represented  in  LPK,  only  four 
of  the  state's  24  salamanders  were 
found  in  southern  Florida  (table  1), 
and  only  three  of  these  were  found  in 
LPK.  The  reason  for  the  low  count  is 
obviously  the  low  elevation  and  poor 
soil  development  of  the  region. 

The  majority  of  Florida's  salaman- 
ders are  members  of  the  family  Ple- 
thodontidae,  and  this  family  is  pri- 
marily distributed  in  the  Appala- 
chian mountains  and  foothills  of  the 
eastern  U.S.  Many  species  are  stream 
dwellers,  others  are  forest  litter  in- 
habitants that  require  a  moist  thick 
leaf  litter  and  soil  development.  The 
mole  salamanders,  family  Ambysto- 
matidae,  also  require  soils  for  bur- 
rowing. Moreover,  salamander  lar- 
vae are  frequently  absent  from 
aquatic  settings  in  which  fish  are 
common. 

One  notable  exception  is  the  newts 
(family  Salamandridae),  but  even  the 
one  member  of  this  family  from  the 
region,  Notophthalmus  viridescens,  is 
rare.  The  only  successful  salaman- 
ders in  the  region  are  fully  aquatic, 
neotenic,  eel-like  animals:  Amphiuma 
means,  Siren  lacertina  and  Pseudobran- 
chus  striatus.  Their  cryptic  life  styles 
and  easy  access  to  the  underground 
aquifer  through  the  porous  limestone 
bedrock  may  be  important  reasons 
for  their  success. 

The  number  of  anuran,  lizard  and 
turtle  species  are  all  rather  low  in 
southern  Florida  (tables  1  and  2). 
Several  species  of  lizards  extend 
southward  past  the  mainland  into 
the  Rorida  Keys,  but  appear  to  have 
completely  by-passed  the  western 
extension  of  the  Miami  Rock  Ridge 
(in  particular  LPK)  e.g.  Eumeces 
egregius  and  Cnemidophorous  sexlinea- 
tus.  Two  species  are  endemic  to  the 
sandhills  and  scrub  habitats  of  Ror- 
ida  (Sceloporous  woodi  and  Neoseps 
reynoldsi)  and  their  absence  in  the 
area  is  again  probably  due  to  the  lack 


of  suitable  soils  and  substrates.  The 
reason  for  the  absence  of  the  other 
two  species  of  Ophisaurus  (O.  attenu- 
atus  and  O.  ventralis)  listed  by  Duell- 
man and  Schwartz  (1958)  is  not  clear, 
although  they  did  note  that  Ophisau- 
rus compressus  was  the  "most  abun- 
dant" of  the  three  species  in  southern 
Florida. 

The  only  notable  introduced  lizard 
was  Anolis  sagrei.  This  species  is  so 
common  in  southern  Florida  now 
that  it  is  no  surprise  that  large  popu- 
lations are  found  in  some  parts  of  the 
current  study  area  (Wilson  and  Por- 
ras  1983).  In  LPK  it  was  generally 
limited  to  areas  where  there  was  a 
greater  rate  of  contact  with  visitors, 
and  in  Disturbed  settings.  In  remote 
Hammocks  anoles  were  rarely  ob- 
served, but  Palma  Vista  I  and  Royal 
Palm  Hammocks  (both  sites  that  are 
popular  with  visitors  and  adjoin 
roads)  Anolis  sagrei  is  extremely  com- 
mon, as  well  as  throughout  the  hole- 
in-the-donut.  At  the  current  time  the 
park  appears  to  have  a  limited 
"load"  of  exotic  lizards.  Hemidactylus 
garnoti  was  observed  at  the  parking 
lot  at  Pahayokee  visitors  site,  and 
there  are  occasional  reports  of  this 
species  and  of  Anolis  equestris  in  the 
LPK  campground  area  and  the  "Pine 
Island"  residential  area  for  park 
staff. 

Of  the  few  specimens  of  Gopherus 
polyphemus  seen  during  the  study,  the 
only  one  from  the  study  area  was 
crossing  the  road  into  the  hole-in-the- 
donut  (several  others  were  seen  in 
the  Pine  Island  residential  area  and 
one  shell  was  near  a  pond,  but  no 
one  is  certain  of  the  source  of  these 
animals,  and  some  visitors  have  been 
known  to  release  gopher  tortoises 
near  the  entrance  to  the  park). 
Whether  the  sighting  within  the 
study  area  (the  turtle  was  measured, 
and  marked)  is  indicative  of  a  small 
population  or  is  a  captive  released  by 
a  visitor  is  not  at  all  clear. 

The  presence  of  a  population  of 
gopher  tortoises  on  Cape  Sable 
(Kushlan  and  Mazzotti  1985)  does 
not  help  in  explaining  the  single 


specimen,  and  Duellman  and 
Schwartz  (1958)  list  only  one  speci- 
men for  Dade  County.  Duellman  and 
Schwartz  (1958:260)  described  Ster- 
notherus  odoratus  as  "the  least  abun- 
dant of  the  three  southern  Florida 
kinosternids,"  and  I  have  found  it  in 
the  Shark  River  Slough  region  but 
not  LPK.  Kinosternon  subrubrum  is  de- 
scribed by  Duellman  and  Schwartz 
(1958:265)  as  avoiding  "the  main  part 
of  the  Everglades,  an  area  where  K. 
bauri  reaches  its  greatest  abundance. 
When  the  above  three  rare  species 
are  noted  the  turtle  list  for  Long  Pine 
Key  is  typical  of  the  southern  Rorida 
region. 

Some  of  the  species  listed  by  Du- 
ellman and  Schwartz  were  not  com- 
mon in  the  southern  everglades,  but 
were  found  in  other  areas  of  south- 
ern Rorida.  There  were  no  species  of 
anurans  that  I  expected  to  find  and 
did  not.  The  burrowing  nature  of 
Scaphiopus  holbrooki  probably  pre- 
vents it  from  being  common  in  LPK, 
and  it  was  never  seen  or  heard  dur- 
ing this  study. 

The  crocodilian  fauna  of  LPK  is 
composed  of  only  one  species,  the 
American  alligator  (although  there 
have  been  rare  occurrences  of  the 
American  crocodile,  Crocodylus 
acutus,  in  the  freshwater  reaches  of 
the  Taylor  Slough  drainage  in  the  vi- 
cinity of  the  study  area,  W.B. 
Robertson,  Jr.  pers.  comm).  The  alli- 
gator is  found  in  almost  every  place 
in  the  everglades  where  there  is  wa- 
ter. We  commonly  found  evidence  of 
alligators  in  the  seasonally  flooded 
Prairie  (alligator  trails)  and  in  the 
willow  heads  and  Hammocks  ("ga- 
tor holes,"  a  few  nests  seen,  juvenile 
and  adult  alligators  observed).  The 
LPK  region  is  certainly  peripheral  to 
the  main  distribution  of  the  species 
in  the  park. 

The  snake  fauna  is  clearly  the  best 
represented  fauna  in  LPK.  Of  the  26 
species  listed  for  southern  Florida,  21 
were  collected  during  the  study.  Of 
the  five  not  found  during  this  study 
only  one  was  expected,  Seminatrix 
pygaea,  and  the  technique  for  trap- 


81 


ping  this  species  described  by  Lo- 
raine  (1985)  will  be  tried  in  the  study- 
area  in  the  future.  Heterodon  platy rhi- 
nos was  described  by  Duellman  and 
Schwartz  (1958)  as  not  being  abun- 
dant in  southern  Florida,  and  there  is 
only  one  report  of  it  from  the  LPK 
area  (Roger  L.  Hammer  pers. 
comm.). 

Masticophis  flagellum  is  still  re- 
ported from  the  pineland  remnants 
of  southwest  Dade  County.  Duell- 
man and  Schwartz  (1958)  had  no  rec- 
ords of  this  species  from  the  park, 
but  since  then  there  has  been  one  rec- 
ord from  the  park. 

Pituophis  melanoleucus  was  repre- 
sented in  the  work  of  Duellman  and 
Schwartz  by  a  single  specimen  from 
Miami,  and  a  single  specimen  of  this 
species  was  collected  in  1984  in 
North  Miami  Beach.  The  snake  was 
probably  a  captive  pet  released  in  the 
area,  since  its  feces  contained  white 
mouse  remains  (Robert  J.  Nodell, 
pers.  comm.).  Tantilla  oolitica  (T. 
coromta  wagneri  of  Duellman  and 
Schwartz)  has  never  been  recorded 
from  the  park,  and  its  range  is  lim- 
ited to  isolated  Atlantic  Coastal 
Ridge  remnants  on  the  eastern  coast 
and  the  Florida  Keys  (Wilson  and 
Porras  1983). 


Habitat  Use  and  Preferences 

Within  the  LPK  region.  Prairie  habi- 
tat has  the  most  diverse  and  abun- 
dant herpetofauna.  The  Prairie  is  a 
broad  transition  zone  or  ecotone  be- 
tween the  longer  hydroperiod  Slough 
habitat  and  the  drier  Uplands,  and 
they  are  seasonally  inhabited  by  most 
species  from  those  two  habitats  as 
well  as  a  semi-aquatic  fauna  of  their 
own. 

Duellman  and  Schwartz  (1958:206- 
213)  characterized  the  habitats  of 
southern  Florida,  as  they  pertain  to 
Long  Pine  Key,  as:  Xeric  (including 
the  rocky  Pineland  of  Long  Pine 
Key),  Mesic  (including  the  tropical 
hardwood  Hammocks  of  Long  Pine 
Key),  and  Altemohygric  (including 


Prairie),  and  their  characterization 
for  each  species  is  given  in  table  2. 

All  of  the  18  species  that  Duellman 
and  Schwartz  (1958:211)  character- 
ized as  generalists  i.e.  "common  to 
all  three"  (i.e.  Prairie,  Pineland,  and 
Hammock)  were  found  in  Long  Pine 
Key.  Seventeen  of  the  21  species 
(81%)  they  characterized  as  inhabi- 
tants of  the  Prairie  (or  Altemohygric 
habitat)  were  found  in  the  study 
area. 

Only  9  of  the  22  species  (40%)  that 
Duellman  and  Schwartz  (1958:210) 
characterized  as  Xeric  or  Pineland 
species  are  found  in  the  region.  Four 
of  these  9  species  were  actually  more 
common  in  Hammocks  (Eleutherodac- 
tylus  planirostris,  Sphaerodactylus 
notatus,  Anolis  sagrei,  and  Micrurus 
fulvius),  one  (Scincella  laterale)  was 
common  in  Prairie,  three  were  rare 
(Gopherus  polyphemus,  Lampropeltis 
triangulum,  and  Cetnophora  coccinea) 
and  only  one  (Crotalus  adamanteus) 
was  actually  most  common  in  Pine- 
land (see  table  2). 

Using  the  species  associations  of 
Duellman  and  Schwartz  (1958),  of 
the  51  species  from  Long  Pine  Key, 
35%  (18)  are  generalists,  33%  (17)  are 
Prairie  species,  18%  (9)  are  Pineland 
or  Xeric  in  habitat  association,  6%  (3, 
Limmoedus  ocularis,  Pseudacris  nigrita 
and  Ophisaurus  compressus)  are  com- 
mon to  Prairie  and  Pineland,  6%  (3, 
Alligator  mississipiensis,  Trionyxferox 
and  Deirochelys  reticularia)  are  pri- 
marily Slough  or  Hygric  (Duellman 
and  Schwartz  1958:212),  and  2%  (1, 
Osteopilus  septentrionalis)  from  Edifi- 
carian-Ruderal  and  Hammock  (Me- 
sic) habitats. 

The  limit  to  the  preservation  of 
overall  diversity  of  the  Long  Pine 
Key  region  is  the  extent  of  rocky  Pi- 
neland habitat,  because  it  is  the  ma- 
jor habitat  type  of  the  area  with  the 
smallest  percentage  (40%)  of  its  her- 
petofauna (as  defined  by  Duellman 
and  Schwartz  1958)  represented.  It  is 
important  to  note  that  the  common 
use  of  interdigitating  finger  glades, 
i.e.  the  local  Prairie,  and  Hammocks 
by  some  of  the  Pineland  species 


makes  it  clear  that  overall  diversity 
depends  upon  continued  manage- 
ment to  preserve  the  current  patch- 
iness  of  the  area. 

Sixty  two  percent  of  the  species 
trapped  in  the  Disturbed  habitat  are 
characterized  as  generalists  by  Duell- 
man and  Schwartz  (1958),  14%  are 
from  Pineland  and  Prairie,  14%  are 
from  Pineland  and  10%  are  from 
Prairie. 

While  the  vast  majority  of  am- 
phibians and  reptiles  were  either 
trapped  and,  or  seen  in  the  Disturbed 
habitat,  a  few  were  rarely  or  never 
seen  in  the  Disturbed  habitat: 
Limnaoedus  ocularis,  Pseudacris  nigrita, 
Scincella  laterale  and  Sphaerodactylus 
notatus.  In  contrast  to  these  native 
species,  which  were  not  common  to 
the  Disturbed  habitat,  the  two  exotic 
species,  Osteopilus  septentrionalis  and 
Anolis  sagrei  were  most  common 
there. 

Species  composition  of  the  Dis- 
turbed habitat  primarily  depends  on 
the  historical  topography  of  the  area. 
The  vast  majority  of  species  there  are 
generalists,  but  the  area  is  large 
enough  that  local  variations  in  hy- 
droperiod attract  a  number  of  species 
more  commonly  associated  with 
drier  or  wetter  conditions  and  future 
analyses  of  this  very  complex  area 
will  involve  a  more  specific  separa- 
tion of  habitat  types  within  the  area. 
Clearly,  most  of  the  species  of  am- 
phibians and  reptiles  are  responding 
to  basic  microhabitat  requirements 
that  have  little  to  do  with  the  actual 
species  composition  of  the  vegetation 
(Campbell  and  Christman  1 982a:  170- 
171). 


Abundance 

It  is  impossible  to  accurately  com- 
pare the  trapping  results  of  this 
study  to  other  studies.  The  methods, 
objectives  and  local  circumstances  of 
each  study  vary  widely.  Perhaps 
most  confounding  is  the  variability  in 
the  number  of  months  per  year  dur- 
ing which  species  are  active,  and  this 


82 


makes  comparisons  based  on  animals 
per  check  day  difficult.  There  are 
also  differences  in  types  of  arrays 
used,  the  purposes  of  the  trapping 
effort,  substrate  characteristics  and 
ability  to  use  pit  traps,  all  of  which 
preclude  valid  comparisons. 
Campbell  and  Christman  (1982b) 
summarized  their  results  from  north- 
ern Florida,  in  which  they  operated 
30  arrays  for  7432  array-days.  They 
collected  1644  animals  of  43  species 
from  11  habitats  for  an  average  of 
0.22  animals  per  array-day.  In  LPK, 
13  arrays  operated  a  total  of  9792  ar- 
ray-days and  collected  1709  animals 
of  37  species  in  4  habitats  for  an  aver- 
age of  0.18  animals  per  array  day,  a 
similar  catch  rate  per  array  day. 

Campbell  and  Christman  (1982b) 
used  both  funnel  traps  and  pit  traps, 
and  they  estimated  that  only  36%  of 
their  collection  came  from  funnel 
traps.  They  also  state  that  69%  of  the 
animals  trapped  were  Eleutherodacty- 
lus  planirostris,  and  that  90%  of  their 
trappings  were  of  E.  planirostris  and 
Gastrophryne  carolinensis.  Both  of 
these  species  were  readily  trapped  in 
their  pit  traps.  If  their  pit  trap  ex- 
cluded, and  look  at  the  percent  from 
funnel  traps,  there  was  a  much  trap 
yield. 

There  are  so  many  differences  in 
the  two  studies  that  the  only  conclu- 
sion to  be  drawn  is  that  the  results 
compare  favorably  with  that  the  LPK 
region  has  a  moderate  diversity  and 
comparable  abundance  of  animals, 
based  upon  similar  trapping  effort. 

Comparisons  to  other  studies  are 
even  more  difficult,  since  studies  in 
more  temperate  climates  are  done 
only  during  the  warmer  months  of 
the  year.  For  example,  Clawson  and 
Baskett  (1982),  in  Missouri,  used  13 
arrays  a  total  of  3159  array  days  in 
the  spring,  summer,  and  fall,  and 
captured  2545  animals,  for  an  aver- 
age of  0.81  animals  per  array  day. 
This  much  higher  figure  may  well  be 
representative  of  the  greater  concen- 
tration of  both  animals  and  resources 
typically  found  in  more  temperate 
climes. 


Species  Diversity 

Species  richness  for  southern  Florida 
was  described  by  Duellman  and 
Schwartz  (1958:205)  as  "depauper- 
ate" and  "impoverished."  They  state 
that  "an  impoverished  herpetofauna 
is  what  might  be  expected  at  the  end 
of  a  long  peninsula,  through  the 
length  of  which  certain  habitats  and 
their  inhabitants  disappear." 

The  difficulty  in  evaluating  this 
statement  arises  from  the  fact  that 
there  is  much  more  involved  in  the 
biogeography  of  the  peninsula  of 
Florida  than  a  simple  "peninsula  ef- 
fect" due  to  reduced  area  and  dis- 
tance from  centers  of  distribution 
(Robertson  and  Kushlan  1984).  There 
is  also  the  recent  geological  origin  of 
the  land  area,  the  poor  development 
of  soils  in  the  area  during  the  time 
since  emergence,  the  lack  of  variation 
in  relief  of  the  area  (Olmsted  and 
Loope  1984),  and  the  severe  human 
disturbance.  All  of  these  factors  need 
to  be  considered  in  evaluating  the 
possible  reasons  for  an  "impover- 
ished" fauna.  Finally  there  is  the  is- 
sue of  deciding  whether  the  fauna 
deserves  the  label  of  "impoverished" 
in  the  first  place. 

A  reduced  species  list  does  not  by 
itself  determine  whether  the  biomass 
of  the  existing  species  is  high  or  low, 
e.g.  while  the  species  list  for  fresh 
water  fish  is  considered  low  for  the 
area  (Loftus  and  Kushlan  1987)  they 
are  the  principal  food  of  an  enor- 
mous biomass  of  wading  birds. 
Robertson  and  Kushlan  (1984:234) 
have  addressed  this  point:  "...the 
nearly  unique  ability  of  the  South 
Florida  ecosystem  to  support  such 
large  numbers  of  14  species  of  super- 
ficially similar  secondary  and  tertiary 
consumers  on  a  resource  base  that  is 
reduced  in  species  diversity  by  bio- 
geographic  factors  is  generally  unap- 
preciated." and  the  nesting  efforts 
(1972  or  1974  numbers)  of  the  White 
Ibis  and  Wood  Storks  alone  are  esti- 
mated to  have  required  "in  excess  of 
3  billion  kilocalories  or  approxi- 
mately 2500  metric  tons  of  food..." 


As  the  impact  of  the  remaining  12 
species  of  wading  birds  is  not  known 
and  the  secondary  productivity  of 
South  Florida  habitats  has  not  yet 
been  studied,  the  meaning  of  this  en- 
ergy requirement  to  the  total  system 
is  undeterminable." 

During  this  study  we  have  col- 
lected data  on  51  species  of  amphibi- 
ans and  reptiles  (table  2).  This  is  not 
a  low  figure  for  an  area  the  size  of 
LPK  (8000  ha). 

Vogt  and  Hine  (1982)  list  34  spe- 
cies of  amphibians  and  reptiles  from 
their  study  area  in  southern  Wiscon- 
sin. Clawson  and  Baskett  (1982)  list 
35  species  from  their  Missouri  study 
area.  Clarke  (1958)  lists  39  species 
from  Osage  County,  Kansas.  In  trap- 
ping studies  in  the  Florida  sandhills 
of  Tampa,  Mushinsky  (1985)  lists  27 
species.  Campbell  and  Christman 
(1982b)  list  60  species  from  their  ex- 
tensive study  in  northern  Florida, 
and  this  number  comes  from  a  vari- 
ety of  sampling  techniques  in,  at 
least,  11  different  habitat  types. 

Gibbons  and  Harrison  (1981)  list 
68  species  from  coastal  mainland 
South  Carolina  and  Gibbons  and  Pat- 
terson (1978)  list  94  species  from  the 
Savannah  River  Plant  in  South  Caro- 
lina. Myers  and  Rand  (1969)  list  100 
species  for  Barro  Colorado  Island, 
Panama.  Crump  (1971)  lists  116  spe- 
cies for  the  Belem  area  of  Brazil. 

From  the  temperate  to  tropic  lati- 
tudes there  is  an  obvious  increase  in 
overall  diversity,  but  the  species  rich- 
ness for  the  LPK  is  not  very  low  for 
its  latitude.  The  presence  of  51  spe- 
cies and  the  fact  that  many  are  abun- 
dant makes  it  clear  that  the  applica- 
tion of  terms  such  as  impoverished 
or  depauperate  must  be  used  in  con- 
text. Rather  than  pondering  the  ab- 
sence of  some  species  (especially 
when  for  the  group  with  the  least 
representation  in  the  area,  the  sala- 
manders, it  is  quite  clear  why  they 
are  not  common,  see  above)  I  find 
myself,  like  Robertson  and  Kushlan 
(1984,  above),  more  impressed  with 
the  actual  abundance  of  animal  life  in 
this  unique  area. 


83 


Conclusions 

1.  The  species  list  for  the  LPK 
includes  at  least  51  species, 
15  species  of  amphibians  and 
36  species  of  reptiles.  The 
most  poorly  represented 
group  is  the  salamanders,  the 
best  represented  group  is  the 
snakes.  The  survey  of  current 
species  composition  is  basi- 
cally the  same  as  reported  30 
years  ago  for  the  area  by  Du- 
ellman  and  Schwartz  (1958). 
The  fact  that  there  has  been 
no  reduction  in  species  rich- 
ness of  the  local  area  should 
be  considered  a  major  benefit 
of  the  preservation  of  the  re- 
gion inside  the  national  park. 

2.  Amphibians  and  reptiles  of 
LPK  are  primarily  habitat 
generalists,  usually  being 
found  in  three  of  the  four 
major  habitat  types  in  the 
area.  The  principal  separa- 
tion by  habitat  is  related  to 
the  characteristics  of  the  sub- 
strate, there  being  a  subset  of 
herptiles  most  commonly 
found  in  areas  with  greater 
soil  development  (Ham- 
mocks and  the  Disturbed  ar- 
eas) and  another  subset  of 
herptiles  that  are  more  com- 
mon in  seasonally  flooded 
Prairie.  The  most  poorly  rep- 
resented group  is  that  de- 
scribed as  primarily  from 
Xeric,  Pineland  habitat,  and 
the  absence  of  sandy  soils  in 
the  rocky  Pineland  makes 
this  the  most  fragile  compo- 
nent of  the  Everglades  herpe- 
tofauna.  The  findings  of  this 
study  do  not  differ  signifi- 
cantly from  those  of  Duell- 
man  and  Schwartz  (1958) 
from  thirty  years  ago.  The 
results  point  out  that  there  is 
a  significant  portion  of  the 
local  herpe  to  fauna  that  relies 
upon  the  preservation  of 
large  contiguous  areas  of  na- 


tive Pineland  interspersed 
with  Hammocks  and  season- 
ally flooded  Prairie  for  its 
continued  success. 

3.  Phenologies  of  amphibians 
and  reptiles  of  the  LPK  can 
be  described  as  modified 
temperate  zone  patterns. 
While  the  subtropical  charac- 
ter of  the  southern  coastal 
portion  of  peninsular  Rorida 
results  in  a  year  long  grow- 
ing season,  with  only  occa- 
sional frosts,  the  seasonality 
of  rainfall  and  the  temperate 
zone  origin  of  the  herpe- 
tofauna  results  in  a  tradi- 
tional spring  emergence  of 
the  herptiles,  tied  to  increas- 
ing day  length,  warmer  tem- 
peratures and  the  onset  of 
heavy  rainfall. 

4.  Estimates  of  density  and 
relative  abundance  remain 
difficult  to  give  at  the  current 
time.  Comparison  of  current 
trapping  results  with  those  of 
Campbell  and  Christman 
(1982a,  1982b)  from  11  habi- 
tats in  northern  Florida  indi- 
cate a  similar  level  of  abun- 
dance for  the  two  areas,  but 
differences  in  the  actual  spe- 
cies lists,  habitat  types  and 
methodologies  make  such 
conclusions  tenuous.  Com- 
parisons of  the  fauna  of  the 
area  with  those  of  a  wide  va- 
riety of  other  regions  indicate 
that  the  herpetofauna  of 
LPK,  with  the  exception  of 
the  salamanders,  has  a  mod- 
erate level  of  diversity. 

Acknowledgnrjents 

I  wish  to  thank  all  the  dedicated  stu- 
dents of  ecology  and  herpetology  at 
F.I.U.  who  gave  their  time  so  will- 
ingly during  the  study.  Doug  Barker, 
Peter  Beck,  Laura  Brandt,  Teresa  De- 
Francesco,  Bob  Dunne,  Ernesto  Her- 


nandez, Liz  Lewis,  Nancy  O'Hare, 
and  Arlene  Sackman  helped  with 
trap  checking  and  collecting.  To 
Frank  S.  Bernardino,  Jr.,  Bob  Nodell, 
Todd  Steiner,  and  Joe  Wasilewski  I 
owe  a  great  debt  for  their  dedication 
to  the  field  work.  I  thank  David  But- 
ler and  SARLON  Industries  for  the 
donation  of  the  shade  cloth  used  to 
make  the  fencing,  and  David  W.  Lee, 
of  F.I.U.,  for  suggesting  the  use  of 
shade  cloth  to  us.  I  thank  the  staffs  of 
the  South  Florida  Research  Center 
and  the  Division  of  Resources  Man- 
agement of  the  park  for  their  pa- 
tience, generosity,  perspectives  and 
spontaneous  collection  of  specimens 
for  our  studies.  Most  of  all  I  wish  to 
thank  Gary  Hendrix  and  William  B. 
Robertson,  Jr.  for  their  interest  and 
support. 

This  research  was  sponsored  by 
the  U.S.  National  Park  Service-Flor- 
ida International  University  Coop- 
erative Agreement  (CA-5000-3-8005, 
Supplemental  Agreement  No.2, 1984) 
and  the  Horida  International  Univer- 
sity Foundation. 

Literature  Cited 

Aleksiuk,  Michael,  and  Patrick  T. 
Gregory.  1974.  Regulation  of  sea- 
sonal mating  behavior  in  Thamno- 
phis  sirtalis  parietalis.  Copeia 
1974:681-689. 

Auffenberg,  Walter.  1982.  Horida 
environments  and  their  herpe- 
tofaunas.  Part  III.  Herpe togeogra- 
phy.  Florida  Herpetologist.4:l-36. 

Brower,  James  E.,  and  Jerrold  H.  Zar. 
1984.  Field  and  ecological  methods 
for  general  ecology.  Wm  C. 
Brown.  Dubuque.  226p. 

Brown,  William  S,  and  William  S. 
Parker.  1976.  A  ventral  scale  clip- 
ping system  for  permanently 
marking  snakes  (Reptilia,  Serpen- 
tes).  Journal  of  Herpetology 
10:247-249. 

Brown,  William  S.,  and  William  S. 
Parker.  1982a.  Niche  dimensions 
and  resource  partitioning  in  a 
Great  Basin  Desert  snake  commu- 


84 


nity.  p.59-81.  In  Herpetological 
Communities.  Norman  J.  Scott,  Jr., 
editor.  U.S.  Fish  and  Wildlife  Serv- 
ice, Wildlife  Research  Report  13. 
Washington  D.C. 

Brown,  William  S.,  and  William  S. 
Parker.  1982b.  Growth,  reproduc- 
tion and  demography  of  the  racer. 
Coluber  constrictor  mormon,  in 
northern  Utah.  In  Contributions  to 
vertebrate  ecology  and  sys- 
tematics:  a  tribute  to  Henry  S. 
Fitch.  Richard  A.  Seigel  and  oth- 
ers, editors.  Museum  of  Natural 
History  University  of  Kansas,  Spe- 
cial Publication  No.  10.  Lawrence. 

Busack,  Stephen  D.  and  S.Blair 
Hedges.  1984.  Is  the  peninsular 
effect  a  red  herring?  American 
Naturalist  123:266-275. 

Campbell,  Howard  W.,  and  Steven  P. 
Christman.  1982a.  The  herpeto- 
logical components  of  Florida 
sandhill  and  sand  pine  scrub  asso- 
ciations, p. 163-171.  In  Herpetologi- 
cal Communities.  Norman  J.  Scott, 
Jr.,  editor.  U.S.  Fish  Wildlife  Serv- 
ice Wildlife  Research  Report  13. 
Washington  D.C. 

Campbell,  Howard  W.,  and  Steven  P. 
Christman.  1982b.  Field  tech- 
niques for  herpetofaunal  commu- 
nity analysis,  p.193-200.  In  Herpe- 
tological Communities.  Norman  J. 
Scott,  Jr.,  editor  U.S.  Fish  and 
Wildlife  Service  Wildlife  Research 
Report  13.  Washington  D.C. 

Carpenter,  Charles  C.  1952.  Com- 
parative ecology  of  the  common 
garter  snake  (Thamnophis  s.  sir- 
talis),  the  ribbon  snake  (Thamno- 
phis s.  sauritus)  and  Butler's  garter 
snake  (Thamnophis  butleri)  in 
mixed  populations.  Ecological 
Monographs  22:235-258. 

Chambers,  John  M.,  William  S. 
Cleveland,  Beat  Kleiner  and  Paul 
A.  Tukey.  Graphical  Methods  for 
Data  Analysis,  p.  395.  Duxbury. 
Boston,  Mass. 

Clark,  Donald  R.  Jr.  1970.  Ecological 
study  of  the  worm  snake,  Car- 
phophis  vermis  (Kenicott).  Museum 
of  Natural  History  Publication, 
University  of  Kansas  19:85-194. 


Clark,  Robert  F.  1958.  An  ecological 
study  of  reptiles  and  amphibians 
in  Osage  County,  Kansas.  The 
Emporia  State  Research  Stud- 
ies.7:l-52. 

Clawson,  Mary  E.,  and  Thomas  S. 
Baskett.  1982.  Herpetofauna  of  the 
Ashland  Wildlife  Area,  Boone 
County,  Missouri.  Transactions, 
Missouri  Acad,  of  Science  16:5-16. 

Cook,  R.E.  1969.  Variation  in  species 
density  in  North  American  birds. 
Systematic  Zoology  18:63-84. 

Crump,  Martha  L.  1971.  Quantitative 
analysis  of  the  ecological  distribu- 
tion of  a  tropical  herpetofauna. 
Museum  of  Natural  History,  Uni- 
versity of  Kansas,  Occasional  Pa- 
per 3:1-62. 

Crump,  Martha  L.  1982.  Amphibian 
reproductive  ecology  on  the  com- 
munity level,  p. 21 -36.  In  Herpeto- 
logical Communities.  Norman  J. 
Scott,  Jr.,  editor.  U.S.Fish  and 
Wildlife  Service  Wildlife  Research 
Report  13.  Washington  D.C. 

Davis,  J. H.  1943.  The  natural  features 
of  southern  Rorida.  Geological 
Bulletin  25,  Rorida  Geological 
Survey. 

Dinnen,  J.  Walter.  1984.  The  fishes  of 
the  everglades,  p.  258-268.  In  Envi- 
ronments of  South  Rorida  Present 
and  Past  II.  Patrick  J.  Gleason,  edi- 
tor. Miami  Geol.  Soc,  Miami. 

Duellman,  William  E.,  and  A. 
Schwartz.  1958.  Amphibians  and 
reptiles  of  southern  Rorida.  Bulle- 
tin of  the  Rorida  State  Museum 
3:181-324. 

Ewel,  John  J.,  Dennis  S.  Ojima,  Dori 
A.  Karl,  and  William  F.  DeBusk. 
1982.  Schinus  in  successional  eco- 
systems of  Everglades  National 
Park.  Report  T-676.  Homestead, 
Rorida:  Everglades  National  Park. 
South  Florida  Research  Center. 

Gibbons,  J.  Whitfield,  and  Karen  K. 
Patterson.  1978.  The  reptiles  and 
amphibians  of  the  Savannah  River 
Plant.  Savannah  River  Ecology 
Plant.  DuPont  Press,  SRO-NERP-2: 
1-24. 

Gibbons,  J.  Whitfield,  and  Julian  R. 
Harrison  III.  1981.  Reptiles  and 


amphibians  of  Kiawah  and  Caper 
Islands,  South  Carolina.  Brimley- 
ana  5:145-162. 

Gibbons,  J.  Whitfield,  and  Raymond 
D.  Semlitsch.  1981.  Terrestrial  drift 
fences  with  pitfall  traps:  an  effec- 
tive technique  for  quantitative 
sampling  of  animal  populations. 
Brimleyana.  7:1-16. 

Hamilton,  W.J.  1951.  The  food  and 
feeding  behavior  of  the  garter 
snake  in  New  York  State.  Ameri- 
can Midland  Naturalist  46:385- 
390. 

Heatwole,  Harold.  1982.  A  review  of 
structuring  in  herpetofaunal  as- 
semblages. p.1-19.  In  Herpetofau- 
nal Communities.  Norman  J.  Scott, 
Jr.,  editor.  U.S.  Fish  and  Wildlife 
Service  Wildlife  Research  Report 
13.  Washington  D.C. 

Horn,  Henry  S.  1966.  Measurement 
of  "overlap"  in  comparative  eco- 
logical studies.  American  Natural- 
ist 100:419-424. 

Johnson,  J.M.,  I.C.  Olmsted,  and  O.L. 
Bass  Jr.  1983.  Vegetation  map  of 
Long  Pine  Key,  Everglades  Na- 
tional Park.  South  Rorida  Re- 
search Center.  U.S.  National  Park 
Service.  Homestead. 

Kiester,  A.  Ross  1971.  Species  diver- 
sity of  North  American  amphibi- 
ans and  reptiles.  Systematic  Zool- 
ogy 20:127-137. 

Kushlan,  James  A.  and  Frank  J.  Maz- 
zotti.  1985.  Environmental  effects 
on  a  coastal  population  of  gopher 
tortoises.  Journal  of  Hepetology 
18:231-239. 

Layne,  James  N.  1984.  The  land 
mammals  of  south  Florida,  p.269- 
296.  In  Environments  of  South 
Rorida  Present  and  Past  II.  Patrick 
J.  Gleason,  editor.  Miami  Geologi- 
cal Society,  Miami. 

Loftus,  William  F.  and  James  A. 
Kushlan.  1987.  Freshwater  fishes 
of  southern  Rorida.  Bulletin  of  the 
Rorida  State  Museum,  Biological 
Sciences.  31:147-344. 

Loraine,  Raymond  K.  1985.  Seasonal 
changes  in  foraging  success  and 
diet  composition  of  Seminatrix 
pygaea.  Abstract  from  1985  com- 


85 


bined  meeting  of  the  Society  for 
the  Study  of  Amphibians  and  Rep- 
tiles and  Herpetologisf  s  League, 
Tampa,  Florida.  August  4-9, 1985. 

MacArthur,  R.H.,  and  Wilson.  E.O. 
1967.  The  theory  of  island  biogeo- 
graphy.  Princeton  University 
Press.  Princeton  N.J. 

Moler,  Paul  E.  1985.  A  new  species  of 
frog  (Ranidae:  Ram)  from  north- 
western Florida.  Copeia. 1985:379- 
383. 

Mushinsky,  Henry  R.  1985.  Fire  and 
the  Rorida  sandhill  herpetofaunal 
community:  with  special  attention 
to  responses  of  Cnemidophorus 
sexlineatus.  Herpetologica  41:333- 
342. 

Myers,  Charles  W.,  and  A.  Stanley 
Rand.  1969.  Checklist  of  amphibi- 
ans and  reptiles  of  Barro  Colorado 
Island,  Panama,  with  comments 
on  faunal  change  and  sampling. 
Smithsonian  Contributions  to  Zo- 
ology. 10:1-11.  Smithsonian  Insti- 
tution Press.  Washington  D.C. 

Olmsted,  Ingrid  C,  Lloyd  L.  Loope, 
and  Charles  E.  Hilsenbeck.  1980. 
Tropical  hardwood  Hammocks  of 
the  interior  of  Everglades  National 
Park  and  Big  Cypress  National 
Preserve.  National  Park  Service, 
South  Florida  Research  Center 
Report  T-604.  Everglades  National 
Park,  Homestead,  FL.  58  p. 

Olmsted,  Ingrid  C,  William  B. 
Robertson  Jr.,  Jill  Johnson,  and 
Oron  L.  Bass  Jr.  1983.  The  vegeta- 
tion of  Long  Pine  Key,  Everglades 
National  Park.  National  Park  Serv- 
ice, South  Horida  Research  Center 
Rep.  Sn<C-83/05.  Everglades  Na- 
tional Park,  Homestead,  FL.  64  p. 

Olmsted,  Ingrid  C.  and  Lloyd  L. 
Loope.  1984.  Plant  Communities 
of  Everglades  National  Park.  p. 
167-184.  In  Environments  of  South 
Horida  Present  and  Past  II.  Patrick 
J.  Gleason,  editor.  Miami  Geologi- 
cal Society,  Miami. 

Preston,  F.W.  1962.  The  canonical 
distribution  of  commonness  and 
rarity.  Ecology .43:185-215. 

Reichenbach,  Norman  G.,  and 
George  H.  Dalrymple.  1986.  En- 


ergy use,  life  histories,  and  the 
evaluation  of  potential  competi- 
tion in  two  species  of  garter  snake. 
Journal  of  Herpetology.  20:131- 
151. 

Robertson,  William  B.  Jr.  1955.  An 
analysis  of  the  breeding-bird 
populations  of  tropical  Florida  in 
relation  to  the  vegetation.  Unpub- 
lished Ph.D.  thesis.  University  of 
Ilhnois.  Champa ign-Urbana. 
599  p. 

Robertson,  William  B.  Jr.,  and  James 
A.  Kushlan.  1984.  The  southern 
Horida  avifauna,  p.  219-257.  In 
Environments  of  South  Horida 
Present  and  Past  II.  P.J.  Gleason, 
editor.  Miami  Geological  Society. 
Miami. 

Rose,  Paul  W.  1977.  Annual  hydrol- 
ogy review:  Everglades  National 
Park  and  Big  Cypress  Preserve 
1976.  Everglades  National  Park 
Report  T-505.  U.S.  National  Park 
Service.  Homestead.  FL.  126  p. 

Simpson,  G.G.  1964.  Species  densities 
of  North  American  recent  mam- 
mals. Systematic  Zoology  13:57-73. 

Snyder,  James  R.  1986.  The  impact  of 
wet  season  and  dry  season  pre- 
scribed fires  on  Miami  Rock  Ridge 
pineland.  Everglades  National 
Park.  National  Park  Service,  South 
Horida  Research  Center  Report 
SFRC-86/06.  Everglades  National 
Park,  Homestead,  Fl.  106  p. 

Taylor,  Dale  L.,  and  Alan  Hemdon. 

1981.  Impact  of  22  years  of  fire  on 
understory  hardwood  shrubs  in 
slash  pine  communities  within  Ev- 
erglades National  Park.  National 
Park  Service,  South  Horida  Re- 
search Center  Report  T-640.  Ever- 
glades National  Park,  Homestead, 
H.  30  p. 

Turner,  F,B.  1977.  The  dynamics  of 
populations  of  squamates,  croco- 
dilians  and  rhyncocephalians.  p. 
157-264.  In  Biology  of  the  Reptilia. 
Carl  Gans  and  Donald  W.  Tinkle, 
editors.  Academic  Press.  New 
York. 

Vogt,  Richard  C,  and  Ruth  L.  Hine. 

1982.  Evaluation  of  techniques  for 
assessment  of  amphibian  and  rep- 


tile populations  in  Wisconsin. 
p.201-217.  In  Herpetological  Com- 
munities. Norman  J.  Scott,  Jr.,  edi- 
tor. U.S.  Fish  and  Wildlife  Service 
Wildlife  Research  Report  13. 
Washington  D.C. 

Wiest,  John  A.  Jr.  1982.  Anuran  Suc- 
cession at  temp)orary  ponds  in  a 
post  oak-savanna  region  of  Texas. 
p.39-47.  In  Herpetological  Com- 
munities. Norman  J.  Scott,  Jr.,  edi- 
tor. U.S.  Fish  and  Wildlife  Service 
Wildlife  Research  Report  13. 
Washington  D.C. 

Wilson,  Larry  David,  and  Louis  Por- 
ras.  1983.  The  ecological  impact  of 
man  on  the  south  Horida  herpe- 
tofauna.  Museum  of  Natural  His- 
tory, University  of  Kansas,  Special 
Publication  9.  89+vi  p. 

Woodbury,  A.M.  1951.  Introduction — 
a  ten  year  study.  Herpetologica. 
7:4-14. 


86 


The  Herpetofaunal 
Community  of  Temporary 
Ponds  in  North  Florida  Sand- 
hills: Species  Composition, 
Temporal  Use,  and 
Management  Implications^ 

C.  Kenneth  Dodd,  Jr.^  and  Bert  G.  Ctiarest^ 


Abstract.— Amphibians  and  reptiles  use  an 
isolated  temporary  wetland  in  a  north  Florida 
sandhills  throughout  the  year  despite  variation  in 
environmental  conditions.  Species  composition  and 
number  of  individuals  varies  seasonally  and 
annually.  Temporal  variation  in  habitat  use  must  be 
considered  in  managing  small  wetlands  and 
assessing  their  importance  to  the  herpetofaunal 
community. 


The  sandhills  and  xeric  live  oak  her- 
petofauna  of  Florida  is  diverse  and 
contains  a  number  of  endemic  spe- 
cies. Whereas  the  terrestrial  herpe- 
tofauna  has  been  described  for  a  few 
sandhills  communities  (Campbell 
and  Christman  1982,  Mushinsky 
1985),  there  have  been  no  long-term 
studies  of  the  ecology  of  species  us- 
ing temporary  ponds.  For  breeding 
amphibians,  sandhills  temporary 
ponds  are  often  the  only  sources  of 
water  that  are  free  of  predatory  fish 
and  many  larger  predatory  insects, 
and  such  ponds  may  be  extremely 
important  for  amphibian  reproduc- 
tive success  (Ma can  1966,  Sexton  and 
Phillips  1986,  Semlitsch  1987,  Moler 
and  Franz  1988).  At  the  same  time, 
the  ephemeral  nature  of  these  breed- 
ing sites  makes  reproductive  success 
uncertain  and  thus  provides  an  op- 
posing selective  pressure  for  their 
use  (Semlitsch  1987). 

Since  January  1985,  we  have  been 
conducting  studies  on  the  herpe- 
tofaunal community  at  a  temporary 
pond  in  a  north-central  Florida  long- 
leaf  pine-turkey  oak  ("high  pine") 

'Paper  presented  at  symposiurr).  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff .  AZ,  July  19-21,  1988). 

'C.  Kenneth  Dodd,  Jr.  is  Zoologist  (Re- 
search), National  Ecology  Research  Center, 
U.S.  Fish  and  Wildlife  Service,  412  N.E.  16th 
Avenue,  Room  250,  Gainesville,  FL  32601. 

^Bert  G.  Charest  is  Biological  Aid  (Wild- 
life), National  Ecology  Research  Center, 
U.S.  Fish  and  Wildlife  Sen/ice,  412  N.E.  16th 
Avenue,  Room  250,  Gainesville,  FL  32601. 


sandhills.  Little  is  known  of  the  com- 
position of  such  Florida  herpetofau- 
nal communities,  although  Moler 
and  Franz  (1988)  reported  16  anuran 
species  breeding  in  various  types  of 
wetlands  surrounded  by  sandhills  on 
the  3750  ha  Katharine  Ordway  Pre- 
serve-Swisher  Memorial  Sanctuary  in 
Putnam  County.  Nothing  is  known 
about  movement  patterns  and  activ- 
ity cycles  of  the  herpetofauna,  or 
about  the  numbers  of  individuals 
breeding  at  such  ponds  and  the  num- 
bers of  offspring  produced. 

The  purposes  of  our  study  are  to 
gain  insight  into  the  structure  of  the 
herpetofaunal  community  using  a 
temporary  pond  in  a  sandhills  eco- 
system, to  assess  variation  in  species 
composition  and  temporal  use  of  the 
pond,  and  to  gather  basic  biological 
information  on  the  species  that  com- 
prise the  community.  This  paper 
presents  findings  based  on  two  years 
of  fieldwork  of  a  projected  five  year 
study. 

Methods 

Breezeway  Pond,  a  0.16  ha  isolated 
temporary  pond  in  a  shallow  1.3  ha 
basin  on  the  Katharine  Ordway  Pre- 
serve-Swisher  Memorial  Sanctuary, 
Putnam  County,  Florida,  was  en- 
circled with  a  230  m  drift  fence 
(mean  height  =  36  cm  above  the  sub- 
strate) following  the  general  proce- 
dure of  Gibbons  and  Semlitsch 
(1982),  reviewed  by  Jones  (1986a). 


Buckets  were  spaced  at  10  m  inter- 
vals and  paired  on  opposite  sides  of 
the  fence,  making  23  stations  of  two 
buckets  each.  Sloping  covers  were 
put  over  the  buckets  and  wet 
sponges  were  placed  in  them  to  mini- 
mize exposure  to  direct  rays  of  the 
sun  and  desiccation,  respectively.  As 
a  result,  mortality  among  captured 
animals  was  <  1.0%  and  was  caused 
primarily  by  invertebrate  predation 
(spiders,  ants,  centipedes,  and 
beetles). 

Breezeway  Pond  is  located  at  an 
ecotone.  To  the  immediate  south  and 
west,  the  predominant  habitat  is 
"high  pine"  sandhills  dominated  by 
longleaf  pine  (Pinus  palustris),  turkey 
oak  (Quercus  laevis)  and  wiregrass 
(Aristida  striata).  A  xeric  hammock 
dominated  by  sand  live  oak  (Q.  gemi- 
nata)  and  laurel  oak  (Q.  laurifolia) 
faces  the  north,  while  a  small  "Pani- 
cum  meadow"  dominated  by 
maidencane  (Panicum  hemitomon), 
lies  to  the  east.  The  distance  from  the 
drift  fence  to  the  nearest  forested 
plant  association  is  no  more  than 
about  50  m  in  any  direction. 

Buckets  were  checked  5  days  per 
week  in  the  morning  (beginning 
0700-0900  h  depending  on  season) 
from  January  16  through  April  12, 
1985,  and  from  October  1,  1985,  until 
September  30, 1987.  For  purposes  of 
discussion  and  analysis,  a  year  refers 
to  a  12-month  period  from  October 
through  the  following  September 
(e.g.  1986  =  October  1985  through 
September  1986)  because  reproduc- 


87 


tion  and  metamorphosis  generally 
cease  in  early  autumn  while  winter 
breeding  has  yet  to  commence. 

All  reptiles  and  amphibians  were 
measured  in  the  field  (snout-vent 
length,  carapace  and  plastron  length 
[for  turtles],  tail  length  [for  snakes 
and  glass  lizards]),  weighed  and 
marked  for  future  identification  us- 
ing a  year  code  (e.g.,  0022  identifies 
animals  marked  in  1986)  or  an  indi- 
vidual identification  number  (all 
turtles,  snakes,  gopher  frogs  [Ram 
areolata],  red-tailed  skinks  [Eumeces 
egregius],  and  ground  skinks  [Scin- 
cella  lateralis]).  Very  small  animals, 
mostly  juvenile  frogs  and  lizards, 
were  not  marked  because  of  their  ex- 
tremely small  toes. 

Notes  were  recorded  on  tail  regen- 
eration and  damage,  breeding  and 
hatchling  coloration,  and  reproduc- 
tive status.  All  animals,  except  liz- 
ards, were  released  on  the  opposite 
side  of  the  fence  from  site  of  capture; 
lizards  were  released  on  the  same 
side  as  captured.  Weather  condi- 
tions, rainfall,  pond  water  level,  and 
maximum  and  minimum  air  and  wa- 
ter temperatures  were  recorded. 
These  data  are  similar  to  those  re- 
corded in  other  long-term  studies 
employing  drift  fences  to  study  am- 
phibian communities  (Gibbons  and 
Bennett  1974,  Gibbons  and  Semlitsch 
1982)  and  are  vital  to  the  inventory 
and  management  of  ecological  com- 
munities and  individual  species 
aones  1986b). 

In  this  paper  we  concentrate  our 
analyses  on  the  two  most  commonly 
captured  amphibians,  the  striped 
newt  (Notophthalmus  perstriatus),  a 
species  listed  as  of  special  concern  in 
Florida  (Christman  and  Means  1978), 
and  the  eastern  narrow-mouthed 
toad  (Gastrophryne  carolinensis),  a 
common  Florida  frog  (Carr  1940). 

We  also  divided  the  year  into  bi- 
weekly sampling  periods  and  plotted 
the  cumulative  number  of  species 
captured  versus  sampling  period. 
The  three  years  were  plotted  sepa- 
rately. Data  from  October  1985 
through  September  1987  were  treated 


two  ways:  (1)  as  if  sampling  began  in 
October,  and  (2)  as  if  sampling  began 
in  April.  This  provided  a  between 
year  comparison  of  how  effective 
sampling  for  species  numbers  would 
be  if  sampling  began  in  the  autumn 
as  opposed  to  the  spring. 

Statistical  Analysis 

Variation  in  the  overall  biweekly  cap- 
ture of  amphibians  and  reptiles  be- 
tween 1986  and  1987  was  compared 
using  a  Chi-square  contingency  table. 
The  Spearman  Rank  Correlation  Ma- 
trix then  was  used  to  compare  bucket 
capture  frequency  between  first  cap- 
ture and  recaptured  individuals,  in 
both  1986  and  1987,  of  G.  carolinensis 
and  N,  perstriatus.  Since  there  were 
no  significant  differences,  captures 
and  recaptures  were  combined  in 
subsequent  analyses. 

We  tested  for  within-year  vari- 
ation in  capture  frequency  inside  and 
outside  the  fence  using  a  one  sample 
Chi-square  goodness  of-fit-test.  The 
Spearman  Rank  Correlation  Matrix 
was  again  used  to  make  the  follow- 
ing comparisons:  (1)  a  within  year 
comparison  of  animals  captured  in- 
side the  fence  with  those  captured 
outside  the  fence  for  both  1986  and 
1987  [both  species],  (2)  a  comparison 
of  juvenile  with  adult  G.  carolinensis 
in  1986,  (3)  a  comparison  of  juvenile 
G.  carolinensis  inside  and  outside  the 
fence,  and  (4)  a  between  year  com- 
parison of  animals  captured  per 
bucket  inside  or  outside  the  fence 
[both  species]. 

To  determine  if  N.  perstriatus  and 
G.  carolinensis  preferentially  oriented 
to  or  from  one  of  the  three  habitat 
types  surrounding  the  pond,  data 
were  collapsed  and  analyzed  using  a 
Kruskal-Wallis  1-way  ANOVA. 
Buckets  1-3  and  21-23  faced  a  xeric 
hammock,  4-6  faced  a  small  open 
field,  and  7-20  faced  sandhills,  thus 
producing  the  three  habitat  catego- 
ries. 

Statistical  analyses  were  carried 
out  using  the  SAS  program  for 


microcomputers  (SAS  Institute  Inc. 
1985)  or  program  ABSTAT  version 
4.09  (Anderson-Bell  1984).  For  all 
analyses,  P  <  0.05  was  considered  in- 
dicative of  statistical  significance. 

Results 

Environmental  Conditions 

Severe  cold  weather  and  a  prolonged 
drought  characterized  the  sampling 
period  from  January  through  April 
1985.  In  Gainesville,  33  km  west  of 
Breezeway  Pond,  low  temperatures 
reached  -12  C  and  rainfall  was  152.4 
mm  below  normal  for  the  three 
month  period.  Breezeway  Pond  was 
dry  throughout  this  period.  Summer 
rains  filled  the  pond  in  mid-July,  and 
water  remained  until  December  16; 
maximum  pond  depth  was  60  cm  but 
declined  steadily  after  September. 
Free  water  was  present  from  January 
10-February  3  and  from  March  14  to 
April  22, 1986.  The  pond  remained 
dry  throughout  the  summer  of  1986 
despite  summer  thunderstorms  and 
did  not  refill  until  February  24, 1987. 
From  then  until  June  20  (115  days), 
up  to  60  cm  of  water  filled  the  pond. 
On  June  20,  the  pond  dried  and  re- 
mained dry  through  September  30. 

Species  Composition 

Thirty-nine  species  (7161  individual 
captures)  used  the  pond  or  its  pe- 
riphery at  some  point  during  the  27 
months  that  the  traps  were  moni- 
tored (table  1).  The  amphibians  cap- 
tured most  often  were  the  winter/ 
spring  breeding  striped  newt,  Noto- 
phthalmus perstriatus,  and  the  spring/ 
summer  breeding  eastern  narrow- 
mouthed  toad,  Gastrophryne  carolinen- 
sis. Only  one  other  salamander  was 
collected  at  Breezeway  Pond,  the 
dwarf  salamander,  Eurycea  quad- 
ridigitata.  Fourteen  species  of  frogs 
visited  the  pond,  and  six  were  pres- 
ent at  virtually  any  time  of  the  year: 
Acris  gryllus,  Bufo  quercicus,  B.  ter- 


88 


catesbeiam  were  caught  mainly  in  the 
summer.  Adult  R.  areolata  were 
caught  in  the  early  spring  as  they 


restris,  G.  carolinensis,  Limnaoedus  ocu- 
laris, and  Scaphiopus  holbrooki.  Adult 
Hyla  femoralis  and  juvenile  Ram 

Table  1.— Species  and  numbers  of  individual  amphibians  and  reptiles  cap- 
tured (first  number)  and  recaptured  (second  number)  at  Breezeway  Pond, 
January  1985  through  September  1987.  *  =  very  small  Individuals  not 
marked. 


1985 

1985-1986 

1986-1987 

Total 

(January- 

(UCIODGr- 

VVJCIODGi- 

A  nril\ 

oaiarnana©rs 

*Eun/cea  quadridigifafa 

5/0 

10/0 

8/0 

23/0 

Nofophfhalmus  persfriafus 

29/5 

558/309 

744/226 

1331/540 

Frogs 

*Acris  gryllus 

5/0 

74/5 

64/1 

143/6 

DU/L/  K^LJCfl  ^I^LJO 

1  /O 

111/31 

96/50 

208/81 

Bufo  ferresfris 

6/2 

65/46 

109/109 

180/157 

Eleufherodaciytus  planirosfris 

0/0 

0/0 

2/0 

2/0 

*  Gasfrophryne  carolinensis 

2/0 

1500/226 

379/274 

1881/500 

Hyla  chrysoscelis 

0/0 

1/0 

0/0 

1/0 

Hyla  femoralis 

0/0 

4/0 

39/2 

43/2 

Hvin  'snuirf^lld 

0/0 

3/0 

0/0 

3/0 

*  Limnaoedus  ocularis 

14/0 

20/0 

49/0 

83/0 

Rana  areolafa 

2/1 

9/5 

46/23 

57/29 

Rana  cafesbelana 

2/4 

9/4 

0/0 

11/8 

Rar^a  gryllo 

1/0 

0/0 

5/0 

6/0 

Rana  sphenocephala 

0/0 

5/0 

15/2 

20/2 

Scaphiopus  holbrooki 

1/1 

66/19 

165/92 

232/112 

Turtles 

Apalone  ferox 

0/0 

6/4 

0/0 

6/4 

Delrochelys  reHcularla 

0/0 

2/0 

0/0 

2/0 

KInosfernon  subrubrum 

9/0 

7/0 

11/4 

27/4 

Pseudemys  Horldana 

0/0 

17/14 

2/2 

19/16 

Lizards 

Cnemidophorus  sexllneafus 

18/7 

140/135 

122/115 

280/257 

*Eumeces  egregius 

14/2 

54/8 

30/4 

98/14 

Eumeces  inexpecfatus 

0/0 

0/0 

1/0 

1/0 

*Ophlsaurus  venfralls 

0/0 

14/2 

15/0 

29/2 

Sceloporus  undulafus 

4/0 

7/2 

2/0 

13/2 

*Scincella  lateralis 

23/0 

217/2 

207/2 

447/4 

Snakes 

Cemophora  cocclnea 

1/1 

2/0 

2/0 

5/1 

Coluber  constrictor 

2/0 

7/0 

8/8 

17/8 

DIadophls  punctatus 

0/0 

2/0 

2/2 

4/2 

MIcrurus  fulvius 

0/0 

6/0 

8/0 

14/0 

Nerodia  fasciata 

3/0 

4/1 

13/1 

20/2 

Nerodia  floridana 

1/0 

6/1 

4/0 

11/1 

Reglna  alleni 

2/1 

1/0 

2/0 

5/1 

Seminatrix  pygaea 

59/14 

18/10 

13/11 

90/35 

SIstrurus  mlliarlus 

0/0 

4/0 

2/1 

6/1 

J 

moved  toward  breeding  ponds,  and 
juvenile  R.  areolata  and  R.  sphenoceph- 
ala were  caught  in  late  summer  and 
early  autumn  presumably  as  they 
emigrated  to  terrestrial  habitats. 

The  most  commonly  captured  rep- 
tiles were  the  lizards  Scincella  later- 
alis, Cnemidophorus  sexlineatus  and 
Eumeces  egregius,  and  the  snake  Semi- 
natrix pygaea  (table  1).  Recent  hatch- 
lings  accounted  for  all  individuals  of 
the  lizards  Ophisaurus  ventralis  and 
most  S.  lateralis,  as  well  as  the  snakes 
Coluber  constrictor,  Nerodia  fasciata 
and  Thamnophis  sirtalis,  and  the 
turtles  Pseudemys  floridana  and  Kinos- 
ternon  subrubrum.  The  only  snake 
caught  in  substantial  numbers  was 
the  swamp  snake,  S.  pygaea,  espe- 
cially as  they  left  the  pond  during  the 
1985  drought. 


Cunnulative  Capture  Rotes 

The  rate  at  which  species  were  cap- 
tured varied  between  1986  and  1987 
(fig.  1).  More  species  were  captured 
at  a  faster  rate  in  1986  than  in  1987 
for  sampling  begun  in  October.  How- 
ever, the  reverse  was  true  for  sam- 
pling begun  in  April.  In  autumn,  the 
number  of  new  species  reached  an 
asymptote  after  about  six  weeks  of 
sampling  in  both  years  but  at  differ- 
ent levels  (25  in  1986,  23  in  1987).  In 
spring,  the  capture  of  new  species 
rose  steadily  both  years;  in  1986  it 
never  leveled  off  whereas  in  1987  it 
leveled  off  (at  31)  only  after  four 
months  of  sampling.  In  1985,  the  rate 
at  which  new  species  were  observed 
rose  rapidly  throughout  the  period 
and  was  beginning  to  level  off  only 
when  the  observations  were  termi- 
nated. 

In  1985,  three  months  of  sampling 
produced  25  of  the  39  (64%)  species 
now  known  to  be  present  at  Breeze- 
way  Pond.  Corresponding  percent- 
ages for  other  years  and  durations  of 
sampling  are  as  follows:  1986  -  6 
months  begun  in  October  =  74%,  6 
months  begun  in  April  =  77%,  12 
months  =  85%;  1987  -  6  months  be- 


89 


gun  in  October  =  59%,  6  months  be- 
gun in  April  =  82%,  12  months  = 
87%. 


Variation  in  Biweekly  Capture 

The  numbers  of  amphibians  and  rep- 
tiles captured  biweekly  varied  and 
was  significantly  different  between 
1986  and  1987  for  both  amphibians 
(X2  =  1366.46, 1  df,  P  <  0.001)  and 
reptiles  (X^  =  128.08, 1  df,  P  <  0.001). 
For  amphibians,  very  few  were 
caught  from  October  1986  through 
January  1987  compared  with  the 
same  period  in  1985-1986.  There  also 
were  many  fewer  individuals  caught 
during  the  summer  of  1987  com- 
pared with  1986.  This  was  due  to  a 
late  summer  drought  which  resulted 
in  the  complete  drying  of  the  pond 
with  subsequent  reproductive  failure 
of  G.  carolinensis.  Successful  repro- 
duction by  this  species  in  the  sum- 
mer of  1985  accounted  for  the  large 
numbers  of  amphibians  captured  in 
1986  (fig.  2).  Even  if  juvenile  narrow- 
mouthed  toads  are  excluded  (N  = 
690),  there  were  still  nearly  1000 
more  amphibians  recorded  in  1986 
compared  with  1987  (3425  in  1986, 
2475  in  1987). 

The  numbers  of  reptiles  recorded 
in  and  around  Breezeway  Pond  were 
very  similar  between  years,  although 
there  was  enough  variation  to  make 
the  patterns  significantly  different. 
As  might  be  expected,  reptile  activity 
decreased  during  the  winter  from 
late  October  through  mid-March  al- 
though some  individuals  were  active 
year  round  (fig.  3).  The  peak  in  num- 
bers in  mid-July  1986  represents  both 
a  large  number  of  species  captured 
as  well  as  an  influx  of  hatchling  S. 
lateralis. 


Temporal  Capture  Variation: 
Nofophthalmus  perstriatus  and 
Seminatrix  pygaea 

An  example  of  annual  variation  in 
numbers  of  individuals  and  dates  of 


Figure  1  .—A  comparison  of  the  rate  at  which  species  were  recorded  for  sampling  from  Janu- 
ary-April 1985  (1985),  October  1985-September  1986  (1986),  and  October  1986  through  Sep- 
tember 1987  (1987).  For  1986  and  1987,  the  data  were  treated  as  if  sampling  began  either  in 
October  or  April. 


Figure  2.— Number  of  amphibiarw  captured  at  Breezeway  Pond  in  1 986  and  1 987  by  2-week 
intervals. 


90 


120  -1 


--'-•--'--oooooooooSSoS-JJ 


o     o  o 


8    S  S 


Figure  3.— Number  of  reptiles  captured  at  Breezeway  Pond  in  1986  and  1987  by  2-week  Inter- 
vals. 


NOTOPHTHALMUS  PERSTRIATUS  -  BREEZEWAY  POND 


capture  is  illustrated  by  comparing 
collecting  data  from  1985  through 
1987  for  striped  newts,  N.  perstriatus 
(fig.  4),  and  swamp  snakes,  S.  pygaea 
(fig.  5).  From  mid-January  through 
mid-April,  the  numbers  of  newts 
captured  varied  from  34  in  1985  to 
364  in  1986  and  449  in  1987.  Most 
captures  occurred  from  the  first 
week  of  February  through  the  latter 
part  of  March,  and  were  associated 
with  rainfall  >  10  mm.  Movements  in 
1985  occurred  despite  bitter  cold  and 
prolonged  drought. 

In  contrast,  striped  swamp  snakes 
did  not  leave  the  pond  during  the 
cold  weather  of  1985,  but  waited  un- 
til temperatures  moderated  in  early 
March  (fig.  5).  Unlike  newts,  how- 
ever, they  did  not  return  in  appre- 
ciable numbers  later  in  1986  or  1987 
despite  favorable  habitat  and  climatic 
conditions. 


Orientation  and  Movement 
Patterns:  Gastrophryne 
carolinensis  and  Notophthalmus 
perstriatus 


<0  - 


M  • 


108B  N-34 


f!'  !•!«  ;! 


A 


-1016  N-364 
■•10B7  N-449 


III 


in!'  i 


rrrpr 


HP 

■ 

I 


101  •» 


i 
i 


TTT 


*l  1  I 


1. 


JAN 


FEB 


MAR 


APR 


—I 


Figure  4.— Comparison  of  thie  numbers  of  striped  newts  (Notophthalmus  persfriafus)  cap- 
tured from  January  16  ttiroughi  April  16,  1985-1987.  The  stars  indicate  days  of  >  10  mm  rain- 
fall. 


The  frequency  of  bucket  capture, 
both  inside  and  outside  the  drift 
fence,  varied  significantly  for  both 
adult  G.  carolinensis  and  N.  perstria- 
tus in  1986  and  1987  (table  2).  These 
data  indicate  non-random  movement 
into  and  out  of  the  pond.  There  was 
no  significant  correlation  between 
inside  and  outside  bucket  capture 
frequency  for  G.  carolinensis  in  1986 
(r^  =  -0.20,  22  df)  or  1987  (r^  =  -0.25, 
22  df).  There  was  significant  correla- 
tion between  inside  bucket  captures 
between  1986  and  1987  (r  =  0.35,  22 

s  ' 

df)  but  not  between  outside  bucket 
captures  between  years  (r^  =  0.06,  22 
df).  These  results  indicate  that  nar- 
row-mouthed toads  left  the  pond  in 
similar  directions  but  entered  it  from 
different  directions. 

Juvenile  G.  carolinensis  entering 
and  exiting  Breezeway  Pond  showed 
distinct  differences  between  capture 
frequency  at  different  stations  (X^  = 
535.73,  df  =  22,  P  <  0.001).  However, 


91 


they  showed  no  correlation  with 
adult  capture  frequency  per  station 
(r^  =  0.09,  22  df).  There  also  was  no 
correlation  in  bucket  capture  fre- 
quencies for  juveniles  caught  inside 
and  outside  the  drift  fence  (r^  =  0.26, 
22  df).  These  data  apply  only  to  1986 
because  no  juveniles  were  observed 
in  1987. 

For  N,  perstriatus,  there  was  like- 
wise no  significant  correlation  in  in- 
side versus  outside  bucket  capture 
frequency  in  1986  (r^  =  0.23,  22  df)  or 
1987  (r^  =  0.03,  22  df).  Capture  fre- 
quencies were  compared  outside  the 
fence  in  1986  versus  1987  (r^  =  0.07, 
22  df,  P  >  0.05)  and  inside  the  fence 
in  1986  versus  1987  (r^  =  0.55,  22  df,  P 
<  0.01).  As  with  Gastrophryne,  these 
results  suggest  that  newts  were  leav- 
ing the  pond  in  similar  directions  be- 
tween years,  but  that  they  were  en- 
tering it  from  different  directions. 


Habitat  Relationships 

Adult  Gastrophryne  did  not  m.ove  to- 
ward specific  habitats  in  either  1986 
(X2  =  2.62,  2  df,  P  =  0.27)  or  1987  (X^ 
=  0.32,  2  df,  P  =  0.85).  On  the  other 
hand,  juvenile  narrow-mouthed 
toads  moved  toward  the  sandhills  at 
a  higher  frequency  than  would  be 
expected  if  movements  were  random 
(X2  =  13.31,  2  df,  P  =  0.001),  but  not 
toward  the  pond  from  any  particular 
direction  (X^  =  2.26,  2  df,  P  =  0.32). 
Striped  newts  showed  non-random 
movement  in  1986  (X^  =  7.79,  2  df,  P 
=  0.02)  toward  the  sandhills  but  in 
1987  moved  toward  the  Panicum 
meadow  more  often  than  would  be 
expected  by  chance  alone  (X^  =  9.42, 
2  df,  P  =  0.009).  Movement  in  rela- 
tion to  nearby  habitat  is  illustrated  in 
figure  6. 


Discussion 

Was  Sampling  Effective? 

Although  we  caught  39  species  in  > 
7,000  captures,  it  is  likely  that  more 


14^ 


II  • 


> 
< 

o 
c 

U  I 

n 
< 

>  • 

o 
z 


4- 


SEMINATRIX  PYQAEA  -  BREEZEWAY  POND 


t 


A 


I 
t 

1 1 

I  ff 

1 1 


O'  —  - "  1985  N-73 


1986  N-2 


*  1987  N-1 


II  ■ 


It 


JAN 


FEB 


a 


ill  I 

ill  I 

III  I 


.  I 

i  i 

I  Mill 

I  IIII.I 


MAR 


APR 


te 


Figure  5— Comparison  of  the  numbers  of  swamp  snal<es  (Seminatrix  pygaea)  captured  from 
January  16  through  April  16,  1985-1987.  The  stars  indicate  days  of  >  10  mm  rainfall. 


species  of  amphibians  and  reptiles 
occasionally  visit  Breezeway  Pond. 
Some  species,  such  as  the  eastern 
coach  whip  snake  (Masticophis  flagel- 
lum),  Florida  pine  snake  (Pituophis 
melanoleucus),  and  gopher  tortoise 
(Gopherus  polyphemus),  are  common 


in  adjacent  sandhills  but  have  not 
been  observed  in  or  near  the  pond. 
Large  snakes  (e.g.,  Pituophis,  Mastico- 
phis) could  easily  go  over  the  fence 
and  thus  avoid  capture.  The  barking 
treefrog  (Hyla  gratiosa)  bred  in  the 
pond  before  the  initiation  of  our 


r 


Table  2.— Is  the  frequency  of  bucket  capture  random  inside  and  outside 
the  drift  fence?  For  ail  analyses,  there  were  23  stations  and  22  df.  A  signifi- 
cant value  indicates  non-random  movement. 


Species 

Year 

Orientation 

X2 

P 

Gosfrophryne 

1986 

Inside 

55.68 

<  0.001 

carolinensis 

1986 

Outside 

81.25 

<  0.001 

1987 

Inside 

84.00 

<  0.001 

1987 

Outside 

100.69 

<  0.001 

Nofophfhalmus 

1986 

Inside 

243.56 

<  0.001 

perstriatus 

1986 

Outside 

93.44 

<  0.001 

1987 

Inside 

88.45 

<  0.001 

1987 

Outside 

145.48 

<  0.001 

92 


study  (R.  Franz,  pers.  comm.),  but 
we  have  never  captured  it  or  heard  it 
calling  from  the  pond. 

Some  species,  particularly 
treefrogs  such  as  Hyla  femoralis, 
might  be  able  to  climb  over  the  fence 
and  thus  go  undetected  (Gibbons 
and  Semlitsch  1982).  Newts  (N. 
viridescens)  are  known  to  scale  drift 
fences  (Semlitsch  and  Pechmann 
1985)  although  we  have  not  observed 
N.  perstriatus  doing  so.  We  have  ob- 
served a  substantial  number  of  un- 
marked newts  inside  the  drift  fence 
even  after  two  years  of  study,  but  we 
do  not  know  if  they  were  residents 
that  were  moving  after  remaining  in 
the  pond  area  for  several  years,  or  if 
they  entered  by  crawling  over  or  un- 
der the  drift  fence.  Harris  et  al.  (1988) 
noted  that  many  adult  N.  viridescens 
burrowed  into  mud  at  the  edge  of 
North  Carolina  sandhills  ponds  as 
the  ponds  dried. 

For  these  reasons,  our  data  proba- 
bly underrepresent  both  the  number 
of  species  and  individuals  using  the 
pond  during  the  two  years  of  obser- 
vation. On  the  other  hand,  it  is  un- 
likely that  some  species  (e.g.,  Bufo, 
Scaphiopus)  are  able  to  climb  the 
fence.  As  such,  capture  results  of 
these  species  may  provide  a  reasona- 
bly accurate  estimate  of  pond  use. 

Activity  Patterns 

It  is  difficult  to  interpret  data  on  ac- 
tivity patterns  of  species  with  only 
two  years  of  data  because  there  are 
many  variables  that  influence  activity 
cycles  and  the  timing  of  reproduc- 
tion. These  variables,  such  as  rainfall 
amount  and  distribution,  maximum 
and  minimum  temperatures,  and 
hydroperiod  (Wiest  1982,  Semlitsch 
1985,  Pechmann  et  al.  1988),  vary 
daily,  seasonally  and  yearly,  and 
may  affect  different  species  in  differ- 
ent ways.  The  subtle  interaction  of 
these  parameters  probably  accounts 
for  the  variation  in  activity  patterns 
observed  between  years  (Semlitsch 
1985,  Semlitsch  and  Pechmann  1985). 


HABITAT  DISTRIBUTION  GC  1986  JUV  :  LEAVING  POND 


GO  1986  ADULT  GC  1987 


NP  1986  NP  1987 


Figure  6.— Diagram  illustrating  the  relationship  between  buckets,  emigration  from  the  pond, 
and  nearby  habitat  for  Gastrophryne  corolinensis  (GC)  and  Notophthalmus  perstriatus  (NP). 


93 


Amphibians  breeding  in  sandhills 
ponds  are  faced  with  substantial  un- 
certainty as  to  whether  or  not  suit- 
able conditions  will  prevail  for  repro- 
duction. Breezeway  Pond  was  cho- 
sen as  the  site  for  our  study  because 
it  had  consistently  held  water  from 
the  spring  of  1983  through  January 
1985  (R.  Franz,  pers.  comm.).  Begin- 
ning in  January,  climatic  conditions 
changed  resulting  in  two  years  of 
drought  with  only  sporadic  free  wa- 
ter. Temporary  ponds  may  allow  re- 
production free  of  certain  predators, 
but  their  use  comes  at  the  cost  of  re- 
productive uncertainty. 

Amphibians  are  active  during  or 
immediately  after  periods  of  rainfall 
or  high  humidities.  However,  the 
interaction  of  moisture  and  tempera- 
ture and  how  they  affect  condensa- 
tion probably  affects  diel  activity 
(Semlitsch  and  Pechmann  1985,  Du- 
ellman  and  Trueb  1986,  Pechmann 
and  Semlitsch  1986)  but  also  seasonal 
activity. 

The  extremely  dry  conditions  at 
Breezeway  Pond  during  the  study 
makes  it  difficult  to  predict  whether 
patterns  observed  in  early  1985  and 
from  late  1985  through  late  1987  are 
"typical"  for  the  amphibian  commu- 
nity using  the  pond.  Observations 
from  other  long-term  studies  of  her- 
petofaunal  communities  suggest  that 
there  is  wide  variation  in  numbers  of 
individuals  at  a  site  and  in  reproduc- 
tive success  from  year  to  year  (Gill 
1978,  Semlitsch  1983, 1985, 1987, 
Pechmann  et  al.  1988). 

Because  of  their  lack  of  depend- 
ence on  standing  water,  temperature 
is  probably  more  important  than  hy- 
droperiod  in  governing  reptile  daily 
and  seasonal  activity,  at  least  for  spe- 
cies in  direct  spatial  proximity  to  the 
pond.  However,  reptile  predators 
that  opportunistically  visit  tempo- 
rary ponds,  such  as  garter  snakes 
(Thamnophis  sp.),  might  increase  the 
number  of  visits  and  duration  of  stay 
if  a  sufficiently  long  hydroperiod  al- 
lows amphibian  reproduction  to  take 
place.  Our  data  are  insufficient  as  yet 
to  answer  this  question. 


Some  individuals  are  active  even 
during  unfavorable  environmental 
conditions  of  drought  and  unseason- 
ally  cold  temperatures.  Amphibians 
and  reptiles  are  generally,  but  not 
always  inactive  during  cold  or  dry 
weather.  For  instance,  Semlitsch 
(1983,  1985)  noted  that  mole  sala- 
manders {Amhystoma  sp.)  in  South 
Carolina  bred  during  the  coldest  but 
not  necessarily  the  wettest  months. 
He  felt  that  most  animals  moved  to 
breeding  ponds  at  this  time  to  allow 
sufficient  time  for  larval  develop- 
ment prior  to  pond  drying  (Semlitsch 
1987).  Such  may  not  explain  winter/ 
early  spring  breeding  in  N.  perstriatus 
because  the  breeding  period  is  ex- 
tended (Bishop  1947)  and  larvae  have 
been  found  from  April  through  De- 
cember (Christman  and  Means  1978). 
The  larval  period  is  unknown,  but  its 
duration  is  critical  to  successful  re- 
production in  temporary  sandhills 
ponds. 

Individuals  moving  at  times  of 
unusually  cold  and  dry  weather  may 
be  searching  for  more  favorable  re- 
treats or  escaping  adverse  condi- 
tions. If  the  onset  of  migration  (sensu 
Semlitsch  1985)  commenced  during 
unusually  adverse  conditions,  and 
the  unfavorable  conditions  extended 
for  a  long  period  of  time,  the  popula- 
tion could  be  vulnerable  to  local  ex- 
tinction via  mortality  or  emigration. 
Prolonged  drought  brought  about 
the  local  extinction,  via  emigration, 
of  the  resident  Semimtrix  population. 

Movement  Patterns  and 
Orientation 

Because  of  the  small  size  of  Breeze- 
way Pond,  it  is  difficult  to  ascribe 
directed  movements  of  individuals 
as  migrating  to,  or  originating  from, 
a  specific  habitat  type.  Because  the 
pond  was  located  in  an  ecotone,  an 
animal  captured  at  buckets  facing  the 
interface  between  sandhills  and  xeric 
hammock  could  move  in  either  direc- 
tion once  beyond  the  fence.  Likewise, 
an  animal  originating  from  one  habi- 


tat type  could  be  misclassified  if  it 
moved  a  relatively  short  distance 
and  fell  into  a  bucket  facing  a  differ- 
ent habitat  type.  The  open  field  was 
also  rather  small  and,  although  we 
did  not  feel  comfortable  assigning 
buckets  4-6  to  sandhills  or  xeric  ham- 
mock, it  is  likely  that  animals  exiting 
or  entering  the  pond  through  these 
buckets  came  from  or  went  to  one  or 
the  other  habitat. 

Given  these  qualifications,  adult 
Gastrophryne  did  not  exhibit  habitat 
preferences,  although  juveniles  left 
the  pond  primarily  toward  sandhills. 
Gastrophryne  are  commonly  recorded 
in  sandhills  (Carr  1940,  Campbell 
and  Christman  1982,  Mushinsky 
1985)  and  have  been  found  in 
sandhills  >  100  m  from  the  nearest 
water  source  (Franz  1986,  Dodd  pers. 
obs.).  Xeric  hammock  or  sandhills 
apparently  provide  narrow-mouthed 
toads  suitable  cover  and  resources 
away  from  the  breeding  pond,  but 
why  juvenile  Gastrophryne  would 
move  toward  sandhills  is  unknown. 

Striped  newts  are  most  commonly 
found  in  flatwoods  ponds  in  pine- 
palmetto  habitats  (Christman  and 
Means  1978)  as  well  as  ponds  in 
sandhills  and  scrub  areas  (Campbell 
and  Christman  1982).  To  what  extent 
they  use  sandhills  habitats  away 
from  ponds  is  unknown.  Carr  (1940, 
reported  as  N.  v.  symmetrica)  re- 
corded efts  in  high  and  mesophytic 
hammocks  in  light,  porous  soil. 
However,  striped  newts  at  Breeze- 
way Pond  moved  toward  sandhills 
or  meadow  rather  than  hammock. 
Migration  distances  of  striped  newts 
are  unknown  although  displaced  N. 
viridescens  can  move  400  m  through 
deciduous  forest  to  return  to  a  resi- 
dent pond  (Gill  1979).  N.  perstriatus 
probably  can  travel  similar  distances 
in  its  migrations. 

Management  implications 

The  Florida  sandhills  are  undergoing 
extensive  habitat  alteration  because 
of  rapid  human  population  growth 


94 


and  associated  development.  In  the 
late  1970's,  Auffenberg  and  Franz 
(1982)  estimated  that  70.6%  of  the 
sand  pine-scrub  oak,  57%  of  the  long- 
leaf  pine,  and  37.7%  of  the  xeric  ham- 
mock communities  had  been  de- 
stroyed by  forest  plantation  agricul- 
ture and  urbanization.  In  Putnam 
County,  the  site  of  our  study,  >  50% 
of  the  land  area  originally  supporting 
such  communities  no  longer  does  so. 
With  projected  human  population 
increases  of  more  than  300%  between 
1972  and  2000  (Auffenberg  and  Franz 
1982),  there  has  been  increasing  con- 
cern for  the  loss  of  sandhills  habitats 
in  northern  and  central  Florida.  Ex- 
tensive loss  of  habitat  is  occurring  in 
other  portions  of  the  state  and  South- 
east, such  that  only  14%  of  the  long- 
leaf  pine  (Pinus  palustris)  forests  re- 
main from  estimates  of  over  70  mil- 
lion acres  that  once  comprised  this 
community  (Means  and  Grow  1985). 

Because  of  habitat  loss,  amphibian 
and  reptile  populations  dependent 
upon  sandhills  probably  are  declin- 
ing. Many  of  the  amphibians,  such  as 
the  Rorida  gopher  frog.  Ram  areolata 
aesopus,  and  the  striped  newt,  N.  per- 
striatus,  are  considered  endangered, 
threatened,  or  rare  (Fogarty  1978, 
Christman  and  Means  1978),  yet 
there  are  few  data  on  their  life  histo- 
ries or  population  dynamics. 

The  paucity  of  information  on  spe- 
cies composition  and  population  dy- 
namics of  amphibians  and  reptiles 
that  use  temporary  ponds  in  xeric 
habitat  masks  the  probable  impor- 
tance of  such  habitats.  Variation  in 
annual  habitat  use,  both  intraspecifi- 
cally  and  inter-specifically,  appears 
to  be  considerable.  Long-term  eco- 
logical studies  of  the  herpetofaunal 
community  are  needed  to  under- 
stand the  magnitude  of  such  vari- 
ation and  its  potential  significance. 

Information  on  the  biology  of  the 
species  comprising  the  sandhills  her- 
petofaunal community  could  be  im- 
portant in  planning  for  the  manage- 
ment of  sandhills  ecosystems  by 
State  and  Federal  agencies.  For  in- 
stance, Florida  Statutes  Section 


373.414  required  Water  Management 
Districts  to  adopt  rules  to  establish 
specific  permitting-criteria  for  small 
isolated  wetlands,  including  size 
thresholds  below  which  impacts  on 
fish  and  wildlife  habitats  would  not 
be  considered.  When  these  rules 
were  adopted,  almost  no  data  were 
available  on  herpetofaunal  communi- 
ties on  which  to  make  recommenda- 
tions for  size  threshold  considera- 
tions. Lack  of  information  led,  in 
part,  to  variation  among  regulations 
adopted  by  the  different  Water  Man- 
agement Districts. 

There  is  considerable  interest 
among  Rorida  biologists,  conserva- 
tionists, and  land  use  planners  in  the 
concept  of  wildlife  corridors  to  main- 
tain biotic  diversity  (Harris  1985). 
Unfortunately,  most  discussions 
have  centered  on  riparian  habitats. 
The  lack  of  data  on  sandhills  habitat 
use,  especially  by  candidate  endan- 
gered or  threatened  species,  could 
hamper  the  long-term  survival  of 
such  species.  Many  sandhills  species 
are  likely  dependent  on  small  iso- 
lated wetlands  for  at  least  a  portion 
of  their  life  cycle.  By  focusing  on  ri- 
parian habitats,  planners  may  be 
overlooking  the  importance  of  up- 
land habitats  and  their  associated 
small  wetlands  to  the  maintenance  of 
biotic  diversity. 

The  following  are  the  most  impor- 
tant implications  of  our  study  for  the 
conservation  and  management  of 
small  isolated  wetlands  and  their  as- 
sociated herpetofaunal  communities 
in  "high  pine"  xeric  habitats  in 
northern  and  central  Rorida.  These 
should  be  kept  in  mind  when  evalu- 
ating impacts  of  habitat  loss  and 
planning  assessment  studies. 

1.  Many  species  use  these  habi- 
tats: some  are  permanent 
residents,  some  are  migrants, 
and  some  wander  through 
the  area  on  an  irregular  ba- 
sis. All  pond-breeding  spe- 
cies live  in  surrounding  ter- 
restrial habitats  during  the 
non-breeding  season.  Thus, 


the  pond  and  a  portion  of  the 
terrestrial  habitat  are  both 
critical  to  species  persistence. 

2.  Such  habitats  are  used  year- 
round  despite  seemingly  un- 
favorable periods  of  drought 
and  cold  weather. 

3.  Species  composition  varies 
within  a  year:  some  species 
are  found  only  in  one  season, 
some  predominate  at  one 
time  but  are  found  com- 
monly at  other  times,  some 
are  very  rarely  observed. 

4.  Reproductive  output  among 
species  varies  considerably: 
in  one  year  spring  breeders 
may  be  successful,  in  other 
years  summer  breeders  may 
be  successful,  in  some  years 
both  probably  produce 
young,  in  other  years  neither 
may  successfully  reproduce. 
The  longer  that  studies  are 
conducted,  the  greater  is  the 
likelihood  that  multiple  pat- 
terns will  emerge. 

5.  Activity  patterns  change  sea- 
sonally and  annually  proba- 
bly in  response  to  environ- 
mental cues,  particularly 
rainfall,  temperature,  and 
hydroperiod. 

6.  To  determine  the  total  num- 
ber of  species  using  such 
wetlands,  spring  and  early 
summer  sampling  produces 
the  best  results,  but  single 
season  or  even  yearly  sam- 
pling will  not  catch  all  spe- 
cies. 

7.  Quick  surveys  underestimate 
both  numbers  of  species  and 
individuals,  as  well  as  an- 
nual variation,  and  thus  un- 
derestimate the  importance 
of  temporary  isolated  wet- 
lands in  sandhills. 


95 


8.  To  adequately  understand 
complex  communities,  long- 
term  studies  are  absolutely- 
essential  for  management 
and  conservation. 


Acknowledgments 

We  thank  H.  1.  Kochman  for  advice 
on  statistical  analyses,  and  R.  Ash- 
ton,  R.  Franz,  J.  Oldemeyer,  J.  H.  K. 
Pechmann,  R.  Seigel  and  R.  D.  Sem- 
litsch  for  their  comments  on  the 
manuscript.  R.  L.  Burke,  K.  M.  Enge, 
and  J.  N.  Stuart  assisted  with  various 
phases  of  fieldwork. 

Literature  Cited 

Anderson-Bell.  1984.  ABSTAT.  Re- 
lease 4.  Anderson-Bell,  Canon 
City,  Colorado. 

Auffenberg,  Walter,  and  Richard 
Franz.  1982.  The  status  and  distri- 
bution of  the  gopher  tortoise  (Go- 
pherus  polyphemus).  p.  95-126.  In 
North  American  Tortoises:  Con- 
servation and  ecology.  R.  Bruce 
Bury,  editor.  Fish  and  Wildlife 
Service,  Wildlife  Research  Report 
12. 

Bishop,  Sherman  C.  1947.  Handbook 
of  Salamanders.  Cornell  Univer- 
sity Press,  Ithaca,  New  York. 

Campbell,  Howard  W.,  and  Steven 
P.Christman.  1982.  The  herpeto- 
logical  components  of  Horida 
sandhill  and  sand  pine  scrub  asso- 
ciations, p.  163-181.  In  Herpeto- 
logical  Communities.  Norman  J. 
Scott,  editor.  Fish  and  Wildlife 
Service,  Wildlife  Research  Report 
13. 

Carr,  Archie  Fairly,  Jr.  1940.  A  contri- 
bution to  the  herpetology  of  Ror- 
ida.  University  of  Florida  Publica- 
tion, Biological  Science  Series 
3(1):1-118. 

Christman,  Steven  P.,  and  D.Bruce 
Means.  1978.  Striped  newt  Notoph- 
thalmus  perstriatus.  p.  14-1 5.  7n  Rare 
and  Endangered  Biota  of  Horida, 
Vol.  3,  Amphibians  &  Reptiles. 


Roy  W.  McDiarmid,  editor.  Uni- 
versity Presses  of  Horida, 
Gainesville. 
Duellman,  William  E.,  and  Linda 
Trueb.  1986.  Biology  of  Amphibi- 
ans. McGraw-Hill  Book  Co.,  New 
York. 

Fogarty,  Michael  J.  1978.  Florida  go- 
pher frog  Rana  areolata  aesopus. 
p.5-6.  In  Rare  and  Endangered  Bi- 
ota of  Florida,  Vol.  3,  Amphibians 
&  Reptiles.  Roy  W.  McDiarmid, 
editor.  University  Presses  of  Flor- 
ida, Gainesville. 

Franz,  Richard.  1986.  Gopherus  pol- 
yphemus  (Gopher  Tortoise).  Bur- 
row commensals.  Herpetological 
Review  17:64. 

Gibbons,  J.  Whitfield,  and  David  H. 
Bennett.  1974.  Determination  of 
anuran  terrestrial  activity  patterns 
by  a  drift  fence  method.  Copeia 
1974:236-243. 

 ,  and  Raymond  D.  Semlitsch. 

1982.  Terrestrial  drift  fences  with 
pitfall  traps:  an  effective  technique 
for  quantitative  sampling  of  ani- 
mal populations.  Brimleyana  7:1- 
16. 

Gill,  Douglas  E.  1978.  The  metapopu- 
lation  ecology  of  the  red-spotted 
newt;  Notophthalmus  viridescens 
(Rafinesque).  Ecological  Mono- 
graphs 48:145-166. 

 .  1979.  Density  dependence  and 

homing  behavior  in  adult  red- 
spotted  newts  Notophthalmus 
viridescens  (Rafinesque).  Ecology 
60:800-813. 

Harris,  Larry  D.  1985.  Conservation 
corridors,  A  highway  system  for 
wildlife.  ENFO  (Florida  Conserva- 
tion Foundation),  November,  10  p. 

Harris,  Reid  N.,  Ross  A.  Alford,  and 
Henry  M.  Wilbur.  1988.  Density 
and  phenology  of  Notophthalmus 
viridescens  dorsalis  in  a  natural 
pond.  Herpetologica  44:234-242. 

Jones,  K.  Bruce.  1986a.  Amphibians 
.  and  reptiles,  p.267-290.  In  Inven- 
tory and  Monitoring  of  Wildlife 
Habitat.  A.  Cooperrider,  R.J.Boyd 
and  H.R.Stuart,  editors.  Bureau  of 
Land  Management  Service  Center, 
Denver,  Colorado. 


 .  1986b.  Data  types,  p.  11-28.  In 

Inventory  and  Monitoring  of 
Wildlife  Habitat.  A.  Cooperrider, 
R.  J.  Boyd  and  H.  R.  Stuart,  edi- 
tors. Bureau  of  Land  Management 
Service  Center,  Denver,  Colorado. 

Macan,  T.  T.  1966.  The  influence  of 
predation  on  the  fauna  of  a 
moorland  fishpond.  Archiv  fur 
Hydrobiologie  61:432-452. 

Means,  D.  Bruce,  and  Gerald  Grow. 
1985.  The  endangered  longleaf 
pine  community.  ENFO  (Horida 
Conservation  Foundation),  Sep- 
tember, 12  p. 

Moler,  Paul  E.,  and  Richard  Franz. 
1988.  Wildlife  values  of  small,  iso- 
lated wetlands  in  the  southeastern 
coastal  plain,  p.  234-241,  In  Pro- 
ceedings of  the  Third  Southeastern 
Nongame  and  Endangered  Wild- 
life Symposium.  R.  R.  Odom,  K.  A. 
Riddleberger,  and  J.  C.  Ozier,  edi- 
tors. Georgia  Department  of  Natu- 
ral Resources,  Athens,  Georgia. 

Mushinsky,  Henry  R.  1985.  Fire  and 
the  Horida  sandhill  herpetofaunal 
community:  with  special  attention 
to  responses  of  Cnemidophorus 
sexlineatus.  Herpetologica  41:333- 
342. 

Pechmann,  Joseph  H.  K.  and  Ray- 
mond D.  Semlitsch.  1986.  Diel  ac- 
tivity patterns  in  the  breeding  mi- 
grations of  winter-breeding  anu- 
rans.  Canadian  Journal  of  Zoology 
64:1116-1120. 

 ,  David  E.  Scott,  J.  Whitfield  Gib- 
bons, and  Raymond  D.  Semlitsch. 
1988.  Influence  of  wetiand  hy- 
droperiod  on  diversity  and  abun- 
dance of  metamorphosing  juvenile 
amphibians.  Wetlands  Ecology 
and  Management  1:1-9. 

SAS  Instihite  Inc.  1985.  SAS  Intro- 
ductory Guide  for  Personal  Com- 
puters, Version  6  Edition.  SAS  In- 
stitute Inc.,  Gary,  North  Carolina. 

Semlitsch,  Raymond  D.  1983.  Struc- 
ture and  dynamics  of  two  breed- 
ing populations  of  the  eastern  ti- 
ger salamander,  Amhystoma  ti- 
grinum.  Copeia  1983:608-616. 

 .  1 985.  Analysis  of  climatic  factors 

influencing  migrations  of  the 


96 


salamander  Amhy stoma  talpoideum. 
Copeia  1985:  477-489. 

 .  1987.  Relationship  of  pond 

drying  to  the  reproductive  success 
of  the  salamander  Amby stoma 
talpoideum.  Copeia  1987:61-69. 

 ,  and  Joseph  H.  K.  Pechmann. 

1985.  Diel  pattern  of  migratory 
activity  for  several  species  of 
pond-breeding  salamanders. 
Copeia  1985:86-91. 

Sexton,  Owen  J.,  and  Christopher 
Phillips.  1986.  A  qualitative  study 
of  fish-amphibian  interactions  in  3 
Missouri  ponds.  Transactions, 
Missouri  Academy  of  Science 
20:25-35. 

Wiest,  John  A.,  Jr.  1982.  Anuran  suc- 
cession at  temporary  ponds  in  a 
post  oak-savanna  region  of  Texas, 
p.  39-47.  In  Herpetological  Com- 
munities. Norman  J.  Scott,  Jr.,  edi- 
tor. Fish  and  Wildlife  Service, 
Wildlife  Research  Report  13. 


97 


Management  of  Annphlbians, 
Reptiles,  and  Snnall  Mammals 
In  Xeric  PInelands  of 
Peninsular  Florida^ 

I.  Jack  $tout,2  Donald  R.  Richardson,^  and 
Richard  E.  Roberts^ 


Abstract.— The  primary  xeric  pinelands  of  peninsu- 
lar Florida  are  longleaf  pine/turi<ey  oak  sandhills  and 
sand  pine  scrub.  Their  management  on  public  lands 
is  largely  confined  to  prescribed  burning  to  maintain 
fire  climax  status  of  the  vegetation.  The  regulation  of 
large-scale  developments  on  private  land  has  stimu- 
lated interest  in  preserve  design  and  management. 
The  suite  of  techniques  used  to  solve  conflict  be- 
tween natural  system  preservation  and  develop- 
ment includes:  (1)  conservation  set  asides  (pre- 
serves) on  site;  (2)  habitat  restoration;  (3)  purchase 
and  dedication  of  off-site  preserves;  (4)  species  relo- 
cation; and  (5)  wildlife  resource  mitigation  fund. 


Xeric  pinelands  seem  incongruent 
with  reference  to  Florida,  a  state  with 
annual  rainfall  that  ranges  from  50-65 
in  (19.6-25.6  cm).  Nonetheless,  the 
Florida  peninsula  contains  thousands 
of  acres  of  sandy  soil  derived  from 
marine  deposits  dating  to  the  Pleisto- 
cene (White  1970).  Two  distinct  plant 
associations,  longleaf  pine  (Pinus 
palustris)/ turkey  oak  (Quercus  laevis) 
sandhill  and  sand  pine  scrub  (Pinus 
clausa),  have  developed  on  these  nu- 
trient deficient  and  excessively  well- 
drained  soils.  Significant  areas  of 
these  plant  associations  occur  at 
higher,  albeit  modest,  elevations  rela- 
tive to  the  surrounding  landscape.  In 
fact,  certain  topographic  features, 
e.g.,  the  Lake  Wales  Ridge  and  the 
Marion  Upland,  were  likely  to  have 
been  true  islands  during  interglacial 
periods  while  the  remainder  of  Flor- 
ida was  covered  by  a  shallow  sea. 
Regardless  of  their  exact  origin,  xeric 
pinelands  support  many  relatively 
unusual  species  of  amphibians,  rep- 
tiles, and  small  mammals. 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff ,  AZ,  July  19-21,  1988). 

'I.  Jack  Stout  is  Professor  of  Biological 
Sciences,  Department  of  Biological  Sci- 
ences, University  of  Central  Florida,  Box 
25000,  Orlando,  FL  32816-0368. 

^Donald  R.  Richardson  is  Adjunct  Profes- 
sor, Department  of  Biology,  University  of 
South  Florida,  Tampa,  FL  33620. 

^Richard  E.  Roberts  is  Biologist,  Division  of 
Recreation  and  Parks,  Florida  Department 
of  Natural  Resources,  Hobe  Sound,  FL 
33455. 


Human  population  growth  (3.3% 
per  year)  and  development  in  Florida 
continues  to  encroach  on  upland 
habitats  and  particularly  on  xeric  pi- 
nelands. Most  of  the  habitat  loss  is  to 
agricultural  uses,  principally  citrus. 
Oddly,  the  state's  excellent  wetlands 
protection  acts  have  forced  develop- 
ment into  the  uplands.  Thus,  xeric 
pinelands  and  their  narrowly 
adapted  fauna  and  flora  are  increas- 
ingly threatened  by  area  reduction, 
fragmentation  and  isolation. 

It  is  our  intent  to  discuss  the  man- 
agement of  these  xeric  pinelands  in 
general  and,  more  specifically,  in  the 
context  of  small  preserves  in  an  oth- 
erwise developed  landscape.  Man- 
agement of  xeric  pinelands  as  ecosys- 
tems is  yet  in  its  infancy,  and  more 
detailed  prescriptions  for  designated 
species  are  unproven.  However, 
progress  is  being  made  (Cox  et  al. 
1987)  and  improvement  and  revision 
of  current  thinking  on  management 
practices  is  anticipated.  This  paper 
summarizes  selected  literature  on 
xeric  pineland  and  the  species  associ- 
ated with  these  communities  to  as- 
sess management  practices.  In  addi- 
tion, unpublished  information  has 
been  used  and  identified  in  the  text. 

Preserve  design  efforts  by  us  have 
been  on  behalf  of  developers  re- 
sponding to  development  orders  pre- 
pared by  governmental  agencies. 
These  designs  are  site  specific  in  de- 
tail, but  nonetheless  point  to  general 
problems  and  solutions.  Our  ap- 
proach has  been  to  focus  on  provid- 


ing the  area  required  to  support 
minimum  viable  populations  of 
"keystone"  or  otherwise  critical  ani- 
mal species  of  a  given  xeric  pineland. 
Once  this  area  is  settled  on,  manage- 
ment should  focus  on  those  species 
whose  minimum  area  requirements 
are  met,  whereas  no  special  efforts 
are  expended  on  species  with  larger 
area  requirements. 

XERIC  PINELAND  HABITATS 

Longleaf  Pine/Turkey  Oak 
Sandhills 

The  longleaf  pine /turkey  oak 
sandhill  association  (LLP/TO)  was 
about  15%  (2, 110,  256  ha)  of  the 
natural  landscape  of  peninsular  Hor- 
ida  in  pre-Columbian  times  (fig.  1) 
(Auffenberg  and  Franz  1982).  This 
xeric  pineland  occupies  rolling  to- 
pography of  several  ridge  systems 
that  run  north-south,  notably  Trail 
Ridge,  the  Lake  Wales  Ridge,  and  the 
Brooksville  Ridge;  numerous  lesser 
ridges  and  hills  are  identified  by 
White  (1970).  These  ridges  consist  of 
deep,  well-drained  soils  of  the 
Lakeland,  Eustis,  and  Blanton  asso- 
ciations (Beckenbach  and  Hammett 
1962).  Laessle  (1942, 1958a)  describes 
the  LLP/TO  plant  association  as  a 
fire  climax  system  dominated  by 
longleaf  pine;  slash  pine  (Pinus  elli- 
otti)  replaces  longleaf  pine  in  the 
community  in  south  Florida.  Turkey 
oak  is  a  minor  tree,  but  can  achieve 


98 


co-dominance  when  fires  are  sup- 
pressed. The  predominant  under- 
story  plant  is  wiregrass,  Aristida 
stricta;  however,  a  rich  assemblage  of 
perennial  herbs  vary  in  prominence 
in  concert  with  seasonal  changes. 
Monk  (1968)  recognizes  two  addi- 
tional phases  of  sandhill  vegetation 
in  north  central  Rorida:  (1)  longleaf 
pine/sand  post  oak  (Quercus  marga- 
retta)  and  (2)  longleaf  pine /southern 
red  oak  (Quercus  falcata).  A  fourth 
phase,  longleaf  pine/scrub  hickory 
(Carya  floridam),  occurs  in  the  south- 
ern portion  of  the  Lake  Wales  Ridge 
(Abrahamson  et  al.  1984).  Veno 
(1976),  Givens  et  al.  (1984),  and  Abra- 
hamson et  al.  (1984)  provide  quanti- 
tative data  on  LLP /TO  community 
structure  and  dynamics.  Myers 
(1985)  suggests  that  longleaf  pine/ 
turkey  oak  and  sand  pine  scrub  asso- 
ciations are  successionally  linked  in 
some  portions  of  their  geographic 
ranges.  Differences  in  physical/ 
chemical  features  of  soils  of  LLP/TO 
and  SPS  communities  in  the  Ocala 
National  Forest  are  not  considered  to 
be  sufficient  to  explain  the  local  dis- 


Figure  1  .—Potential  geographic  distribution 
of  longleaf  pine/turkey  oak  sandhill  and 
sand  pine  scrub  xeric  pinelands  in  Florida. 
Light  shading  indicates  the  sandhills  and 
darker  shading  indicates  the  scrub.  These 
distributions  are  based  on  Davis  (1980)  and 
do  not  reflect  nninor  sites  of  either  commu- 
nity due  to  the  scale  of  the  illustration. 


tribution  of  the  communities  (Kalisz 
and  Stone  1984). 

Prior  to  settlement  by  European 
man,  ground  fires  occurred  in  LLP/ 
TO  sandhills  at  intervals  of  1-5  years. 
These  relatively  "cool"  fires  favor 
regeneration  of  longleaf  pine,  flower- 
ing by  grasses  and  herbs,  and  sup- 
press growth  of  woody  plants 
(Myers  1985). 

Sand  Pine  Scrub 

Compared  with  the  LLP/TO  sandhill 
association,  sand  pine  scrub  (SPS) 
has  less  area  (250,000  ha)  and  a  far 
more  limited  distribution  (fig.  1). 
Scrub  is  associated  with  old  shore- 
lines, lake  margins,  and  stream 
courses  where  extremely  well 
washed,  nutrient  deficient  sands 
were  deposited  during  Pleistocene 
times  (Kurz  1942;  Laessle  1958a,  b, 
1967).  The  most  widespread  soils 
supporting  SPS  are  the  St.  Lucie, 
Lakewood,  and  Pomello  associations 
(Beckenbach  and  Hammett  1962). 

Sand  pine  scrub  is  a  two-layered 
community.  Sand  pine  (Pinus  clausa) 
normally  occurs  as  a  relatively  even- 
aged  overstory  species.  The  under- 
story  is  comprised  of  10-20  species  of 
evergreen  shrubs  l-5m  in  height. 
Four  species  of  oaks  comprise  the 
bulk  of  the  biomass,  Quercus  gemi- 
nata,  Q.  myrtifolia,  Q.  chapmanii,  and 
Q.  inopina.  Lesser  numbers  of  other 
species  including  Ceratiola  ericoides, 
Lyonia  ferruginea,  and  Osmanthus 
americanus  add  to  local  diversity. 
Sand  pine  scrub  is  a  fire  climax  com- 
munity (Laessle  1958a,  Abrahamson 
et  al.  1984).  In  contrast  with  LLP/TO, 
SPS  burns  at  intervals  of  20-70  years; 
a  combination  of  ground  and  crown 
fires  destroys  all  the  above-ground 
vegetation.  Most  of  the  woody 
plants,  with  the  notable  exception  of 
the  sand  pine  and  Ceratiola,  readily 
sprout  from  root  crowns  following 
fires.  Laessle  (1958a),  Veno  (1976), 
and  Richardson  (1977)  provide  data 
on  plant  community  structure  of 
scrubs.  Recent  quantitative  studies 


include  those  of  Abrahamson  et  al. 
(1984)  and  Latham  (1985). 

Outstanding  examples  of  SPS  in- 
clude the  "Big  Scrub,"  part  of  the 
Ocala  National  Forest,  scrubs  of  the 
Lake  Wales  Ridge,  e.g.,  the  Archbold 
Biological  Station,  and  stands  along 
the  Atlantic  Coastal  Ridge. 

SMALL  VERTEBRATE  SPECIES 
ASSEMBLAGES 

Longleaf  Pine/Turkey  Oak 
Sandhills 

Amphibians  and  Reptiles 

At  least  47  species  of  herptiles,  in- 
cluding 2  newts,  13  toads  and  frogs, 
3  turtles,  10  lizards,  1  amphis- 
baenian,  and  18  snakes,  are  reported 
to  occur  in  LLP/TO  habitats  (table  1). 
Campbell  and  Christman  (1982)  list  5 
categories  of  reptile  and  amphibian 
species  that  occur  in  LLP/TO  and 
SPS:  (1)  characteristic  (18  species);  (2) 
associated  with  tortoise  burrows  (3 
species);  (3)  frequent  (8  species);  (4) 
occasional  (14  species);  and  (5)  asso- 
ciated with  aquatic  habitats  (21  spe- 
cies). Of  the  characteristic  species,  7 
are  regarded  as  adapted  to  xeric  con- 
ditions, 3  as  sand  swimmers,  viz., 
Neoseps  reynoldsi;  Eumeces  egregius, 
and  Tantilla  relic ta,  and  the 
remainder  (Sceloporus  woodi,  Mastico- 
phis  flagellum,  Stilosoma  extenuatum, 
Cnemidophorus  sexlineatus)  to  other 
physical  features  of  the  habitats. 

The  gopher  tortoise  (Gopherus  pol- 
yphemus)  is  a  terrestrial  turtle  that 
digs  deep  burrows  in  the  well- 
drained  sandhill  soils  (Auffenberg 
and  Franz  1982).  Stout  (1981)  and 
Eisenberg  (1983)  recognized  the  go- 
pher tortoise  was  the  keystone  spe- 
cies in  xeric  pinelands.  Some  80  spe- 
cies of  animals  may  be  classified  as 
burrow  commensals  (Cox  et  al.  1987); 
however,  the  number  of  obligatory 
commensals  is  much  smaller.  Herp- 
tiles particularly  associated  with  go- 
pher tortoise  burrows  include  Rana 


99 


areolata,  Pituophis  melanoleucus,  and 
Drymarchon  corais. 

The  snake  fauna  of  LLP/TO 
sandhills  is  species  rich  (>  18  spe- 
cies). This  diversity  includes  large 
forms,  e.g.,  Drymarchon  corais  couperi 
and  Cro talus  adamanteus,  and  small, 
specialized  species  like  Stilosoma  ex- 
tenuatum.  This  latter  ophiophagous 
species  feeds  largely  on  Tantilla 
relicta;  Tantilla,  is  in  turn  specialized 
on  Tenebrionidae  larvae  (Mushinsky 
1984). 

Small  Mammals 

At  least  19  species  of  small  mammals 
with  body  masses  less  than  6.0  kg 


may  be  anticipated  in  LLP/TO  sand- 
hills (table  2).  Two  are  fossorial, 
Scalopus  aquations  and  Geomys  pinetia, 
1  semi-fossorial,  P.  polionotus,  and  2 
occur  in  the  surface  litter,  Blarina 
carolinensis  and  Cryptotis  parva. 

Arboreal  species  include  Sciurus 
carolinensis,  S.  niger,  Glaucomys  volans, 
P.  gossypinus,  and  Ochrotomys 
nuttalli.  Podomys  floridanus  nests  in 
the  burrows  of  the  gopher  tortoise 
and  the  pocket  gopher  (Layne  1969); 
it  may  enlarge  other  openings  in  the 
soil  to  establish  burrows  independ- 
ently of  the  gopher  tortoise  (R.  E. 
Roberts,  personal  observation). 
Dasypus  novemcinctus  is  the  only  ex- 
otic species  of  mammal  that  is  clearly 


established  in  the  sandhill  commu- 
nity. 

Sand  Pine  Scrub 

Amphibians  and  Reptiles 

Campbell  and  Christman  (1982) 
listed  64  species  of  reptiles  and  am- 
phibians that  may  be  found  in  LLP/ 
TO  sandhills  and  SPS.  Pitfall  trap- 
ping in  six  different  even-aged 
stands  of  SPS  on  the  Ocala  National 
Forest  by  Christman  et  al.  (unpub- 
lished manuscript  and  personal  com- 
munication) revealed  27  species 
(table  1).  Of  1,624  individuals 


Table  1.— Herpetofauno  trapped  or  observed  within  the  xeric  pinelonds  of  peninsula  Florida.  Standard  herp  arrays 
were  used  in  each  study  to  sample  for  a  period  of  at  least  one  year. 


Species  Long  leaf  pine/       Sand  pine  Species  Long  leaf  pine/       Sand  pine 

turkey  oak  scrub  turkey  oak  scrub 

Campbell    Mushinsky  Stout    Christman  Campbell    Mushinsky  Stout  Christman 

&  Christman      1965       etaL       etai.  &  Christman      1985       etal.  etal. 

1982  unpubl.    unpubl.  1982  unpubl.  unpubl. 


Nofophfhalmus 

viridescens 
N.  persfriafus 
ScopNopus  holbrookii 
Bufo  ferresfris 
Bufo  quercicus 
Eleufh  ero  dac  fylus 

planirosfris 
HyJa  fern  or  alls 
Hyla  grafiosa 
Hyla  squirella 
Hyla  cinerea 
Acris  gryllus 
Rana  gryllo 
Rana  areolata 
Rana  ufricularia 
Gasfrophryne 

carolinensis 
Kinosfernon  bauri 
Terrapene  Carolina 

bauri 

Gopherus  polyphemus 
Anolis  carolinensis 
Anolis  sagrei 
Sceloporus  undulafus 
Sceloporus  wood! 
Ophisaurus  compressus 
Cnemldophorus 
sexlineatus 


X 
X 


X 
X 
X 

X 
X 


X 
X 
X 

X 

X 


X 
X 

X 
X 
X 
X 
X 
X 


X 
X 
X 
X 


X 
X 


Scincella  lateralis  x 

Eumeces  inexpectatus  x 

E,  egregius  lividus  — 

E.  egregius  onocrepis  x 

Neoseps  reynoldsi  x 

Rhineura  floridgna  x 

Nerodia  fasciafa  — 

Th  amn  ophis  saurifus  — 

Riiadinaea  flavilafa  — 

Diadophis  punc  tafus  x 

Paranoia  abacura  — 

Coluber  constrictor  x 

Masticophis  flagellum  x 

Opheodrys  aestivus  x 

Drymarclion  corais  — 

Baphe  guttata  — 

Pituophls  melanoleucus  x 
Lampropeltis 

triangulum  x 

Stilosoma  extenuatum  x 

Cemophiora  coccinea  x 

Tantilla  relicta  x 

Heterodon  platyrhinos  x 

Heterodon  simus  — 

Micrurus  fulvius  fulvius  x 

Sistrurus  millarius  x 

Crotalus  adamanteus  — 

Totals  29 


X 
X 
X 


X 

27 


X 
X 
X 
X 
X 
X 


X 
X 
X 
X 
X 


33 


27 


100 


trapped,  the  common  species  were 
Bufo  terrestris  (n=332),  Cnemidophorus 
sexlineatus  (n=329),  and  Sceloporus 
woodi  (n=216);  five  species  were  rep- 
resented by  single  captures.  Christ- 
man  et  al.  concluded  that  the  herpe- 
tofaunal  diversity  declined  with  in- 
creasing age  of  SPS  stands. 

Gopherus  polyphemus  is  the  key- 
stone species  in  SPS  but  is  less  com- 
mon there  than  in  LLP /TO  (Auf fen- 
berg  and  Franz  1982).  Many,  if  not 
most,  of  the  burrow  commensals  are 
common  in  SPS  (Cox  et  al.  1987). 
Podomys  floridanus  is  an  example. 

[   Small  Mammals 

Fourteen  species  of  small  mammals 
commonly  inhabit  SPS  (table  2).  Pod- 


omys floridanus  is  a  predictable  mem- 
ber of  the  assemblage  throughout  the 
range  of  scrubs  in  peninsular  Florida 
(Layne  1978).  Three  subspecies  of 
Peromyscus  polionotus  occur  in  scrubs 
of  the  interior  and  east  coast  portions 
of  the  peninsula.  Common  small 
mammals  in  central  peninsular  Flor- 
ida scrubs  include  Podomys  floridanus, 
Peromyscus  gossypinus,  Ochrotomys 
nuttalli,  and  Glaucomys  volans  (Swin- 
dell 1987).  Podomys  floridanus  is  the 
predominate  small  mammal  in 
scrubs  of  southeast  Florida 
(Richardson  etal.  1986). 

Limited  data  suggest  Spilogale 
putorius  is  a  major  predator  on  small 
mammals  in  scrubs  with  lesser  roles 
played  by  Mephitis  mephitis  and 
Mustela  frenata  (Stout  and  Roberts, 
personal  observations). 


Table  2.— Small  mammal  community  structure  In  sandhill  and  sand  pine 
scrub  plant  associations  of  peninsular  Florida.  The  upper  limit  of  body  mass 
of  small  mammals  was  arbitrarily  set  at  6.0  kg. 

Mammal  Species  Longleaf  pine/turkey  oak' -^-^  Sand  pine  scrub* 


Didelphis  virginiana 

X 

X 

Crypfofis  parva 

X 

Blarina  carolinensis 

X 

X 

Scalopus  aquaficus 

X 

Dasypus  novemcinctus 

X 

X 

SyMlagus  floridanus 

X 

X 

Sciurus  carolinensis 

X 

X 

Sciurus  niger 

X 

Glaucomys  volans 

X 

X 

Geomys  pinefis 

X 

Peromyscus  polionotus 

X 

X 

Peromyscus  gossypinus 

X 

X 

Podomys  floridanus 

X 

X 

Ochrotomys  nuttalli 

X 

X 

Sigmodon  hispidus 

X 

X 

Urocyon  cinereoargenfeus 

X 

Procyon  lotor 

X 

X 

Mustela  frenata 

? 

X 

Spilogale  putorius 

X 

X 

Mephitis  mephitis 

X 

No.  Species 

19 

14 

^  St  out  et  al.,  unpublished 
^Arata  1959 
^Humphrey  et  al.  1985 
"Stout  1982 


ENDANGERED  AND  THREATENED 
SPECIES 

Ten  species  of  amphibians,  reptiles, 
and  small  mammals  associated  with 
xeric  pineland  are  currently  listed  as 
having  some  level  of  threatened,  en- 
dangered, or  sensitive  status  by  ei- 
ther the  state  of  Florida  or  the  De- 
partment of  Interior  (table  3).  The 
extensive  overlap  in  species  composi- 
tion between  the  two  pineland  com- 
munities results  from  the  high  num- 
ber of  species  common  to  both  types. 
The  Endangered  Species  Act  charges 
federal  agencies  with  the  responsibil- 
ity to  manage  federally  listed  species 
on  federally  owned  lands.  At  the 
state  level,  preservation  of  these 
listed  species  is  of  major  concern 
when  they  occur  on  parcels  of  land 
scheduled  for  large-scale  develop- 
ment. Preserve  design  and  manage- 
ment practices  for  these  species  have 
largely  evolved  on  an  ad  hoc  basis 
without  adequate  time  for  an  evalu- 
ation of  the  management  or  the  long- 
term  implications  for  the  species. 

MANAGEMENT  OF  XERIC 
PINELANDS  ON  PUBLIC  LANDS 

Of  three  national  forests  in  Rorida, 
only  the  Ocala  National  Forest  is  lo- 
cated in  the  peninsula.  It  totals 
153,846  ha  of  which  85,020  ha  are  SPS 
and  18,219  ha  LLP/TO.  The  National 
Forest  Management  Act  (1976)  and 
pursuant  regulations  (36  CFR  219) 
require  that  each  forest  be  managed 
to  maintain  well-distributed  and  vi- 
able populations  of  wildlife  species, 
including  species  that  are  endan- 
gered or  threatened  (Norse  et  al. 
1986). 

Silvicultural  systems  differ  be- 
tween the  two  pineland  communi- 
ties. On  the  Ocala  National  Forest 
sand  pine  scrub  is  routinely  har- 
vested in  patchy  clearcuts  that  range 
from  16-24  ha  in  area.  Scrub  under- 
story  vegetation  is  allowed  to  regen- 
erate naturally;  however,  sand  pine 
is  seeded  following  site  preparation 


101 


by  a  single  roller  chopping.  The  har- 
vest rotation  length  is  about  50  years. 
In  contrast,  LLP /TO  is  ostensibly 
managed  on  a  80-100  year  rotation 
and  shelterwood  cutting  favors  natu- 
ral regeneration  of  the  longleaf  pine 
(Don  Bethancourt,  personal  commu- 
nication). In  practice,  harvesting  of 
longleaf  pine  may  occur  in  60  years. 

Effectiveness  of  ecosystem  man- 
agement in  the  SPS  community  v;^ill 
be  judged  by  the  response  of  desig- 
nated indicator  species,  such  as  go- 
pher tortoises  and  scrub  jays  (Aphelo- 
coma  coerulescens)  (table  3).  The  go- 
pher tortoise  is  also  a  designated  in- 
dicator species  for  the  LLP/TO  com- 
munity. The  significance  of  the  go- 
pher tortoise  as  a  keystone  species 
was  emphasized  in  1986  when  har- 
vesting of  the  species  on  national  for- 
ests in  Florida  was  made  illegal 
through  an  agreement  between  the 
U.S.  Forest  Service  and  the  Florida 
Game  and  Fresh  Water  Fish  Commis- 
sion. Other  species-specific  manage- 
ment practices  involving  amphibians, 
reptiles,  or  small  mammals  have  not 


been  deemed  necessary  to  carry  out 
on  the  Ocala  National  Forest  (Don 
Bethancourt,  personal  communica- 
tion). In  fact,  the  impact  of  timber 
harvesting  on  small  vertebrates  of 
LLP/TO  and  SPS  communities  is 
simply  not  known. 

Public  lands  in  Florida  supporting 
xeric  pinelands  include,  but  are  not 
limited  to,  state  forests  and  state 
parks.  State  forests  with  large  acre- 
ages of  LLP/TO,  e.g.,  the  Withla- 
coochee  State  Forest,  are  managed  at 
the  ecosystem  level.  Prescribed  burn- 
ing is  done  every  3-8  years  and  fu- 
ture timber  sales  will  follow  a  rota- 
tion length  of  80-120  years;  currently 
rotation  lengths  are  about  60  years 
and  are  not  regarded  as  favorably  for 
endemic  wildlife.  Wildlife  manage- 
ment areas  overlap  the  state  forest 
holdings  and  are  managed  for  sus- 
tained yields  of  wildlife  by  the  Flor- 
ida Game  and  Fresh  Water  Fish 
Commission  based  on  a  memoran- 
dum of  understanding  between 
agencies  (Cathy  Ryan,  personal  com- 
munication). 


Table  3.— Endangered  and  potentially  endangered  amphibians,  reptiles, 
and  small  mammals  (Wood  1 987)  inhabiting  xeric  pinelands  of  peninsular 
Florida. 


Species  group 


Xeric  pineland     Designated  status' 


LLP/TO      SPS     FGFWFC2  USFWS^ 


Amphibians  and  Reptiles 

Drymarchon  cords  couperi 

Eumeces  egregius  lividus 

Gopherus  polyphemus 

Neosepsreynoldsi 

Pifuophis  melanoieucus  mugifus 

Rona  areolafa 

Sfilosoma  exfenuafum 
Mammals 

Geomys  pinefis  goffi 

Podomys  floridanus 

Sciurus  niger  shermani 


X 
X 
X 
X 
X 
X 
X 

X 
X 
X 


X 
X 
X 
X 
X 
X 
X 


X 


T 

T 

I 

ssc 

T 

SSC 
SSC 

T 

I 

E 

SSC 
SSC 


T 
T 

UR2 

T 
UR2 
UR2 
UR2 

UR3 
UR2 
UR2 


'f=  Endongeredj  T=Jhreatened;  SSC=  Species  of  Special  Concern:  UR2=  Under  re- 
view for  listing,  but  substantial  evidence  of  biological  vulnerability  and/or  threat  is 
lacking:  UR3  =  Still  formally  under  review  for  listing,  but  no  longer  being  considered  for 
listing  due  to  existing  pervasive  evidence  of  extinction. 

^Florida  Game  and  Fresh  Water  Fish  Commission 


^United  States  Fish  and  Wildlife  Service 


State  parks  are  managed  by  the 
Division  of  Recreation  and  Parks  of 
the  Rorida  Department  of  Natural 
Resources  (FDNR).  An  ecosystem 
approach  is  taken  in  the  restoration 
and  management  of  xeric  pinelands 
on  state  park  lands  (Jim  Stevenson, 
personal  communication).  Prescribed 
burning  has  been  used  since  1969  to 
control  hardwood  invasion  of  LLP/ 
TO  stands  and  to  stimulate  growth 
and  flowering  of  grasses  and  herbs. 
Burning  in  spring  and  early  summer 
appears  to  best  duplicate  the  historic 
timing  of  lightning  initiated  fires  in 
xeric  pinelands.  The  impact  of  these 
management  practices  on  the  plant 
community  has  been  documented 
(Davis  1984);  the  response  of  reptiles, 
amphibians,  and  small  mammals  is 
currently  under  study  (Stout  et  al. 
unpublished).  Generally,  mature 
stands  of  SPS  have  not  been  burned 
until  recently,  due  to  the  unpredict- 
able behavior  of  fire  in  the  commu- 
nity; however,  a  prescription  for 
burning  this  fuel  type  has  been  writ- 
ten and  tested  on  private  land  and 
state  parks  (Doran  et  al.  1987).  Early 
recovery  stages  of  SPS  appear  to  sup- 
port the  greatest  diversity  of  reptiles 
and  amphibians.  However,  as  can- 
opy closure  occurs  in  SPS,  ground 
cover  diminishes  and  habitat  quality 
for  gopher  tortoises  declines  (Cox  et 
al.  1987).  In  contrast,  similar  numbers 
of  Podomys  have  been  observed  in 
early  (R.  E.  Roberts,  unpublished 
data,  J.  Dickinson  State  Park);  inter- 
mediate (Stout  1982);  and  old  growth 
SPS  (James  N.  Layne,  unpublished 
data,  Archbold  Biological  Station). 

State  parks,  reserves,  and  pre- 
serves appear  to  be  ideal  lands  to  ex- 
plore species-specific  management 
measures  for  herp tiles  and  small 
mammals.  For  example,  sand  swim- 
ming herptiles  (Smith  1982)  require 
openings  that  are  relatively  root  free 
in  LLP/TO  and  SPS  habitats.  The 
natural  occurrence  of  such  openings 
may  have  been  due  to  "hot"  spots 
associated  with  the  combustion  of 
high  fuel  loads,  e.g.,  fallen  trees  (Ron 
Myers,  personal  communication). 


102 


Concentrarion  of  natural  fuels  prior 
to  prescription  bums  in  SPS  would 
offer  a  means  to  create  microhabitat 
conditions  favorable  for  the  sand 
swimmers. 


MANAGEMENT  OF  XERIC 
PINELAND  ON  PRIVATE  LAND 

Development  of  Regional  Impact 

Concern  with  management  of  am- 
phibians, reptiles,  and  small  mam- 
mals on  private  lands  in  Rorida  de- 
rives from  state  and  federal  protec- 
tion of  endangered  species  and  the 
development  guidelines  promul- 
gated during  the  Development  of 
Regional  Impact  (DRI)  process.  "The 
Rorida  Environmental  Land  and 
Water  Management  Act  of  1972" 
(Chapter  380,  Florida  Statutes)  de- 
fines developments  of  regional  im- 
pact in  Section  380.06(1),  Rorida  Stat- 
utes, as  "...any  development  which, 
because  of  its  character,  magnitude, 
or  location,  would  have  a  substantial 
effect  upon  the  health,  safety,  or  wel- 
fare of  citizens  of  more  than  one 
county  (Anonymous  1976)."  Large 
scale  development  projects  in  penin- 
sular Florida  commonly  involve  hun- 
dreds to  several  thousand  acres  of 
relatively  natural  landscape.  The  DRI 
process  requires  bona  fide  studies  of 
wildlife  populations  and  their  associ- 
ated habitats;  emphasis  is  placed  on 
listed  species.  Developers  must  pre- 
pare viable  management  strategies  to 
accommodate  wildlife  resources  de- 
pendent upon  their  lands  (Cox  et  al. 
1987;  Richardson  et  al.  1986). 

Management  strategies  of  devel- 
opers with  xeric  pinelands  generally 
follow  one  of  two  somewhat  overlap- 
ping approaches  to  preserve  habitat 
and/or  species  values:  (1)  conserva- 
tion set  asides  or  (2)  mitigation.  Con- 
servation set  asides  are,  in  principle, 
the  preferred  solution.  In  practice 
some  habitat  is  dedicated  in  perpetu- 
ity as  a  nature  preserve;  preserve  de- 
sign currently  is  a  somewhat  ad  hoc 
process  and  will  be  discussed  more 


completely  in  a  subsequent  section  of 
this  paper.  Very  high  land  values 
may  dictate  mitigation  rather  than  on 
site  preservation  of  habitat. 

Mitigation  may  take  many  forms 
to  compensate  for  development  of 
xeric  pinelands.  Restoration  of  de- 
graded land  (Humphrey  et  al.  1985), 
not  necessarily  xeric  pinelands,  is  one 
method.  Another  tactic  is  to  purchase 
comparable  land  or  some  other  type 
of  land  of  equivalent  natural  value 
elsewhere  and  dedicate  it  to  preser- 
vation. A  formal  process  for  accom- 
plishing this  option  is  presently  un- 
der study  by  the  Florida  Game  and 
Fresh  Water  Fish  Commission. 

Preservation  of  habitat  is  the  basic 
purpose  of  conservation  set  asides 
and  mitigations.  The  value  of  these 
efforts  depends  on  the  proximity  to 
larger,  undeveloped  tracts  of  land, 
travel  corridors,  area  of  preserves, 
and  future  management  options. 

Another  form  of  mitigation  is  the 
relocation  of  sensitive  species  from 
tracts  of  land  to  be  developed  to  land 
dedicated  to  purposes  that  are  con- 
sistent with  the  long-term  survival  of 
the  relocated  species.  In  Rorida,  the 
gopher  tortoise  has  been  the  focus  of 
numerous  relocation  efforts.  Diemer 
(1984)  discussed  the  advantages  and 
disadvantages  of  relocation  of  go- 
pher tortoises  as  a  species  manage- 
ment strategy.  Formal  research  on 
gopher  tortoise  relocation  was  re- 
cently reported  (Proced.  Gopher  Tor- 
toise Relocation  Symp.,  27  June  1987, 
Gainesville,  FL,  in  press).  The  Rorida 
Game  and  Fresh  Water  Fish  Commis- 
sion regulates  relocations  by  a  permit 
system  based  on  a  standardized  relo- 
cation protocol. 

Preserve  Design 

Preserve  design  is  an  evolving  and 
controversial  area  of  conservation 
biology  (Diamond  1975, 1978;  Gilbert 
1980;  Higgs  1981;  Margules  1982; 
Pickett  and  Thompson  1978;  Pyle 
1980;  Soule  and  Simberioff  1986). 
Large  preserves  encompassing  a  mo- 


saic of  xeric  pinelands,  mesic  forests, 
and  seasonal  and  f>ermanent  wet- 
lands would  perhaps  offer  the  ideal 
landscape  unit  for  long-term  preser- 
vation of  amphibians,  reptiles,  and 
small  mammals  in  peninsular  Ror- 
ida. Because  preserves  on  private 
lands  must  be  justified  and  dedicated 
through  the  DRI  process,  economics 
dictates  preserve  units  of  minimal 
size.  Rarely  do  we  have  the  opportu- 
nity to  cluster  or  juxtapose  these 
small  units  to  take  advantage  of  the 
so  called  "rescue  effect"  (Brown  and 
Kodric-Brown  1977). 

In  practice,  conservation  set  asides 
tend  not  only  to  be  small  in  acreage 
but  also  only  of  one  habitat  type.  The 
latter  presents  a  dilemma  for  species 
whose  requirements  often  include 
two  or  more  contrasting  habitats.  For 
example,  the  gopher  frog  lives  in  tor- 
toise burrows  in  LLP/TO  sandhills 
during  late  spring,  summer  and  early 
fall  and  migrates  to  temporary  wet 
season  depressions  to  breed  in  win- 
ter and  early  spring  (Moler  and 
Franz  1987).  Thus  a  mosaic  of  up- 
land-wetland habitats  in  close  prox- 
imity are  essential  to  maintain  viable 
populations  of  this  species.  Other 
species  such  as  the  indigo  snake  have 
home  range  requirements  that  in- 
clude 122-202  ha  of  several  upland- 
wetland  habitat  types  (Moler  1985; 
Moler  unpublished  data).  It  is  obvi- 
ous that  large  landscape  units  are 
necessary  to  preserve  viable  popula- 
tions of  these  animals. 

We  have  prepared  a  detailed  pre- 
serve design  for  a  SPS  community 
within  the  city  of  Boca  Raton,  Rorida 
(Richardson  et  al.  1986;  Stout  et  al. 
1987;  manuscript  in  preparation). 
The  approach  taken  anticipated 
Soule  and  Simberioff  (1986)  and  rec- 
ommended the  area  of  the  preserve 
be  sufficient  to  support  a  minimum 
viable  population  (Franklin  1980)  of 
gopher  tortoises  because  of  their 
status  as  the  keystone  species.  Al- 
though biologically  reasonable,  this 
basis  for  determining  preserve  size  is 
often  economically  unrealistic  from 
the  view  point  of  the  private  land- 


103 


owner.  A  consortium  of  public  land- 
owners would,  however,  permit  the 
purchase  and  long-term  management 
of  the  preserve  as  recommended. 

Cox  et  al.  (1987)  offer  guidelines 
for  the  design  of  preserves  on  private 
lands  to  maintain  gopher  tortoise 
populations.  They  employed  the 
computer  simulation  model 
POPDYN  (Perez-Trejo  and  Samson 
manuscript)  to  determine  population 
viability  based  on  different  initial 
sizes.  Populations  of  40-50  individu- 
als were  found  to  be  likely  (>90%)  to 
persist  200  years.  Based  on  existing 
literature  on  home  range  require- 
ments. Cox  et  al.  (1987)  recom- 
mended a  minimum  preserve  of  10- 


20  ha,  depending  on  habitat  quality, 
to  support  40-50  tortoises. 

Another  approach  to  determining 
the  area  of  a  preserve  employs  ''inci- 
dence functions"  (Diamond  1978). 
Incidence  functions  are  species  spe- 
cific and  derived  from  data  sets 
which  reveal  the  fraction  of  plots 
(discrete  habitats)  of  different  areas 
that  actually  support  the  species.  It  is 
a  matter  of  judgement  as  to  the 
probability  of  occurrence,  e.g.  0.5  as 
opposed  to  0.7,  that  would  set  a 
lower  limit  to  area  for  an  acceptable 
preserve.  Data  sets  useful  for  evalu- 
ating this  approach  with  respect  to 
amphibians,  reptiles,  and  small 
mammals  in  xeric  pinelands  are  pres- 


Table  4.— Incidence  of  Gopherus  polyphemus  a  keystone  species,  and 
Podomys  fioridanus  in  xeric  pinelands  of  peninsular  Florida.  Presence  (+)  or 
absence  <-)  Is  indicated.  Study  sites  are  ranked  according  to  area  within 
the  xeric  pinelands.  Quantitative  sampling  of  the  12  LLP/TO  study  sites  con- 
sisted of  5  days  of  live-trapping  and  observation  at  invervals  of  3  months 
over  a  period  of  18  months  (1986-1988).  Study  sites  In  SPS  were  sampled  by 
live -trapping  and  observation  a  minimum  of  3  consecutive  days,  often  in 
the  same  season  of  consecutive  years  (Stout  et  al.  unpublished). 


Incidence  of  species  in  xeric  pineland 


Study  sites 


Area 


LLP/TO 


SPS 


(ha) 

Gopherus  Podomys  Gopherus  Podor 

Lake  Mary 

1.2 

+  — 

Morningside  Nature  Center 

2.0 

+  — 

San  Felasco 

4.1 

+  — 

Spruce  Creek 

4.1 

+  — 

Orange  City 

5.6 

+  — 

Bok  Tower 

9.3 

+  — 

Wekiwa  Springs 

9.7 

+  + 

Suwannee  River 

10.1 

+  + 

O'Lena 

10.5 

+  + 

J.  Butterfield  Brooks 

15.8 

+  + 

Starkey  Well  Field 

16.2 

+  — 

Sandhill  Boy  Scout  Camp 

16.2 

+  — 

Interlachen 

21.8 

+  — 

Yamato  Plaza 

2.8 

+  — 

Yamato  Scrub,  B 

3.2 

+  + 

Quantum  Park,  A 

4.4 

+  + 

Quantum  Park,  B 

4.4 

+  + 

Quantum  Park,  C 

4.8 

+  + 

Yamato  Scrub,  A 

8.5 

+  + 

Summit  Place 

10.5 

+  — 

Potomac  Road 

17.8 

+  + 

Cedar  Grove 

21.5 

+  + 

J.  Dickinson 

256.2 

+  + 

ently  lacking.  Table  4  provides  data 
we  have  gathered  on  area  of  discrete 
habitats  and  the  presence  or  absence 
of  gopher  tortoises  and  Florida  mice. 
It  is  apparent  that  tortoises  are  less 
area  sensitive  than  Florida  mice  and 
that  Florida  mice  are  patchy  in  occur- 
rence in  LLP /TO,  perhaps  only  sec- 
ondarily related  to  area. 

Incidence  functions  do  not  neces- 
sarily reveal  the  minimum  area  re- 
quired to  support  minimum  viable 
populations  (Franklin  1980).  We  be- 
lieve preserve  area  should  be  based 
on  providing  this  requirement,  par- 
ticularly when  preserves  are  isolated 
relative  to  average  dispersal  dis- 
tances of  keystone  species.  However, 
clusters  of  preserves  within  dispersal 
distances  of  keystone  species  may  be 
of  less  area  per  preserve  due  to  a 
high  likelihood  of  reinvasion  from 
nearby  populations  following  local 
population  extirpations  (Noss  and 
Harris  1986). 


Monagennent  of  Preserves  in  Xeric 
Pinelands 

The  future  viability  of  preserves  de- 
pends largely  on  their  ownership  af- 
ter development  of  the  surrounding 
landscape.  It  is  unlikely  that  home- 
owners associations  will  assume  the 
cost  of  management  if  preserves  re- 
main as  a  part  of  the  overall  develop- 
ment's "commons."  Public  owner- 
ship is  an  alternative  and  might  rest 
with  a  city,  county,  or  state.  Local 
governments  seem  more  appropriate; 
however,  funds  and  expertise  to 
manage  may  be  lacking.  One  pre- 
serve in  south  Rorida  is  designed  to 
border  a  city  park,  thus  allowing  its 
maintenance  and /or  management 
costs  to  be  assumed  over  time  as  part 
of  the  existing  park  system 
(Richardson,  personal  observation). 
Regardless  of  the  ownership,  a  com- 
mitment to  long-term  management 
must  be  achieved  if  a  preserve  is  to 
retain  natural  values. 

Management  options  for  nature 
preserves  range  from  a  decision  1)  to 


104 


do  nothing  and  let  nature  take  its 
course;  2)  to  manage  for  maintenance 
of  a  viable  ecosystem,  which  implies 
the  natural  biota,  including  amphibi- 
ans, reptiles,  and  small  mammals, 
will  be  present  in  proportion  to  their 
normal  abundance;  or  3)  to  focus 
management  on  the  needs  of  one  or 
more  species.  White  and  Bratton 
(1980)  have  exposed  the  folly  of  the 
first  management  option.  The  deci- 
sion to  emphasize  ecosystem  or  spe- 
cies management  depends  on  the  en- 
tity responsible  for  management, 
type  of  preserve,  management  objec- 
tives, area  of  the  preserve,  nature  of 
the  surrounding  lands,  relative  over- 
all or  regional  rarity  of  particular 
species,  and  the  resources  available 
for  management. 

Management  objectives  of  any 
preserve  should  focus  on:  1)  mainte- 
nance of  normal  ecosystem  proc- 
esses; 2)  conservation  of  soil;  3) 
maintenance  or  restoration  of  normal 
hydrologic  conditions;  4)  prevention 
of  establishment  of  exotic  species.;  5) 
and  prevention  of  human  encroach- 
ment (e.g.,  dumping,  ATVs,  etc.)  Be- 
yond these  generalities,  management 
of  preserves  is  an  idiosyncratic  proc- 
ess that  may  concern  endemic  spe- 
cies, genetics  of  inbred  populations, 
or  restoration  of  periodic  wild  fires. 

Xeric  pinelands  of  peninsular  Flor- 
ida depend  on  periodic  fires  to  main- 
tain their  structure  and  function 
(Laessle  1958a;  Abrahamson  1984). 
Thus  a  burning  program  is  essential 
in  the  management  of  LLP /TO  or 
SPS  preserves.  Spring  or  early  sum- 
mer prescribed  bums  are  routinely 
used  to  maintain  LLP/ TO  communi- 
ties on  state  parks.  Doran  et  al.  (1987) 
have  documented  prescribed  burns 
of  SPS  preserves  in  an  urban  setting 
based  on  rather  esoteric  fire  models 
developed  by  the  U.S.  Forest  Service. 
Gopher  tortoises  respond  favorably 
to  the  bums  (Stout  et  al.  1988).  A  mo- 
saic of  recovery  stages  in  SPS  may 
favor  beta  diversity  of  herptiles  and 
small  mammals.  Mushinsky  (1985) 
has  carefully  documented  the  re- 
sponse of  the  herpetofauna  to  a  vari- 


ety of  buming  schedules  in  LLP/TO. 
Diversity  and  abundance  of  amphibi- 
ans and  reptiles  was  increased  on 
experimental  plots  relative  to  un- 
bumed  controls.  Re-establishment  of 
the  pine  overstory  may  be  necessary 
to  produce  needle  cast  for  carrying 
fire  (Landers  and  Speake  1980). 

Management  of  conservation  set 
asides  and /or  easements  may  focus 
on  particular  species  or  combinations 
of  species.  The  smaller  the  preserve 
the  more  likely  that  a  reduced  suite 
of  species  will  be  present 
(Richardson  et  al.  1986).  Given  that  a 
fixed  area  is  available  for  manage- 
ment, major  efforts  to  enhance  or 
maintain  habitat  should  target  those 
species  that  can  maintain  viable 
populations  within  the  preserve 
(Shaffer  1986).  A  species  whose  mini- 
mum area  requirements  for  a  mini- 
mum viable  population  exceeds  the 
preserve  area  should  not  be  of  major 
concern  (Shaffer  and  Samson  1985); 
nonetheless,  such  species  can  benefit 
from  the  preserves  if  travel  corridors 
exist  (Harris  1984). 

DISCUSSION 

Xeric  pinelands  of  peninsular  Florida 
support  a  species-rich  assemblage  of 
reptiles,  amphibians,  and  small 
mammals.  (Growth  and  development 
continues  to  diminish  LLP/ TO  and 
SPS  habitats  to  the  detriment  of  the 
associated  biota.  Land  in  public  own- 
ership, e.g.  state  parks  and  forests, 
national  forests,  and  private  hold- 
ings, e.g.,  the  Archbold  Biological 
Station,  and  institutional  lands  such 
as  the  Ordway  and  Swisher  Pre- 
serves, jointly  owned  and  managed 
by  the  University  of  Florida  and  The 
Nature  Conservancy,  will  be  increas- 
ingly valuable  as  other  xeric  pine- 
lands are  converted  to  land  uses  not 
favorable  to  the  biota.  Thus,  manage- 
ment of  these  xeric  pinelands  will 
become  more  important  in  the  fu- 
ture. At  present  management  is 
largely  limited  to  prescribed  bums  to 
maintain  what  were  historically  fire 


climax  communities.  Thus,  fire  man- 
agement is  tantamount  to  small  ver- 
tebrate management. 

In  the  future  as  air  quality  stan- 
dards are  modified,  prescribed  bum- 
ing, particularly  in  or  near  urbanized 
areas,  will  be  restricted  or  eliminated 
as  a  management  option.  Alternative 
means  of  habitat  manipulation  need 
to  be  developed,  particularly  for  SPS. 

Basic  information  on  the  life  his- 
tory of  many  amphibians,  reptiles, 
and  small  mammals  of  xeric  pine- 
lands is  lacking.  The  Nongame  Wild- 
life Program  of  the  Rorida  Game  and 
Fresh  Water  Fish  Commission  has 
initiated  and  funded  rather  large 
scale  studies  of  SPS  and  LLP/TO 
communities.  These  studies  are  at  the 
community  level  and  largely  obser- 
vational. Management  needs  of  indi- 
vidual species  may  be  derived  only 
secondarily  from  this  research.  Stud- 
ies that  focus  on  particular  species 
will  ultimately  lead  to  more  refined 
habitat  management  guidelines.  The 
report  by  Cox  et  al.  (1987)  will  likely 
serve  as  a  model  for  the  preparation 
of  habitat  protection  guidelines;  man- 
agement follows  protection  (White 
and  Bratton  1980). 

Management  alternatives  at  the 
ecosystem  and  species  level  are 
needed  now  for  xeric  pinelands  on 
private  lands  undergoing  develop- 
ment. Regulation  of  development  in 
these  habitats  as  currently  practiced 
will  result  in  a  patchwork  of  small, 
isolated  nature  preserves.  Preserva- 
tion of  natural  habitat  in  a  developed 
landscape  is,  of  course,  desirable. 
However,  several  problems  remain: 
(1)  who  will  own  the  preserves,  (2) 
how  will  a  management  plan  be  pre- 
pared, and  (3)  who  will  be  respon- 
sible for  management?  Even  another 
decade  of  rapid  growth  in  peninsular 
Florida  may  result  in  a  few  hundred 
nature  preserves,  which  will  not  nec- 
essarily be  restricted  to  xeric  pine- 
land  habitat.  Ignoring  the  question  of 
ownership,  no  public  land  manage- 
ment agency  is  currently  capable  of 
assuming  the  charge  of  managing 
these  preserves.  Lack  of  manage- 


105 


ment,  e.g.,  failure  to  conduct  pre- 
scribed burning,  will  allow  succes- 
sional  changes  to  occur  to  the  detri- 
ment of  many  small  vertebrates  nar- 
rowly adapted  to  xeric  pinelands. 
Loss  of  habitat  and  species  values 
originally  used  by  jurisdictional 
agencies  to  secure  preserve  set  asides 
provides  a  potential  basis  for  private 
land  owners  to  request  development 
rights  on  the  land.  This  action  would 
defeat  the  entire  purpose  of  having 
conservation  set  asides. 

An  alternative  to  on  site  habitat 
protection  is  offered  by  Cox  et  al. 
(1987)  in  regard  to  preserving  habitat 
for  the  gopher  tortoise.  The  alterna- 
tive, a  Wildlife  Resource  Mitigation 
Fund  (WRMF),  allows  a  developer  to 
contribute  money  to  the  fund  to  miti- 
gate losses  of  valuable  wildlife  habi- 
tat on  lands  being  developed.  The 
collective  monies  of  several  develop- 
ment projects  would  allow  an  inde- 
pendent group  such  as  the  Trust  For 
Public  Lands  to  assist  in  the  purchase 
of  commensurate  lands  to  expand  an 
existing  public  park,  preserve  or  for- 
est. Management  is  more  likely  to  be 
applied  to  these  lands  and  ultimately 
the  resources  are  better  served  by  the 
public  agencies. 

ACKNOWLEDGMENTS 

We  thank  the  authors  of  the  papers 
cited  herein  for  their  efforts  and 
dedication  to  science.  Biologists  who 
contributed  to  our  knowledge  of 
xeric  pineland  include  but  are  not 
limited  to  the  following  individuals: 
Dan  Austin,  Don  Bethancourt,  Russ 
Burke,  Steve  Christman,  David  Cook, 
David  Corey,  Jim  Cox,  Joan  Diemer, 
Dick  Franz,  Larry  Harris,  Randy 
Kautz,  Jim  Layne,  Wayne  Marion, 
Paul  Moler,  Ron  Myers,  Reed  Noss, 
Cathy  Ryan,  Jim  Stevenson,  and  Don 
Wood.  Support  for  research  on  the 
ecology  of  sandhill  communities  was 
provided  by  the  Nongame  Wildlife 
Program,  RFP86-003,  of  the  Rorida 
Game  and  Fresh  Water  Fish  Commis- 
sion. Development  groups  that 


funded  work  by  the  authors  on  xeric 
pinelands  include  Hardy-Lieb  Devel- 
opment Corporation,  The  Adler 
Group,  and  Deutsch-Ireland  Proper- 
ties. The  Division  of  Recreation  and 
Parks  (FDNR),  Department  of  Biol- 
ogy, University  of  South  Florida,  and 
the  Department  of  Biological  Sci- 
ences, University  of  Central  Florida 
assisted  in  our  studies  in  a  variety  of 
ways.  We  thank  Beverly  Bamekow, 
Rita  Greenwell,  Barbara  Erwin,  and 
Nancy  Small  for  typing  the  manu- 
script. Lastly,  we  thank  Paul  E. 
Moler,  James  N.  Layne  and  Robert  C. 
Szaro  for  providing  excellent  sugges- 
tions to  improve  the  paper. 

LITERATURE  CITED 

Abrahamson,  Warren  G.  1984.  Post- 
fire  recovery  of  Florida  Lake 
Wales  ridge  vegetation.  American 
Journal  of  Botany  71:9-21. 

Abrahamson,  Warren  G.,  Ann  F. 
Johnson,  James  N.  Layne,  and  Pa- 
tricia A.  Peroni.  1984.  Vegetation 
of  the  Archbold  Biological  Station, 
Florida:  an  example  of  the  South- 
ern Lake  Wales  Ridge.  Florida  Sci- 
entist 47:209-250. 

Anonymous.  1976.  Developments  of 
regional  impact:  guidebook  for 
preparation  of  the  application  for 
development  approval.  65  p.  Divi- 
sion of  State  Planning,  Department 
of  Administration,  Tallahassee, 
Florida. 

Arata,  Andrew  A.  1959.  Effects  of 
burning  on  vegetation  and  rodent 
populations  in  a  longleaf  pine  tur- 
key oak  association  in  north  Cen- 
tral Florida.  Quarterly  Journal 
Florida  Academy  Sciences  22:94- 
104. 

Auffenberg,  Walter,  and  Richard 
Franz.  1982.  The  status  and  distri- 
bution of  the  gopher  tortoise  (Go- 
pherus  -polyphemus).  p.  95-126.  In 
North  American  tortoises:  conser- 
vation and  ecology.  R.  Bruce  Bury, 
editor.  Wildlife  Research  Report 
12. 126  p.  U.S.  Fish  and  Wildlife 
Service,  Washington,  D.C. 


Beckenbach,  J.  R.,  and  J.  W.  Ham- 
mett.  1962.  General  soil  map  of 
Florida.  1  p.  Florida  Agricultural 
Experiment  Stations  and  the  Soil 
Conservation  Service. 

Brown,  James  H.,  and  Astrid  Kodric- 
Brown.  1977.  Turnover  rates  in 
insular  biogeography:  effects  of 
immigration  on  extinction.  Ecol- 
ogy 58:445-449. 

Campbell,  Howard  W.,  and  Steven  P. 
Christman.  1982.  The  herpetologi- 
cal  components  of  Horida  sandhill 
and  sand  pine  scrub  associations, 
p.  163-171.  In  Herpetological  com- 
munities: a  symposium  of  the  So- 
ciety for  the  Study  of  Amphibians 
and  Reptiles  and  the  Herpetolo- 
gists'  League,  August  1977.  Nor- 
man J.  Scott,  Jr.,  editor.  239  p.  U.S. 
Fish  and  Wildlife  Service,  Wildlife 
Research  Report  13. 

Christman,  Steven  P.,  Howard  I. 
Kochman,  Howard  W.  Campbell, 
Charles  R.  Smith,  and  Wallace  S. 
Lippincott,  Jr.  Successional 
changes  in  community  structures: 
amphibians  and  reptiles  in  Horida 
sand  pine  scrub.  Unpublished 
manuscript. 

Cox,  James,  Douglas  Inkley,  and 
Randy  Kautz,  1987.  Ecology  and 
habitat  protection  needs  of  gopher 
tortoise  (Gopherus  polyphemus) 
populations  found  on  lands  slated 
for  large-scale  development  in 
Horida.  Horida  Game  and  Fresh 
Water  Fish  Commission  Nongame 
Wildlife  Program  Technical  Report 
no.  4.  75  p.  Tallahassee,  FL. 

Davis,  John  H.  1980.  General  map  of 
natural  vegetation  of  Florida. 
Agric.  Exp.  Sta.  Institute  of  Food 
and  Agricultural  Sciences  Circular 
S-178. 1  p.  University  of  Horida, 
Gainesville. 

Davis,  Mary  M.  1984.  Phenology  of 
flowering  plants  on  experimental 
fire  plots,  St.  Marks  National 
Wildlife  Refuge.  M.  S.  Thesis.  54  p. 
Horida  State  University.  Tallahas- 
see, FL. 

Diamond,  Jared  M.  1975.  The  island 
dilemma:  lessons  of  modem  bio- 
geographic  studies  for  the  design 


106 


of  natural  reserves.  Biological 
Conservation  7:129-146. 

Diamond,  J.  M.  1978.  Critical  areas 
for  maintaining  viable  populations 
of  species,  p.  27-40.  In  The  break- 
down and  restoration  of  ecosys- 
tems. M.  W.  Holdgate  and  M.  J. 
Woodman,  editors.  Plenum  Press, 
New  York,  N.  Y. 

Diemer,  Joan  E.  1984.  Tortoise  reloca- 
tion in  Florida:  solution  or  prob- 
lem? Proc.  1984  Symposium  Des- 
ert Tortoise  Council:  131-135. 

Doran,  R.  F.,  Donald  R.  Richardson, 
and  Richard  E.  Roberts.  1987.  Pre- 
scribed burning  of  the  sand  pine 
scrub  community:  Yamato  Scrub, 
a  test  case.  Florida  Scientist  50:184- 
192. 

Eisenberg,  J.  F.  1983.  The  gopher  tor- 
toise as  a  keystone  species,  p.  1-4. 
In  The  gopher  tortoise:  a  keystone 
species.  R.  J.  Bryant  and  R.  Franz, 
editors.  Proceedings  of  the  4th 
Annual  Meeting  of  the  Gopher 
Tortoise  Council.  Rorida  State 
Museum,  Gainesville,  FL. 

Franklin,  Ian  Robert.  1980.  Evolution- 
ary change  in  small  populations, 
p.  135-149.  In  Conservation  biol- 
ogy: an  evolutionary-ecological 
perspective.  Michael  E.  Soule  and 
Bruce  A.  Wilcox,  editors.  395  p. 
Sinauer  Assoc.,  Sunderland,  MA. 

Gilbert,  F.  S.  1980.  The  equilibrium 
theory  of  island  biogeography: 
fact  or  fiction?  Journal  of  Biogeo- 
graphy 7:209-235. 

Givens,  Kerry  T.,  James  N.  Layne, 
Warren  G.  Abrahamson,  and 
Susan  C.  White-Schuler.  1984. 
Structural  changes  and  succes- 
sional  relationships  of  five  Florida 
Lake  Wales  Ridge  plant  communi- 
ties. Bulletin  Torrey  Botanical 
Club  111:8-18. 

Harris,  Larry  D.  1984.  The  frag- 
mented forest.  211  p.  The  Univer- 
sity of  Chicago  Press,  Chicago,  IL. 

Higgs,  A.  J.  1981.  Island  biogeogra- 
phy theory  and  nature  reserve  de- 
sign. Journal  of  Biogeography 
8:117-124. 

Humphrey,  Stephen  R.,  John  F.  Eis- 
enberg, and  Richard  Franz.  1985. 


Possibilities  for  restoring  wildlife 
of  a  longleaf  pine  savanna  in  an 
abandoned  citrus  grove.  Wildlife 
Society  Bulletin  13:487-496. 

Kalisz,  Paul  J.,  and  Earl  L.  Stone. 
1984.  The  longleaf  pine  islands  of 
the  Ocala  National  Forest,  Florida: 
a  soil  study.  Ecology  65:1743-1754. 

Kurz,  Herman.  1942.  Florida  dunes 
and  scrub  vegetation  and  geology. 
Florida  Geological  Survey  Bulletin 
23:15-154. 

Laessle,  Albert  M.  1942.  The  plant 
communities  of  the  Welaka  Area. 
University  of  Florida  Publication, 
Biological  Science  Series  4:1-143. 

Laessle,  Albert  M.  1958a.  The  origin 
and  successional  relationship  of 
sandhill  vegetation  and  sand  pine 
scrub.  Ecological  Monographs 
28:361-387. 

Laessle,  Albert  M.  1958b.  A  report  on 
succession  studies  of  selected 
plant  communities  on  the  Univ.  of 
Fla.  conservation  reserve,  Welaka, 
Florida.  Quarterly  Journal  Florida 
Academy  Science  21:101-112. 

Laessle,  A.  M.  1967.  Relationship  of 
sand  pine  scrub  to  former  shore 
lines.  Quarterly  Journal  Florida 
Academy  Science  30:269-286. 

Landers,  J.  Larry,  and  Dan  W. 
Speake.  1980.  Management  needs 
of  sandhill  reptiles  in  Southern 
Georgia.  Annual  conference 
Southeastern  Association  Fish  and 
Wildlife  Agencies.  34:515-529. 

Latham,  Pamela  J.  1985.  Structural 
comparisons  of  sand  pine  scrubs 
of  east-central  Florida.  M.  S.  The- 
sis. 106  p.  University  of  Central 
Florida,  Orlando. 

Layne,  James  N.  1969.  Nest-building 
behavior  in  three  species  of  deer 
mice,  Peromyscus.  Behaviour 
35:288-303. 

Layne,  James  N.  1978.  Peromyscus  flo- 
ridanus  (Chapman),  p.  21-22.  In 
Rare  and  endangered  biota  of 
FLorida.  Vol.  1:  mammals.  James 
N.  Layne,  editor.  52  p.  University 
Presses  of  Florida,  Gainesville,  Fl. 

Margules,  C,  A.  J.  Higgs,  and  R.  W. 
Rafe.  1982.  Modern  biogeographic 
theory:  are  there  any  lessons  for 


nature  reserve  design?  Biological 
Conservation  24:115-128. 

Moler,  Paul  E.  1985.  Distribution  of 
the  eastern  indigo  snake,  Dry- 
marchon  corais  couperi,  in  Rorida. 
Herpetological  Review  116:37-38. 

Moler,  Paul  E.,  and  Richard  Franz. 
1987.  Wildlife  values  of  small,  iso- 
lated wetlands  in  the  southeastern 
coastal  plain,  p.  234-241.  In  Proc. 
3rd  S.  E.  Nongame  and  Endan- 
gered Wildlife  Symposium.  R.  R. 
Odom,  K.  A.  Riddleberger  and  J. 
C.  Ozier,  editors.  Georgia  Depart- 
ment Natural  Resources,  Game 
and  Fish  Division. 

Monk,  Carl  D.  1968.  Successional  and 
environmental  relationships  of  the 
forest  vegetation  of  North  Central 
Rorida.  The  American  Midland 
Naturalist  79:441-457. 

Mushinsky,  Henry  R.  1984.  Fire  and 
the  Rorida  sandhill  herpetofaunal 
community:  with  special  attention 
to  responses  of  Cnemidophorus 
sexlineatus.  Herpetologica  41:333- 
342. 

Myers,  Ronald  L.  1985.  Fire  and  the 
dynamic  relationship  between 
Rorida  sandhill  and  sand  pine 
scrub  vegetation.  Bulletin  Torrey 
Botanical  Club  112:241-252. 

Norse,  Elliott  A.,  Kenneth  L.  Rosen- 
baum,  David  S.  Wilcove,  Bruce  A. 
Wilcox,  William  H.  Romme, 
David  W.  Johnston,  and  Martha  L. 
Stout.  1986.  Conserving  biological 
diversity  in  our  national  forests. 
116  p.  The  Wilderness  Society. 

Noss,  Reed  F.,  and  Larry  D.  Harris. 
1986.  Nodes,  networks,  and 
MUMS:  preserving  diversity  at  all 
scales.  Environmental  Manage- 
ment 10:299-309. 

Pickett,  S.  T.  A.,  and  John  N.  Th- 
ompson. 1978.  Patch  dynamics 
and  the  design  of  nature  reserves. 
Biological  Conservation  13:27-37. 

Pyle,  Robert  Michael.  1980.  Manage- 
ment of  nature  reserves,  p.  319- 
327.  In  Conservation  Biology:  An 
evolutionary-ecological  perspec- 
tive. Michael  E.  Soule  and  Bruce 
A.  Wilcox,  editors.  395  p.  Sinauer 
Associates,  Sunderland,  Mass. 


107 


Richardson,  Donald  Robert.  1977. 
Vegetation  of  the  Atlantic  coastal 
ridge  of  Palm  Beach  County,  Flor- 
ida. Florida  Scientist  40:281-330. 

Richardson,  Donald  R.,  I.  Jack  Stout, 
Richard  E.  Roberts,  Daniel  F. 
Austin,  and  Taylor  R.  Alexander. 
1986.  Design  and  management 
recommendations  for  a  sand  pine 
scrub  preserve:  the  Yamato  scrub. 
124  p.  Ecological  Consultants, 
Tampa,  Florida. 

Shaffer,  Mark  L.  1986.  The  met- 
apopulation  and  species  conserva- 
tion: the  special  case  of  the  north- 
ern spotted  owl.  p.  86-99.  In  Ecol- 
ogy and  management  of  the  spot- 
ted owl  in  the  Pacific  Northwest, 
Ralph  J.  Gutierrez,  and  Andrew  B. 
Carey,  tech  editors.  USDA  Forest 
Service.  General  Technical  Report 
PNW-185, 119  p.  Pacific  North- 
west Forest  and  Range  Experiment 
Station,  Portland,  Oregon. 

Shaffer,  Mark  L.,  and  F.  B.  Samson. 
1985.  Population  size  and  extinc- 
tion: a  note  on  determining  critical 
population  sizes.  American  Natu- 
ralist 125:144-152. 

Smith,  Charles  R.  1982.  Food  re- 
source partitioning  of  fossorial 
Rorida  reptiles,  p.  173-178.  In  Her- 
petological  communities:  a  sympo- 
sium of  the  Society  for  the  Study 
of  Amphibians  and  Reptiles  and 
the  Herpetologists'  League,  Au- 
gust 1977.  Norman  J.  Scott,  Jr., 
editor.  239  p.  U.S.  Fish  and  Wild- 
life Service,  Wildlife  Research  Re- 
port 13. 

Soule,  Michael  E.,  and  Daniel  Sim- 
berloff.  1986.  What  do  genetics 
and  ecology  tell  us  about  the  de- 
sign of  nature  reserves?  Biological 
Conservation  35:19-40. 

Stout,  I.  Jack.  1981.  Conservation  bi- 
ology of  Peromycus  floridanus.  Flor- 
ida Scientist  44  (Supplement  1):37. 

Stout,  I.  Jack.  1982.  Small  mammal 
community  structure  in  sand  pine 
scrub.  Bulletin  Ecological  Society 
America  63:68-69. 

Stout,  I.  Jack,  Donald  R.  Richardson, 
and  Richard  E.  Roberts.  1988.  Re- 
sponse of  resident  and  relocated 


gopher  tortoises  to  a  prescribed 
bum  in  a  sand  pine  scrub  commu- 
nity. Proced.  Gopher  Tortoise  Re- 
location Symp.,  27  June  1987, 
Gainesville,  FL. 

Stout,  I.  Jack,  Donald  R.  Richardson, 
Richard  E.  Roberts,  and  Daniel  F. 
Austin.  1987.  Design  of  a  nature 
preserve  in  a  subtropical  urbaniz- 
ing landscape:  application  of  ecol- 
ogic  and  genetic  principles.  Bulle- 
tin Ecological  Society  of  America 
68:423-424. 

Swindell,  C.  E.  1987.  Multivariate 
analysis  of  competition  and  habi- 
tat use  by  small  mammals  in  Flor- 
ida Sand  Pine  Scrub.  M.  S.  Thesis. 
59  p.  University  of  Central  Florida, 
Orlando. 

Veno,  Patricia  A.  1976.  Successional 
relationships  of  five  Florida  plant 
communities.  Ecology  57:498-508. 

White,  Peter  S.,  and  Susan  P.  Bratton. 
1980.  After  preservation:  philo- 
sophical and  practical  problems  of 
change.  Biological  Conservation 
18:241-255. 

White,  W.  A.  1970.  The  geomorphol- 
ogy  of  the  Florida  peninsula.  Bu- 
reau of  Geology,  Tallahassee.  164 
p.  Geological  Bulletin  No.  51. 

Wood,  Don  A.  1987.  Official  lists  of 
endangered  and  potentially  en- 
dangered fauna  and  flora  in  Flor- 
ida. 19  p.  Florida  Game  and  Fresh 
Water  Fish  Commission. 


108 


Distribution  and  Habitat 
Associations  of  Herpetofauna 
in  Arizona:  Comparisons  by 
Habitat  Type^ 


Abstract. —Between  1977  and  1981 ,  the  Bureau  of 
Land  Management  conducted  extensive  surveys  of 
Arizona's  herpetofauna  in  16  different  habitat  types 
on  approximately  8.5  million  acres  of  public  lands. 
This  paper  describes  results  of  one  of  the  most  exten- 
sive surveys  ever  conducted  on  amphibian  and  rep- 
tile communities  in  North  America. 


K.  Bruce  Jones^ 


With  the  passage  of  the  Federal  Land 
PoUcy  and  Management  Act  in  1976, 
the  Bureau  of  Land  Management 
(ELM)  was  mandated  to  keep  an  in- 
ventory of  resources  on  public  lands. 
Information  collected  during  inven- 
tories or  surveys  was  then  to  be  used 
to  identify  issues  for  land  use  plan- 
ning and  opportunities  for  land  man- 
agement. The  BLM  made  a  decision 
to  collect  data  on  all  major  wildlife 
groups  and  their  habitats 

Early  in  the  development  of  its  in- 
ventory program,  the  BLM  recog- 
nized a  need  to  devise  a  strategy  that 
would  compare  animal  distributions 
and  abundance  to  habitats.  This 
strategy  was  important  since  the 
BLM  manages  wildlife  habitats  and 
not  wildlife  populations. 

In  1977  the  BLM  initiated  invento- 
ries of  wildlife  resources  on  public 
lands.  At  that  time,  considerable  in- 
formation was  already  available  on 
game  species.  However,  data  on 
nongame  species  were  mostly  lack- 
ing. As  a  result,  priority  was  given  to 
collecting  data  on  nongame  species 
and  their  habitats. 

Amphibians  and  reptiles  are  im- 
portant members  of  the  nongame 
fauna.  They  use  a  wide  range  of  habi- 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff ,  Arizona,  July  19-21.  1988). 

^K.  Bruce  Jones  is  a  Research  Ecologist 
with  the  Environmental  Protection  Agency. 
Environmental  Monitoring  Systems  Labora- 
tory. Las  Vegas.  Nevada  89193. 


tats  and  are  often  good  indicators  of 
habitat  conditions  (Jones  1981a). 
Therefore,  in  order  to  obtain  infor- 
mation on  these  animals,  principally 
for  land-use  planning,  the  BLM  con- 
ducted extensive  inventories  of  am- 
phibians and  reptiles  by  habitat  type. 
This  inventory  included  a  scheme 
whereby  associations  between  am- 
phibians and  reptiles  and  certain  n\i- 
crohabitats  could  be  determined.  The 
inventory,  conducted  between  1977 
and  1981,  was  one  of  the  most  com- 
prehensive surveys  of  herpetological 
communities  ever  conducted  in 
North  America  (27,885  array-nights 
in  16  habitat  types  over  a  five-year 
period).  It  also  represents  the  first 
large-scale  effort  to  quantitatively 
compare  herpetofaunas  associated 
with  ecosystems.  This  paper  reports 
the  results  of  these  surveys,  includ- 
ing species  distributions  and  associa- 
tions with  microhabitats  and  habitat 
types  (plant  communities). 


STUDY  AREA 

The  study  area  consisted  of  approxi- 
mately 3,441,296  ha  (8.5  million 
acres)  of  public  lands  located  in  cen- 
tral, west-central,  southwestern,  and 
northwestern  Arizona  (fig.  1).  Sixteen 
different  habitat  types  were  deline- 
ated within  this  area,  primarily  from 
an  existing  map  of  vegetation  asso- 
ciations (Brown  et  al.  1979).  Field  re- 
connaissance allowed  more  local  as- 
sociations to  be  recognized  within 


Figure  1.— The  study  area. 


those  presented  by  Brown  et  al. 
(1979).  For  example,  because  of  the 
scale  of  their  map.  Brown  et  al.  (1979) 
failed  to  recognize  several  small,  rel- 
ict stands  of  chaparral  woodland, 
although  Brown  (1978)  had  noted  the 
presence  of  chaparral  woodland 
vegetation  at  several  small  sites  (see 
Jones  et  al.  1985  for  the  importance  of 
small  woodland  stands  to  certain 
herpetofauna).  Therefore,  the  habitat 
type  map  used  to  allocate  samples  in 
this  study  drew  upon  the  Brown 
(1978)  and  Brown  et  al.  (1979)  maps. 


109 


and  results  of  field  reconnaissance. 
For  detailed  descriptions  of  these 
habitat  types  see  Jones  (1981b)  and 
Buse  (1981). 


SAMPLING  METHODS 

Amphibian  and  reptile  distribution 
and  abundance  by  habitat  type  were 
determined  by  on-the-ground  sam- 
pling efforts  between  October,  1977, 
and  July,  1981.  Samples  were  ob- 
tained by  three  methods.  The  most 
extensive  sampling  was  accom- 
plished with  a  pit-fall  trapping 
method  (array)  consisting  of  a  series 


of  18.3 1  (5  gal)  plastic  containers  bur- 
ied in  the  ground  and  connected  by 
0.41  m  (8  inches)  high  aluminum 
drift  fence;  one  trap  was  located  in 
the  center  with  three  evenly  dis- 
persed (120°)  peripheral  traps  7.14  m 
(25  ft)  from  the  center  (Jones  1981a, 
Jones  1986).  This  modified  array 
method  was  designed  specifically  for 
sampling  amphibians  and  reptiles  in 
desert  habitats  (see  Jones  1986  for  a 
comparison  of  this  procedure  with 
the  original  array  trapping  scheme 
designed  by  Christman  and 
Campbell  1982).  A  total  of  183  arrays 
were  used  to  sample  16  different 
habitat  types  (see  table  1  for  sum- 


Table  1  .—Sampling  effort  in  each  habitat  type. 


#of 
arrays 


#  of  trap 
nights 


#  of  road 
riding  road 
transects 


#  of  field 
searches 


Elevation 
range  (m  (ff.)) 


Ponderosa  Pine  Woodland  (PP) 

5  745  10  15  1677-2531  (5500-8300) 
Pinyon-Juniper  Woodland  (PJ) 

9  945  14  20        1311-1921  (4300-6300) 

Sagebrush  (Great  Basin  Desert)  (SB) 

3  270  12  12        1311-1830  (4300-6000) 

Closed  Chaparral  (CC) 

18  2168  18  20       1250-2287  (4100-7500) 

Open  Chaparral  (OC) 

13  1950  22  25         762-1311  (2500-4300) 

Desert  Grassland  (DO) 

11  1155  15  14       1006-1525  (3300-5000) 

Disclimax  Desert  Grassland  (DD) 

3  300  11  10         884-1311  (2900-4300) 

Mixed  Broadleaf  Riparian  (MB) 

6  784  8  18  884-2287  (2900-7500) 
Cottonwood-Wiilow  Riparian  (CW) 

13  3145  23  28         549-1372  (1800-4500) 

Juniper  Woodland  (mixed  shrub)  (JM) 

9  1080  19  22        793-1342  (2600-4400) 

Canotia  Mixed  Shrub  (CA) 

3  265  11  16        884-1189  (2900-3900) 

Mesquite  Bosque  (floodplain  woodland)  (ME) 

15  3025  18  22  213-915  (700-3000) 
Mixed  Riparian  Scrub  (Xeroriparian)  (MR) 

16  2640  23  18         229-1220  (750-4000) 
Mojave  Desertscrub  (MD) 

15  1803  25  24         610-1220  (2000-4000) 

Sonoran  Desertscrub  (Arizona  Upland)  (SD) 

22  3970  33  27         335-1189  (1100-3900) 

Creosotebush  (Lower  Colorado)  (CB) 

22  3640  32  18  213-915  (700-3000) 


mary  of  sampling  effort  in  each  habi- 
tat typ>e).  Arrays  were  placed  so  that 
microhabitat  variability  within  each 
habitat  type  was  sampled.  The  num- 
ber of  arrays  used  to  sample  habitat 
types  was  partially  influenced  by  the 
size  of  habitats;  generally,  more  ex- 
tensive habitats  received  prop)ortion- 
ally  larger  samples.  However,  certain 
habitats  (e.g.,  riparian)  were  known 
to  be  great  sources  of  diversity 
within  desert  regions;  therefore,  pri- 
ority was  given  to  obtaining  larger 
samples  within  these  habitats.  Once 
placed  into  the  ground,  arrays  were 
continuously  open  for  a  minimum  of 
60  days.  Some  arrays  (60)  were  open 
for  9  months.  Generally,  samples 
were  taken  during  the  spring,  sum- 
mer, and  fall.  However,  some  arrays 
(17)  were  open  only  during  spring 
months  and  others  only  in  the  fall 
(12).  The  opening  of  new  arrays  at 
different  locations,  and  the  closing  of 
other  arrays,  were  often  dictated  by 
BLM's  predetermined  resource  plan- 
ning schedule. 

Since  some  amphibians  and  many 
snakes  could  not  be  effectively 
sampled  by  pit-fall  traps,  it  was  nec- 
essary to  use  two  other  field  tech- 
niques. Road  riding,  consisting  of 
traveling  roads  from  dusk  to  ap- 
proximately 2300  h  throughout  de- 
lineated habitat  types,  was  used  to 
determine  the  occurrence  of  amphibi- 
ans and  medium  and  large  snakes 
(see  table  1  for  sampling  effort  within 
each  habitat  type). 

Time-constraint  searches  (Bury 
and  Raphael  1983),  consisting  of 
walking  along  permanent  and  tem- 
porary water  sources  (natural  and 
man-made)  at  night,  were  used  to 
verify  the  presence  of  frogs  and 
toads  at  waters  within  habitat  types 
(see  table  1  for  sampling  effort  within 
each  habitat  type). 

Finally,  to  get  an  idea  of  the 
known  distribution  of  amphibians 
and  reptiles  within  the  study  area,  I 
obtained  records  from  7  museums 
known  for  their  outstanding  collec- 
tions of  amphibians  and  reptiles  from 
the  Southwest:  the  University  of 


110 


Michigan,  Arizona  State  University, 
the  University  of  New  Mexico, 
Northern  Arizona  University,  the 
University  of  Arizona,  the  Los  Ange- 
les County  Museum,  and  the  Univer- 
sity of  California  at  Berkeley.  In  addi- 
tion, these  data  were  used  to  com- 
pare the  past  distribution  of  amphibi- 
ans and  reptiles  within  the  study 
area  with  that  obtained  during  the 
BLM's  inventories. 

Microhabitat  data  were  collected 
on  each  array  site  and  along  roads  by 
a  modified  point-intercept  method 
consisting  of  100  sample  points  sepa- 
rated by  8  m  (26  ft)  along  a  randomly 
determined  compass  line;  on  array 
sites,  the  center  of  the  line  crossed 
over  the  array.  At  each  point,  the  fol- 
lowing measurements  were  taken:  (1) 
vertical  distribution  of  vegetation  be- 
tween 0-0.6  m  (0-2  ft),  0.6-1.7  m  (2-6 
ft),  1.7-6.0  m  (6-20  ft),  and  >  6  m  (20 
ft)  (each  time  vegetation  occurred  in 
a  height  class  above  the  point,  a  con- 
tact or  "hit"  was  recorded);  (2)  pene- 
tration to  the  nearest  cm  into  the  soil 
by  a  pointed  metal  rod  (1  cm  in  di- 
ameter); (3)  depth  of  leaf  litter  (if 
present);  (4)  depth  of  other  litter  such 
as  debris  heaps  (piles  of  logs,  leaves 
and  other  dead  vegetative  material) 
and  rotting  logs;  (5)  characterization 
of  surface  rock  into  size  classes  of 
sand,  gravel  (<  1  cm  or  0.4  inches  in 
diameter),  cobble  (1  to  5  cm  or  0.4  to 
2  inches  in  diameter),  stone  (>  5  cm 
or  2  inches  in  diameter),  and  bed- 
rock. Vegetation  cover  and  percent- 
age of  the  surface  occupied  by  each 
rock  and  litter  size  class  was  deter- 
mined by  comparing  the  number  of 
"hits"  in  each  category  (e.g.,  litter) 
with  the  total  number  of  sample 
points  (100).  Plant  species  were  also 
recorded  along  each  100  point 
transect  (see  table  1  for  the  number 
of  microhabitat  samples  taken  in 
each  habitat  type). 

DATA  ANALYSIS 

I  calculated  relative  abundance  of 
each  amphibian  and  reptile  species  as 


the  total  number  of  any  species 
caught  during  a  24-hour  period  (ar- 
ray-night). Relative  abundance  was 
determined  for  each  species  on  array 
sites  by  taking  the  greatest  number  of 
individuals  of  a  species  trapped  dur- 
ing a  30-day  period  and  dividing  by 
the  number  of  days.  This  calculation 
was  used  because  of  monthly  differ- 
ences in  species'  activity  patterns. 
The  number  of  arrays  in  which  a  spe- 
cies was  trapped  in  each  habitat  type 
also  was  compiled  to  determine  how 
widespread  a  species  was  within  in- 
dividual habitat  types. 

A  principal  components  analysis 
(Pimental  1979)  was  performed  to 
compress  microhabitat  data  into  a 
smaller,  depictable  subset.  Mean  fac- 
tor scores  of  compressed  microhabi- 
tat data  were  computed  for  each 
habitat  type  and  plotted  on  a  3  vector 
(axis)  graph.  Similarly,  mean  factor 
scores  of  compressed  microhabitat 
data  were  computed  for  each  am- 
phibian and  reptile  species  (turtles 
were  excluded  because  aquatic  mi- 
crohabitats  were  not  measured). 
These  scores  were  calculated  for  each 
species  by  averaging  mean  factor 
scores  for  microhabitats  on  which  a 
species  occurred. 

Species  richness  (total  number  of 
species)  and  species  diversity  were 
calculated  for  each  habitat  type.  Two 
calculations  of  species  richness  for 
habitats  were  used;  one  that  used 
only  array  data  and  one  that  used  all 
data  (array,  road-riding,  and  field- 
search  data).  In  addition,  the  average 
number  of  species  collected  per  array 
(30-day  period)  was  calculated  and 
compared  to  overall,  array-deter- 
mined, species  richness.  Species  di- 
versity of  each  habitat  was  deter- 
mined from  a  Shannon-Weaver  di- 
versity index  (Hair  1980):  H'  =  E  p, 
logjQ  Pi/  where  s  =  the  number  of  spe- 
cies and  Pj  is  the  proportion  of  the 
total  number  of  individuals  consist- 
ing of  the  i**^  species.  Average  species 
diversity  per  array  was  calculated  for 
each  habitat  type.  Because  road-rid- 
ing and  field  searches  did  not  yield 
estimates  of  relative  abundance  simi- 


lar to  arrays,  only  array  data  were 
used  to  calculate  species  diversity. 

Two  types  of  cluster  analysis  were 
used  to  determine  similarities  among 
habitat  types.  The  first  cluster  analy- 
sis was  performed  only  on  array 
data,  and  it  was  based  on  euclidean 
distances  (Pimental  1979).  Calcula- 
tion of  euclidean  distances  between 
habitats  were  based  on  a  combina- 
tion of  species'  presence  or  absence 
on  a  site  and  similarity  in  species' 
dominance  (relative  abundance)  be- 
tween habitats.  Since  medium  and 
large  snakes  (>  0.5  m  or  1.5  ft)  are  not 
readily  caught  in  pit-fall  traps,  their 
relative  abundances  could  not  be  cal- 
culated accurately.  To  compare  the 
overall  herpetofaunas  of  habitat 
types,  a  second  cluster  analysis  was 
performed.  This  procedure  involved 
calculation  of  Simpson  similarity  co- 
efficients (Pimental  1979).  These  coef- 
ficients were  then  submitted  to  a 
cluster  analysis.  Unlike  the  analysis 
of  array  data  via  euclidean  distances, 
the  use  of  Simpson  similarity  coeffi- 
cients in  a  cluster  analysis  did  not 
consider  relative  dominance  in  calcu- 
lating distances  between  habitats. 

Several  thousand  site  specific  dis- 
tributional records  were  obtained  for 
amphibians  and  reptiles  within  the 
study  (to  16.2  ha  or  40  acre  accu- 
racy). These  individual  records  were 
too  numerous  to  report  here;  detailed 
locality  records  for  each  species  are 
kept  at  the  Bureau  of  Land  Manage- 
ment's Phoenix  District  Office. 


RESULTS 

Microhabitats 

A  principal  components  analysis 
(PCA)  of  microhabitats  yielded  3 
compressed  habitat  components 
(axes),  and  the  cumulative  propor- 
tion of  eigenvalues  was  <  1 .0  with 
83%  of  the  variability  accounted  for 
by  the  matrix  (p  <  .05).  This  analysis 
revealed  large  differences  in  the  mi- 
crohabitat among  habitat  types  (fig. 
2).  Desert  grassland,  disclimax  desert 


111 


grassland,  and  creosotebush  habitats 
had  open  canopies  and  low-height 
vegetative  structure,  whereas 
pinyon-juniper,  mixed  riparian 
scrub,  cottonwood-willow  riparian, 
mixed  broadleaf  riparian,  and  pon- 
derosa  pine  had  tree  canopies  and 
large  amounts  of  vegetative  debris, 
such  as  leaf  litter  and  logs,  on  their 
surfaces  (fig.  2).  Closed  and  open 
chaparral  habitats  consisted  of 
shrubs  with  rocky  surfaces,  and 
Sonoran  Desert  had  a  combination  of 
trees  and  shrubs  and  rocky  surfaces 
(fig.  2). 

Species  Distributions  and 
Abundances 

A  total  of  28  species  of  lizards,  30 
snakes,  4  turtles,  9  toads,  3  frogs,  and 
1  salamander  were  observed  or 
trapped  during  the  study.  Sceloporus 


magister,  Urosaurus  ormtus,  Uta 
stansburiam,  and  Cnemidophorus  tigris 
were  the  most  widely  distributed 
and  abundant  lizards  throughout  the 
study  area's  habitat  types  (table  2). 
These  lizards  also  consistently  oc- 
curred on  a  large  number  of  sites 
within  each  habitat  type  (table  2). 
Certain  lizards,  such  as  Gambelia  wis- 
lizeni,  Phrynosoma  solare,  and  Dip- 
sosaurus  dorsalis  occurred  only  on 
lower  elevation  (<  915  m  or  3000  ft), 
desert  habitats,  and  other  lizards, 
such  as  Sceloporus  undulatus,  Gerrhon- 
otus  kingi,  and  Phrynosoma  douglassi 
occurred  only  on  higher  elevation  (> 
1220  m  or  4000  ft)  habitats  (table  2). 
Some  species,  such  as  Eumeces  gilberti 
and  Cophosaurus  texam,  were  princi- 
pally found  on  higher  elevation  habi- 
tats, but  also  inhabited  cottonwood- 
willow  riparian  habitats  at  lower  ele- 
vations (549-915  m  or  1800-3000  ft) 
(table  2).  Certain  lizards,  such  as 


1.5 
Trees 


Component  II 


Figure  2.— Mean  factor  scores  of  microhabitats  for  habitat  types.  (Abbreviations  correspond 
to  thiose  listed  for  hiabitats  in  table  1 .) 


Cnemidophorus  burti  and  Eumeces  ob- 
soletus,  had  limited  distributions 
within  the  study  area  (table  2);  C. 
burti  is  principally  distributed  in  the 
Sonoran  Desert  and  Desert  Grass- 
land habitats  in  extreme  southern 
Arizona  and  Mexico,  and  E.  obsoletus 
only  occurs  in  the  chaparral  habitat 
type  in  the  extreme  eastern  portion 
of  the  study  area.  Although  re- 
stricted to  higher  elevation  and  ripar- 
ian habitats  throughout  most  of  the 
study  area,  C.  texana  was  found  in 
Sonoran  Desert  in  the  extreme  east- 
ern portion  of  the  study  area.  Most 
lizards  occurred  throughout  the 
study  area  where  suitable  habitat 
was  present  and  were  not  restricted 
by  geographic  range. 

A  PCA  revealed  that  lizards  dif- 
fered in  their  associations  with  cer- 
tain microhabitats  (fig.  3).  Some  of 
the  widely  distributed  species,  such 
as  Cnemidophorus  tigris  and  Uta 
stansburiana,  showed  little  association 
with  any  of  the  principal  components 
(fig.  3),  although  the  distribution  of 
other  common  species,  such  as  Sce- 
loporus magister  and  Urosaurus  ornatus 
was  highly  correlated  with  the  pres- 
ence of  vegetation  debris  (fig.  3). 
More  than  half  of  the  lizards  oc- 
curred on  sites  with  relatively  open 
canopies  and  shrubs  or  grasses,  and 
many  also  preferred  rocky  substrates 
(fig.  3).  Dipsosaurus  dorsalis,  Callisau- 
rus  draconoides,  and  Gambelia  wislizeni 
occurred  on  sites  with  sand 
substrate.  Gerrhonotus  kingi  and 
Eumeces  gilberti  occurred  on  sites 
with  large  amounts  of  vegetative  de- 
bris, medium  to  high  canopies,  and 
rocky  substrates,  and  Xantusia  vigilis 
on  sites  with  similar  substrate  but 
with  a  more  open  canopy  (fig.  3). 
Crotaphytus  collaris  and  Sauromalus 
obesus  occurred  on  sites  that  were 
open,  rocky,  and  shrubby  or  grassy 
(fig.  3). 

Snakes  showed  similar  distribu- 
tional patterns  to  lizards.  Some 
snakes,  such  as  Lampropeltis  getulus, 
Pituophis  melanoleucus,  Rhinocheilus 
leconti,  Crotalus  atrox,  and  Crotalus 
molossus,  occurred  in  many  habitat 


112 


Table  2.— Relative  abundance  of  lizards  by  habitat  type.  Relative  abundance   the  number  of  an  individual  species 
caught  in  an  array  per  24  h  period.  *  Indicates  species  verified  in  a  habitat  type  via  road-riding  and  searches.  The 
number  below  the  Habitat  Type  in  ( >  =  the  total  number  of  arrays.  The  number  In  ( )  to  the  right  of  the  species'  relative 
abundance  =  the  number  of  arrays  in  which  the  species  was  trapped. 


PP      PJ        SB  CC 

OC 

DG 

DD 

MB 

cw 

JM 

CA 

ME 

MR 

MD 

SD  CB 

(5)     (9)       (3)  (18) 

(13) 

(11) 

(3) 

(6) 

(13) 

(9) 

(3) 

(15) 

(16) 

(15) 

(22)  (22) 

Gerrhonotus  kingi 

-       -        -  .03(1) 

.03(1) 

Coleonyx  vahegafus 

03(1)  -  - 

.03(1) 

.03(1) 

.01(1) 

.05(2) 

.01(5) 

.03(8) 

.02(6) 

.04(11)  .06(11) 

Heloderma  suspectum 

♦ 

» 

.03(1) 

♦ 

♦ 

.03(1) 

.03(1) 

.03(1) 

.03(2) 

Callisaurus  draconoides 

-  -  -  .06(1)  -  -  -  -  .10(3)  .01(6)  05(2)  .03(2)  .05(7)  .08(4)  .06(10)  .04(6) 
Cophosaurus  fexona 

.07(1).09(5)     -     .10(5)  .03(1)      -        -     .08(2)   .10(4)      -        -     .01(1)   .03(2)      -     .02(2)  - 
Crofaphyfus  collaris 

-  *         *      .03(2)  •      .10(5)  .03(2)   .03(1)   .04(2)  - 
Dipsosaurus  dorsalis 

Gambelia  wislizenii 


-------     .01(1)      -     .03(1)  .08(9) 

-     .07(1)      -        ♦         _        _     .01(1)   .03(2)   .01(2)   .02(3)  .02(3) 


Holbrookia  maculafa 

-       -       -     .08(1)      -  ,03(1) 

Phrynosoma  douglassi 

.06(3).04(3)  .13(1)  .04(6)  _  _  _ 
Rirynosoma  plafyrhinos 


Phrynosoma  sol  are 
Sauromalus  obesus 


-        '  -  .07(1)      -        -  .01(1)   .11(3)      -     .03(1)   .02(3)   .02(3)  .05(7) 

----  ----  -     .03(1)      -     .03(2)  .02(1) 

-  -        -        -     .03(1)  _  _        _     .06(1)  .01(7)   .02(1)      -     .03(1)      *      .02(1)  - 
Sceloporus  clarki 

-  .03(1)      _        _        _  _  .      .03(2)   .03(2)  -        -        -     .03(1)   .03(1)   .03(1)  - 
Sceloporus  magister 

-  .05(5)      -     .05(7)   .03(3)  .03(2)  -     .11(4)   .23(7)  .03(8)   .19(3)  .13(10)  .1 1(16)  .10(15)  .07(14)  .03(6) 
Sceloporus  undulafus 

.13(3).  13(4)  .17(3)  .07(13)     -  .10(3)  -     .02(1)   .04(2)  -        -        -        -        -  - 

Urosaurus  graciosus 

----------  -     .07(7)  .07(11)  .01(3)   .04(2)  .07(13) 

Urosaurus  omafus 

.03(1).04(4)     -     .04(6)   .03(7)  .05(3)  -     .15(4)   .20(5)  .03(1)      -     .08(5)   .04(5)   .03(3)   .06(7)  ,02(3) 
Ufa  sfansburiana 

-  .03(2)      -     .04(4)   ,04(7)  .05(1)  .10(1)      -     .11(7)  .05(8)   .05(2)   .08(5)  .1 1(13)  .05(12)  .13(17)  .09(15) 
Eumeces  gilberfi 

.03(1).06(3)     -     .05(9)  .11(10)  .03(1)  -     .02(2)   .04(4)  .03(1)  ______ 

Eumeces  obsolefus 

Cnemidophorus  burfi 

Cnemidophorus  flagellicaudus 

-  .05(3)      -     .04(5)      -  .07(2)  -     .08(1)  .02(1) 
Cnemidophorus  inornatus 

-  .03(2)      -        *         -  .03(1) 

(continued) 


113 


c 

Table2.— (continued). 

PP      PJ       SB      CC      OC  DG 
(5)     (9)       (3)       (18)      (13)  (11) 

DD 
(3) 

MB      CW      JM  CA 
(6)      (13)      (9)  (3) 

ME 
(15) 

MR 
(16) 

MD 
(15) 

SD 
(22) 

CB 
(22) 

Cnemidophorus  figris 

nom  10  f/^^     —      07(6)    05(3)  09(4) 

23(2) 

10(3)    07(7)   .14(9)   .25(3)  . 

14(9) 

.25(16) 

.13(15) 

.17(21) 

.15(21) 

{^noniioopnoius  ui  ]ii~juit3i  lo 

_       _        _     .04(1)      -  .03(1) 

— 

— 

— 

— 

— 

— 

Cnemidophorus  velox 

-       *      .49(3)   .14(5)      -  .01(1) 

.05(2)   .02(2)      -  - 

— 

— 

— 

— 

Xanfusia  vigilis 

-     .01(1)  - 

.07(1) 

-    .02(1)      -     .05(1)    .08(7)  - 

— 

Total  Number  of  Species  (includes  species  verified  by  road-riding  and  searclnes) 
7        14        4        20        12        14        4        10        16        11        9  10 

16 

14 

17 

12 

Mean  Relative  Abundance 

.37     .69       .79       .96       .43  .74 

.47 

.67      1.06      .54  .72 

,59 

.91 

.58 

.78 

.63 

Species  Diversity  (H') 

.56    1.00      .40      1.18      .89  1.07 

.54 

.91      1.00      .76  .72 

.86 

1.05 

1.00 

1.09 

.95 

J 


15 
.75 


Component  I 


-.75 


Sknitn/ 
Grasses 


Component  III 


1  -  EwMcd  obsoWM 

2  -  Ctleonyi  foritfohis 

]  -  Mcdemn  ml^mian 

4  -  CdtMM  *ocanoid«l 

5  -  Cophosounn  kenun 

6  -  OuttphyfaiS  cdbnt 

7  -  D^jMnwva  towh 

I  -  GonMio  (WiOii 

$  -  Hotmcldo  mooioto 
10  -  Rnynosoni  dougkns 

I I  -  Rfjbosoto  plotjrtinos 
12  -  Rf)iwswno  vim 


13  -  Sann«lut  cbsw 

U  -  Scelopann  darti 
15  -  Scikpana  najim 
IS  -  Scilcpmf  uiduMif 

17  -  Umwvn  fTOCWSus 

18  -  UoNwvs  omotn 

19  -  Uto  ilmlMioM 

20  -  EiiffltcM  ffbvti 

21  -  DinMcts  obsolekis 

22  -  OMnUai^mi  barb' 

23  -  Ca«nidg|]hona  9aj/kKin 


24  -  Omidofiham  •wmtba 

25  -  CfwrMoftow  ttf* 

26  -  Cneiitfoftow  u^mn 

27  -  Cn«n4«|iharui  lim 
21  -  bnhso  liglil 


Opsn 


-.75 


-.50 


-J5 


J5 


JO 


Component  II 


Figure  3.— Mean  factor  scores  of  microhabitats  for  lizards. 


types.  Others,  such  as  Chilomeniscus 
cinctus,  Chiomctis  occipitalis,  Phyl- 
lorhynchus  browni,  Phyllorhynchus  de- 
curtatus,  and  Crotalus  cerastes,  oc- 
curred primarily  on  lower  elevation 
(<  915  m  or  3000  ft),  desert  habitats, 
and  some,  such  as  Lampropeltis  py- 
romelam  and  Crotalus  viridis  cerberus, 
occurred  only  on  higher  elevation 
(>1525  m  or  5000  ft)  habitats  (table  3). 
Lichanura  trivirgata  and  P.  browni  oc- 
cur primarily  outside  the  study  ar- 
eas, and  their  distributions  only 
overlap  the  extreme  southern  and 
southwestern  portions  of  the  study 
area.  Therefore,  they  were  limited  to 
the  small  number  of  sites  with  suit- 
able habitat.  Thamnophis  cyrtopsis  and 
Thamnophis  marcianus  were  restricted 
to  sites  with  water,  with  the  former 
occurring  on  a  large  number  of  habi- 
tats and  the  latter  only  in  a  mesquite 
bosque  habitat  along  the  Gila  River 
south  of  Phoenix.  Similar  to  Copho- 
saurus  texana,  Tantilla  hobartsmithii 
was  found  on  higher  elevation 
(>1220  m  or  4000  ft)  and  riparian 
habitats  throughout  most  of  the 
study  area,  but  also  in  Sonoran  Des- 
ert in  the  eastern  portion  of  the  study 
area. 

A  PC  A  of  microhabitats  on  which 
snakes  occurred  revealed  that,  simi- 


114 


Table  3.— Relative  abundance  of  snakes  by  habitat  type.  Relative  abundance  =  the  number  of  an  Individual  species 
caught  In  an  array  per  24  h  period.*  Indicates  species  verified  In  a  habitat  type  via  road-riding  and  searches.  The 
number  below  the  Habitat  Type  in  ( )  =  the  total  number  of  arrays.  The  number  in  ()  to  the  right  of  the  species'  relative 
abundance  =  the  number  of  arrays  in  which  the  species  was  trapped. 


Arizona  elegans 

  *    ♦ 

Chilomeniscus  cine f us 

Chionacfis  occipitalis 

Diadophis  puncfafus 

Hypsiglena  for  quota 

Lampropeltis  getulus 
  «         »  ♦ 

Lampropeltis  pyromelana 

♦       

Lichanura  trivirgata 


-  -  -  -  •  •  -  *  .02(1)  *  .03(2) 
-  -  -  -  .05(3)  -  -  .07(7)  .08(7)  -  .02(2)  .03(1) 
-------  .03(3)  .12(2)      -  .05(4)  .06(5) 

.02(2)  .02(1)  - 

-  -     .03(2)      *      .03(2)   .02(1)  .02(1)      *  ,02(1)      *  - 


Masficophis  bilineatus 

-  -        -        -  .03(2)   .03(1)      -        •         *         *         _        _        _        _        _  _ 

Masticophis  flagellum 

-  ,03(1)  -  *  *  -  -  -  .02(1)  *  .02(1)  .02(1)  .02(1)  ,02(1)  ,02(1)  .03(2) 
Masficophis  taeniatus 

,03(2)   .03(2)      -     ,02(1)      *         _        _        _        _        _  _ 

Pituophis  melanoleucus 


.02(1)      *      .02(1)      *      .02(1)      *         *         •  .01(1) 


Pliyllorliynclius  browni 
Phyilorhynclius  decurfafus 
Rhinoclieilus  leconfi 


-  —       -        -        -        -     ,02(2)      -        *  .03(2) 

-  .02(1)   .02(1)      *         *         *         *         *      .02(1)  .02(1) 


Salvadora  liexalepis 

-  -  -  .02(1)  .02(1)  *  -  -  .03(2)  •  .03(1)  .03(1)  .02(1)  .04(2)  .02(1) 
Sonora  semiannulafa 

-  -  -  *  *  .10(2)  -  -  *  .03(1)  *  ♦  _  *  .05(3) 
Tanfilla  hobarf smith ii 

-  -        -     .05(5)  .08(8)  ,07(2)      -     .02(1)   .05(4)   .03(1)      -     .03(2)      -  - 
Thamnophis  cyrtopsis 

*->»»•  »•♦♦♦♦»»♦ 

Thamnophis  marcianus 

              «   

Trimorphodon  biscutafus  lambda 
  _      »       

Crofalus  afrox 
Crofalus  cerastes 
Crofalus  mifchelli 


.02(1) 


.02(2) 


(continued) 


115 


> 

Table  3.— (continued). 

PP      PJ       SB  CC 

OC 

DG 

DD 

MB 

CW 

JM 

CA 

ME 

MR 

MD 

SD 

CB 

(5)     (9)       (3)  (18) 

(13) 

(11) 

(3) 

(6) 

(13) 

(9) 

(3) 

(15) 

(16) 

(15) 

(22) 

(22) 

Crofalus  molossus 

* 

« 

• 

• 

• 

• 

♦ 

♦ 

Crofalus  scututatus 

.02(1) 

« 

» ■■ 

♦ 

Crofalus  figris 

Crofalus  viridis  cerberus 
•        *  « 

Micruroides  euryxanfhus 

Leptotyphlops  humilis 
-    .04(3)      -  - 


.05(2)   .05(3)  .05(3) 


.03(2)      -        -        *  - 
.09(6)   .03(2)   .02(1)   .03(2)  .06(6) 


Total  Number  of  Species  (includes  species  verified  by  road-riding  and  searches) 

2 


4        11        6  17 

22 

12 

Mean  Relative  Abundance 

-      .07        -  .07 

.18 

.23 

Species  Diversity  (H') 
-      .30        -  .26 

.63 

.54 

12 

20 

18 

13 

17 

18 

16 

25 

16 

.18 

.24. 

16 

.07 

.36 

.28 

.12 

.22. 

.29 

.75 

.81 

.68 

.47 

.93 

.68 

,68 

.86 

.90 

15 
TrMt 

75  _ 


Component  I 


Crane* 


29  -  Arizona  ejegans .       4-6  • 

30  -  Chilomeniscus  cinctus  47  • 

31  -  Chionactis  occipitalis  4-8  • 

32  -  Diodophis  punclatus    4-9  • 

33  -  Hypsiglena  torquata    50  ■ 

34  -  Lompropeltis  qetuius  51  ■ 

35  -  Lompropeltis  pyromelaiM  • 

36  -  Lichanura  trivirqata    53  ■ 

37  -  Mosticophis  bilineatus  54  • 

38  -  Mosticophis  flagellum  55  • 

39  -  Masticophis  taeniatus  56  • 

40  -  Pituophis  melanoleucu57  ■ 

41  -  Phyllorhynchus  browni  58  • 

42  -  Phyllorhynchus  decurtotus 

43  -  Rhinocheilus  leconti 

44  -  Salvadoro  hexolepis 

45  -  Sonora  semiannulqja 


Tontilla  hobortsmithii 
Thomnophis  cyrtopsis 
Thomnobhis  mcrcianus 
Trimorpnodon  biscutctus  lambdc 
Crotolus  otrox 
Crotolus  cerastes 
Crotalus  mitchelli 
Crotolus  molossus 
Crotalus  scutulatus 
Crotalus  tigris 
Crotolu?  viridis  cerberus 
Micruroides  euryxonthus 
Leptotyphlops  humilis 


"  51  54 


Component 


50 


A2  M 


5:i 


51) 


57 


49 


58 


48 


46 


56 


Open 
Canopy 


Component 


Figure  4.— Mean  factor  scores  of  microhabitats  for  snakes. 


lar  to  those  of  lizards,  microhabitat 
associations  differed  among  snakes 
(fig.  4).  Many  of  the  widely  distrib- 
uted snakes,  such  as  Hypsiglena 
torquata,  Lampropeltis  getulus,  Mastico- 
phis  flagellum,  and  Pituophis  melano- 
leucus,  showed  no  strong  relationship 
with  any  of  the  compressed  habitat 
components  (fig.  4).  Conversely, 
most  species  with  limited  distribu- 
tions showed  a  strong  relationship 
with  certain  components  (fig.  4). 
Chionactis  occipitalis,  Crotalus  cerastes, 
Crotalus  scutulatus,  and  Phyllorhyn- 
chus browni  consistently  occurred  on 
open,  sandy  sites,  and  Chilomeniscus 
cinctus  occurred  on  sites  with  sandy 
substrate  but  taller  canopy  (fig.  4). 
Other  species,  such  as  Crotalus  mitch- 
elli and  Sonora  semiannulata,  were 
found  on  sites  with  open  canopies 
but  rocky  substrates  (fig.  4).  Thamno- 
phis  marcianus  and  Tantilla 
hobartsmithii  occurred  on  sites  with 
sandy  substrates  but  closed  canopies 
and  large  amounts  of  vegetative  de- 
bris, and  Lampropeltis  pyromelana  oc- 
curred only  on  sites  with  high 
amounts  of  vegetative  debris  (fig.  4). 
Other  species,  such  as  Diadophis 


116 


pundatus,  Thamnophis  a/rtopsis,  and 
Crotalus  viridis  cerherus,  occurred  on 
rocky  sites  with  high  amounts  of 
vegetative  debris  (fig.  4). 

Except  for  a  single  Gopherus  agas- 
sizii  captured  in  an  array,  all  turtle 
records  came  from  road-riding  and 
field  searches.  Four  species  of  turtles 
were  recorded  within  the  study  area, 
three  aquatic  and  one  terrestrial 
(table  4).  Of  these,  G.  agassizii  was  the 
most  widely  distributed  (verified  in  9 
habitat  types,  table  4).  A  more  thor- 
ough account  of  this  turtle's  distribu- 
tion is  described  by  Burge  (1979, 
1980).  Pseudemys  scripta,  an  intro- 
duced species,  was  limited  to  a 
stretch  of  the  Gila  River  from  the 
99th  Street  bridge  in  southwest  Phoe- 
nix to  Gillespie  Dam,  located  ap- 
proximately 24  km  (15  miles)  south 
of  Buckeye.  Trionyx  spiniferus  oc- 
curred at  Alamo  Lake  (confluence  of 
the  Big  Sandy  and  Santa  Maria  rivers 
in  western  Arizona)  and  along  peren- 
nial stretches  of  the  Gila  River  from 
Phoenix  to  Yuma.  Kinosternon  sonori- 
ense  occurred  on  several  permanent 
streams  and  rivers  throughout  the 
study  area. 

In  contrast  to  the  observed  distri- 
bution patterns  among  lizards  and 
snakes,  the  distribution  of  amphibi- 
ans did  not  shown  an  elevational  pat- 
tern. Although  certain  species  such 


as  Bufo  punctatus  and  Scaphiopus 
couchi  occurred  in  a  large  number  of 
habitat  types,  most  species  were 
found  in  at  least  one  lower  (<  915  m 
or  3000  ft)  and  one  higher  (>  1220  m 
or  4000  ft)  elevation  site  (table  5). 
Similar  to  lizards  and  snakes,  there 
are  some  amphibians  whose  ranges 
are  principally  outside  the  study  area 
and  are,  therefore,  found  only  on  a 
few  sites  (table  5).  The  ranges  of  Bufo 
debilis,  Bufo  retiformes,  and  Gastro- 
phyrne  olivacea  are  primarily  in  north- 
ern Mexico,  or  east  and  south  of  the 
study  area  in  the  Chihuahuan  Desert; 
within  the  study  areas,  their  ranges 
are  limited  to  desert  grassland  habi- 
tats in  the  extreme  southern  portion 
(Vekol  Valley,  48  km  or  30  mi  west- 
southwest  of  Casa  Grande).  All 
populations  of  Ambystoma  tigrinum 
were  located  at  earthen  stock  tanks 
(dirt  tanks).  Presumably,  all  of  these 
populations  were  introduced. 

A  PCA  demonstrated  correlations 
between  occurrence  of  amphibian 
species  and  particular  microhabitats 
(fig.  5).  Bufo  debilis,  B.  retiformes,  and 
Gastrophyrne  olivacea  occurred  on 
sandy,  grassy  sites,  and  Bufo  cognatus 
on  sandy,  shrubby  sites  (dg.  5).  Bufo 
microscaphus  and  B.  punctatus  oc- 
curred on  rocky  sites,  and  Hyla  areni- 
color  on  rocky  sites  generally  occu- 
pied by  trees  and  large  amounts  of 


vegetation  debris  (fig.  5).  Certain 
species,  such  as  Scaphiopus  couchi, 
Bufo  alvarius,  and  Bufo  woodhousei  oc- 
curred on  sites  with  a  wide  variety  of 
substrates  (fig.  5). 

The  occurrence  and  frequency  of 
water  was  not  quantitatively  meas- 
ured at  each  site;  therefore,  the  influ- 
ence of  water  was  not  considered  in 
the  development  of  figure  5.  How- 
ever, all  sites  with  amphibians  had 
surface  water  during  some  part  of 
the  year,  especially  during  summer 
months.  AH  sites  with  Bufo  mi- 
croscaphus, Rana  pipiens,  R.  catesbe- 
iana,  and  Hyla  arenicolor  had  perma- 
nent water  (e.g.,  springs,  creeks,  and 
rivers). 

At  the  start  of  the  survey  in  1977, 
populations  of  Bufo  microscaphus  and 
B.  woodhousei  sympatric  on  major 
drainages,  such  as  the  Hassayampa, 
Santa  Maria,  Agua  Fria,  and  New 
rivers,  could  be  easily  distinguished 
from  one  another.  By  1981,  popula- 
tions on  all  of  these  drainages  were 
indistinguishable. 


Range  Extensions 

Thirty-five  range  extensions  were 
recorded  for  amphibians  and  reptiles 
within  the  study  area.  Except  for  the 
following  discussion,  range  exten- 


Table  4.— Distribution  of  turtles  by  habitat  type  C).  Records  are  entirely  from  road-riding  and  searches  (except  where 
otherwise  Indicated.  All  turtles  except  Gopherus  agassizii  occurred  only  at  sites  with  permanent  water  within  habitat 
types  listed  below. 


PP  PJ 


SB 


CC  OC 


DG 


DD 


MB 


CW  JM 


CA 


ME 


MR  MD 


SD 


CB 


Gopherus  agassizii 

    « 

Pseudemys  scripta 
Trionyx  spiniferus 
Kinosternon  sonoriense 


"  Trapped  in  an  array 
Number  of  species 


117 


— -  ..  ,  ,, 

Table  5  -Relative  abundance  of  amphibians  by  habitat  type.Relative  abundance  =  the  number  of  an  individual  spe- 
cies caught  in  an  array  per  24  h  period. '  indicates  species  verified  in  a  habitat  type  via  road-riding  and  searches. 
The  number  below  the  Habitat  Type  in  ( )  =  the  total  number  of  arrays.  The  number  in  ( )  to  the  right  of  the  species 
relative  abundance  =  the  number  of  arrays  in  which  the  species  was  trapped. 


PP 
(5) 


PJ 
(9) 


SB 
(3) 


CC 
(18) 


OC 
(13) 


DG 
(11) 


DD 
(3) 


(6) 


CW 
(13) 


JM 
(9) 


CA 
(3) 


ME 
(15) 


MR 
(16) 


MD 
(15) 


SD 
(22) 


CB 
(22) 


Bufo  olvorius 


-  .07(1) 


Bufo  cognatus 

*         —  — 

Bufo  debilis 

Bufo  mlcroscaphus' 

-  .05(2)  -• 
Bufo  puncfafus 

.03(1).03(1)  *  .15(8)  .11(8) 
Bufo  refiformis 

Bufo  woodhouseP 

Hyla  orenicolor 

-  .07(1)  .03(1) 
Gastrophyrne  olivacea 


.06(2) 
.14(3) 
.03(1) 


.03(2) 


.03(4) 

.09(3)  - 


.06(4) 


18(2) 


.13(3)  .06(3)  V  -  -07(4)  _  -  -  - 
.20(1)   .16(3)   .12(2)   .23(1)   .28(6)   .05(2)   .06(2)   .10(7)  - 

_        •         _        -        -     .03(1)  - 


Scaphiopus  couchi 

Rana  pipiens 
•        ♦  ♦ 

Rona  cafesbeiana 
Ambysfoma  figrinum 


.12(2) 


.05(2) 
.10(3) 


•  .20(1) 


.15(6)   .06(2)   .07(6)   .06(5)  .11(6) 


.03(1) 


'95%  of  these  were  a  cross  between  the  two  species  (B,  microscaphus  x  B.  woodhousei) 
Total  Number  of  Species  (includes  species  verified  by  road-riding  and  searches) 
4        4         3         6         6         8         2         4         6         5  5 
Mean  Relative  Abundance 

.03     .03        -       .34      .26      .56       -       .33       .45      .12  .23 
Species  Diversity  (H') 

_       -        -       .56       .42       .71        -       .29       .51  - 


8 

6 

5 

6 

3 

.65 

.14 

.13 

.16 

.17 

.65 

.46 

.30 

.29 

.28 

sions  discovered  during  this  study 
have  been  described  elsewhere  (Jones 
et  al.  1981,  Jones  et  al.  1982,  Buse 
1983,  Jones  et  al.  1983,  Jones  et  al. 
1985).  The  southernmost  distribution 
of  Tantilla  hobartsmithii  was  extended 
from  the  Salt  River  east  of  Phoenix, 
southwest  in  the  mesquite  bosque 
habitat  along  the  Gila  River  to  56  km 
(35  miles)  east-northeast  of  Yuma 
(fig.  6).  A  population  of  T.  hobart- 
smithii was  also  discovered  in  a  10  ha 


(25  acres)  open  chaparral  habitat  in 
the  Eagletail  Mountains  (fig.  6).  The 
westernmost  distribution  of  Cnemido- 
phorus  burti  was  extended  from  the 
Tucson  area  northwest  by  discovery 
of  isolated  populations  in  desert 
grassland  habitats  on  summits  of  the 
Tabletop  and  Estrella  mountains  (fig. 
6). 

An  isolated  population  of  Mastico- 
phis  bilineatus  lineolatus  was  discov- 
ered on  the  summit  of  Tabletop 


Mountain  in  a  relict  desert  grassland 
habitat  (fig.  6).  This  population  ex- 
tends the  known  distribution  of  this 
subspecies  approximately  100  km  (62 
mi)  to  the  north  of  the  only  other 
known  population  (Ajo  Mountains). 

Finally,  an  isolated  population  of 
Diadophis  punctatus  was  discovered 
in  a  relict  desert  grassland  commu- 
nity on  the  summit  of  the  Estrella 
Mountains  southwest  of  Phoenix  (fig. 
6). 


118 


Comparison  of  Habitat  Types 

Based  on  data  compiled  from  pit-fall 
trapping,  road-riding,  and  searches, 
the  Sonoran  Desert  habitat  had  the 
greatest  species  richness  (49  species, 
fig.  7).  Closed  chaparral  and  cotton- 
wood-willow  riparian  habitats  were 
the  second  richest  habitats  (44  spe- 
cies), and  open  chaparral  and  mixed 
riparian  scrub  were  third  (41  species, 
fig.  7). 

Disclimax  desert  grassland  had 
the  fewest  species  (8),  and  sagebrush 
and  ponderosa  pine  had  the  second 
and  third  fewest  species  (13  and  15 
species,  respectively,  fig.  7).  All  other 
habitats  had  at  least  27  species  but 
not  more  than  39  (fig.  7).  Although 
Sonoran  Desert  had  the  richest  lizard 
and  snake  faunas,  mesquite  bosque 
and  desert  grassland  habitats  had  the 
richest  amphibian  fauna  (fig.  7).  The 
mesquite  bosque  habitat  type  had  the 


greatest  number  of  turtle  species 
(four  species,  fig.  7). 

When  only  array  data  are  com- 
piled, disclimax  desert  grassland, 
sagebrush,  and  ponderosa  pine  habi- 
tats still  had  by  far  the  lowest  num- 
ber of  species,  but  Sonoran  Desert 
and  mesquite  bosque  had  the  great- 
est number  of  species  (fig.  8).  As 
when  all  data  were  taken  into  ac- 
count, mixed  riparian  scrub,  cotton- 
wood-willow  riparian,  closed  chap- 
arral, and  open  chaparral  had  high 
species  richness  (fig.  8).  However, 
desert  grassland  was  relatively  more 
diverse  using  only  array  data  (fig.  8). 

The  difference  between  array  vs. 
all  data  appears  to  result  from  the 
inability  of  arrays  to  consistently  ver- 
ify (trap)  turtles  and  medium  and 
large  snakes,  although  many  larger 
snake  species  were  verified  because 
young-of-the-year  were  easily 
trapped. 


A  more  revealing  statistic  is  the 
average  number  of  species  verified 
by  an  array  (fig.  8).  This  analysis  re- 
veals which  habitats  consistently  had 
the  largest  number  of  species  at 
sample  sites.  Certain  habitats,  such 
as  desert  grassland,  although  high  in 
overall  species  richness,  had  rela- 
tively few  species  verified  at  each 
array  site  (fig.  8).  Other  habitats, 
such  as  ponderosa  pine,  sagebrush, 
and  disclimax  desert  grassland,  had 
the  lowest  number  of  total  species 
and  the  lowest  average  number  of 
species  per  array  site  (fig.  8).  Many  of 
the  habitats  that  had  high  overall 
species  richness  also  had  high  overall 
richness  at  each  array  site;  however, 
cotton  wood-willow  had  a  higher  av- 
erage number  of  species  per  array 
site  than  did  Sonoran  Desert  (fig.  8). 

Species  diversity  indices  (H')  cal- 
culated from  array  data  reveal  pat- 
terns similar  to  those  described 
above  (fig.  9).  Disclimax  desert  grass- 
land, sagebrush,  and  ponderosa  pine 
continue  to  exhibit  low  diversity,  and 
Sonoran  Desert,  closed  chaparral, 
cotton  wood-willow  riparian,  mixed 
riparian  scrub,  and  desert  grassland 
continue  to  be  diverse  (fig.  9).  How- 
ever, as  in  the  previous  analysis,  the 
average  diversity  per  array  site  is 
low  when  compared  to  total  diver- 
sity for  individual  habitats  (fig.  9).  Of 
the  habitats  with  high  overall  diver- 
sity, mixed  broadleaf  riparian  and 
cottonwood-willow  riparian  had 
relatively  high  average  diversity  per 
array  site  (fig.  9). 

A  comparison  of  herpetofaunas  of 
each  habitat  type  by  cluster  analyses 
revealed  that  all  desert  habitats,  such 
as  creosotebush,  Sonoran  Desert, 
Mohave  Desert,  and  mixed  riparian 
scrub  had  very  similar  herpetofaunas 
(figs.  10  and  11).  In  both  cluster 
analyses,  open  and  closed  chaparral 
had  similar  herpetofaunas,  and  sage- 
brush and  disclimax  desert  grassland 
had  a  herpetofauna  different  from 
any  other  habitat.  However,  there 
were  differences  in  results  of  the  two 
cluster  analyses  for  other  habitats. 
Whereas  the  cluster  analysis  of  array 


1.5 
Tro« 

75 


Component  I 


-.75 


Grauei 


Component 


59  -  aufo  olvoriui 

60  -  Bufo  cognotm 

61  -  Biifo  deUD] 

62  -  SjTo  mlcroicaphui 

63  -  auto  punctatui 
-  Buto  rotltormS 

66  -  Buf  0  woodhomel 

66  -  Hyla  aronlcolof 

67  -  Gaitrophyme  ollvocea 

68  -  ScopmopLa  coucni 

69-  nana  plplem 

70-  nono  coteiDelana 
71  -  AmbyitDtno  ttgrtnum 


iJ  71 


Open 
Canopy 


0  .25 


60 


VeQerotlva 


Component  I 


Figure  5.— Mean  factor  scores  of  microhabitats  for  amphibians. 


119 


data  revealed  large  differences  be- 
tween the  herpetofaunas  of  cotton- 
wood-willow  and  desert  habitats, 
such  as  Sonoran  and  Mohave  Des- 
erts, these  habitats  had  a  relatively 
moderate  degree  of  overlap  when  all 
data  were  analyzed  (figs.  10  and  11). 
Additionally,  ponderosa  pine  and 
pinyon-juniper  habitats  were  similar 
when  array  data  were  analyzed  and 
relatively  dissimilar  when  all  data 
were  submitted  to  cluster  analysis 
(figs.  10  and  11). 


DISCUSSION 

Overall,  western  Arizona  has  an  ex- 
tremely diverse  herpetofauna,  pri- 
marily because  of  its  large  variety  of 
habitats  zoogeographic  location.  The 
Hualapai  Mountains,  located  in 
northwestern  Arizona,  are  adjacent 
to  three  major  deserts:  the  Mohave 
Desert  to  the  northwest,  the  Great 
Basin  Desert  to  the  northeast,  and  the 
Sonoran  Desert  to  the  south.  No- 
where else  on  the  North  American 
continent  does  such  a  phenomenon 
exist.  The  diversity  of  habitat  in  this 
area  is  also  enhanced  by  the  occur- 
rence of  several  woodland  islands. 


Number  of  Species 


50 


40 


30  _ 


20  - 


10  _ 


Tuttes 


Amphibians 


Snakes 


Uzofds 


PP    PJ    SB  CC  OC  DG  DO  MB  CW  JM  CA  rvE  fv«  MD  SO  CB 


Habitat  Type 


Figure  7.— Number  of  species  by  taxonomic  group  by  habitat  type.  (Abbrev.  correspond  to 
ttiose  listed  for  liabitats  In  table  1 .) 


Number  of  Species 


30 


25  — 


20 


15  - 


10  — 


6  — 


#Of 

Species 

Ave  #  of 
species/ 
array 


PP    PJ     SB    CC  OC  DG   DD    MB  CW   JM   CA    ME    MR   MD    SD  CB 


Figure  6.— Map  of  range  extensions. 


Habitat  Type 


Figure  8.— Total  number  of  species  caughit  in  arrays  by  habitat  type  vs.  the  average  number 
of  species  caught  per  array  by  habitat  type.  (Abbrev.  correspond  to  those  listed  for  habitats 
In  table  1 .) 


120 


Species  Diversity  (H') 


Patterns  of  Species  Distributions 


Habitat  Type 


Figure  9.— Total  species  diversity  (H')  by  hiabitat  type  vs.  average  species  per  array  by  tiabl- 
tat  type.  (Abbrev.  correspond  to  thiose  listed  for  htabitats  in  table  1 .) 


Similarity 


MR  SD  MD  CA  OC  JM  CW  CC  ME  CB  DG  PJ  MB  PP  SB  DD 


1.0 


Figure  10.— Cluster  analysis  (dendrogram)  of  array  data  Illustrating  similarities  in  tiabitat  type 
tierpetofaunas.  (Abbrev.  correspond  to  thiose  listed  for  tiabitats  in  table  1 .) 


This  survey  reveals  that  certain  spe- 
cies are  widespread,  occurring  in 
several  habitats,  but  many  species 
are  limited  to  specific  habitat  types. 
Also,  some  species  occur  on  most 
sample  sites  within  a  habitat  type 
and  others  on  only  a  few.  There  ap- 
pear to  be  at  least  3  major  factors 
contributing  to  distributional  pat- 
terns of  amphibians  and  reptiles  in 
the  study  area. 

Geographic  Limitations 

The  ranges  of  certain  species  only 
peripherally  occur  in  western 
Arizona.  Cnemidophorus  burti,  Phyl- 
lorhynchus  browni,  Masticophis  bilinea- 
tus  lineolatus,  and  Bufo  retiformis  oc- 
cur principally  in  northern  Mexico 
whereas  others  such  as  Holbrookia 
maculata,  Eumeces  obsoletus,  Gastro- 
phyrne  olivacea,  and  Bufo  debilis  are 
mostly  east  and  north  of  the  study 
area  (Stebbins  1985).  Bufo  retiformis, 
Gastrophyrne  olivacea,  and  Bufo  debilis 
are  associated  with  low  elevation 
(457-915  m  or  1500-3000  ft)  desert 
grassland  (Jones  et  al.  1983),  and 
these  habitats  are  mostly  absent  in 
the  central  and  northern  portions  of 
the  study  area.  However,  habitat 
suitable  for  other  species  listed  above 
appears  to  be  available  throughout 
most  of  the  study  area. 

Physical  barriers,  such  as  topogra- 
phy, elevation,  and  climate  may  have 
presented  these  species  from  coloniz- 
ing or  immigrating  into  suitable  habi- 
tats to  the  north  and  west  (see  Con- 
nor and  Simberloff  1979,  Case  1983, 
Jones  et  al.  1985  for  discussion  of  the 
influence  of  physical  barriers  on  colo- 
nization/immigration). In  addition, 
competition  between  species  may 
have  limited  individual  species' 
ranges  during  initial  and  subsequent 
colonization  of  suitable  habitats  (e.g., 
during  periods  of  large  climatic 
changes).  Perhaps  the  best  example 
of  this  is  the  distributional  relation- 
ship between  Eumeces  gilberti  and  E. 


121 


ohsoletus.  E.  gilberti  belongs  to  the 
skiltonianus  group  of  skinks,  whose 
evolutionary  center  is  the  western 
United  States  (Taylor  1935,  Rogers 
and  Fitch  1947). 

Conversely,  E.  obsoletus  evolved  in 
the  Great  Plains  region  (Fitch  1955). 
Both  of  these  lizards  occupy  seem- 
ingly identical,  but  separate,  habitats 
in  central  Arizona,  and  their  distribu- 
tions come  together  in  chaparral  and 
desert  grassland  habitat  types  near 
Cordes  Junction;  the  westernmost 
range  of  E.  obsoletus  is  just  east  of 
Interstate  Highway  17  and  the  east- 
ernmost range  of  E.  gilberti  is  just 
west  of  the  highway.  These  lizards 
are  similar  in  appearance,  with  E.  ob- 
soletus averaging  slightly  larger  in 
size. 

Although  subtle  differences  in  mi- 
crohabitat  cannot  be  ruled  out  as  fac- 
tors influencing  their  ranges,  it  ap- 

Similarity 


pears  that  these  lizards  are  mutual 
exclusive  (competitive  exclusion). 

Several  remnant  stands  of  chapar- 
ral and  desert  grassland  occur  in 
western  and  northwestern  Arizona  at 
or  near  the  summits  of  mountain 
ranges.  These  relict  stands  or  habitat 
islands  are  isolated  within  creo- 
sotebush  and  Sonoran  Desert  habi- 
tats as  a  result  of  the  retreat  of  the 
last  Ice  Age  (see  Van  Devender  and 
Spaulding  1977).  Data  collected  in 
my  study  show  that  several  reptiles 
typically  found  in  "upland"  habitats 
(e.g.,  large  continuous  stands  of  des- 
ert grassland  and  woodlands  associ- 
ated with  the  Colorado  Plateau  of 
central  and  northern  Arizona)  inhabit 
these  isolated  mountain  stands,  al- 
though the  number  and  composition 
of  these  upland  species  vary  among 
mountains.  Habitat  island  size  ap- 
pears to  be  of  primary  importance  in 


determining  the  number  of  upland 
present  species  (see  Jones  et  al.  1985). 

The  turtles  Pseudemys  scripta  and 
Trionyx  spiniferus  are  present  along 
the  Gila  River  as  a  result  of 
introductions.  P.  scripta  is  a  popular 
pet,  and  specimens  have  been  re- 
leased along  the  Gila  River  in  south- 
west Phoenix.  T.  spiniferus  was  intro- 
duced along  the  Colorado  River  in 
the  early  1900's  (Stebbins  1985);  pre- 
sumably, these  populations  ex- 
panded into  the  Gila  River  at  the 
confluence  of  the  Gila  and  Colorado 
rivers  near  Yuma. 


Microhabitats  and  Physical 
Characteristics  of  Habitat 

Many  studies  have  shown  a  strong 
relationship  between  the  distribution 
and  abundance  of  amphibians  and 
reptiles  and  the  presence  and  amount 
of  certain  microhabitats  (Norris  1953, 
Pianka  1966,  Zweifel  and  Lowe  1966, 
Fleharty  1967,  Pianka  and  Parker 
1972).  The  distribution  of  a  number 
of  species  within  western  Arizona 
area  appears  to  be  influenced  by  the 
presence  of  microhabitats  on  sites, 
although  most  of  the  widespread 
species,  such  as  Cnemidophorus  tigris, 
Pituophis  melanoleucus,  and  Lam- 
propeltis  getulus  show  no  strong  rela- 
tionship with  any  specific  habitat 
components,  others  (e.g.,  Urosaurus 
ornatus  and  Sceloporus  magister)  occur 
on  sites  with  trees  and  downed  litter. 
Many  sites  in  the  study  area,  includ- 
ing desert  and  upland  habitat  types, 
have  trees  and  downed  logs,  and  this 
probably  accounts  for  these  species' 
wide  distributions.  The  habitat  analy- 
sis revealed  that  several  species  are 
associated  with  specific  substrate 
types  (e.g.,  rock),  density  or  height  of 
the  vegetation  canopy,  type  of  vege- 
tation (shrubs  or  grasses  vs.  trees),  or 
presence  of  downed  litter. 

Species'  associations  with  certain 
microhabitats  may  reflect  their  physi- 
cal or  behavioral  limitations.  For 
example,  Eumeces  gilberti  may  be  re- 
stricted to  sites  with  large  amounts 


MR  SD  MD  CA  OC  JM  CW  CC  ME  CB  DG  PJ  MB  PP  SB  DD 


1.0 


Figure  1 1  .—Cluster  analysis  (dendrogram)  of  all  data  illustrating  similarities  in  tiabitat  type 
hterpetofaunos.  (Abbrev.  correspond  to  thiose  listed  for  hiabitats  in  table  1 .) 


122 


of  downed  litter  (primarily  leaves 
and  logs)  because  of  its  low  preferred 
body  temperature  and  feeding  habits 
(Jones  1981b,  Jones  and  Glinski  1985). 
Large  amounts  of  surface  litter  on 
certain  riparian  sites  may  explain  the 
occurrence  of  this  lizard  in  cotton- 
wood-willow  riparian  sites  within 
desert  regions  (down  to  549  m  or 
1800  ft)  (see  Jones  and  Glinski  1985). 
Several  other  species  typically  found 
on  upland  habitats  (e.g.,  chaparral), 
such  as  Tantilla  hobartsmithii,  Copho- 
saurus  texana,  Masticophis  bilineatus, 
and  Diadophis  punctatus,  also  may 
persist  on  riparian  habitats  within 
deserts  because  of  the  high  moisture 
regime  associated  with  surface  litter, 
higher  humidity,  and  surface  water 
(Jones  and  Glinski  1985). 

A  similar  relationship  appears  to 
exist  in  desert  habitats  occupied  by 
Xantusia  vigilis.  This  lizard  also  has  a 
low  preferred  body  temperature,  and 
it  only  occurs  on  Mojave  Desert  sites 
occupied  by  agaves  (Agave  spp.)  and 
yuccas  (Yucca  spp.  and  Nolim  spp.); 
these  plants  create  cool,  moist  mi- 
crohabitats  within  desert  habitats.  In 
the  southern  part  of  its  range,  X.  vig- 
ilis only  occupies  Sonoran  Desert  on 
steep  slopes  in  mountain  canyons,  or 
on  top  of  mountains  (>  1220  m  or 
4000  ft)  in  chaparral  habitats.  This 
shift  in  habitat  association  may  re- 
flect increased  average  temperature 
and  aridity  associated  with  decreas- 
ing latitude;  canyons  and  mountain 
summits  may  be  the  only  sites  mod- 
erate enough  to  support  this  lizard. 

A  similar  moisture  or  temperature 
relationship  may  also  account  for  dif- 
ferences observed  in  habitat  type  as- 
sociations of  Tantilla  hobartsmithii, 
Cophosaurus  texana,  and  Diadophis 
punctatus  in  the  eastern  and  western 
portions  of  their  ranges.  In  the  west- 
ern portion  of  the  study  area,  these 
reptiles  occur  only  in  chaparral  or 
riparian  habitat  types  (excluding 
mixed  riparian  scrub  habitats).  In  the 
eastern  and  southeastern  portions  of 
the  study  area,  these  species  also  oc- 
cur in  the  Sonoran  Desert  habitat 
type.  Eastern  and  southeastern  Sono- 


ran Desert  habitats  within  the  study 
area  are  more  extensive  than  those  to 
the  west  and  northwest,  and  they  are 
not  interrupted  by  large  creo- 
sotebush  habitats;  western  and 
northwestern  sites  are  restricted 
mostly  to  mountain  slopes,  separated 
by  extensive  creosotebush  flats.  In 
addition,  eastern  and  southeastern 
sites  appear  to  have  more  springs 
and  perennial  creeks  than  western 
and  northwestern  sites,  and  this  ad- 
ditional moisture  might  contribute  to 
the  presence  of  these  species  on  these 
sites. 

The  presence  of  surface  water  also 
has  a  profound  affect  on  the  distribu- 
tion and  abundance  of  certain  species 
within  the  study  area.  Kinosternon 
sonoriense,  Trionyx  spiniferus,  Thamno- 
phis  cyrtopsis,  Bufo  alvarius,  Bufo  mi- 
croscaphus,  Bufo  woodhousei,  Rana  pipi- 
ens,  Rana  catesbeiana,  Hyla  arenicolor, 
and  Ambystoma  tigrinum  occur  only 
on  sites  with  permanent  water 
(springs,  creeks,  rivers,  dirt  tanks). 
All  of  these  species  are  restricted  to 
permanently  watered  sites  because  of 
a  combination  of  physiological 
(Walker  and  Whitford  1970),  mor- 
phological (Mayhew  1968),  reproduc- 
tive (Justus  et  al.  1977),  or  behavioral 
(Hulse  1974)  limitations.  In  addition 
to  occurring  near  permanent  water, 
Bufo  punctatus  also  occurs  in  rock- 
bound  canyons  with  intermittent  wa- 
ter, and  Bufo  cognatus,  B.  debilis,  B. 
retiformis,  and  Gastrophyrne  olivacea 
occur  on  sites  with  clay  and  clay- 
loam  soils  that  accumulate  surface 
water  during  summer  convectional 
rainstorms.  All  of  these  species  pos- 
sess adaptations,  such  as  a  rapidly 
developing  embryo,  that  are  condu- 
cive to  survival  in  areas  with  inter- 
mittent surface  water  (Creusere  and 
Whitford  1976). 

A  number  of  species  were  verified 
on  fewer  than  half  of  the  array  sites 
within  habitat  types.  These  low  per- 
centages may  reflect  species'  associa- 
tion with  specific  microhabitats  and 
the  abundance  and  distribution  of 
microhabitats  within  habitat  types. 
For  example,  Chilomeniscus  cinctus 


occurred  on  less  than  half  of  the  cot- 
tonwood-willow  and  mixed  riparian 
scrub  array  sites.  The  habitat  analysis 
shows  that  this  species  is  associated 
with  sandy  and  fine  gravel  soils,  but 
many  of  the  cotton  wood-willow  ri- 
parian and  mixed  riparian  scrub 
sample  sites  have  rocky  substrates. 
Therefore,  the  substrate  type  limits 
this  species'  range  within  these  habi- 
tat types. 

However,  there  were  other  spe- 
cies, especially  snakes  in  excess  of  0.5 
m  (1.5  ft),  that  were  not  readily 
caught  in  pit-fall  traps,  although  a 
small  percentage  of  arrays  captured  a 
few  large  snakes;  these  snakes  were 
feeding  on  small  rodents  at  the  bot- 
tom of  traps.  Therefore,  the  paucity 
of  large  snakes  on  samples  sites 
within  habitats  probably  reflects  the 
ability  of  larger  snakes  to  escape 
from  pit-fall  traps  rather  than  the  dis- 
tribution and  abundance  of  mi- 
crohabitats within  habitat  types.  Ad- 
ditionally, amphibians  and  reptiles 
with  restricted  activity  patterns  (e.g., 
toads)  or  home  ranges  (Xantusia  vig- 
ilis) also  were  rarely  trapped  and, 
therefore,  verified  on  few  sites  within 
a  habitat.  The  limited  number  of 
mixed  broadleaf  and  chaparral  array 
sites  with  Gerrhonotus  kingi  probably 
reflect  a  low  sampling  effort  in  these 
habitats  during  the  fall;  this  lizard's 
peak  activity  is  during  its  breeding 
season  in  the  fall  (Robert  Bowker 
personal  comm.). 

Habitat  Conditions 

The  condition  of  habitats  may  play 
an  important  role  in  determining  the 
distribution  and  abundance  of  am- 
phibians and  reptiles.  In  Arizona,  the 
large  variety  of  land  uses  within  the 
area  may  affects  the  distribution  and 
abundance  of  certain  microhabitats 
and  may  account  for  variation  in  spe- 
cies composition  within  habitats.  A 
number  of  studies  have  shown  the 
effects  of  land  uses  on  amphibians 
and  reptiles  and  their  habitats.  These 
include  grazing  (Bury  and  Busack 


123 


1974,  Jones  1981a,  Szaro  et  al  1985), 
off-road  vehicle  use  (Bury  et  al.  1977, 
Bury  1980),  forest  management  (Ben- 
nett et  al.  1980),  and  stream  modifi- 
cation resulting  from  water  im- 
poundments (Jones,  this  volume). 
Generally,  these  affect  habitat  struc- 
ture. For  example,  excessive,  long- 
term  livestock  grazing  reduces  the 
abundance  and  diversity  of  forbs  and 
perennial  grasses.  Many  former  des- 
ert grassland  habitats  are  now  domi- 
nated by  shrubs  such  as  creosotebush 
(Larrea  tridentata)  and  mesquite 
(Prosopis  glandulosa)  (York  and  Dick- 
Peddie  1969).  Jones  (1981a)  showed 
large  differences  in  the  presence  and 
abundance  of  certain  lizards  on  heav- 
ily vs.  lightly  grazed  sites,  especially 
on  riparian,  desert  grassland,  and 
woodland  habitats,  attributable  to 
differences  in  lizard  ecology  and  dif- 
ferences in  habitat  structure  between 
heavily  vs.  lightly  grazed  areas.  Cer- 
tain lizards,  such  as  Cnemidophorus 
tigris,  prefer  open,  shrubby  sites; 
these  lizards  are  more  abundant  on 
heavily  grazed  sites  where  shrubs 
have  replaced  grasses  and  forbs 
(Jones  1981a).  Conversely,  certain 
lizards,  such  as  Eumeces  gilberti,  pre- 
fer grassy,  moist  sites,  and  are,  there- 
fore, less  abundant  on  or  absent  from 
sites  where  grazing  has  reduced  tree 
reproduction  (e.g.,  cottonwoods, 
Populus  fremontii  on  riparian  sites)  or 
suppressed  grasses  (e.g.,  on  desert 
grassland  sites)  (Jones  1981a). 

The  reduction  of  naturally-occur- 
ring water  and  the  modification  of 
river  and  stream  habitats  has  been 
shown  to  affect  the  composition  of 
amphibians  and  reptiles  within  habi- 
tats, especially  riparian  sites  (Jones 
1988).  Platz  (1984)  attributes  the  ex- 
tinction of  Ram  onca  to  modification 
of  stream  habitats  along  the  Virgin 
River.  Species  that  prefer  lentic  or 
pool  habitats  should  increase  on  sites 
with  water  impoundments,  whereas 
species  that  prefer  lotic  or  running 
water  should  decrease. 

Natural  phenomena,  such  as  fire, 
also  affect  species  composition 
within  habitats  (Kahn  1960,  Simovich 


1979).  Simovich  (1979)  showed  that 
fire  set  back  succession  within  chap- 
arral habitats  (grass/forb  succes- 
sional  stage),  and  that  these  changes 
resulted  in  increases  in  certain  spe- 
cies and  decreases  in  others.  As  suc- 
cession proceeded  to  shrubs  and 
trees,  reptiles  that  were  abundant  in 
the  grass/forb  successional  stage 
(e.g.,  Phrynosoma  coromtum)  became 
less  abundant,  and  others  that  pre- 
ferred wooded  sites  (e.g.,  Sceloporus 
occidentalis)  became  more  abundant. 


Historical  vs.  Present  Distributions 

Prior  to  this  study,  records  of  am- 
phibians and  reptiles  on  the  study 
area  were  limited;  one  of  the  primary 
reasons  for  which  this  study  was 
conducted  was  to  assemble  basic  dis- 
tribution information.  Therefore, 
range  expansions  or  reductions  were 
hard  to  document.  This  study  re- 
sulted in  range  extensions  of  ap- 
proximately 35  species,  and  clarified 
the  relationship  of  Arizona  habitats 
to  habitats  in  adjacent  geographic 
regions.  Many  species,  such  as  Helod- 
erma  suspectum,  Eumeces  gilberti,  Sce- 
loporus clarki,  Tantilla  hobartsmithii, 
and  parthenogenic  whiptail  lizards 
(Cnemidophorus  flagellicaudus,  C.  uni- 
parens,  and  C.  velox)  proved  to  be 
considerably  more  widespread  than 
previous  records  indicated — not  sur- 
prising since  many  areas  had  never 
been  intensively  sampled.  The  expan- 
sion of  E.  gilberti' s  range  results  from 
the  discovery  of  the  California 
subspecies,  E.  g.  rubricaudatus,  in 
chaparral  and  pinyon-juniper  habi- 
tats; the  distribution  of  E.  g.  ari- 
zonenis  is  limited  to  a  cottonwood- 
willow  riparian  habitat  along  an  18 
km  (11  mi)  stretch  of  the  Has- 
sayampa  River  immediately  south  of 
Wickenburg  (see  Jones  et  al.  1985, 
Jones  and  Glinski  1985). 

Only  one  species  demonstrated  a 
range  reduction.  Pure  populations  of 
Bufo  microscaphus  have  apparently 
been  reduced  due  to  hybridization 
with  Bufo  woodhousei,  especially  on 


major  drainages.  Water  impound- 
ment and  diversion-associated 
changes  in  aquatic  habitats  from  per- 
manent riffles  and  runs  to  pools  may 
have  caused  the  immigration  of  B. 
woodhousei  into  areas  formerly  occu- 
pied by  only  B.  microscaphus  (Brian 
Sullivan  personal  comm.). 

There  is  considerable  taxonomic 
confusion  about  a  population  of 
Kinosternon  sonoriense  on  the  Big 
Sandy  River  near  Wikieup.  Because 
specimens  with  raised  9th  marginal 
scales  had  been  taken  from  this  area, 
Stebbins  (1966)  considered  this  popu- 
lation to  be  Kinosternon  flavescens,  but 
Iverson  (1978)  considered  it  to  be  K. 
sonoriense,  based  on  specimens  with- 
out 9th  marginals.  Of  the  12  indi- 
viduals observed  during  this  study,  6 
had  raised  9th  marginals  and  6  did 
not.  Based  on  its  large  separation 
from  the  nearest  population  of  K. 
flavescens,  Iverson  (personal  comm.) 
considers  this  population  to  be  an 
aberrant  form  of  K.  sonoriense. 


Similarity  of  Habitats  Types 

It  is  possible  to  discern  definite  pat- 
terns in  the  diversity  of  and  similari- 
ties between  the  herpetofaunas  of 
different  habitat  types  within  the 
study  area.  There  is  an  apparent  ele- 
vational  gradient  affecting  species 
diversity.  Desert  habitats  between 
610  and  1067  m  (2000-3500  ft),  ripar- 
ian habitats  between  549  and  1220  m 
(1800-4000  ft),  and  chaparral  habitats 
between  1067  and  1525  m  (3500-5000 
ft)  had  greater  species  richness  than 
higher  elevation  woodland  (>  1677  m 
or  5500  ft,  e.g.,  Ponderosa  pine)  and 
desert  habitats  (>  1220  m  or  4000  ft, 
e.g.,  sagebrush).  Additionally,  low 
elevation  desert  habitats  (>  610  m  or 
2000  ft,  e.g.,  creosotebush),  had  rela- 
tively low  species  diversity.  Higher 
species  diversity  on  middle  elevation 
habitat  types  may  reflect  these  habi- 
tats' moderate  environmental  and 
climatic  conditions,  whereas  higher 
and  lower  elevation  habitats  possess 


124 


extreme  environmental  and  climatic 
conditions  (e.g.,  temperature).  For 
example,  low  elevation  creosotebush 
habitats  have  sparse  canopies,  and 
temperatures  often  exceed  60  C  near 
the  surface  in  summer  (Costing 
1956).  High  elevation  sites  are  cold 
and  are  often  snowcovered  until  late 
April  so  that  the  growing  season  is 
short.  Although  possessing  relatively 
low  species  richness,  low  elevation 
creosotebush  habitats  are  more  di- 
verse than  high  elevation  sites.  These 
differences  in  diversity  may  reflect 
thermal  conditions  at  these  eleva- 
tional  extremes.  Many  of  the  species 
that  occur  within  creosotebush  are 
nocturnal,  and,  therefore,  these  ani- 
mals avoid  exposure  to  extreme  sur- 
face heat.  On  higher  elevation  habi- 
tats, the  problem  is  not  avoiding  heat 
but,  rather,  gaining  heat  for  activity. 
Other  than  along  rock  outcrops, 
rapid  heating  is  difficult  for  reptiles 
at  higher  elevations.  Differences  be- 
tween diversity  and  species  composi- 
tion on  medium  elevation  habitat 
types  probably  reflect  differences  in 
microhabitat  abundance  and  diver- 
sity on  habitat  types  (see  earlier  dis- 
cussion on  microhabitats).  Lack  of 
diversity  on  disclimax  desert  grass- 
land sites  probably  reflects  the  lack 
of  vegetation  structure  on  these  sites. 

There  was  similarity  in  the  herpe- 
tofaunas  of  certain  habitat  types.  All 
desert  habitats,  except  sagebrush, 
had  very  similar  herpetofaunas,  as 
did  most  moderate  elevation  habitats 
(e.g.,  chaparral,  pinyon-juniper,  and 
mixed  riparian  scrub).  This  is  pre- 
dictable because  all  of  these  habitats 
occur  in  close  proximity  and  are 
structurally  similar.  There  was  a 
moderate  degree  of  similarity  be- 
tween cottonwood-willow  riparian 
and  desert  habitats,  chaparral  and 
cottonwood-willow  riparian,  and 
chaparral  and  desert  habitats.  Be- 
cause cottonwood-willow  riparian 
habitats  traverse  through  both  desert 
habitats  and  upland  habitats,  many 
of  the  species  associated  with  the 
surrounding  habitats  also  frequent 
riparian  sites;  riparian  sites  are  im- 


portant sources  of  food  and  cover 
(Ohmart  and  Anderson  1986).  Simi- 
larities between  chaparral  and  desert 
habitat  types,  such  as  Mohave  Des- 
ert, Sonoran  Desert,  and  mixed  ripar- 
ian scrub,  result  from  occurrence  of 
typical  desert  species  (e.g.,  Callisau- 
rus  draconoides)  on  upland  sites  rather 
than  the  occurrence  of  upland  spe- 
cies (e.g.,  E.  gilberti)  on  desert  sites. 

The  diversity  of  and  similarities 
among  amphibian  and  reptile  com- 
munities of  habitat  types  also  may 
have  been  affected  by  the  proximity 
of  habitat  types  to  evolutionary  cen- 
ters. Because  of  the  many  new  rec- 
ords for  herpetofauna  generated  by 
this  study,  we  now  have  a  better  pic- 
ture of  the  sources  of  diversity  for 
this  area.  Many  of  the  amphibians 
and  reptiles  occurring  in  the  Sonoran 
and  Mohave  Deserts  evolved  in  Baja 
California  and  along  the  western  sec- 
tion of  mainland  Mexico;  these  areas 
were  linked  until  their  separation  13 
million  years  ago  (Murphy  1983). 
With  the  retreat  of  pleistocene  glacia- 
tion  and  spread  of  xerophyllous  and 
desert  habitats,  amphibians  and  rep- 
tiles moved  northward  into  southern 
California  and  southwestern  Ari- 
zona; hence,  Sonoran  and  Mohave 
Desert  habitat  types  have  similar  her- 
petofaunas. Although  many  species 
immigrated  into  what  is  today  the 
Sonoran  and  Mohave  Deserts,  only  a 
few  species  immigrated  as  far  north 
as  the  Great  Basin  Desert.  Higher  ele- 
vations may  have  precluded  many  of 
these  species  from  colonizing  the 
Great  Basin  desert  habitat  types  and, 
hence,  it's  herpetofauna  is  different 
from  and  less  rich  than  those  of  the 
other  two  deserts. 

The  discovery  of  the  subspecies 
Eumeces  gilberti  rubricaudatus,  for- 
merly unknown  in  Arizona,  suggests 
that  Arizona  chaparral  was  closely 
associated  with  (Zalifomia  chaparral 
during  Pleistocene  glaciation;  E.  g. 
rubricaudatus  evolved  in  California 
sclerophyll  woodland  (Taylor  1935). 
That  parthenogenic  whiptail  lizards, 
such  as  Cnemidophorus  flagellicaudis, 
C.  uniparens,  and  C.  velox,  are  absent 


from  California  chaparral  suggest 
that  these  species  evolved  after  Pleis- 
tocene glaciation. 

There  were  a  few  inconsistencies 
in  the  results  of  the  two  analyses 
used  to  determine  similarity  between 
habitats  (the  cluster  analysis  of  all 
data  vs.  the  cluster  analysis  of  only 
array  data).  These  inconsistences  par- 
tially result  from  the  inconsistency  of 
arrays  to  capture  turtles  and  medium 
and  large-sized  snakes,  and  partially 
from  the  analyses  themselves  (see  the 
Methods  Section  for  a  more  detailed 
explanation). 

Conclusions  and 
Recommendations 

This  survey  indicates  that  most  spe- 
cies present  within  western  Arizona 
are  widespread,  and  that  few  war- 
rant special  management  considera- 
tion. However,  it  is  evident  that  cer- 
tain species  are  more  vulnerable  to 
range  or  population  reduction  than 
others.  Generally,  these  species  are 
those  that  require  microhabitats  that 
are  easily  affected  by  land  uses. 

It  appears  that  habitat  moisture 
and  moderated  surface  temperatures 
are  of  primary  importance  to  many 
species  in  western  Arizona.  Downed 
and  dead  surface  litter  (debris),  such 
as  logs  and  leaves,  play  a  major  role 
in  moderating  surface  temperature 
and  enhancing  moisture  (Dauben- 
mire  1974).  Horizontal  and  vertical 
vegetation  structure  also  help  moder- 
ate temperatures  and  increase  mois- 
ture. In  developing  management 
schemes,  priority  should  be  given  to 
maintaining  or  enhancing  surface  lit- 
ter and  vegetation  structure.  It  is  im- 
portant to  maintain  tree  reproduc- 
tion, and  to  leave  litter  on  the  surface 
rather  than  piling  and  burning  it.  The 
latter  practice  is  especially  important 
on  cottonwood-willow  riparian  sites 
within  deserts,  since  many  species  in 
riparian  sites  are  totally  dependent 
on  surface  litter  for  their  survival 
(Jones  and  Glinski  1985).  Many  ripar- 
ian sites  within  the  study  area  have 


125 


reduced  amounts  of  trees  and  sur- 
face litter,  principally  because  live- 
stock have  greatly  reduced  the  repro- 
duction of  Cottonwood  trees  by  re- 
ducing the  survival  of  seedlings 
(Jones  1981a).  Management  prescrip- 
tions are  needed  on  these  sites  to  in- 
crease the  survivorship  of  seedling 
and  young  cottonwood  trees. 

Populations  of  ''upland"  species 
(e.g.,  Eumeces  gilberti)  on  habitat  is- 
lands are  more  vulnerable  to  impacts 
associated  with  certain  land  uses 
than  populations  occurring  on  major, 
continuous  stands.  Jones  et  al.  (1985) 
described  these  habitat  islands,  some 
only  10  ha  (25  acres)  in  size.  Loss  or 
fragmentation  of  any  portion  of  these 
islands  could  result  in  the  local  extir- 
pation of  one  or  several  upland  spe- 
cies (see  Bury  and  Luckenbach  1983 
and  Harris  1984  for  the  effects  of 
habitat  fragmentation  and  habitat 
loss  on  species  occurring  on  habitat 
islands).  Because  even  small  modifi- 
cations to  island  habitats  can  result  in 
the  extirpation  of  upland  species, 
proposed  projects  should  be  moved 
to  alternative  sites  whenever  pos- 
sible; mitigation  strategies  should  be 
used  only  as  a  last  resort.  Top  prior- 
ity should  be  given  to  protecting 
these  sites  in  land-use  and  on-the- 
ground  activity  plans  (see  Jones  et  al. 
1985  for  specific  locations  of  these 
sites). 

Although  all  amphibians  in  the 
study  area  (excluding  Bufo  mi- 
croscaphus)  appear  to  be  stable,  water 
in  many  habitats  continues  to  be  de- 
veloped. In  addition,  new  informa- 
tion (Bruce  Bury  personal  comm. 
Com  and  Fogleman  1984)  suggest 
that  several  populations  of  ranid 
frogs  have  been  extirpated  from 
western  North  America,  although 
there  is  no  apparent  cause  for  their 
extirpation.  Considering  the  heavy 
use  of  spring  and  creek  water,  and 
the  reported  loss  of  many  ranid 
populations  in  the  West,  high  prior- 
ity should  be  given  to  monitoring 
amphibian  populations  at  springs 
and  creeks  in  Arizona.  Additionally, 
high  priority  should  be  given  to  de- 


termining the  extent  of  hybridization 
between  the  toads  B.  microscaphus 
and  Bufo  woodhousei.  Pure  popula- 
tions of  B.  microscaphus  should  be  lo- 
cated and  protected  against  hybridi- 
zation with  B.  woodhousei.  If  only  a 
few  pure  populations  are  found,  the 
Arizona  Game  and  Fish  Department 
and/or  the  U.S.  Fish  and  Wildlife 
Service  should  set  up  a  captive 
breeding  program  to  reduce  this 
toad's  risk  of  extinction. 

Although  I  obtained  distributional 
records  of  Gopherus  agassizii,  Burge 
(1979, 1980)  and  Schneider  (1980) 
provide  considerably  more  detail  on 
the  needs  of  this  species.  However, 
many  biologists  consider  G.  agassizii 
to  be  declining  throughout  most  of 
its  range.  The  U.S.  Fish  and  Wildlife 
Service  (1987)  continues  to  list  G. 
agassizii  as  a  species  that  needs  fur- 
ther study  to  determine  its  status, 
although  it  has  determined  that  the 
Federal  listing  of  the  tortoise 
throughout  its  range  is  warranted 
but  precluded  by  species  needing 
more  immediate  listing  (e.g.,  species 
in  more  eminent  danger  of  extinc- 
tions). The  BLM  should  continue  to 
give  high  priority  to  the  study  and 
management  of  this  species  in  Ari- 
zona. 

If  the  few  measures  suggested  in 
this  paper  are  implemented,  western 
Arizona  should  continue  to  support 
one  of  North  America's  most  diverse 
herpetofaunas. 

ACKNOWLEDGMENTS 

I  am  indebted  to  several  people  for 
the  completion  of  this  project.  Don 
Seibert,  Bob  Furlow,  and  Ted  Cor- 
dery  were  instrumental  in  obtaining 
funding,  equipment,  and  personnel 
for  this  study.  Lauren  Kepner,  Tim 
Buse,  Dan  Abbas,  Terry  Bergstedt, 
Kelly  Bothwell,  William  Kepner, 
Dave  Shaffer,  Bob  Hall,  Ted  Cordery, 
Scott  Belfit,  Ted  Allen,  Ken  Relyea, 
Becky  Peck,  Brian  Millsap,  Jim  Zook, 
Jim  Harrison,  and  Greg  Watts  helped 
collect  both  animal  and  habitat  data. 


Special  thanks  to  W.L.  Minckley  and 
M.J.  Fouquette  for  technical  contribu- 
tions to  this  project's  study  design, 
and  to  the  Bureau  of  Land  Manage- 
menf  s  line  managers  and  supervi- 
sors. Bill  Barker,  Roger  Taylor,  Barry 
Stallings,  Dean  Durfee,  Gary 
McVicker,  and  Malcolm  Schnitkner, 
for  their  continuous  support  of  re- 
source inventories  on  public  lands.  I 
thank  John  Fay,  Scott  Belfit,  R.  Bruce 
Bury,  and  Robert  Szaro  for  review  of 
this  manuscript.  Finally,  all  of  us 
who  strive  for  the  conservation  of 
nongame  wildlife  on  public  lands  are 
indebted  to  Gary  McVicker,  Bill 
McMahan,  and  Don  Seibert  for  their 
tireless  efforts  in  getting  top-level 
management  to  support  nongame 
programs. 

LITERATURE  CITED 

Bennett,  Stephen  H.,  J.  Whitfield  Gib- 
bons, and  Jill  Glanville.  1980.  Ter- 
restrial activity,  abundance,  and 
diversity  of  amphibians  in  differ- 
ently managed  forest  types. 
American  Midland  Naturalist 
103:412-416. 

Brown,  David  E.  1978.  The  vegeta- 
tion and  occurrence  of  chaparral 
and  woodland  flora  on  isolated 
mountains  within  the  Sonoran  and 
Mohave  Deserts  in  Arizona.  Jour- 
nal of  Arizona  Academy  Sciences 
13:1-12. 

Brown,  David  E.,  Charles  H.  Lowe, 
and  Charles  P.  Pase.  1979.  A  digit- 
ized classification  system  for  the 
biotic  communities  of  North 
America,  with  community  (series) 
and  association  examples  from  the 
Southwest.  Journal  of  Arizona 
Academy  of  Science  14,  Suppl.  1, 
p.  16. 

Burge,  Betty  L.  1979.  A  survey  of  the 
present  distribution  of  the  desert 
tortoise  (Gopherus  agassizii)  in  Ari- 
zona. U.S.  Bureau  of  Land  Man- 
agement, Denver,  Colorado.  Con- 
tract No.  YA-512-CTB-108. 

Burge,  Betty  L.  1980.  Survey  of  the 
present  distribution  of  the  desert 


126 


tortoise,  Gopherus  agassizii,  in  Ari- 
zona. U.S.  Bureau  of  Land  Man- 
agement, Denver,  Colorado.  Con- 
tract No.  YA-512-CT8-108. 

Bury,  R.  Bruce.  1980.  What  we  know 
and  do  not  know  about  off-road 
vehicle  impact  on  wildlife.  In  R.N. 
Andrews  and  P.F.  Nowah  (eds.). 
Off-road  vehicle  use:  a  manage- 
ment challenge.  U.S.D.A.  Office  of 
Environmental  Quality,  748  p. 

Bury,  R.  Bruce  and  Stephen  D. 
Busack.  1974.  Some  effects  of  off- 
road  vehicles  and  sheep  grazing 
on  lizard  populations  in  the 
Mojave  Desert.  Biological  Conser- 
vation 6:179-183.  Bury,  R.  Bruce, 
Roger  A.  Luckenbach,  and 
Stephen  D.  Busack.  1977.  Effects  of 
off-road  vehicles  on  vertebrates  in 
the  California  Desert.  U.S.  Fish 
and  Wildlife  Service  Wildlife  Re- 
search Report  No.  8. 

Bury,  R.  Bruce  and  M.G.  Raphael. 
1983.  Inventory  methods  for  am- 
phibians and  reptiles.  Proceedings 
of  the  International  Conference  on 
Renewable  Resources,  Inventories 
for  monitoring  changes  and 
trends.  Oregon  State  University, 
Corvallis. 

Bury,  R.  Bruce  and  Roger  A.  Lucken- 
bach. 1983.  Vehicular  recreation  in 
arid  lands  drives:  biotic  responses 
and  management  alternatives,  p. 
217-221.  In  R.H.  Webb  and  H.G. 
Wilshire  (eds.).  Environmental  ef- 
fects of  off-road  vehicles:  impacts 
and  management  in  arid  regions. 
Springer- Verlag,  New  York,  New 
York. 

Buse,  Timothy  C.  1981.  Distribution, 
ecology,  and  habitat  management 
of  the  reptiles  and  amphibians  of 
the  Cerbat  planning  unit,  Mohave 
County,  Arizona.  U.S.  Bureau  of 
Land  Management,  Kingman,  Ari- 
zona. Unpubl.  Man. 

Buse,  Timothy  C.  1983.  Herpetologi- 
cal  records  from  northwestern 
Arizona.  Herpetol.  Rev.  14:53-54. 

Campbell,  Howard  W.  and  Stephen 
P.  Christman.  1982.  Field  tech- 
niques for  herpetofaunal  commu- 
nity analysis,  p.  193-209.  In  Norm 


J.  Scott  (ed.),  Herpetological  com- 
munities, U.S.  Fish  and  Wildlife 
Service,  Wildlife  Research  Report 
Number  13. 

Case,  Thomas  J.  1983.  The  reptiles: 
ecology.  In  T.J.  Case  and  M.L. 
Cody  (eds.).  Island  biogeography 
in  the  Sea  of  Cortez.  University  of 
California  Press,  Berkeley. 

Connor,  Edward  F.  and  Daniel  Sim- 
berloff.  1979.  The  assembly  of  spe- 
cies communities:  chance  or  com- 
petition? Ecology  60:1132-1140. 

Com,  Paul  Stephan  and  James  C. 
Fogleman.  1984.  Extinction  of 
montane  populations  of  the  north- 
ern leopard  frog  (Rana  pipiens)  in 
Colorado.  Journal  of  Herpetology 
18:147-152. 

Creusere,  F.  Michael  and  Walter  G. 
Whitford.  1976.  Ecological  rela- 
tionships in  a  desert  anuran  com- 
munity. Herpetologica  32:7-18. 

Daubenmire,  Rexford  F.  1974.  Plants 
and  environment:  a  textbook  of 
autecology.  John  Wiley  and  Sons, 
New  York,  New  York.  3rd  Ed. 

Fitch,  Henry  S.  1955.  Habits  and  ad- 
aptations of  the  Great  Plains  skink 
(Eumeces  obsoletus).  Ecological 
Monographs  25(3):59-83. 

Reharty,  Eugene  D.  1967.  Compara- 
tive ecology  of  Thamnophis  elegans, 
Thamnophis  cyrtopsis,  and  Thamno- 
phis rufipunctatus  in  New  Mexico. 
Southwestern  Naturalist  12(3):207- 
230. 

Hair,  Jay  D.  1980.  Measurements  of 
ecological  diversity,  p.  269-275.  In 
S.D.  Schemnitz  (ed.).  Wildlife 
Management  Techniques  Manual. 
The  Wildlife  Society,  Washington, 
D.C. 

Harris,  Larry  D.  1984.  Island  biogeo- 
graphy applied:  old  growth  is- 
lands and  wildlife  conservation  in 
the  western  Cascades.  University 
of  Chicago  Press,  Chicago,  Illinois. 

Hulse,  Arthur  C.  1974.  An  autecol- 
ogy study  of  Kinosternon  sonoriense 
leconte  (Chelonia:  Kinosternidae). 
Phd.  Dissertation,  Arizona  State 
University,  Tempe. 

Iverson,  John.  1978.  Distributional 
problems  of  the  genus  Kinosternon 


in  the  American  Southwest. 
Copeia  1978:476-479. 
Jones,  K.  Bruce.  1981a.  Effects  of 
grazing  on  lizard  abundance  and 
diversity  in  western  Arizona. 
Southwestern  Naturalist  26(2): 107- 
115. 

Jones,  Kenneth  Bruce.  1981b.  Distri- 
bution, ecology,  and  habitat  man- 
agement of  the  reptiles  and  am- 
phibians of  the  Hualapai- Aquarius 
planning  areas,  Mohave  and 
Yavapai  Counties,  Arizona.  U.S. 
Bureau  of  Land  Manage.  Technical 
Note  No.  353,  Denver,  Colo. 

Jones,  K.  Bruce.  1987.  Amphibians 
and  reptiles,  p.  267-290.  In  A.Y. 
Cooperrider,  R.J.  Boyd,  and  H.R. 
Stuart  (eds.).  Inventory  and  moni- 
toring of  wildlife  habitat.  U.S.  Bu- 
reau of  Land  Management,  Den- 
ver, Colorado  xviii,  858  p. 

Jones,  K.  Bruce.,  Dan  R.  Abbas,  and 
Terry  A.  Bergstedt.  1981.  Herpeto- 
logical records  from  central  and 
northwestern  Arizona.  Herpeto- 
logical Review  12(1):16. 

Jones,  K.  Bruce  and  Patricia  C. 

Glinski.  1985.  Microhabitats  of  liz- 
ards in  a  southwestern  riparian 
community,  p.  355-358.  In  R.  Roy 
Johnson  et.  al..  Riparian  ecosys- 
tems and  their  management:  rec- 
onciling conflicting  uses.  First 
North  American  riparian  confer- 
ence. Rocky  Mountain  Forest  and 
Range  Experimental  Station,  Gen- 
eral Technical  Report  Number 
RM-120.,  Fort  Collins,  Colo. 

Jones,  K.  Bruce,  Lauren  P.  Kepner, 
and  William  G.  Kepner.  1983. 
Anurans  of  Vekol  Valley,  central 
Arizona.  Southwestern  Naturalist 
28(4):469-470. 

Jones,  K.  Bruce,  Lauren  P.  Kepner, 
and  Thomas  E.  Martin.  1985.  Spe- 
cies of  reptiles  occupying  habitat 
islands  in  western  Arizona:  a  de- 
terministic assemblage.  Oecologia 
66:595-601. 

Jones,  K.  Bruce,  Lauren  M.  Porzer, 
and  Kelly  J.  Bothwell.  1982.  Her- 
petological records  from  westcen- 
tral  Arizona.  Herpetological  Re- 
view. 13(2):54. 


127 


Justus,  J.  T.,  Mark  Sandomir,  Tom 
Urquhart,  and  Barbara  Orgel 
Evan.  1977.  Developmental  rates 
of  two  species  of  toads  from  the 
desert  Southwest.  Copeia 
1977:592-594. 

Kahn,  Walter  C.  1960.  Observations 
on  the  effect  of  a  bum  on  a  popu- 
lation of  Sceloporus  occidentalis. 
Ecology  41(2):358-359. 

Mayhew,  William  W.  1968.  Biology 
of  desert  amphibians  and  reptiles, 
p.  195-356.  In  G.W.  Brown,  Jr. 
(ed.).  Desert  biology.  Academic 
Press,  New  York,  New  York. 

Murphy,  Robert  W.  1983.  Paleobio- 
geography  and  genetic  differentia- 
tion of  the  Baja  California  herpe- 
tofauna.  Occasional  Papers  Cali- 
fornia Academic  Sciences  137:1-48. 

Norris,  Kenneth  S.  1953.  The  ecology 
of  the  desert  iguana,  Dipsosaurus 
dorsalis.  Ecology  34:265-287. 

Ohmart,  Robert  D.  and  Bertin  W. 
Anderson.  1987.  Riparian  habitats, 
p.  169-200.  In  A.Y.  Cooperrider, 
R.J.  Boyd,  and  H.R.  Stuart  (eds.). 
Inventory  and  monitoring  of  wild- 
life habitat.  U.S.  Bureau  of  Land 
Management,  Denver,  Colorado 
xviii,  858  p. 

Oosting,  Henry  J.  1956.  The  study  of 
plant  communities.  W.H.  Freeman 
and  Company,  San  Franciso,  Cali- 
fornia. 

Pianka,  Eric  R.  1966.  Convexity,  des- 
ert lizards,  and  spatial  heterogene- 
ity. Ecology  47:1055-1059. 

Pianka,  Eric  R.  and  William  S. 
Parker.  1972.  Ecology  of  the  igua- 
nid  lizard,  Callisaurus  draconoides. 
Copeia  1972:493-508. 

Pimental,  Richard  A.  1979.  Mor- 
phometries: the  multivariate 
analysis  of  biological  data.  Ken- 
dall/Hunt Publishing  Company, 
Dubuque,  Iowa. 

Platz,  John  E.  1984.  Status  report  for 
Rana  onca  Cope.  Status  report  pre- 
pared for  the  U.S.  Fish  and  Wild- 
life Service,  Albuquerque,  New 
Mexico.  28  p. 

Rogers,  Thomas  L.  and  Henry  S. 
Fitch.  1947.  Variation  in  the  skinks 
(Reptilia:  Lacertilia)  of  the  skil- 


tonianus  group.  University  of 
California  Publications  in  Zoology 
48(4):169-220. 

Schneider,  Paul  B.  1980.  A  popula- 
tion analysis  of  the  desert  tortoise 
in  Arizona  during  1980.  U.S.  Bu- 
reau of  Land  Management,  Phoe- 
nix, Arizona.  Contract  No.  AZ- 
950-279-0014. 

Simovich,  Marie  A.  1979.  Post  fire 
reptile  succession.  Cal-Neva 
1979:104-113. 

Stebbins,  Robert  C.  1966.  A  field 
guide  to  western  reptiles  and  am- 
phibians. Houghton  Mifflin  Co., 
Boston,  1st  Edition. 

Stebbins,  Robert  C.  1985.  A  field 
guide  to  western  reptiles  and  am- 
phibians. Houghton  Mifflin  Co., 
Boston,  2nd  Edition. 

Szaro,  Robert  C,  Scott  C.  Belfit,  and 
J.  Kevin  Aitkin.  1985.  Impact  of 
grazing  on  a  riparian  garter  snake, 
p.  359-363.  In  R.  Roy  Johnson  et. 
al..  Riparian  ecosystems  and  their 
management:  reconciling  conflict- 
ing uses.  First  North  American 
riparian  conference.  Rocky  Moun- 
tain Forest  and  Range  Experimen- 
tal Station,  General  Technical  Re- 
port Number  RM-120.,  Fort 
Collins,  Colorado. 

Taylor,  Edward  H.  1935.  A  taxo- 
nomic  study  of  the  cosmopolitan 
scincoid  lizards  of  the  genus 
Eumeces  with  an  account  of  the 
distributions  and  relationships  of 
its  species.  University  of  Kansas 
Science  Bulletin  23:1-643. 

U.S.  Fish  and  Wildlife  Service.  1987. 
Notice  of  findings  on  petitions  and 
initiation  of  status  review.  Federal 
Register  52(126):24485-24488. 

Van  Devender,  Thomas  R.  and  Wil- 
liam G.  Spaulding.  1977.  Develop- 
ment of  vegetation  and  climate  in 
the  southwestern  United  States. 
Science  204:701-710. 

Walker,  Robert  F.  and  Walter  G. 
Whitford.  1970.  Soil  water  absorp- 
tion capabilities  in  selected  species 
of  anurans.  Herpetologica 
26(4):411-416. 

York,  John  C.  and  William  A.  Dick- 
Peddie.  1969.  Vegetation  changes 


in  southern  New  Mexico  during 
the  past  hundred  years,  p.  157-166. 
In  W.G.  McGinnes  and  B.J. 
Goldman  (eds.).  Arid  lands  in  per- 
spective. University  of  Arizona 
Press,  Tucson. 
Zweifel,  Richard  G.  and  Charles  H. 
Lowe.  1966.  The  ecology  of  a 
population  of  Xantusia  vigilis,  the 
desert  night  lizard.  American  Mu- 
seum Novitates  2247:1-57. 


128 


Multivariate  Analysis  of  the 
Summer  Habitat  Structure  of 
Rana  pipiens  Schreber,  in 
Lac  Saint  Pierre  (Quebec, 
Canada)^ 

N.  Beauregard^  and  R.  Leclair  Jr^ 


Abstract.— Thirty  stations  representing  various  ripar- 
ian habitats  typical  of  the  Lac  Saint  Pierre  area  were 
sampled  with  a  system  of  drift  fences  and  funnel 
traps  to  characterize  the  summer  habitat  structure 
of  a  leopard  frog  population.  A  discriminant  analysis 
indicates  that  habitats  with  high  frog  density  (1)  ore 
close  to  the  marsh  line,  (2)  have  a  tall  herbaceous 
stratum  with  high  richness  and  (3)  have  a  low  moss 
cover.  A  stepwise  multiple  regression  model  used  5 
of  the  vegetation  structure  variables,  and  explains 
CO.  70%  of  the  variability  associated  with  frog  density 
among  stations. 


The  leopard  frog,  Rana  pipiens,  is  the 
most  abundant  frog  species  in  the 
Lac  Saint  Pierre  area  (Leclair  1985, 
Leclair  and  Baribeau  1982,  Paquin 
1982),  and  also  one  of  the  most  com- 
mon vertebrates  in  aquatic  communi- 
ties in  North  America  (Dole  1965a). 
Despite  this  apparent  abundance, 
many  herpetological  surveys  made  in 
the  last  fifteen  years  have  shown  dra- 
matic reductions  in  leopard  frog  den- 
sities. Gibbs  et  al.  (1971)  estimated  a 
50%  drop  in  the  global  population  of 
leopard  frogs  in  the  USA,  during  the 
1960's.  Many  other  workers  have  re- 
ported population  reductions  and 
local  extinctions  in  Canada  and  the 
USA  (Collins  and  Wilbur  1979,  Cook 
1984,  Degraaf  and  Rudis  1983,  Froom 
1982,  Hayes  and  Jennings  1986,  Hine 
et  al.  1981). 

In  area  where  hypothesis  of  preda- 
tion  or  competition  by  introduced 
species  (Bullfrogs  or  predatory 
fishes)  (Hayes  and  Jennings  1986) 
does  not  apply,  two  major  causes 
have  been  invoked  as  responsible  for 

^  Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  t^ammals  in  North  America.  (Flag- 
staff .  AZ,  July  19-21,  1988). 

^Norman  Beauregard  is  a  graduate  stu- 
dent in  Environmental  Sciences,  Universite 
du  Quebec  a  Trois-Rivieres,  Department  of 
chimie-biologie,  CP.  500  Trois-Rivieres,  Que- 
bec, Canada,  G9A  51-17. 

^Raymond  Leclair  Jr.  is  Professor  of  Her- 
pefology  and  Ecology.  Universite  du 
Quebec  a  Trois-Rivieres,  Department  of 
chimie-biologie,  CP.  500  Trois-Rivieres,  Que- 
bec, Canada,  G9A  5H7. 


this  situation  (1)  overexploitation  of 
natural  stocks,  and  (2)  loss  or  altera- 
tion of  habitat  rendering  it  unsuitable 
for  R.  pipiens  (Cook  1984,  Frier  and 
Zappalorti  1984,  Leclair  1985,  Mar- 
cotte  1981,  Rittschof  1975).  Riparian 
habitats  have  been  especially  affected 
by  human  activities  (Sarrazin  et  al. 
1983,  MLCP  1985).  In  Canada,  50%  of 
the  wetlands  that  once  supported 
wildlife  have  now  been  reclaimed  for 
agricultural,  industrial  or  urban  de- 
velopment, or  have  been  altered  by 
pollution  (SCF  1980).  Even  greater 
riparian  habitat  has  occurred  along 
the  St.Lawrence  river,  where  70%  of 
the  riparian  habitats  have  been  elimi- 
nated. 

According  to  Ministere  des 
Loisirs,  de  la  Chasse  et  de  la  Peche 
(MLCP)  (1985),  essential  habitats  are 
those  vital  to  population  or  species 
survival,  whether  these  habitats  are 
used  temporarily  or  permanently. 
This  definition  emphasizes  several 
crucial  aspects  of  amphibian  habitat 
use,  i.e.  the  use  of  aquatic  as  well  as 
terrestrial  habitats,  and  of  migratory 
routes  between  the  two.  Up  to  now, 
quantitative  studies  of  habitat  re- 
quirements for  anuran  species  have 
focused  mostly  on  the  aquatic  habi- 
tats (Beebee  1977,  Clark  and  Euler 

1982,  Dale  et  al.  1984,  Gascon  and 
Planas  1986,  Hine  et  al.  1981).  This 
situation  largely  results  from  the  lack 
of  appropriate  quantitative  sampling 
method  for  amphibian  populations  in 
terrestrial  habitats  (Bury  and  Raphael 

1983,  Clawson  et  al.  1984).  Recently, 


Campbell  and  Christman  (1982),  and 
Vogt  and  Hine  (1982)  have  devel- 
oped adequate  techniques  that  help 
overcome  this  situation. 

The  aims  of  the  present  study 
were  (1)  to  characterize  the  structural 
aspects  (biotic  and  abiotic)  of  the  ter- 
restrial habitats  of  Rana  pipiens  and 
(2)  to  develop  a  model  relating  frog 
abundance  to  habitat  descriptors. 

Study  Area 

The  study  area  is  a  30  X  0.9  km  strip 
extending  from  Trois-Rivieres  to  Ber- 
thierville  (Quebec,  Canada),  on  the 
north  shore  of  Lac  Saint  Pierre  (73*30' 
W  X  46°05'  N).  The  Lac  Saint  Pierre 
covers  about  300  km^  and  is  formed 
by  a  widening  of  the  St.-Lawrence 
river  (fig.  1).  The  lake  flood  plain  is 
extensive  (Tessier  et  al.  1984)  and 
consequently,  spawning  sites  for 
amphibians  are  abundant  in  spring. 
The  habitats  most  frequently  used  by 
Rana  pipiens  (based  on  mating  call 
frequencies)  are  flooded  fields  of 
reed  phalaris  (Phalaris  arundinacea) 
and  of  purple  loosestrife  (Lythrum 
salicaria),  mixed  with  willow  {Salix 
sp.)  (Leclair  1983).  From  these  fields, 
numerous  bays,  small  rivers,  drain- 
ing canals  and  natural  or  man-made 
pools  facilitate  movement  of  frog  to- 
wards adjacent  terrestrial  habitats. 

According  to  the  maps  produced 
by  Denis  Jacques  (1986)  and  by 
Tessier  and  Caron  (1980)  on  the  ri- 
parian vegetation  of  Lac  Saint  Pierre, 


129 


at  least  ten  plant  communities  may 
be  recognized  on  the  criterion  of 
dominant  species.  These  plant  com- 
munities can  be  grouped  in  six  differ- 
ent physionomic  types  (table  1). 
Thirty  stations  were  selected  in  order 
to  sample  the  diversity  of  habitats. 
From  the  maps,  sampling  sites  were 
located  in  habitat  patches  not  having 
less  than  2500  m^  of  homogeneous 
vegetation.  The  final  choice  of  sites 
was  determined  by  physical  and  le- 
gal accessibility. 

Materials  and  Methods 


Sampling  Technique 

At  each  station,  frogs  were  sampled 
with  12  funnel  traps  placed  on  each 
side  of  two  15  m  drift  fences  made  of 
polyethylene  and  forming  a  right 
angle  (fig.  2).  Dirt  and/or  litter  was 
brushed  into  the  mouth  of  each  fun- 
nel to  simulate  a  natural  entrance 


(Clawson  and  Baskett  1982).  This  de- 
sign has  been  shown  to  allow  for 
sampling  in  various  kinds  of  terres- 
trial habitats,  and  to  provide  data  for 
the  estimation  of  demographic  para- 
meters and  for  comparison  between 
various  habitats  (Campbell  and 
Christman  1982,  Clawson  et  al.  1984). 

Funnel  traps  were  opened  for  at 
least  10  consecutive  days  in  each  pe- 
riod (10  days  in  May,  10  in  June,  12 
in  July,  10  in  august,  22  in  Fall)  and 
were  checked  every  other  day. 

Data  recorded  for  each  capture 
were:  date,  station  number,  direction 
of  capture  (N,  S,  E  or  W),  species,  sex 
and  snout-urostyle  length.  Captured 
frog  were  marked  by  clipping  the 
fourth  digit  of  the  hindfoot.  Clipped 
phalanges  were  kept  for  age  determi- 
nation trough  skeletochronological 
examination  (Leclair  and  Castanet 
1987). 

Because  of  the  way  the  arrays 
were  used  for  sampling,  captures  re- 
flected the  relative  abundance  of 
frogs  among  stations,  not  their  abso- 
lute density. 


Environmental  Variables^ 

Each  station  was  characterized  by  6 
spatial  variables:  distance  to  the 
marsh  line  (DMARSH),  to  the  nearest 
permanent  pool  (DWATERP),  to  the 
nearest  temporary  pool  (DWATERP), 

^See  appendix  1  for  all  abbreviations 
used  in  fhe  text. 

^Variable  measured  monthly. 


^  7m 


h   15  m  -I 


Figure  2.— Schematic  representation  of  the 
trapping  arrays. 


Figure  1  .—General  location  of  the  study  area  (upper  map)  and  the  study  area's  relationship  to 
lake  St.  Pierre  (Quebec,  Canada)  (lower  map). 

130 


to  the  nearest  human  aUeration 
(road,  path,  residence,  crop)  (DHU- 
MAN),  and  to  the  nearest  open  habi- 
tat without  shrub  or  tree  canopy 
(DOPEN)  or  closed  habitat  with  can- 
opy (DCLOSE).  All  distances  were 
measured  in  the  field  with  a  topofil 
marker  (lost  thread  measure  appara- 
tus), except  for  some  measures  of 
DMARSH  taken  from  a  1:10  000  to- 
pographical map.  Elevation  from  the 


marsh  ground  (ALTREL)  was  taken 
with  a  Keuffel  and  Essel  altimeter. 
Water  table  level  (WTABLE^)  was 
measured  with  a  piezometer,  placed 
1  m  deep. 

Edaphic  variables  measured  were: 
soil  moisture  (MOIST^),  from  oven- 
dried  soil  samples  (80  °C,  24  hrs);  soil 
fractions  (SAND,  SILT,  CLAY),  as 
determined  by  the  Bouyoucos 
method  (Bouyoucos  1936);  soil  water 


Table  1  .—Characteristics  of  the  sampiing  stations  according  to  physiog- 
nomic type  and  to  major  plant  species. 


Sta. 

Physionomic  type 

Code 

1 

Open  dry  field 

O 

2 

Brushy  dry  field 

B 

3 

Wooded  swamp 

F 

4 

Riparian  marsh 

M 

5 

Shrub  swamp 

S 

6 

Wet  prairie 

P 

7 

Open  dry  field 

O 

8 

Wooded  swamp 

F 

9 

Wet  prairie 

P 

10 

Wooded  swamp 

F 

11 

Wet  prairie 

P 

12 

Shrub  swamp 

S 

13 

Wet  prairie 

P 

14 

Shrub  swamp 

S 

15 

Riparian  marsh 

M 

16 

Wet  prairie 

P 

17 

Brushy  dry  field 

B 

18 

Wet  prairie 

P 

19 

Brushy  dry  field 

B 

20 

Open  dry  field 

O 

21 

Open  dry  field 

O 

22 

Wet  prairie 

P 

23 

Shrub  swamp 

S 

24 

Riparian  marsh 

M 

25 

Wooded  swamp 

F 

26 

Shrub  swamp 

S 

27 

Wooded  swamp 

F 

28 

Riparian  marsh 

M 

29 

Wet  prairie 

P 

30 

Wet  prairie 

P 

Code       Major  plant  species 


Solidago  canadensis,  Aster  umbellafus 
Spirea  lafifolia,  Populus  tremuloides 
Acer saccharinum,  Laporfea  canad- 
ensis 

Sparganium  eurycarpum,  Scirpus  flu- 
viafilis 

Spirea  lafifolia.  On  ode  a  sensibilis 
Calamagrosfis  canadensis,  Phalaris 
arundinacea 

Solidago  rugosa.  Aster  umbellafus 
Acersaccharium,  Laporfea  canaden- 
sis 

Carex  lacustris,  Lyft-)rum  salicaria 
Salix  nigra,  Laporfea  canadensis 
Typha  lafifolia,  Onoclea  sensibilis 
Salix  spp .  .  Myrica  gale 
Calamagrosfis  canadensis 
Salix  cordafa,  Phalaris  arundinacea 
Sparganium  eurycarpum,  Equisetum 
fluviafile 

Phalaris  arundinacea 
Spirea  lafifolia,  Populus  tremuloides 
Carex  lacustris,  Lythrum  salicaria 
Spirea  lafifolia,  Salix  ssp. 
Solidago  canadensis.  Aster  umbellafus 
Phleum  pratense,  Agrostis  alba 
Calamagrosfis  canadensis,  Phalaris 
arundinacea 

Salix  ssp . ,  Rorippa  amphibia 
Sparganium  eurycarpum,  Sagittaria 
lafifolia 

Acer  saccharinum .  Populus  deltoides 
Salix  ssp . ,  Spirea  lafifolia 
Acer  saccharinum,  Onoclea  sensibilis 
Sparganium  eurycarpum,  Rorippa  am- 
phibia 

Carex  lacustris,  Lythrum  salicaria 
Calamagrosfis  canadensis,  Lythrum 
salicaria 


pH  (PH),  as  determined  with  a  Fisher 
pH-meter,  and  soil  temperature 
(TEMP2).  The  soil  temperature  vari- 
able used  in  the  statistics  is  ex- 
pressed as  the  sum  (5  reading  per 
month)  of  the  deviations  from  the 
daily  mean  taken  over  all  stations. 

Percent  of  ground  covered  by  lit- 
ter (LITTER5),  dead  wood  (DEAD- 
WOOD),  mosses  (MOSSCOV^),  her- 
baceous plants  (HERBCOV^),  and 
percent  bare  ground  (BAREGRND^) 
was  estimated  by  two  independent 
observers  in  5  X  5  meters  quadrat, 
and  the  mean  was  recorded.  Litter 
thickness  (LITTHICK^)  and  height  of 
the  herbaceous  stratum 
(HERBHGHT^)  represented  the  mean 
of  5  measurements  taken  with  a  me- 
terstick. 

Quantitative  assessment  of  vegeta- 
tion structure  was  represented  by 
Fox's  photometric  index  (Fox  1979) 
as: 

=  In  (la/lb) 
H  (b-a) 

where     represents  the  photometric 
index  for  the  amount  of  vegetation 
present  in  a  layer  between  two  levels, 
when  la  and  lb  are  the  light  intensi- 
ties immediately  above  and  below 
the  layer  and  H(b-a),  the  layer  thick- 
ness. Readings  of  light  intensity  were 
taken  with  a  Sekonic  light  meter,  at  0, 
20,  50,  and  100  cm  above  ground, 
above  the  herbaceous  canopy  and  in 
the  open  field  adjacent  to  stahon 
having  closed  canopy.  At  each  site, 
measurements  were  taken  at  five 
points  which  were  then  averaged  to 
provide  one  value.  Five  photometric 
index  were  computed:  vegetation  in- 
dex in  the  0-20  cm  layer  (PHOT20^); 
from  20  to  50  cm  (PHOT505);  from  50 
to  100  cm  (PHOTIOO^);  herb  layer 
above  100  cm  (PHOT+s);  and  shrub 
and  tree  strata  (PHOTCAN^). 

Vegetation  structure  was  also  de- 
scribed in  8  growth-form  categories: 
(TREE)  woody  plants  >  10  cm  diame- 
ter; (SHRUBHI)  woody  plants  >  2.5 
m  tall);  (SHRUBLO)  woody  plants  < 
2.5  m  tall;  (HGH')  high  graminoid 
herbs  >  100  cm  tall;  (MGH^)  medium 


131 


size  graminoid  herbs  from  20  to  100 
cm  tall;  (HBLH^)  high  broad-leaf 
herbs  >  100  cm  tall;  (MBLH^)  me- 
dium size  broad-leaf  herbs  from  20  to 
100  cm  tall;  (SMALL^)  herbs  layer 
below  20  cm  tall.  Basal  area  (BA- 
SARE A)  was  calculated  by  measur- 
ing tree  diameter  at  breast  height 
with  a  caliper.  Richness  in  herbs  spe- 
cies (NSPHERB),  shrubs 
(NSPSHRUB),  and  trees  (NSPTREE) 
was  determined  in  a  400  m^  quadrat. 
Minimal  area  of  homogeneous  vege- 
tation patch  (MINAREA)  was  esti- 
mated according  to  the  graphical 
method  of  Braun-Blanquet  (1964). 


Statistics 

Spearman  rank  correlations  and  chi- 
square  tests  were  used  to  test  for 
non-random  distribution  of  captured 
frogs  among  age  class  and  among 
periods  of  sampling.  Chi-square  tests 
were  also  used  to  detect  a  significant 
movement  of  frogs.  Because  some 
variables  were  not  normally  distrib- 
uted (Kolmogorov-Smirnov  test), 
they  were  square-root  transformed 
before  analysis  (indicated  on  appen- 
dix 1). 

For  final  analyses,  the  number  of 
variables  was  reduced  by  screening 
an  initial  principal  component  analy- 
sis (PCA),  and  by  using  Pearson  rank 
correlations  (Green  1979).  Because  of 
heterogeneity  in  the  variables  meas- 
ured, the  correlation  matrix  was  used 
to  extract  the  principal  components 
that  explained  the  greatest  propor- 
tion of  variability.  A  second  PCA 
with  the  22  extracted  variables 
served  to  define  the  structural  differ- 
ences among  stations,  and  to  reduce 
the  data  set  to  a  few  important  di- 
mensions that  could  identify  most  of 
the  structural  variability  among 
measured  habitats.  To  construct  a 
classification  model  for  potential 
habitats,  a  discriminant  analysis 
(DFA)  was  done  on  three  groupings 
of  stations  based  on  frog  abundance. 
The  model  was  validated  through  a 
simulation. 


A  stepwise  multiple  regression 
was  used  to  identify  which  habitat 
characteristics  account  for  most  of 
the  variability  in  the  analyzed  data 
(Clawson  et  al.  1984).  An  independ- 
ent variable  was  included  in  the 
model  when  its  partial  F-value  was 
significant  (a  =  0.05).  Partial  correla- 
tion coefficients  were  used  to  verify 
the  statistical  relation  between  the 
dependent  variables  and  the  inde- 
pendent one.  This  analysis  has  been 
identified  as  the  most  appropriate  to 
study  the  combined  effects  of  various 


habitat  variables  on  wildlife  density 
(Legendre  and  Legendre  1984).  Inter- 
pretation of  the  models  obtained 
from  such  analyses  takes  into  ac- 
count combinations  of  variables,  but 
not  variables  taken  individually  (Sch- 
errer  1984).  Statistical  analyses  were 
performed  with  SPSS  (Nie  et  al. 
1975). 

Results 

A  total  of  798  individuals  represent- 
ing 4  species  of  anurans  (Ram  pipiens, 


r 


Table  2.— Capture  data  by  sampling  period,  and  by  age  class. 

Number  of  captures 


By  month 

By  age  class 

Sum 

Sum^ 

WV4I  1  1 

Station 

M 

J 

J 

A 

S-O 

Adult  Juv.  NMY' 

adjusted 

1 

0 

1 

0 

2 

2 

4 

0 

1 

5 

5 

2 

0 

0 

3 

2 

6 

5 

3 

3 

11 

11 

3 

1 

0 

0 

1 

3 

2 

2 

1 

5 

5 

4 

8 

n 

5 

3 

7 

11 

17 

6 

34 

34 

5 

6 

3 

4 

2 

2 

15 

1 

1 

17 

17 

6 

7 

13 

3 

3 

9 

20 

11 

4 

35 

35 

7 

0 

1 

1 

3 

1 

5 

0 

1 

6 

6 

8 

0 

0 

1 

0 

2 

0 

2 

1 

3 

3 

9 

8 

8 

4 

2 

5 

14 

7 

6 

27 

27 

10 

3 

9 

3 

2 

2 

12 

3 

4 

19 

19 

11 

5 

3 

1 

2 

8 

7 

4 

8 

19 

19 

12 

11 

9 

4 

4 

3 

21 

3 

7 

31 

31 

13 

15 

8 

7 

12 

5 

28 

13 

4 

47 

47 

14 

22 

7 

10 

39 

47 

16 

12 

78 

107 

15 

12 

15 

8 

12 

39 

51 

21 

11 

86 

86 

16 

3 

0 

2 

0 

3 

1 

1 

5 

7 

17 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0 

18 

0 

0 

0 

0 

0 

0 

0 

0 

0 

19 

0 

0 

3 

0 

0 

3 

0 

0 

3 

3 

20 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0 

21 

0 

0 

3 

1 

1 

0 

0 

5 

5 

5 

22 

5 

2 

6 

2 

0 

8 

4 

0 

15 

15 

23 

10 

10 

2 

1 

0 

9 

13 

0 

23 

23 

24 

26 

21 

3 

2 

7 

28 

23 

1 

59 

59 

25 

8 

3 

1 

4 

3 

17 

2 

0 

19 

19 

26 

5 

1 

1 

3 

7 

12 

1 

3 

17 

17 

27 

0 

1 

4 

9 

6 

3 

5 

14 

19 

28 

2 

10 

12 

10 

20 

3 

n 

34 

47 

29 

5 

2 

0 

1 

1 

3 

5 

0 

9 

9 

30 

3 

5 

6 

7 

11 

3 

7 

21 

29 

Total 

135 

150 

86 

98 

178 

362 

161 

103 

647 

704 

^ Newly  metamorphosed  young. 

'Sum  adjusted  for  stations  not  inventoried  in  May. 


132 


R.  catesheiana,  R.  sylvatica,  Bufo  ameri- 
canus)  were  captured  during  the 
study.  Many  small  rodents  (n  =  188) 
and  a  few  weasels  (Mustela  ermim) 
were  also  captured.  The  results  pre- 
sented here  relate  only  to  R.  pipiens. 

Table  2  presents  the  capture  data 
for  the  various  stations  and  sampling 
periods,  along  with  data  on  popula- 
tion age  structure.  The  mean  capture 
rate  is  0.35  capture/day/ station;  sta- 
tions range  from  0  to  1.77  captures/ 


day/ station.  Preliminary  trials  on 
three  stations  in  fall  1986  had  given 
4.8  captures/day/station. 

Spearman's  correlation  coefficients 
(table  3)  from  among  all  possible  age 
groups  and  sampling  period  pairs 
were  all  significant  except  those  be- 
tween captures  at  period  1  and 
newly  metamorphosed  young  (R  = 
0.3471,  P  =  0.097).  We  also  compute  a 
contingency  table  (table  4)  to  check 
for  independence  of  the  two  vari- 


Table  3.— Spearman  rank  correlations  and  signifiance  level  between  cap- 
tures for  all  possible  age  groups  and  sampling  periods  pairs. 


Periods 


Age 


Period/Age 

M 

J 

J 

A 

s-o 

Adult 

Juv 

NMY' 

May 

«*« 

•♦» 

NS 

June 

0,88 

» 

July 

0.59 

0.57 

«« 

*** 

August 

0.63 

0.56 

0.64 

♦  ♦• 

**• 

Sep.-Oct, 

0.52 

0.47 

0.52 

0.75 

••• 

Adult 

0.88 

0.83 

0,74 

0.81 

0.67 

Juvenile 

0.82 

0.79 

0.59 

0.49 

0.60 

0.70 

♦ 

NMYl 

0.35 

0.44 

0.62 

0,70 

0,81 

0.56 

0.46 

'Newly  metamorphosed  youngs. 

'P<0.05. 

"P<0.01 

'"P<  0.001. 

NS  =  non  significant. 


Table  4.--Contingency  table  for  non  random  distribution  of  age  group  cap- 
tures among  the  physlonomic  types  of  habitat. 


Physionomic  types  of  habitat 


Age 

Dry 

Shrub 

Wooded 

Wet 

Riparian 

groups 

habitat' 

swamp 

swamp 

prairie 

marsh 

NMY 

10 

23 

11 

30 

29 

Count 

4.9 

26.5 

9.9 

28.3 

33.4 

Exp.  vol. 

Juvenile 

3 

34 

12 

48 

64 

Count 

7.7 

41.4 

15.4 

44.2 

52.2 

Exp.  vol. 

Adult 

17 

104 

37 

94 

no 

Count 

17.3 

93.1 

34.7 

99,5 

117.4 

Exp.  vol. 

For  all  habitats:  D.F.  =  8.  X' =  16.64,  0.025  <P<  0.05. 
Without  dry  habitats:  D.F.  =  6.     =  6.04.  0.10  <P<  0.25. 

'Open  dry  field  and  Brushy  dry  field  were  joined  to  respect  chi-square  require- 


ments. 


ables  ''age  group"  and  "physionomic 
type  of  habitat."  There  was  a  weak 
relationship  (D.F.  =  8,     =  16.64, 
0.025  <  P  <  0.5)  created  mostly  by  the 
capture  of  a  few  young  (n  =  5)  at  a 
dry  open  field  station  (#  21,  see  table 
2).  Otherwise,  all  other  habitats 
shared  proportional  distribution  for 
the  different  age  groups  (D.F.  =  6, 
=  8.04, 0.10  <  P  <  0.25).  Further  gen- 
eral PCA  and  DFA  models  used  the 
number  of  total  captures  per  station, 
irrespective  of  sampling  periods  or 
age  groups. 

Following  preliminary  screening, 
we  removed  variables  that  were  not 
normally  distributed  (DOPEN, 
DCLOSE,  DEADWOOD,  HBLH, 
SMALLH),  those  correlated  with 
other  variables  (DWATERP,  LITTER, 
SAND,  WTABLE),  and  those  related 
to  the  tree  and  shrub  strata 
(NSPTREE,  NSPSHRUB,  TREE, 
SHRUBHI,  SHRUBLO,  BASAREA, 
PHOTCAN)  which  diluted  the  re- 
sults of  PCA. 

Figure  3  illustrates  the  distribution 
of  the  remaining  22  variables  along 
the  first  two  PCA  axes,  based  on  data 
of  table  5.  The  first  axis  explains 
22.3%  of  the  variation  and  is  corre- 
lated to  descriptors  of  vegetation 
structure,  such  as  density  of  grami- 
noids  (HGH),  vegetation  height 
(HERBHGHT),  photometric  index 

2  «s») 


.  M.TRB. 

PHOT. 

MOSSCOV 

H — 1 — ^ 

'  PH071» 

your 

Ohuuan' 

Figure  3.— Projection  of  the  22  biophysical 
variable  vectors  onto  plane  defined  by  the 
first  two  principal  components.  The  circle  at 
the  origin  has  a  radius  of  0.30. 


133 


(PHOT20,  PHOT+),  litter  thickness 
(LITTHICK)  and  moss  cover 
(MOSSCOV).  The  second  axis  ex- 
plains 15.2%  of  the  variability  and  is 
correlated  to  marsh  distance 
(DMARSH,  ALTREL),  number  of 
herb  species  (NSPHERB),  mid-height 
graminoids  (MGH),  short  distance  to 
human  alteration  (DHUMAN)  and 
bare  ground  (BAREGRND).  The 
third  axis  explains  10.5%  of  the  data 
variability,  which  is  significant  ac- 
cording to  the  broken  stick  model 
(Frontier  1976,  Legendre  et  Legendre 
1984).  It  is  related  to  edaphic  factors 
such  as:  pH  (PH),  silt  fraction  (SILT) 
and  soil  moisture  (MOIST)  (table  5). 
The  forth  and  subsequent  axes  are 
not  significant. 

Figure  4a  gives  the  relative  posi- 
tion of  stations  according  to  the  first 
two  axes  of  the  PCA.  Five  groups 


may  be  easily  circled  at  best,  accord- 
ing to  their  physionomic  type.  Dry 
habitats  (open  and  brushy  fields)  are 
at  the  top  of  the  figure  and  are  char- 
acterized by  a  greater  distance  to  the 
marsh  line,  a  higher  moss  cover  and 
a  plant  cover  which  is  meager  but 
has  a  high  species  diversity.  The  dry 
open  fields  with  high  PH(DT+  are  dis- 
tinct from  the  dry  brushy  fields 
which  have  a  lot  of  bare  ground, 
those  two  variables  being  in  opposite 
direction  (fig.  3).  Although  the  vari- 
ables on  tree  and  shrub  strata  were 
removed  from  PCA,  wooded  and 
shrub  swamps  appear  distinct  from 
the  other  habitats.  They  are  clustered 
along  the  BAREGRND  and  MBLH 
vectors  (fig.  3)  opposed  to  variables 
describing  vegetation  structure  and 
positively  correlated  to  the  first  axe. 
Wet  prairies  and  riparian  marshes 


can  be  differentiated  from  the  other 
three  habitats  along  the  first  axis  by  a 
more  elaborated  herbaceous  struc- 
ture. Stations  positioned  in  the  Spar- 
ganium  eurycarpum  community, 
which  occupies  approximately  the 
first  100  m  of  the  riparian  marsh 
(Tessier  et  a.  1984),  have  a  very  wet 
soil,  the  water  receding  only  about 
the  end  of  May.  Wet  prairies  are  dis- 
tinguished from  the  preceding  habi- 
tat by  the  conjugated  differences  of 
many  variables  related  to  axe  2. 

Figure  4b  shows  the  position  of 
the  stations  as  in  figure  4a  but  are 
best  circled  by  classes  of  frog  abun- 
dance. This  figure  emphasizes  the 
relationship  between  habitat  aridity 
and  frog  density,  the  lowest  frog 
densities  occurring  in  the  driest  habi- 
tats (open  dry  field,  brushy  dry 
field).  Higher  frog  density  stations 
include  those  from  the  marsh  line 
and  those  from  the  wet  fields.  Inter- 
mediate frog  densities  occur  in  forest 
and  shrub  sites. 

A  DFA  of  frog  density  classes  al- 
lowed us  to  identify  a  few  variables 
that  were  easy  to  quantify  and  also  to 
classify  habitats  according  to  their 
potential  use  by  leopard  frogs.  Table 
6  presents  the  standardized  coeffi- 
cients (computed  with  z-score)  of  the 
variables  for  each  DFA  axis,  and 
non-standardized  coefficients  associ- 
ated with  classification  function. 
Four  such  variables  were  retained 
from  DFA.  DMARSH  alone  allows 
for  60%  of  the  stations  to  be  correctly 
classified.  Addition  of  the  NSPHERB 
variable  adds  another  13%.  When 
PHOT+  and  MOSSCOV  variables 
were  used,  90%  of  the  stations  were 
correctly  classified. 

Figure  5  integrates  information 
about  habitat  and  density  by  indicat- 
ing the  position  of  each  station  and 
group  centroids  of  frog  density 
classes  along  the  two  canonical  axes. 
As  for  PCA,  the  value  of  the  stan- 
dardized coefficients  for  each  vari- 
able associated  to  each  DFA  axis  is 
proportional  to  the  length  of  each  ar- 
row. The  first  axis,  which  represents 
the  major  part  of  the  interclass  vari- 


r  \ 

Table  5.— Sorted  factor  loadings  for  ttie  principal  component  analysis  of 
habitat  variables. 


Factor 


1 

2 

3 

(22.3%)' 

(15.2%) 

(10.5%) 

HGH 

0.845 

0.094 

0.271 

LITTHICK 

0.841 

0.102 

-0.119 

MOSSCOV 

-0.772 

0.180 

0.139 

HERBHGHT 

0.604 

0.144 

0.178 

MINAREA 

-0.602 

0.158 

0.184 

PHOT20 

0.586 

0.266 

-0.016 

HERBCOV 

0.536 

0.040 

-0.413 

PHOT+ 

0.845 

0.400 

0.250 

ALTREL 

-0.347 

0.753 

0.087 

DMARSH 

-0.394 

0.738 

-0.166 

NSPHERB 

-0.006 

0.654 

0.464 

MGH 

-0.067 

0.578 

-0.125 

DHUMAN 

-0.146 

-0.566 

0.050 

BAREGRND 

-0.444 

-0.537 

0.143 

SILT 

-0.098 

0.042 

0.696 

PH 

0.097 

0.264 

-0.640 

MOIST 

0.481 

-0,313 

0.509 

TEMP 

-0.108 

0.404 

-0.208 

PHOTIOO 

0.490 

-0.090 

-0.384 

PHOT50 

0.452 

0.030 

0.409 

MBLH 

-0.337 

-0.405 

-0.077 

CLAY 

-0.088 

0.214 

0.314 

'Percentage  of  total  variance  explained  by  each  component. 

V  :  :     :      .  .       ■  -  -  ^^^^^^^^^^^^^^^^^  ^^^  ^^  ^  ^ 


134 


Table  6.— Summary  statistics  for  discriminant  function  analysis  of  habitat 
characteristics  according  to  three  classes  of  frog  abundance  (as  defined 
In  fig.  4b). 


Variable 


Wilk's 
lambda 


%  correct  Standardized 
classification  coefficients^ 
total 


Unstandardized 
coefficients'* 


Axe  1 
88.8% 


Axe  2 

n.2% 


Axe  1 


Axe  2 


DMAf^H 

0,505 

0.0001 

60.0 

1.220 

0.065 

0.00613 

0.00033 

NSPHERB 

0.328 

<0.0001 

73.3 

-1.024 

0.505 

-0.143 

0.071 

PHOT+ 

0,255 

<0.0001 

73.3 

0.537 

0.709 

0.584 

0.771 

MOSSCOV^ 

0,214 

<0.0001 

90.0 

0.500 

-0.185 

0.480 

-0.179 

(CONSTANT) 

-0.908 

-1.522 

'DFA  useszscore  data  and  gives  the  relative  contribution  of  each  variable  to  final 
discrimination. 

^Classification  fonctlon  uses  original  data  and  allows  to  know  to  which  group  sam- 
pling stations  belong,  DMARSH  expressed  in  meter.  NSPHERB  espressed  in  number  of 
herb  species,  MOSSCOV  espressed  in  %  ground  cover. 


ability  (88.8%),  is  mostly  related  to 
DMARSH,  NSPHERB  and 
MOSSCOV.  The  second  axis  (11.2% 
of  intergroup  variation)  reflects  pri- 
marily variation  in  the  photometric 
index  above  1  m  (PHOT-»-)  and 
NSPHERB. 

To  validate  our  discriminant 
model,  we  randomly  drew  103 
samples  of  three  groups  of  stations, 
and  ran  a  DFA.  The  distribution  of 
the  103  samples  does  not  depart  sig- 
nificantly from  normality 
(Kolmogorow-Smirnov  test  =  1.233,  P 
=  0.096).  The  results  give  a  mean  of 
correct  classifications  of  68.6%  with  a 
maximum  of  83.3%  and  a  standard 
error  of  7.2%.  A  t-test  (T  =  2.98,  P(_l) 
=  0.0025)  indicates  that  the  probabil- 
ity of  obtaining  a  value  equal  to  90% 
is  less  than  0.0025. 

Finally,  using  stepwise  multiple 
regression  analyses,  we  identified 
those  variables  used  in  models  that 
best  predict  frog  abundance.  For 
such  modelling,  Clawson  et  al.  (1984) 
have  pointed  out  the  importance  of 


L.  \  1  U 

-2.0         -1.0  0  1.0  2.0 

FACTOR  1 


Figure  4a.— Ordination  of  the  sampling  sta- 
tions in  the  plane  defined  by  the  first  two 
principol  components  according  to  station 
physiognomy. 


incorporating  phenological  aspects  of 
habitat  utilization.  Directions  of  cap- 
tures (table  2)  were  then  analyzed  in 
order  to  group  the  capture  data  in 
different  periods  of  activity  based  on 


■  \  \ —  * 

•2.0        -1.0  0  1.0  2.0 


FACTOR  1 

Figure  4b.— Ordination  of  the  sampling  sta- 
tions in  the  plane  defined  by  the  first  two 
principle  components  according  to  frog 
abundance.  1:  number  of  capture  <  9;  2:  8 
<  number  of  capture  <  26;  3:  number  of 
capture  >  25. 


seasonal  patterns  of  movement  (i.e. 
movement  away  from  aquatic  over- 
wintering sites  in  Spring,  movement 
within  a  summer  foraging  range,  and 
movement  towards  aquatic  overwin- 
tering sites  in  Fall). 

Chi-square  values  (table  7a) 
showed  significant  movement  for 
period  1,2  and  5.  Individuals  cap- 
tured in  the  Fall  seem  to  move  back 
towards  the  lake  where  they  pre- 
sumably overwinter.  A  stepwise  re- 
gression model  associated  with  this 
period  (model  3)  would  thus  charac- 
terize habitat  used  during  Fall  migra- 
tion. Although  we  got  significant  chi- 
square  in  early  season  (sampling  pe- 
riods 1  and  2),  interpretation  is 
doubtful  whether  or  not  there  was  a 
migration  movement  from  the  over- 
wintering site  (i.e.  from  south  and 
east).  To  test  for  an  actual  movement, 
we  associated  the  two  compass  di- 
rections in  the  general  direction  to- 
wards the  overwintering  site  and  we 
tested  them  against  the  two  compass 
directions  in  the  general  direction 
away  from  the  overwintering  site 
(i.e.  north  and  west).  No  significant 
movement  was  then  noted  (table  7b). 
Consequently,  we  referred  to  the 
phenology  of  the  leopard  frog  de- 
scribed by  Dole  (1967)  and  Rittschof 


135 


(1975)  to  decide  for  grouping  of  sam- 
pling periods. 

In  May,  as  leopard  frogs  remained 
at  proximity  of  their  reproductive 
site  and  because  we  had  only  24  sam- 
pling stations  at  that  time,  data  from 
period  1  were  analyzed  separately 
(model  1).  Data  from  June,  July  and 
August  (periods  2, 3  and  4)  were 
grouped  together  to  construct  a 
single  model  (model  2)  because  in 
June  individuals  normally  tend  to 
disperse  in  their  summer  foraging 
habitats  (Rittschof,  1975),  and  in  July 
and  August  no  definite  movement 
direction  was  observed  (that  is  typi- 
cal when  foraging  habitat  is  occu- 
pied). We  also  analyzed  the  data  for 
all  periods  in  two  general  models 
(models  4  and  5). 

Model  1  (table  8)  explains  ca.  82% 
of  the  variation  in  frog  density  for 
the  month  of  May  using  6  variables. 
The  first  one  is  distance  to  marsh 


F 
U 
N 
C 
T 


Table  7a.— Capture  data  by  sampling  period  and  by  direction  and  chi- 
square  values  for  tests  of  goodness  of  fit.  P  values  „  0.05  are  considered 


significant. 

Month 

North  West 

South 

East 

Exp.vaiue 

P 

May 

45  24 

24 

41 

33.50 

9.42 

<0.025 

June 

42  28 

27 

52 

37.25 

11.56 

<0.010 

July 

26  T9 

15 

25 

21.25 

3.80 

>0.25 

August 

19  24 

31 

23 

24.25 

3.08 

>0.25 

Sep.-Oct. 

106  13 

26 

33 

44.50 

117.96 

<0.001 

Table  7b.- 

Results  of  test  for  nonrandom  distribution  of  captures  among  the 

two  general  directions  of  movement  from  and  away  overwintering  sites. 

Month 

North  +  West 

South  +  East 

P 

May 

69 

65 

0.119 

0.067 

>0.75 

June 

70 

79 

0.272 

0.215 

>0.50 

July 

45 

40 

0.294 

0.188 

>0.50 

August 

43 

54 

1.247 

1.031 

>0.25 

Sep.-Oct. 

119 

59 

20.224 

19.556 

<0.001 

X'.  =  Chi-square  wifh  Yates  correction  for  cor^tinuity. 


0 
N 

S 

c 

0 
R 
E 


2.0-- 


1 .0-- 


0  -- 


-1,0- 


NSPHERB 


DM  ARS 


,'22 


3.0     -  2.0     -  1,0       0  1.0       2.0  3.0 

FUNCTION   SCORE  1 


Figure  5.— Localization  of  the  sampling  stations  (represented  by  tfieir  abundance  class)  in 
\he  discriminant  space  according  to  their  function  score.  The  relative  contribution  of  each 
variable  (NSPHERB,  PHOT+,  DMARSH  AND  MOSSCOV)  involved  in  the  two  discriminant  func- 
tions is  indicated  by  the  length  of  each  vector.  Class  centroids  are  represented  by  *.  Mis- 
classified  stations  are  circled. 


line.  Four  of  the  five  other  variables 
are  related  to  soil  characteristics: 
temperature,  moisture,  silt  fraction 
and  bare  ground.  In  model  2  (sum- 
mer feeding  habitats),  about  70%  of 
variation  in  frog  density  is  explained 
by  only  three  variables:  distance  to 
marsh  line,  number  of  herb  species 
and  clay  fraction.  The  third  model, 
for  the  month  of  September  and  Oc- 
tober, explains  only  34.6%  of  vari- 
ation in  Fall  captures  with  two  vari- 
ables: DMARSH  and  NSPHERB.  It 
should  be  noted  that  the  same  two 
variables  explain  61.5%  of  the  vari- 
ation in  model  2. 

In  the  next  two  models  (table  8) 
the  seasonal  captures  were  corrected 
to  account  for  the  lower  number  of 
stations  sampled  in  May.  Model  4 
includes  five  variables:  DMARSH 
and  NSPHERB  again,  and  three  vari- 
ables related  to  vegetation  structure 
(PHOT+,  PHOT20,  PHOT50).  These 
last  three  variables  explain  an  addi- 
tional 21.6%  of  the  variation  in  frog 
density  in  the  model. 

Hooding  of  St.  Lawrence  river 
over  our  study  sites  is  a  major  mani- 
festation in  the  Lac  Saint  Pierre  area 


136 


having  a  strong  impact  on  frog  distri- 
bution as  indicated  by  the  presence 
of  the  variable  DMARSH  in  all  previ- 
ous models.  However,  when  water 
recesses,  we  get  a  mosaic  of  habitats 
that  can  be  found  elsewhere  in  North 
America  but  independently  of  the 
presence  of  such  marsh  line.  That  is 
the  reason  why  we  ran  another  mul- 
tiple regression  (model  5)  after  hav- 


ing removed  DMARSH.  This  last 
model  emphasizes  the  significance  of 
vegetation,  all  5  variables  included 
being  related  to  vegetation  structure. 
This  model  explains  69.2%  of  the 
variation  in  total  captured  frogs. 

To  facilitate  the  understanding  of 
our  interpretation,  we  present  in  ap- 
pendix 2  the  significant  level  of  the 
Pearson  rank  correlations  between 


Table  8.— Multiple  regression  models  for  frog  captures. 

Variable 

Coefficient 

Probability 

Adjusted 

(p  ±SE) 

(a  value  for  F) 

Model  1  Capture  in  May  (24  stations) 

(Intercept) 

-2.51     ±  3.60 

0.4950 

0,4950 

DMARSH 

-0.0116    ±  0.0027 

0.0005 

0,484 

TEMP 

0.176    ±  0.050 

0.0027 

0,574 

MOIST 

0.230    ±  0.053 

0.0004 

0.636 

PHOT50 

-1.430    ±  0.429 

0.0039 

0.691 

SILT 

0.284    ±  0.074 

0.0013 

0.770 

BAREGRND 

-0.999    ±  0.419 

0.0290 

0,818 

Model  2  Captures  in  June,  July  and  August 

(Intercept) 

6.05    ±  2.99 

0.0532 

DMARSH 

-0.0355    ±  0.0044 

0.0000 

0.298 

NSPHERB 

0.649    ±  0.153 

0.0002 

0.615 

CLAY 

0.151     ±  0.051 

0.0068 

0.700 

Model  3  Captures  in  September  and  October 

(Intercept) 

1.21     ±  3.54 

0.7358 

DMARSH 

-0.0196    ±  0.0056 

0.0016 

0.121 

NSPHERB 

0.663    ±  0.204 

0.0030 

0.346 

Model  4  Adjusted  total  captures 

(Intercept) 

-6.80    ±  8.03 

0.4053 

DMARSH 

-0.0600    ±  0.0099 

0.0000 

0.351 

NSPHERB 

1.745    ±  0.362 

0.0001 

0.553 

PHOT+ 

-14.859    ±  3.001 

0.0000 

0.607 

PHOT20 

4.100    ±  1.274 

0.0037 

0.693 

PHOT50 

4.804    ±  1.576 

0.0055 

0.769 

Model  5  Adjusted  total  captures* 

(Intercept) 

17.64    ±  11.37 

0,1339 

HGH 

20.307    ±  2.732 

0.0000 

0.196 

LITTHICK 

-6.275    ±    1 .264 

0.0000 

0,362 

PHOT+ 

-14.060    ±  3.081 

0.0001 

0.450 

PHOTCAN 

-5.234    ±  1.374 

0,0009 

0.631 

MBLH 

5.195    ±  2.135 

0,0228 

0.692 

"DMARSH  removed  from  the  model  4. 


the  variables  used  in  the  models  (1  to 
5  and  DFA)  and  all  other  variables 
measured  in  the  field. 


Discussion 

Model-Related  Assumptions 

In  order  to  use  density  (estimated  by 
captures)  as  the  dependent  variable 
in  multivariate  analysis  to  model  sea- 
sonal habitat  structure  selected  by 
leopard  frogs,  certain  assumptions 
must  be  made.  Moreover,  we  cannot 
recommend  the  use  of  the  models 
presented  in  table  8  to  predict  den- 
sity for  leopard  frog  populations  for 
which  the  pattern  of  seasonal  fluctua- 
tion and  causes  of  those  fluctuations 
are  unknown  (Clawson  et  al.  1984, 
Hine  et  al.  1981). 

1.  Density  as  estimated  by  cap- 
ture reflects  density  in  the 
sampled  habitats  as  regards 
to  immigration  or  emigration 
to  or  from  neighboring  habi- 
tats (Collins  and  Wilbur 
1979).  Ram  pipiens  is  known 
to  be  very  mobile  (Merrell 
1977,  Rittschof  1975),  and  is 
capable  of  nocturnal  excur- 
sions of  100  m  or  more  (Dole 
1965a).  Nevertheless,  leopard 
frogs  rarely  move  more  than 
10  m  away  from  their  home 
range,  estimated  by  Dole 
(1965b)  to  vary  between  68 
and  503  m^. 

2.  Favorable  habitats  are  char- 
acterized by  frog  densities 
that  are  higher  than  those  in 
unfavorable  habitats  (Par- 
tridge 1978).  However,  if 
density  is  low  (as  observed 
on  our  study  site  in  1987 
when  compared  to  1986),  all 
favorable  habitats  may  not 
be  occupied  (Partridge  1978). 

3.  Multivariate  analyses  are 
based  on  matrices  of  linear 
correlation  between  environ- 


137 


mental  variables  and  an  in- 
dex of  abundance  (Legendre 
and  Legendre  1984),  which 
neglects  saturation  and  nega- 
tive feedback  effects,  as  well 
as  non-linear  patterns  in  the 
species  response  to  environ- 
mental factors. 

4.  Competition  and  predation 
or  the  presence  of  sites  for 
reproduction  may  control 
frog  distribution  patterns  but 


active  habitat  selection  with 
respect  to  vegetation  struc- 
ture also  plays  an  important 
role.  Dole  (1971)  has  ob- 
served that  newly  metamor- 
phosed young  do  not  neces- 
sarily select  the  first  suitable 
site  during  dispersal. 

Finally,  in  models,  it  is  apparently 
essential  to  assume  that  factors  vital 
for  species  survival,  i.e.  those  vari- 
ables actively  selected  by  individu- 


Figure  6.— Ordination  of  the  stations  in  relation  to  marsti  distance  (DMARSH)  and  number  of 
hierb  species  (NSPHERB).  Stations  withi  thie  same  abundance  class  are  circled  by  an  ellipsoid 


als,  and  those  identified  by  the  analy- 
sis do  not  necessarily  coincide.  In 
fact,  apparent  cause-and-effect  rela- 
tionships are  not  often  testable  and 
require  specific  study  on  the  func- 
tional responses  of  species  to  the  se- 
lected variables.  Weller  (1978)  indi- 
cates that  the  study  of  habitat  stimuli 
as  attractants  for  wildlife  remains  to 
be  done.  The  approach  used  in  this 
study  is  valuable  when  variables  de- 
scribing favorable  habitat  are  re- 
quired (Clark  and  Euler  1982,  Green 
1971,Grier  1984). 


Classification  of  Habitats 

The  PCA  analysis  facilitated  under- 
standing of  the  multidimensional 
models,  and  so  allowed  for  system- 
atic description  of  the  various  habi- 
tats found  in  the  Lac  Saint  Pierre 
floodplain.  We  found  that  our  pre- 
established  groupings  were  not  an 
analytical  artefact  but  rather  con- 
firms that  there  is  a  structure  that  can 
be  defined  by  environmental  vari- 
ables not  related  to  species  specific 
local  vegetation. 

Our  results  have  shown  that  dif- 
ferent age  groups  of  R.  pipiens  are  not 
differently  distributed  among  habi- 
tats (tables  3  and  4).  This  conclusion 
have  been  drawn  with  recently  meta- 
morphosed young  representing  only 
16%  of  total  captures  but  is  sup- 
ported by  others  studies  describing 
the  habitats  used  by  young  (Dole 
1971,  Hine  et  al.  1981,  Rittschof  1975, 
Whitaker  1961).  Our  proposed  mod- 
els are  those  independent  of  age  or 
size  groups.  This  might  not  be  the 
same  however,  for  other  species  as 
Clark  and  Euler  (1982)  and  Roberts 
and  Lewin  (1979)  have  noted  for 
Ram  clamitans  and  for  R.  sylvatica, 
respectively. 

The  models  presented  in  this  pa- 
per reveal  the  importance  of  distance 
to  marsh  line  in  habitat  classification. 
This  variable  has  a  high  degree  of 
predictive  power  as  to  the  extent 
habitat  will  be  utilized  by  leopard 
frogs,  in  the  Lac  Saint  Pierre 


138 


floodplain.  However,  systematic 
sampling  in  habitats  of  unknown 
value  indicates  the  presence  of  a  sig- 
nificant number  of  leopard  frogs  in 
some  wooded  and  shrub  swamps 
stations  far  from  the  marsh  (fig.  6). 
The  DFA  model  is  then  relevant  to 
show  the  importance  of  variables  re- 
lated to  structural  components  of 
habitat  such  as  herbaceous  vegeta- 
tion (PHOT+,  NSPHERB)  and  moss 
cover  (MOSSCOV).  In  a  similar 
analysis  on  Missouri  herpetofauna, 
Clawson  et  al.  (1984)  concluded  that 
proximity  to  water  appeared  to  over- 
ride other  variables  in  determining 
the  abundance  of  amphibians. 

Other  multivariate  studies  (Beebee 
1977,  Clark  and  Euler  1982,  Dale  et 
al.  1985  and  Gascon  and  Planas  1986) 
on  anuran  species  habitat  have 
shown  that  bio-physico-chemical 
variables  related  only  to  the  body  of 
water  cannot  give  predictive  infor- 
mation about  the  absence  or  presence 
of  a  respective  amphibian  species. 

Frog  Abundance  Models 

In  spring,  before  the  growing  season, 
frog  distribution  is  related  to  soil 
characteristics,  such  as  temperature. 
This  variable  is  not  significantly  cor- 
related with  any  other  variable  meas- 
ured. It  results  from  the  interaction 
of  many  variables  and  may  be  a  key 
element  in  habitat  selection  during 
that  period.  The  activity  of  ectoth- 
erms  is  known  to  be  related  to  ambi- 
ent temperatures  (Putnam  and  Ben- 
nett 1981),  by  selecting  warmer  habi- 
tat, ectotherms  might  improve  their 
mobility,  thus  escaping  more  easily 
to  predators.  Soil  moisture  is  the 
third  most  important  variable  in  the 
first  model  and  appears  only  in  this 
model.  In  spring,  soil  moisture  re- 
flects the  speed  of  water  recess  after 
snowmelt  and  obviously  is  a  variable 
linked  with  the  proximity  of  over- 
wintering and  spawning  sites. 

The  model  proposed  for  the  sum- 
mer period  is  the  simplest  of  the 
models  presented  in  this  paper  with 


only  3  descriptors  (DMARSH, 
NSPHERB,  CLAY).  Soil  moisture  is 
not  included  into  this  model  al- 
though it  has  been  shown  to  be  the 
major  factor  limiting  the  distribution 
of  anuran  species  in  terrestrial  habi- 
tats (Clark  and  Euler  1982,  Dole 
1965a,  1971,  Rittschof  1975,  Roberts 
and  Lewin  1979).  It  may  be  that  this 
variable  contains  an  information  al- 
ready carried  in  DMARSH  variable; 
its  presence  in  the  summer  model 
would  then  be  a  redundancy.  Clay, 
on  the  other  hand,  is  a  variable 
known  to  play  an  important  role  in 
soil  water  retention  (Ramade  1984). 

Sampling  during  Fall  migration 
have  shown  a  significant  movement 
towards  aquatic  overwintering  sites. 
Model  3  however,  with  two  variables 
explaining  only  34.6  %  of  frog  abun- 
dance, did  not  allow  identification  of 
preferred  migratory  corridors.  It 
seems  that  leopard  frogs  en  route  to 
overwintering  sites  do  not  select  any 
particular  pathway. 

The  last  two  models  use  data  from 
all  sampling  periods.  Model  4,  which 
improves  on  model  2  (summer 
model),  is  interesting  because  its 
photometric  variables  are  signifi- 
cantly correlated  (appendix  2)  with 
many  of  other  variables  describing 
the  habitat  structure.  This  suggests 
the  value  of  such  indices  (Fox  1979) 
in  habitat  modeling  to  quantify  vege- 
tation structure  since  they  can  be 
measured  with  an  instrument  (light 
meter)  easy  to  use. 

The  last  model,  with  69.2%  vari- 
ability explained,  is  of  more  general 
interest  because  the  local  variable 
DMARSH  has  been  removed.  In 
model  5,  the  importance  of  vegeta- 
tion structure  in  habitat  selection  is 
obvious,  and  the  model  can  be  ap- 
plied to  the  entire  distributional 
range  of  R.  pipiens.  HGH  indicates 
the  importance  of  graminoids 
(grasses,  sedges,  etc.)  usually  abun- 
dant in  open  wetlands.  This  vegeta- 
tion cover  provides  a  refuge  from 
many  predators  and  may  thus  con- 
tribute to  maintaining  an  abundant 
frog  population  (Whi taker  1961).  Lit- 


ter thickness  has  a  negative  coeffi- 
cient in  the  model,  but  is  positively 
correlated  with  HGH,  which  sug- 
gests the  existence  of  an  optimum 
foliage  density.  Dole  (1965b,  1967, 
1971)  mentions  that  litter  may  pre- 
clude direct  contact  betv-^een  the  in- 
dividual and  the  moist  substrate  and 
thus  cause  higher  cutaneous  evapo- 
ration. The  three  other  descriptors 
summarize  the  information  on  vege- 
tation structure.  PHOT+  corresponds 
to  the  presence  of  broad-leaf  herbs  > 
100  cm  tall  (Rp  =  .4066,  P  =  0.026), 
and  graminoids  (Rp  =  .3765,  P  = 
0.040);  PHOTCAN  represents  tree 
and  shrub  cover;  MBLH  indicates 
broad-leaf  plant  obstruction  between 
20  and  100  cm  from  the  ground. 

These  results  seem  to  indicate  that 
vegetation  structure,  more  than  spe- 
cific species  composition,  is  an  im- 
portant factor  in  habitat  selection  for 
Rana  pipiens.  This  finding  is  similar 
to  that  of  MacArthur  and  MacArthur 
(1961)  who  have  demonstrated  that 
bird  species  occupying  forests  and 
prairies  choose  their  habitat  on  the 
basis  of  foliage  density  at  different 
levels  from  the  ground,  irrespective 
of  plant  species  composition. 

Conclusion 

In  summary,  we  present  three  types 
of  complementary  analysis  dealing 
with  wet  habitats  used  by  the  leop- 
ard frog  during  Summer.  First,  a 
PCA  gives  a  qualitative  description 
of  five  kinds  of  habitats  typical  to  the 
St.  Lawrence  river  floodplain  and 
offering  potential  supports  to  leop- 
ard frog  populations.  Second,  a  DFA 
model  with  four  easily  measured 
variables  allows  classification  of 
habitats  into  three  groups  of  frog 
abundance.  This  is  a  very  helpful 
way  to  map  potential  frog  species 
habitats  for  protective  purpose.  Fi- 
nally, five  regression  models  (accord- 
ing to  each  phenological  periods  or 
whole  active  season)  explain  frog 
abundance  variations  with  only  a 
few  important  structural  variables. 


139 


Although  the  models  described  in 
this  paper  cannot  fully  demonstrate 
functional  relationships  between 
model  variables  and  frog  density, 
suitable  modifications  of  some  of 
these  variables  (litter  thickness,  for 
instance)  may  increase  frog  popula- 
tion. Refinement  of  these  models  will 
require  experimental  studies  on  func- 
tional responses  of  leopard  frogs  to 
specific  habitat  features. 

Acknowledgments 

Thanks  are  due  to  Benoit  Levesque, 
Sylvain  Cote  and  Jean-Louis  Benoit 
for  field  assistance,  to  Bernard 
Robert  for  graphical  art,  to  Gille 
Houle  for  the  English  version  of  the 
text  and  to  Marc  P.  Hayes  and  Gary 
K.  Meefe  for  their  very  constructive 
comments  on  the  first  draft  of  this 
paper.  Financial  support  came 
through  grants  to  N.B.  from  National 
Research  Council  of  Canada,  Cana- 
dian Wildlife  Federation,  and  Centre 
d'Etude  Universitaire  (Quebec)  and 
from  direct  funding  from  Ministere 
Quebecois  du  Loisir,  de  la  Chasse  et 
de  la  Peche,  and  Universite  du 
Quebec  a  Trois-Rivieres. 


Literature  Cited 

Beebee,  Trevor  J.C.  1977.  Habitats  of 
the  British  amphibians  (l):Chalk 
uplands.  Biological  Conservation 
12:179-193. 

Bouyoucos,  G.J.  1936.  Directions  for 
making  Mechanical  Analysis  of 
Soils  by  the  Hydrometer  Method. 
Soil  Science  32:225-228. 

Braun-Blanquet,  Josias.  1964.  Plant 
sociology.  McGraw-Hill,  New 
York.  865  p. 

Bury,  R.  Bruce  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles,  p. 
416-419  in  J.  F.  Bell  and  T.  Atter- 
bury,  eds.  Renewable  resource  in- 
ventories for  monitoring  changes 
and  trends.  Oregon  state  Univer- 
sity, Corvallis. 


Campbell,  Howard  W.,  and  Steven  P. 
Christman.  1982.  Field  techniques 
for  herpetofaunal  community 
analysis,  p.  193-200;  in  Scott 
N.J.Jr.(ed.),  Herpetological  com- 
munities. United  States  Depart- 
ment of  the  interior.  Fish  Wildlife 
Service,  Wildlife  Research  Report 
13.  239p. 

Clark,  Karen  and  David  Euler.  1982. 
The  importance  of  pH  and  habitat 
disturbance  in  amphibian  distribu- 
tions in  Central  Ontario.  Technical 
Report  Wildlife  Habitat  compo- 
nent, Lakeshore  Capacity  Study, 
Ontario  Ministry  of  Natural  Re- 
sources, Toronto,  Ontario. 

Clawson,  Mary  E.  and  Thomas  S. 
Baskett.  1982.  Herpetofauna  of  the 
Ashland  Wildlife  Area,  Boone 
County,  Missouri.  Transactions  of 
the  Missouri  Academy  Science 
16:5-16. 

Clawson,  Mary  E.,  Thomas  S.  Baskett 
and  Michael  J.  Armbruster.  1984. 
An  approach  to  habitat  modeling 
for  herpetofauna.  Wildlife  Society 
Bulletin  12:61-69. 

Collins,  James  P.  and  Henry  M. 
Wilbur.  1979.  Breeding  habits  and 
habitats  of  the  amphibians  of  the 
Edmin  S.  George  Reserve,  Michi- 
gan with  notes  on  the  local  distri- 
bution of  fishes.  Occasional  Papers 
of  the  Museum  of  Zoology  Uni- 
versity of  Michigan  no  686.  34  p. 

Cook,  Francis  R.  1984.  Introduction 
aux  amphibiens  et  reptiles  du 
Canada.  Musees  Nationaux  du 
Canada,  Ottawa,  211  p. 

Denis  Jacques,  Enr.  1986.  Vegetation 
des  terres  humides  du  lac  St-Pi- 
erre.  Cartographic  1:10  000.  Min- 
istere du  Loisir,  de  la  Chasse  et  de 
la  Peche  du  Quebec,  Corporaton 
pour  la  mise  en  valeur  du  lac  St- 
Pierre.  Feuillet  7  a  12. 

Dale,  J.M.,  B.  Freedman  and  J.  Kere- 
kes.  1984.  Acidity  and  associated 
water  chemistry  of  amphibian 
habitats  in  Nova  Scotia.  Canadian 
Journal  of  Zoology  63:97-105. 

Degraaf,  R.M.  and  D.D.  Rudis.  1983. 
Amphibians  and  reptiles  of  New 
England  Habitats  and  natural  his- 


tory. The  University  of  Massachu- 
setts. 85  p. 

Dole,  Jim  W.  1965a.  Spatial  relations 
in  natural  populations  of  leopard 
frog,  Rana  pipiens  Schreber,  in 
northern  Michigan.  American 
Midland  Naturalist  74:464-478. 

Dole,  Jim  W.  1965b.  Sommer  move- 
ments of  adult  leopard  frogs,  Rana 
pipiens  Schreber,  in  northern 
Michigan.  Ecology  46(3):236-255. 

Dole,  Jim  W.  1967.  Spring  move- 
ments of  leopard  frogs,  Rana  pipi- 
ens Schreber,  in  northern  Michi- 
gan. American  Midland  Naturalist 
78(1):167-181. 

Dole,  Jim  W.  1971.  Dispersal  of  re- 
cently metamorphosed  leopard 
frogs,  Rana  pipiens.  Copeia 
1971(2):221-228. 

Fox,  Barry  J.  1979.  An  objective 
method  of  measuring  the  vegeta- 
tion structure  of  animal  habitats. 
Australian  Wildlife  Research 
6:297-303. 

Frier,  Jo  Ann  and  Robert  T.  Zap- 
palorti.  1984.  Reptile  and  amphib- 
ian management  techniques. 
Transactions  of  the  N.E.  Section 
Wildlife  Society  40:142-148. 

Frontier,  Serge.  1976.  Etude  de  la 
decroissance  des  valeurs  propres 
dans  une  analyse  en  composantes 
principales:  comparaison  avec  le 
modele  du  baton  brise.  Journal  of 
experimental  Marine  Biology  and 
Ecology  25:67-75. 

Froom,  Barbara  1982.  Amphibians  of 
Canada.  McClelland  and  Stewart, 
Toronto.  120  p. 

Gascon,  Claude  and  Dolors  Planas. 
1986.  Spring  pond  water  chemis- 
try and  the  reproduction  of  the 
wood  frog,  Rana  sylvatica.  Cana- 
dian Journal  of  Zoology  64:543- 
550. 

Gibbs,  E.L.,  G.W.  Nace,  and  M.B. 
Emmons.  1971.  The  life  frog  is  al- 
most dead.  Bioscience  21:1027- 
1034. 

Green,  Roger  H.  1971.  A  multivariate 
statistical  approach  to  the  Hutch- 
insonian  niche:  Bivalve  molluscs 
of  central  Canada.  Ecology,  52 
(4):543-556. 


140 


Green,  Roger  H.  1979.  Sampling  de- 
sign and  statistical  methods  for 
environmental  biologists.  Wiley 
Inter-science  Pub.,  Wiley  and 
Sons,  Inc.  257  p. 

Grier,  James.W.  1984.  Biology  of  ani- 
mal behavior.  Times  Mirror/ 
Mosby  College  Publishing,  St. 
Louis. 

Hayes,  Marc  P.  and  Mark  R.  Jen- 
nings. 1986.  Decline  of  Ranid  Frog 
Species  in  Western  North  Amer- 
ica: Are  Bullfrog  (Ram  catesbeiam) 
responsible?  Journal  of  Herpetol- 
ogy  20(4):490-509. 

Hine,  Ruth  L.,  Betty  L.  Les  and  Bruce 
F.  Hellmich.  1981.  Leopard  frog 
populations  and  mortality  in  Wis- 
consin 1974-76.  Technical  Bulletin 
No.  122,  Department  of  Natural 
Resources,  Wisconsin.  39  p. 

Leclair,  Raymond  Jr.  1983.  Utilisation 
de  differents  types  d'habitats  par 
la  grenouille  leopard  pour  fin  de 
reproduction  dans  la  region  de 
Bainville  au  lac  St-Pierre.  Rapport 
technique,  Ministere  du  Loisirs,  de 
la  Chasse  et  de  la  Peche  du 
Quebec.  24  p. 

Leclair,  Raymond  Jr.  1985.  Les  am- 
phibiens  du  Quebec:  Biologie  des 
especes  et  problematique  de  con- 
servation des  habitats.  Ministere 
du  Loisirs,  de  la  Chasse  et  de  la 
Peche  du  Quebec.  119  p. 

Leclair,  Raymond  Jr.  and  L.  Bar- 
ibeau.  1982.  Problematique  des 
anoures  au  Quebec.  Ministere  du 
Loisirs,  de  la  Chasse  et  de  la  Peche 
du  Quebec.  69  p. 

Leclair,  R.  Jr.,  and  J.  Castanet.  1987. 
A  skeletochronological  assessment 
of  age  and  growth  in  the  Frog 
Ram  pipiens  Schreber  (Amphibia, 
Anura)  form  Southwestern  Que- 
bec. Copeia  1987(2):361-369. 

Legendre,  Pierre  and  Louis  Legen- 
dre.  1984.  Ecologie  numerique  (2^ 
ed.).  Tome  1:  Le  traitement  mul- 
tiple des  donnees  ecologiques; 
Tome  2:  La  structure  des  donnees 
ecologiques.  Presses  de  I'Univer- 
site  du  Quebec  et  Masson  Cie.  260; 
335  p. 


MacArthur,  Robert  H.,  and  John  W. 
MacArthur.  1961.  On  bird  species 
diversity.  Ecology.  42:594-598. 

Marcotte,  Alice.  1981.  L'exploitation 
des  grenouilles  au  Quebec.  Rap- 
port technique,  Ministere  du 
Loisirs,  de  la  Chasse  et  de  la  Peche 
du  Quebec,  Service  de  I'Amenage- 
ment  de  la  Faune,  Memoire  Uni- 
versite  du  Quebec  a  Montreal. 
75  p. 

Merrell,  David  J.  1977.  Life  history  of 
the  leopard  frog.  Ram  pipiens,  in 
Minnesota.  University  of  Minne- 
sota Bell  Museum  Natural  History 
Occasionnal  Paper  15:  23  p. 

MLCP.  1985.  Des  habitats  fauniques 
a  conserver.  CoUoque  national  sur 
la  faune.  Ministere  du  Loisirs,  de 
la  Chasse  et  de  la  Peche  du 
Quebec.  81  p. 

Nie,  Norman  H.,  C.  Hadlai  Hull,  Jean 
G.  Jenkins,  Karin  Steinbrenner  and 
Dale  H.  Bent.  1975.  Statistical 
package  for  the  social  sciences. 
McGraw-Hill,  New-York.  675  p. 

Paquin,  Diane.  1982.  L'exploitation 
de  trois  expeces  d'anoures  dans  la 
region  du  lac  St-Pierre.  Rapport 
technique,  Ministere  du  Loisirs,  de 
la  Chasse  et  de  la  Peche  du 
Quebec.  64  p. 

Partridge,  Linda.  1978.  Habitat  selec- 
tion. Pages  351-376  in  J.R.  Krebs 
and  N.B.  Davies.  (ed.).  Behavioral 
ecology:  an  evolutionary  ap- 
proach. Blackwell  Scientific  Publi- 
cations, Oxford. 

Putnam,  Robert  W.  and  Albert  F. 
Bennett.  1981.  Thermal  depend- 
ence of  behavioral  performance  of 
anuran  amphibians.  Animal  be- 
haviour. 29(2):502-509. 

Ramade,  Francois.  1984.  Elements 
d'ecologie;  ecologie  fondamentale. 
McGraw-Hill,  Paris.  397  p. 

Rittschof,  Daniel  1975.  Some  aspects 
of  the  natural  history  and  ecology 
of  the  leopard  frog.  Ram  pipiens. 
University  of  Michigan,  Ann  Ar- 
bor. Ph.  D.  Thesis.  212  p. 

Roberts,  Wayne  and  Victor  Lewin. 
1979.  Habitat  utilization  and  pop- 
ulation densities  of  the  amphibi- 


ans of  northeastern  Alberta.  Cana- 
dian Field-Natur.  93(2):144-154. 

Sarrazin,  R.,  M.  Cantin,  A.  Gagnon, 
C.  Gauthier  and  G.  Lefebvre.  1983. 
La  protection  des  habitats 
fauniques  au  quebec.  Gouv.  du 
Quebec.  Dir.  Gen.  Faune.  Min. 
Loisir  Chasse  et  Peche.  256  p. 

Scherrer,  B.  1984.  Biostatistique. 
Gaetan  Morin  (ed.).  Chicoutimi, 
Quebec.  850  p. 

Service  Canadien  De  La  Faune.  1980. 
Plan  de  gestion  des  oiseaux  aqua- 
tiques  au  Canada,  Environnement 
Canada,  26  p. 

Tessier,  Claude  and  Paul  Caron. 
1980.  Cartographie  ecologique  de 
la  vegetation  de  la  rive  nord  du  lac 
Saint-Pierre.  Rapport  technique 
Service  de  I'amenagement  et  de 
l'exploitation  de  la  faune,  Min- 
istere du  Loisirs,  de  la  Chasse  et 
de  la  Peche  du  Quebec.  19  p.  + 
carte. 

Tessier,  Claude,  Antoine  Aubin  and 
Daniele  Chenevier.  1984.  Les 
facteurs  elevation  et  distance  dans 
la  structure  et  la  productivite 
d'une  communaute  riveraine  de 
macrophytes.  Canadian  Journal  of 
Botany  62:2260-2266. 

Vogt,  Richard  C.  and  Ruth  K.  Hine. 
1982.  Evaluation  of  techniques  for 
assessment  of  amphibian  and  rep- 
tile populations  in  Wisconsin,  p. 
201-217;  in  Scott,  N.J.Jr.  (ed.).  Her- 
petological  communities.  United 
States  Department  of  the  interior. 
Fish  Wildlife  Service,  Wildlife  Re- 
search Report  13.  239p. 

Weller,  M.W.  1978.  Management  of 
freshwater  marshes  for  wildlife,  p. 
267-284  in  Freswater  wetlands; 
ecological  processes  and  manage- 
ment potential.  New  York,  365  p. 

Whitaker,  John  0.,Jr.  1961.  Habitat 
and  food  of  mousetrapped  young 
Ram  pipiens  and  Ram  clamitans. 
Herpetologica  17(3):173-179. 


141 


Appendix  1. 


Abbreviations  for  variables  used  in  \he  text,  figures  and  tables. 


Abbreviations 

Variables 

Abbreviations 

Variables 

Abbreviations 

Variables 

DOPEN 

Distance  to  nearest 

^  BASAREA 

Basal  area 

PHonoo* 

Photometric  index, 

open  habitat 

MINAREA,, 
PH 

Minimal  area 

between  50  et  100 

DCLOSE 

Distance  to  nearest 

pH  of  soil  solution 

cm 

closed  habitat 

SAND 

Sand  fraction  in  soil 

PHOT+* 

Photometric  index. 

DMARSH 

Distance  to  marsh 

SILT 

Silt  fraction  in  soil 

for  herbs  >  100  cm 

line 

CLAY 

Clay  fraction  in  soil 

PHOTCAN* 

Photometric  index 

DWATERP 

Distance  to  nearest 
permanent  pool 

DWATER* 

Distance  to  nearest 
temporary  pool 

under  shrub  and 
tree  strata 

DHUMAIN 

Distance  to  nearest 

MOIST* 

%  soil  moisture 

HERBHGHT* 

Height  of  herb  stra- 

human artefact 

TEMP* 

Temperature  at  the 

tum 

ALTREL 

Altitude  relative  to 

soil  surface 

HERBCOV* 

%  herb  cover 

shore  line 

WTABLE* 

%  bare  ground 

HGH* 

Cover  class  for  high 

NSPHERB 

Number  of  herba- 

BAREGRNDj^* 

Water  table  level 

graminoid  herbs  ( > 

ceous  species 

LITTER* 

%  ground  covered 

100  cm  tall) 

NSPSHRUB 

Number  of  shrub 

with  litter 

HBLH* 

Cover  class  for  high 

species 

MOSSCOVj^* 

%  moss  cover 

broad-leaf  herbs 

NSPTREE 

Number  of  tree  spe- 

DEADWOOD* 

%  ground  covered 

MGH/ 

Cover  class  for  me- 

cies 

by  dead  wood 

dium  graminoid 

TREE 

Cover  class  for  tree 

LITTHICK* 

Litter  thickness 

herbs  (20  to  100  cm) 

stratum 

PHOT20* 

Photometric  index. 

MBLH* 

Cover  class  for  me- 

SHRUBHI 

Cover  class  for 
High  shrub  stratum 

between  0  and  20 
cm 

dium  broad-leaf 
herbs 

SHRUBLO 

Cover  class  for  lov^ 
shrub  stratum 

PHOT50* 

Photometric  index, 
between  20  and  50 

SMALLH* 

Cover  class  for 
herbs  <  20  cm  tall 

cm 


N:  Variable  normalized  by  square-root  transformation. 
':  Variable  measured  monthly. 


142 


Appendix  2. 


Significance  levels  of  Pearson  rank  correlations  between  ttie  variables  included  in  ttie  nnodels  and  all  vari- 
ables measured.  (Significance  levels  1 :  P  <  0.05;  2:  P  <  0.01 ;  3:  P  <  0.001 ;  4:  P  <  0.0001 ;  +:  positive;  negative.) 


4 


Models 

1 

2 

4 

5 

D 

D 

T 

M 

P 

s 

B 

N 

C 

P 

P 

P 

H 

L 

P 

M 

M 

M 

E 

o 

H 

I 

A 

s 

L 

H 

H 

H 

G 

I 

H 

B 

o 

A 

M 

I 

o 

L 

R 

P 

A 

o 

o 

o 

H 

T 

o 

L 

s 

R 

P 

s 

T 

T 

E 

H 

Y 

T 

T 

T 

T 

T 

H 

s 

s 

T 

5 

G 

E 

+ 

2 

5 

H 

c 

c 

H 

0 

R 

R 

0 

0 

I 

A 

o 

N 

B 

c 

N 

V 

D 

K 

DOPEN 

-1 

-1 

-2 

-1 

+1 

+1 

DCLOSE 

+1 

DMARSH 

-1 

-1 

DHUMAN 

ALTREL 

+3 

-1 

+1 

NSPHERB 

NSPSHRUB 

+3 

-1 

-1 

-1 

-1 

+1 

+3 

NSPTREE 

+2 

-1 

-1 

-2 

-1 

-3 

-1 

+1 

+2 

+2 

BASAREA 

+2 

-2 

-2 

-3 

-2 

+1 

+2 

+1 

MINAREA 

-1 

-3 

+2 

+2 

PH 

SAND 

-2 

-4 

-1 

SILT 

-2 

CLAY 

BAREGRND 

-2 

+1 

+1 

LITTER 

-2 

+1 

+1 

+2 

-2 

MOSSCOV 

-1 

-2 

-1 

DEADWOOD 

-1 

-1 

+1 

LITTHICK 

-4 

+2 

+3 

-2 

-2 

-3 

TREE 

+1 

-1 

-1 

-1 

-3 

-2 

+2 

+2 

SHRUBHI 

-2 

-1 

-1 

-2 

+3 

+1 

SHRUBLO 

-1 

_i 

+2 

+2 

MOIST 

-1 

+1 

-1 

-1 

TEMP 

PHOT20 

+1 

+1 

+2 

+2 

-1 

PHOT50 

+2 

+1 

+2 

-1 

PHOTIOO 

-1 

-1 

+1 

-2 

PHOT+ 

+1 

+2 

+1 

-2 

PHOTCAN 

-1 

-2 

-1 

-1 

-2 

+1 

HERBHGHT 

+1 

+3 

+1 

-2 

WTABLE 

+3 

D.WATER 

HERBCOV 

+1 

-2 

-2 

HGH 

-1 

+1 

+2 

+2 

+3 

-1 

-2 

-2 

HBLH 

-1 

+1 

+1 

MGH 

+2 

MBLH 

-1 

-2 

SMALLH 

-2 

-2 

D:  DM  mode/. 


143 


Habitat  Correlates  of 
Distribution  of  the  California 
Red -Legged  Frog  (Rana 
aurora  draytoriii)  and  the 
Foothill  Yellow-Legged  Frog 
(Rana  boylii):  Implications  for 
Management^ 

Marc  P.  Hayes  and  Mark  R.  Jennings^ 


Abstract.— We  examined  features  of  the  habitat 
for  the  California  red-legged  frog  and  foothill  yellow- 
legged  frog  from  the  Central  Valley  of  California. 
Limited  overlap  exists  in  habitat  use  between  each 
frog  species  and  introduced  aquatic  macrofaunal 
predators.  Temporal  data  implicate  aquatic  preda- 
tors that  restrict  red-legged  frogs  to  intermittent 
stream  habitats  as  explaining  limited  overlap.  Identi- 
fication of  responsible  predators  is  currently  pre- 
vented because  the  alternative  of  limited  overlap 
simply  due  to  differential  habitat  use  between  frogs 
and  any  one  putative  predator  cannot  be  rejected. 
Until  the  predators  causing  the  negative  effects  are 
identified,  efforts  should  be  made  to  isolate  these 
frogs  from  likely  predators  and  minimize  alteration  of 
key  features  in  frog  habitat. 

Wright  1920).  Despite  this  history  of 
exploitation,  few  attempts  have  been 
made  to  link  species-specific  habitat 
requirements  of  ranid  frogs  to  their 
management  (but  see  McAuliffe 
1978;  Treanor  1975a,  b;  Treanor  and 
Nicola  1972).  Most  ''management" 
literature  has  either  simply  reviewed 
the  biolog}''  of  selected  ranid  frog 
species  or  indicated  vulnerable  life 
history  stages  needing  study  (Baker 
1942,  Bury  and  Whelan  1984,  Storer 
1933,  Willis  et  al.  1956,  Wright  1920). 

In  this  report,  we  examine  the 
habitat  features  of  two  "non-game" 
species,  the  California  red-legged 
frog  (Rana  aurora  draytonii)  and  the 
foothill  yellow-legged  frog  (Rana 
boylii),  two  ranid  frogs  found  in  low- 
land California.  Each  species  has  dis- 
appeared from  sizable  areas  of  its 
historic  range  (Hayes  and  Jennings 
1986,  Sweet  1983).  Although  histori- 
cal disappearance  of  red-legged  frogs 
has  been  linked  to  its  exploitation  as 
food  (Jennings  and  Hayes  1985), 
causal  factors  in  the  continuing  de- 
cline of  both  species  remain  poorly 
understood.  Insufficient  documenta- 
tion of  the  habitat  requirements  of 
each  species  has  especially  impeded 
identification  of  the  causes  of  decline 
(Hayes  and  Jennings  1986).  In  this 
report,  we  reduce  this  gap  by  identi- 
fying the  habitat  requirements  that 
characterize  each  frog.  We  then  use 
these  data  to  suggest  the  direction 
for  management  of  these  two  species 


The  application  of  habitat 
analysis  to  management  has  a 
long,  complex  history.  The  Greek 
philosopher  Aristotle  inferred  that 
seasonal  variation  in  the  distribu- 
tion of  certain  commercially  ex- 
ploited fishes  was  related  to  changes 
in  their  food  resources  and  habitat 
temperatures  (Cresswell  1862).  In  the 
13th  century,  the  Mongol  emperor 
Kublai  Khan  encouraged  the  gather- 
ing of  data  on  foraging  patterns  of 
sport-hunted  birds  to  facilitate  ma- 
nipulating their  populations  (Leo- 
pold 1931).  Since  these  efforts,  many 
individuals  have  used  diverse  habitat 
data  to  help  understand  factors  that 
influence  the  distribution  and  success 
of  various  species.  Most  often,  such 
data  have  been  used  to  address  com- 
mercially important  or  game  species, 
usually  to  identify  management  al- 
ternatives intended  to  enhance  exist- 
ing populations  or  avert  population 
declines  (Bailey  1984,  Leopold  1933). 
This  emphasis  has  resulted  in  most 
studies  addressing  selected  birds, 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles  and  Small 
Mammals  in  North  America.  (Flagstaff.  AZ. 
July  19-21.  1988.) 

^Environmental  Scientist,  Gaby  &  Gaby, 
Inc..  6832  SW  68th  Street,  Miami,  FL  33 143- 
3 1 15  and  Department  of  Biology,  P.O.  Box 
249118,  University  of  Miami,  Coral  Gables. 
FL  33 124-91 18;  Research  Associate,  Depart- 
ment of  Herpetology,  California  Academy 
of  Sciences,  Golden  Gate  Park,  San  Fran- 
cisco. CA  94118-9961. 


fishes,  and  large  mammals.  In  con- 
trast, species  historically  having  lim- 
ited economic  importance  (i.e.,  "non- 
game"  species)  have  been  largely  ne- 
glected (Bury  1975;  Bury  et  al.  1980a, 
b;  Pister  1976).  Only  over  the  last  15 
years  has  an  appreciation  been 
broadly  realized  that  non-game  spe- 
cies are  also  in  need  of  management. 
Non-game  species  are  often  linked  to 
economically  important  ones,  and  as 
such,  provide  significant  direct  and 
indirect  benefits  to  humans  (Kellert 
1985,  Neill  1974).  Although  this  ap- 
preciation has  led  to  greater  empha- 
sis in  their  study  (Bury  et  al.  1980a, 
Pister  1976),  a  broader  understand- 
ing of  the  biology  of  non-game  spe- 
cies is  increasingly  urgent  because  of 
widespread  habitat  modification  in- 
fluencing declines  among  ever-great- 
er numbers  of  such  species  (Dodd 
1978,  Hayes  and  Jennings  1986,  Hine 
et  al.  1981,  Honegger  1981). 

Amphibians  are  prominent  among 
groups  of  organisms  given  a  non- 
game  label  (Bury  et  al.  1980a).  For 
ranid  frogs,  among  the  most  familiar 
of  amphibian  groups,  non-game  is 
really  a  misnomer  (Brocke  1979)  be- 
cause they  have  a  history  of  human 
exploitation  which  has  its  roots  in 
European  and  aboriginal  cultural  tra- 
ditions (Honegger  1981,  Zahl  1967) 
and  has  included  significant  com- 
mercial enterprises  (Abdulali  1985, 
Chamberlain  1898,  Husain  and  Rah- 
man 1978,  Jennings  and  Hayes  1985, 


144 


until  experiments  can  identify  the 
causes  of  decline. 


METHODS 

Our  analysis  draws  upon  two  data 
sets,  one  addressing  R.  a.  draytonii 
and  the  other,  R.  hoylii.  The  former  is 
based  on  all  known  occurrences  of  R. 
a.  draytonii  (n  =  143)  from  the  Central 
Valley  of  California,  which  we  define 
as  the  collective  drainage  area  of  the 
Kaweah,  Kern,  Sacramento-San 
Joaquin  (to  Carquinez  Strait),  and 
Tule  River  systems.  We  assembled 
these  data  from  museum  records  and 


field  notes  or  direct  observations  of 
the  many  investigators  listed  in  the 
acknowledgments  or  whose  data  are 
cited  in  Childs  and  Howard  (1955), 
Cowan  (1979),  Fitch  (1949),  Grinnell 
and  Storer  (1924),  Grinnell  et  al. 
(1930),  Hallowell  (1854, 1859),  Ingles 
(1932a,  b;  1933;  1936),  Storer  (1925), 
Walker  (1946),  Williamson  (1855), 
and  Wright  and  Wright  (1949).  We 
used  records  not  authenticated  by 
museum  specimens  if  they  were  cor- 
roborated by  at  least  two  sources. 
We  then  determined  the  subset  (n  = 
131)  of  records  that  could  be  both 
mapped  (i.e.,  where  we  could  iden- 
tify the  aquatic  system  likely  to  be 


Table  1  .—Habitat  variables  recorded  for  the  California  red-legged  frog 
(Rana  aurora  draytor)ll)  data  set.  Subset  scored  refers  to  the  subset  of  lo- 
calities for  which  we  were  able  to  score  each  variable.  Percent  scored  re- 
fers to  the  percentage  of  the  entire  data  set  (n  =  143)  for  which  we  were 
able  to  score  each  variable.  See  text  regarding  further  details  concerning 
the  method  of  data  collection  for  each  variable. 


Variable 


Subset  scored  %  scored 

(n=) 


DefirtJtion 


1. 

Habitat  type 

140 

98 

As  (1)  stream  or  (2)  pond 

2. 

Temporal  status 

137 

96 

As  (1)  perennial  or  (2)  inter- 
mittent 

3. 

Drainage  area 

129 

90 

In  km^ 

4. 

Local  gradient 

139 

97 

In  angular  degrees  C)  from 
horizontal 

5. 

Water  depth 

74 

52 

As  (1)  presence  or 

(2)  absence  of  water 
>0.7  m  deep 

6. 

Vegetation  matrix 
(emergent  or  shoreline) 

44 

31 

As  (1)  dense  (area  >25% 
thickly  vegetated) 

(2)  limited  (some,  but 
<25%  of  area) 

(3)  absent 

7. 

Native  fishes 

56 

39 

As  (1)  present  or  (2)  absent 

8. 

Introduced  fishes 

32 

22 

As  (1)  present  or  (2)  absent 

9. 

Introduced  bullfrogs 

115 

80 

As  (1)  present  or  (2)  absent 

10. 

Substrate  alteration 

113 

79 

As  (1)  present  or  (2)  absent 

n. 

Vegetation  reduction 

106 

74 

As  (1)  present  or  (2)  absent 

12. 

Stream  order 

127 

89 

As  defined  by  Strahler 

(1957) 


the  site  of  origin  of  the  source  popu- 
lation upon  which  the  record  was 
based),  and  identified  as  being  from 
different  "point"  localities  i>OA  km 
apart).  Although  our  data  set  was 
developed  primarily  from  this  sub- 
set, we  used  a  few  data  from  the  re- 
maining 12  localities  for  the  habitat 
variables  described  below.  We  used 
this  additional  data  because  they 
were  either  available  with  the  origi- 
nal records  or  could  be  determined 
independent  of  accurate  mapping. 

For  each  locality,  we  recorded  as 
many  of  12  habitat  variables  as  pos- 
sible (table  1).  For  aquatic  habitat 
type,  we  used  the  term  "stream"  for 
localities  with  both  a  well-defined 
drainage  inflow  and  outflow, 
whereas  we  used  "pond"  for  locali- 
ties lacking  a  well-defined  inflow  and 
little  or  no  outflow.  Temporal .  .atus 
of  the  aquatic  habitat  was  scored  as 
perennial  or  intermittent  based  on 
7.5'and  15'  United  States  Geological 
Survey  (USGS)  topographic  maps, 
but  the  status  of  some  localities  was 
modified  based  on  field  reconnais- 
sance or  data  provided  by  other  in- 
vestigators. For  many  localities,  lack 
of  change  in  the  temporal  status  of 
the  aquatic  habitat  during  the  time  R. 
a.  draytonii  was  recorded  was  veri- 
fied by  examining  USGS  topographic 
maps  bracketing  the  frog  record 
date(s).  We  used  the  designation 
intermittent  to  describe  the  interrup- 
tion of  surface  flow  in  streams  or 
complete  dry-down  in  ponds,  either 
occurring  at  least  once  seasonally. 
Drainage  area  indicates  the  size  of 
the  hydrographic  basin  influencing 
the  recorded  locality.  The  drainage 
area,  local  gradient,  and  stream  or- 
der were  largely  estimated  from  7.5' 
USGS  topographic  maps.  We  esti- 
mated large  drainage  areas  (>130 
km^)  by  extrapolation  to  the  recorded 
locality  on  topographic  maps  from 
either  the  drainage  area  for  the  near- 
est upstream  gauging  station  (United 
States  Geological  Survey  1970a,  b)  or 
section  counts  on  United  States  For- 
est Service  and  county  maps.  Local 
gradient  was  estimated  from  map 


145 


distances  of  0.5-1.0  km  across  the  re- 
corded locality  except  in  the  few 
cases  where  pronounced  local  relief 
required  reduction  of  this  distance 
for  an  accurate  estimate. 

Data  for  the  remaining  variables 
(water  depth,  vegetation  matrix,  na- 
tive and  introduced  fishes,  intro- 
duced bullfrogs  [Ram  catesbeiam], 
substrate  alteration,  and  vegetation 
reduction)  were  obtained  for  subsets 
of  the  larger  data  set  from  the 
sources  indicated  earlier  supple- 
mented by  Leidy  (1984),  Moyle  and 
Nichols  (1973),  Moyle  et  al.  (1982), 
and  Rutter  (1908).  The  exact  values 
used  to  partition  water  depth  and 
vegetation  matrix  variables  are  arbi- 
trary. However,  we  chose  their  gen- 
eral dimensions  with  the  intent  of 
identifying  whether  the  habitat  re- 
quirements of  red-legged  frogs  sug- 
gested by  anecdotal  data  (moderately 
deep  water  associated  with  dense 
vegetation;  see  Hayes  and  Jennings 
1986)  were  supported  by  this  data 
set.  Variation  in  the  collective  data 
set  required  scoring  the  fish  and  in- 
troduced bullfrog  data  as  presence/ 
absence,  but  we  also  used  available 
data  on  which  fish  species  were  pres- 
ent to  interpret  the  habitat  require- 
ments of  red-legged  frogs.  Substrate 
alteration  and  vegetation  reduction 
variables  indicate  alteration  of 
aquatic  habitats  that  was,  directly  or 
indirectly,  human-effected.  We 
scored  substrate  alteration  as  present 
if  evidence  existed  that  the  shoreline 
or  substrate  topography  of  the 
aquatic  habitat  had  been  markedly 
altered  (e.g.,  dams,  rip-rap,  bank- 
trampling  by  cattle).  Marked  altera- 
tion meant  that  at  least  25%  of  the 
area  of  substrate  of  a  locahty  ap- 
peared altered.  We  scored  vegetation 
as  being  reduced  when  data  indi- 
cated that  at  least  25%  of  pre-existing 
shoreline  or  emergent  vegetation  had 
been  removed. 

We  also  gathered  current  data  on 
a  subset  of  the  described  localities 
through  field  reconnaissance  and 
some  information  provided  by  others 
(data  gathered  during  the  interval 


1980-1987  represented  ''current'' 
data).  We  used  these  data  to  help 
identify  temporal  changes  that  may 
have  occurred  at  sites  or  in  drainage 
systems  for  which  we  had  historical 
data.  For  this  analysis,  we  used 
"drainage  system"  to  mean  only  the 
primary  and  highest-order  (fide 
Strahler  1957)  secondary  tributaries 
of  the  Sacramento-San  Joaquin  drain- 
age system.  These  data  were  particu- 
larly important  for  indicating  where 
red-legged  frogs  were  probably  ex- 
tinct. 

The  data  set  addressing  R.  hoylii 
consists  of  data  published  by  Moyle 
(1973)  and  Moyle  and  Nichols  (1973) 
from  which  we  re-examined  selected 
elements.  Collection  methods  for 
these  data  are  thoroughly  described 
therein.  Our  reanalysis  used  most  of 
the  variables  described  by  Moyle 
(1973)  with  some  modifications.  We 
used  the  original  estimates  of  the 
numbers  of  each  fish  species  rather 
than  the  coded  values;  the  numbers 
of  yellow-legged  frogs  and  bullfrogs 
remained  coded  because  the  original 
data  were  recorded  as  coded. 
Moyle' s  stream  type  variable  was 
reduced  to  two  categories  by  com- 
bining his  three  intermittent  and 
three  perennial  stream  categories. 
We  also  added  two  variables,  one 
which  combines  Moyle's  cobble  and 
boulder/bedrock  substrate  catego- 
ries. The  other  describes  the  stream 
morphology  category  designated  in 
Moyle's  original  data  as  smooth  wa- 
ter and  fits  the  definition  of  a  run 
(Armour  et  al.  1983).  For  correlations 
between  yellow-legged  frogs  and 
other  species,  we  used  only  the  sub- 
set of  localities  where  either  or  both 
of  yellow-legged  frogs  and  the  spe- 
cies being  compared  was  present. 

We  re-examined  these  data  for 
four  reasons.  First,  Moyle  (1973) 
summarized  data  from  only  some  of 
the  sites  where  yellow-legged  frogs 
were  not  found.  We  were  equally 
interested  in  habitat  variation  among 
all  sites  sampled  where  yellow- 
legged  frogs  had  not  been  found  as 
well  as  sites  where  they  were  found. 


Second,  Moyle  (1973)  found  that  the 
collective  abundance  of  all  fish  spe- 
cies was  inversely  correlated  with 
that  of  yellow-legged  frogs,  but  also 
commented  that  yellow-legged  frogs 
were  most  abundant  where  native 
fishes  were  present.  Because  original 
estimates  of  the  numbers  of  each  fish 
species  were  available  and  an  inverse 
relationship  between  the  abundance 
of  native  frogs  and  introduced  fishes 
had  already  been  identified  (Hayes 
and  Jennings  1986),  we  were  espe- 
cially interested  in  relationships  be- 
tween the  abundance  of  specific  na- 
tive and  introduced  fishes  and  that  of 
yellow-legged  frogs.  Third,  Moyle 
(1973)  coded  fish  abundance  when 
the  data,  as  originally  recorded,  per- 
mit at  least  ranking,  so,  where  pos- 
sible, we  analyzed  the  original  data 
directly  to  minimize  bias  that  can  re- 
sult from  coding  (Sokal  and  Rohlf 
1981).  Lastly,  the  fish  abundance  data 
displayed  skewed  distributions  for 
several  species,  so  we  used  non-par- 
ametric analyses  to  avoid  having  to 
make  any  assumptions  about  sample 
distributions. 

Statistical  treatments  used  are  de- 
scribed in  Sokal  and  Rohlf  (1981)  and 
Zar  (1974).  All  contingency  table 
comparisons  performed  had  one  de- 
gree of  freedom  (df),  so  all  Chi- 
square  values  were  calculated  with 
the  correction  for  continuity  (X^^).  For 
those  analyses  that  required  more 
than  one  comparison  using  some  of 
the  data,  alpha  (a)  was  evaluated 
based  on  the  number  of  comparisons 
to  a  level  equivalent  to  0.05  using  Si- 
dak's  multiplicative  inequality  (Sokal 
and  Rohlf  1981). 


RESULTS 

California  Red-Legged  Frog 

Ram  aurora  draytonii  was  recorded 
primarily  from  aquatic  habitats  that 
were  intermittent  streams  which  in- 
cluded some  area  with  water  at  least 
0.7  meters  deep,  had  a  largely  intact 
emergent  or  shoreline  vegetation. 


146 


Table  S.—Frequency  of  fish  species  co-occurrence  with  Rana  aurora  dray- 
tonii.  Percentage  Is  the  number  of  sites  respective  fish  species  were  re- 
corded as  a  function  of  all  sites  where  fishes  were  recorded  as  co-occur- 
ring with  R.  a.  drayfonii.  An  asterisk  (*)  indicates  introduced  species. 


Co-occurrence  Percentage 


Species 

(n  => 

<%) 

California  roach  (Lavinia  symmefricus) 

19 

47 

Mosquitofish  (Gambusia  affinis)* 

10 

25 

Hitch  (Lavinia  exilicauda) 

6 

15 

Green  sunfish  (Lepomis  cyanellus)* 

6 

15 

Threespine  stickleback  (Gasferosfeus  aculeafus) 

3 

8 

Sacramento  squowfish  (Piychocheiius  grandis) 

2 

5 

Sacramento  sucker  (Cafosfomus  occidenfalis) 

2 

5 

Prickly  sculpin  (Coffus  asper) 

1 

3 

Hardhead  (Mylopliarodon  conoceplialus) 

1 

3 

Rainbow  trout  (Salmo  gairdnerii) 

1 

3 

Brown  trout  (Salmo  fruffa)* 

1 

3 

and  lacked  introduced  bullfrogs 
(table  2).  We  found  descriptions  ade- 
quate to  characterize  vegetation  for 
77%  (33)  of  sites  where  the  emergent 
or  shoreline  vegetation  variable 
could  be  scored.  With  three  excep- 
tions, descriptions  indicated  that  ei- 
ther, or  both  of,  an  emergent  vegeta- 
tion of  cattails  {Typha  spp.)  or  tules 
{Scirpus  spp.),  or  a  shoreline  vegeta- 
tion of  willows  (Salix  spp.)  were 
present.  Shrubby  willows  were  re- 
corded at  67%  (22)  of  the  sites  with 
vegetative  descriptions,  and  were 
identified  as  arroyo  willow  (Salix  la- 
siolepis)  in  the  eight  instances  where  a 
species  name  was  provided.  Only 
juvenile  frogs  were  recorded  at  five 
of  the  six  sites  where  a  limited  emer- 
gent vegetation  was  present  and  at 
the  only  site  that  lacked  a  water 
depth  greater  than  0.7  m.  We  found 
no  significant  difference  in  the  num- 
bers of  intermittent  versus  perennial 


sites  with  red-legged  frogs  that  had  a 
dense  vegetation  and  a  water  depth 
of  >0.7  m  (X2^  =  0.338,  p  =  0.561,  for 


vegetation;  X^^  =  0.017,  p  =  0.897,  for 
water  depth;  X^^^^  ^^^^s  =  5-024  for 
both). 

Rana  aurora  draytonii  was  also 
more  frequently  recorded  at  sites 
with  native  fishes  and  with  substrate 
alteration,  but  less  frequently  re- 
corded at  sites  with  introduced 
fishes.  Fishes  were  present  at  69%  (40 
of  58)  of  sites  where  data  as  to  their 
occurrence  were  recorded;  26  sites 
had  only  native  fishes,  seven  had 
only  introduced  fishes,  and  seven 
had  both.  Only  four  fish  species, 
California  roach  (Lavinia  symmet- 
ricus),  hitch  (Lavinia  exilicauda),  green 
sunfish  (Lepomis  cyanellus),  and 
mosquitofish  (Gambusia  affinis),  were 
recorded  as  co-occurring  with  R.  a. 
draytonii  at  more  than  three  sites 
(table  3),  and  only  California  roach 
was  recorded  at  more  than  25%  (10) 
of  sites.  Sixty  of  the  70  sites  described 
as  being  substrate-altered  at  the  time 
R.  a.  draytonii  was  recorded  were 
small  impoundments. 

California  red-legged  frogs  were 
also  most  frequently  recorded  at  sites 
influenced  by  a  small  drainage  area, 
having  a  low  local  gradient,  and  in 
streams  having  a  low  stream  order. 
Drainage  areas  of  sites  from  which  R. 
a.  draytonii  was  recorded  vary  from 
0.02  km2  to  over  9000  km^,  but  two- 


Table  2.— Variation  among  habitat  variables  for  California  red-legged  frogs 
(Rana  aurora  draytonii).  Number  of  localities  (percentages  of  localities)  in 
each  category  are  indicated.  See  table  1  and  text  for  explanation  of  vari- 
able categories. 


Variable 


Variable  categories 


1. 

Aquatic  habitat  type 

(a)  stream 

129 

(92%) 

(b)  pond 

10 

(8%) 

2. 

Temporal  status  of 

(a)  perennial 

49 

(36%) 

aquatic  site 

(b)  intermittent 

88 

(64%) 

3. 

Water  depth 

(a)  >  0.7  meters 

73 

(99%) 

(b)  <  0.7  meters 

1 

(1%) 

4. 

Emergent  and 

(a)  absent 

0 

(0%) 

shoreline  vegetation 

(b)  limited 

9 

(20%) 

(c)  dense 

35 

(80%) 

5. 

Native  fishes 

(a)  present 

33 

(65%) 

(b)  absent 

18 

(35%) 

6. 

Introduced  fishes 

(a)  present 

14 

(44%) 

(b)  absent 

18 

(56%) 

7. 

Introduced  bullfrogs 

(a)  present 

13 

(1 1%) 

(b)  absent 

102 

(89%) 

8. 

Significant  substrate 

(a)  present 

70 

(62%) 

alteration 

(b)  absent 

43 

(38%) 

9. 

Significant  removal 

(a)  present 

1 

(2%) 

vegetation  (see  #4) 

(b)  absent 

44 

(98%) 

10. 

Current  status 

(a)  probably  extant 

86 

(72%) 

(among  localities) 

(b)  probably  extinct 

34 

(28%) 

11. 

Current  status 

(a)  probably  extant 

18 

(42%) 

(among  drainages) 

(b)  probably  extinct 

25 

(58%) 

147 


thirds  (n  =  83)  are  from  localities 
with  drainage  areas  <40  km^  (fig.  1). 
Local  gradient  (slope)  at  California 
red-legged  frog  localities  varies  from 
0.04°  to  12.8°  from  horizontal,  al- 
though 87%  (n  =  100)  occur  at  sites 
with  slopes  <2°.  California  red- 
legged  frogs  have  been  recorded  in 
1st  to  6th  order  streams,  but  94%  (n  = 
119)  of  these  localities  are  4th-  or 
lesser-order  streams  and  42%  are  1st- 
order  streams  (fig.  2). 

Based  on  the  subset  for  which  cur- 
rent data  were  available  (n  =  120), 
California  red-legged  frogs  are 
probably  extinct  at  >25%  of  the  lo- 
calities where  they  were  historically 
recorded.  When  clustered  into  a 
sample  representing  drainage  sys- 
tems (n  =  43;  see  methods),  this  sub- 
set indicates  that  California  red- 
legged  frogs  are  probably  extinct  in 
over  50%  of  the  drainage  systems  in 
the  Central  Valley  area.  Three  habitat 
variables  (temporal  status  of  aquatic 
habitat,  drainage  area,  and  intro- 
duced bullfrogs)  showed  a  signifi- 
cant relationship  to  the  probability  of 
survival  of  local  populations  of  C!ali- 
fornia  red-legged  frogs  (table  4).  We 
found  that  R.  a.  draytonii  is  likely  ex- 
tant at  82%  (n  =  70)  of  localities  with 
an  intermittent  aquatic  habitat, 
whereas  it  is  probably  extinct  at  71% 
(n  =  22)  of  the  sites  with  a  perennial 
aquatic  habitat.  Grouping  localities 
based  on  drainage  area,  R.  a.  dray- 
tonii is  probably  extant  at  83%  (n  = 


>«01  1 3 

3001-4000  1 1 

MOt-3000  |2 

ATM  ISOI-MOO  2 
(H  km) 

lOOl-ISOO  t 

SOI-tOOO  1 


iKillUMbc) 

Figure  1  .—Frequency  distribution  of  locali- 
ties wtiere  Rana  aurora  draytonii  \nas  been 
recorded  in  thie  Central  Valley,  California 
based  on  drainage  area.  Thie  inset  details 
\he  frequency  distribution  of  localities  witti 
drainage  areas  <  280  kpn*. 


7A\-2tO 

201-240 

12 

161-200 

121-160 

81-120 

13 

41-60 
0-40 

I  I  I  I  I 
0       20      40       60      60  100 


 ■  \ 

Table  4.— Contingency  analysis  relating  selected  habitat  variables  to  an 
estimate  of  the  lilceiihood  that  historically  recorded  California  red-legged 
frog  populations  are  extant.  Status  of  frog  populations  at  recorded  locali- 
ties are  indicated  as  extant  (=  probably  extant)  and  extinct  <=  probably 
extinct).  A  double  asterisk  (")  denotes  significant  contingency  tables, 
based  a  critical  ^^^^y^^^^^Q^^j  =  7.3,  a  adjusted  for  seven  comparisons  (see 
methods). 


Locality  Status 


Variable 

Condition 

extcnt 

extinct 

x% 

Probability 

1. 

Temporal  status 

Perennial 

9 

22 

27.326 

o.ooor* 

Intermittent 

70 

15 

2. 

Drainage  area 

>300  km2 

0 

11 

31.466 

o.ooor* 

<300  km^ 

85 

18 

3. 

Native  fishes 

■f 

13 

6 

0.276 

0.5991 

14 

11 

4 

Introduced  bullfrogs 

+ 

0 

10 

27.140 

0.0001" 

70 

16 

5. 

Substrate  alteration^ 

25 

14 

0.983 

0.3215 

47 

14 

6. 

Introduced  fishes 

+ 

5 

9 

0.003 

0.9524 

7 

10 

7. 

Substrate  alteration'' 

-1- 

21 

3 

<0.001 

0.9944 

26  5 


"Analysis  with  all  localities. 

^Analysis  with  subset  of  localities  having  a  drainage  area  <25  km'. 

\  

affected  by  the  largest  drainage  areas 
(n  =  10).  Similarly,  R.  a.  draytonii  is 
probably  extant  at  81%  (n  =  70)  of 
localities  lacking  introduced  bull- 
frogs and  is  probably  extinct  at  all 
localities  (n  =  10)  where  it  has  been 
recorded  with  bullfrogs.  Remaining 
variables  either  failed  to  show  a  sig- 
nificant relationship  to  the  probabil- 
ity of  California  red-legged  frog  sur- 
vival (table  4),  or  one  of  the  variable 
categories  was  so  rare  that  this  analy- 
sis was  not  applicable  (see  table  2). 

Foothill  Yellow- Legged  Frog 

Rana  boylii  was  recorded  primarily 
from  shallow,  partly  shaded  stream 
sites  with  riffles  and  at  least  a  cobble- 
sized  substrate.  All  29  stream  sites  at 
which  either  post-metamorphic  or 
larval  R.  boylii  were  recorded  were 
<0.6  m  in  average  water  depth  (fig.  3) 
and  had  at  least  some  shading  (fig. 
4).  Rana  boylii  was  recorded  more 


85)  of  sites  influenced  by  a  small 
(<300  km^)  drainage  area,  whereas  it 
is  probably  extinct  at  all  recorded 
localities  (n  =  11)  influenced  by  a 
large  (>300  km^)  drainage  area. 
Moreover,  available  data  indicate 
that  R.  a.  draytonii  is  extinct  at  all  re- 
corded localities  on  the  Central  Val- 
ley floor,  which  includes  all  localities 


0  10  »  M  4Q  SC  M 


Figure  2.— Frequency  distribution  of  locali- 
ties wt^ere  Rana  aurora  draytonii  has  been 
recorded  in  the  Central  Valley,  California 
based  on  stream  order. 


148 


frequently  at  sites  with  a  stream  area 
that  was  >20%  shaded  than  at  sites 
with  >20%  shading.  Only  one  of  29  R. 
hoylii  sites  lacked  riffle  habitat  and  R. 
boylii  was  recorded  significantly 
more  frequently  at  sites  with  >40% 
riffle  area  than  at  sites  with  a  riffle 
area  of  <40%  [X^  =  8.680,  p  =  0.003, 
X'df=,a(2)=o.o25  =  5.024;  fig.  5].  Only  four 
of  29  R.  boylii  sites  lacked  at  least  a 
cobble-sized  substrate  and  R.  boylii 
was  recorded  most  frequently  (20  of 
29)  at  sites  with  >40%  of  the  sub- 
strate that  was  at  least  cobble-sized 
(fig.  6).  Few  other  patterns  could  be 
identified  from  among  the  environ- 
mental variables  that  we  re-analyzed. 
Rana  boylii  was  recorded  more  fre- 
quently from  perennial  streams  (n  = 
19)  than  from  intermittent  ones  (n  = 
10),  but  the  difference  was  not  sig- 
nificant when  compared  to  the  total 
number  of  perennial  (n  =  71)  and 
intermittent  (n  =  59)  stream  sites 
sampled  [X^'^  =  1.268,  p  =  0.260, 
X'df=i,a(2)=o.o25  =  5-024].  Of  13  environ- 
mental variables  that  we  re-exam- 
ined, only  the  percentage  of  stream 
area  in  riffles  was  significantly  corre- 
lated with  the  abundance  of  R.  boylii 
(table  5). 

Rana  boylii  occurred  with  1-5  Cx  = 
2.5)  of  the  vertebrate  members  of  the 
aquatic  macrofauna  at  26  of  the  29 
localities  where  it  was  recorded. 


Figure  3.— Histogram  of  the  proportion  of 
sites  in  stream  depthi  categories  whiere 
Rana  boylii  hias  been  recorded  in  thie  Sierra 
Nevada  foothiiils,  California.  Sample  sizes  as 
a  function  of  thie  total  sample  in  eachi 
stream  deptti  category  are:  <0.20  (n=8/24), 
0.21=0.40  (n=9/43),  0.41-0.60  (n=12/57),  and 
>0.60  (n=0/18). 


0  I-»       Jl-«      <!■«  6'-«0 

^Ntifi  or  sirtM  Am  m  (arriH 

Figure  5.— Histogram  of  thie  proportion  of 
sites  in  riffle  categories  wtiere  Rana  boylii 
has  been  recorded  in  the  Sierra  Nevada 
foothills,  California.  Sample  sizes  as  a  func- 
tion of  the  total  sample  in  each  ritfle  cate- 
gory are:  0%  <n=l/36),  1  -20%  (n=5/31 ),  21  - 
407,  <n=4/21).  41-60%  (n=l  1/28).  61-80% 
(n=7/19),  and  81-100%  (n=2/6>. 


Foothill  yellow-legged  frogs  were 
recorded  as  occurring  with  12  differ- 
ent species,  but  co-occurrence,  ex- 
pressed as  the  percentage  of  total 
sites  at  which  either  R.  boylii  or  the 
co-occurring  species  were  recorded, 
did  not  exceed  31%  (table  6).  Intro- 
duced species  (n  =  6)  occurred  with 
R.  boylii  less  frequently  Tx  =  2, 1-3) 
than  native  species  \x  =  9.3, 1-17)  and 
native  species  had  a  significantly 
higher  percentage  of  co-occurrence 
(3-31%,  X  =  16.5%)  than  introduced 
species  [n  =  6;  2-9%,  x  =  3.7%;  Mann- 
Whitney  test,  U'  =  32.5,  p  =  0.0275, 
U 


criHcala(2)=0.05 


=  31].  Only  four  native 


0         1-20      21-«      4l-«0      6l-«  SI-100 
PrcirUgi  if  SkiM  Slmm  Vm 

Figure  4.— Histogram  of  the  proportion  of 
sites  in  stream  shading  categories  where 
Rana  boylii  has  been  recorded  in  the  Sierra 
Nevada  foothills,  California.  Sample  sizes  as 
a  function  of  the  total  sample  In  each 
sh'eam  shading  category  are:  0%  (n=0/5), 
1  -207o  (n=3/37>,  2 1  -40%  (n=7/38),  41  -607* 
(n=8/30),  61  -807.  (n=9/23),  and  81  -1007, 
(n=2/8). 


0  1-20       2l-«       41-M      i\-m  81-100 

Figure  6.— Histogram  of  the  proportion  of 
sites  in  substrate  categories  where  Rana 
boylii  has  been  recorded  In  the  Sierra  Ne- 
vada foothills,  California.  Sample  sizes  as  a 
function  of  the  total  sample  in  each  sub- 
strate category  are:  07,  (n=4/19),  1-207, 
(n=3/32),  21  -407,  (n=2/23),  41  -607,  (n=7/29), 
61-807,  (n=9/26).  and  81-1007,  (n=4/12). 


fishes,  California  roach,  Sacramento 
sucker  (Catostomus  occidentalis),  Sac- 
ramento squawfish  (Ptychocheilus 
grandis),  and  rainbow  trout  (Salmo 
gairdnerii),  occurred  with  R.  boylii  at 
more  than  three  of  the  29  sites  where 
the  latter  was  recorded,  and  of  these, 
only  California  roach  occurred  with 
R.  boylii  at  more  than  50%  of  the  sites 
where  R.  boylii  was  recorded.  Only 
one  species  assemblage,  that  consist- 
ing of  California  roach,  Sacramento 
squawfish,  and  Sacramento  sucker, 
occurred  with  R.  boylii  more  often 
than  expected  by  chance  alone  (table 
7).  Correlation  analysis  indicated  that 
the  abundance  of  10  of  the  12  co-oc- 
curring species  was  significantly  in- 
versely correlated  with  the  abun- 
dance of  R.  boylii  (table  8). 


DISCUSSION 

Habitat  Variation 

California  Red-Legged  Frog 

A  dense  vegetation  close  to  water 
level  and  shading  water  of  moderate 
depth  are  habitat  features  that  ap- 
pear especially  important  to  Califor- 
nia red-legged  frogs.  Previous  au- 
thors have  suggested  or  implied  the 
occurrence  of  at  least  one  of  these 
habitat  features.  Storer  (1925)  noted 


149 


that  R.  a.  draytonii  in  streams  was  re- 
stricted to  large  pools,  which  implies 
a  moderate  water  depth.  Stebbins 
(1966, 1985)  emphasized  vegetative 
cover  as  important  to  red-legged 
frogs,  but  his  comments  confound 
habitat  characteristics  that  may  be 
attributable  to  northern  versus  Cali- 
fornia (southern)  red-legged  frogs; 
data  on  these  two  forms  should  re- 
main partitioned  until  it  is  well-es- 
tablished that  they  are  not  different 
species  (Hayes  and  Miyamoto  1984, 
Hayes  and  Krempels  1986).  Zweifel 
(1955)  coupled  the  water  depth  and 
vegetation  features  of  California  red- 
legged  frog  habitat,  but  he  empha- 
sizes a  herbaceous  shoreline  vegeta- 
tion. Chir  data  indicate  that  a  more 
complex  vegetation  is  a  feature  of 
sites  where  R.  a.  draytonii  occurs. 
Cattails,  bulrushes,  and  shrubby  wil- 


/fdble  5. —Spearman  rankcorrelatlorr 
between  selected  environmental 
variables  and  the  coded  abun- 
dance of  R.  boylil  as  measured  by 
Moyie  (1973).  Sample  size  for  each 
variable  is  n  =  1 30.  A  double  asterisk 
(**)  Indicates  significant  correlations; 
based  on  a  critical  r,  =  0,267  at  an 
aCtwo -tailed)  =  0.002,  adjusted  for  24 
comparisons  (13  below  and  1 1  In  | 
table  8;  see  methods). 


Variable 

Correlation 
coefficient  <r,  =) 

Human  alteration 

-0.160 

Vegetation 

Aquatic  vegetation  (%) 

-0.157 

Floating  vegetation  (%) 

-0.169 

Shade  (%) 

0.219 

Stream  morphology 

Pools  (%) 

-0.205 

Riffles  (%) 

0.304** 

Runs (%) 

-0.020 

Stream  substrate 

Mud(%) 
Sand  (%) 
Gravel  (%) 
i^ubble  (%) 
Boulder/ Bedrock  (%) 


-0.035 
-0.085 
-0.032 
0.071 
0.192 


yRubble/Boulder/Bedrock  (%)  0. 1 72| 


Table  6.— Occurrences  of  aquatic  macrofaunal  species  among  the  1 30 
stream  sites  sampled  by  t^oyle  (1973)  and  Moyie  and  Nichols  (1973).  Co- 
occurrences is  the  number  of  sites  Rana  boy/// was  found  to  co-occur  with 
each  species.  Percentage  of  co-occurrences  is  co-occurrences  as  the 
percentage  of  those  sites  at  which  either  R.  boylil  or  the  state  species  oc- 
cur. An  asterisk  (*)  indicates  introduced  species.  Ten  other  fish  species 
(Goldfish  (Carassius  auratus),  Prickly  scuipln  (Coitus  asper),  Common  carp 
(Cyprlnus  carplo),  Threadfln  shad  /Dorosoma  pe/enense;,  Threespine  stick- 
leback (Gasterosteus  ocu/ea/us),  Yellow  bullhead  (Ictalurus  nebulosus), 
Redear  sunfish  (Lepomis  microlophus),  Chinook  salmon  (Onchorhynchus 
tshawytscha),  Brown  trout  (Salmo  trutta))  were  recorded  at  low  numbers  of 
stations  (<8);  none  were  recorded  as  co-occurring  with  R,  boylil 


Species 


Occurrences  Co-occur-  %of 
rences  co-occur- 
(n  =)  <n  =)  rences 


Bullfrog  ('/?ono  catesbe/ono/  68 

Green  sunfish  C/.epom/s  c/ane//us)*  61 
Sacramento  sucker  (Cafosfomus  occidenfalis)  55 
Sacramento  squawfish  (Ptychocheilus  grandis)  48 

California  roach  CLov/n/a  symmefncus)  43 

largemouth  bass  (Micropterus  salmoides)*  41 

Mosquitofish  (Gambusia  affinisT  37 

B\ueQi\\  (Lepomis  macrochirus)*  33 

Rainbow  trout  (Salmo  galrdnerii)  27 

White  catfish  (7c/o/afus  cofus)*  13 

Golden  shiner  (Nofemigonus  crysoleucas)*  13 

Hitch  (Lavinia  exilicauda)  12 

Hardhead  (Mylopharodon  conocephalus)  1 1 

Smallmouth  bass  (Micropterus  dolomieui)*  9 


2 
2 
13 
12 
17 
0 
1 
3 
11 
1 
0 
1 
2 
3 


2 
2 
18 
18 
31 
0 
2 
5 
24 
2 
0 
3 
5 
9 


Table  7.— Frequencies  of  species  assemblages  of  aquatic  macrofaunal 
vertebrates  co-occurring  with  R.  boy/// from  data  recorded  by  Moyie 
(1973).  Assemblages  listed  include  only  combinations  of  species  recorded 
as  co-occurring  wlth^.  boy/// at  least  seven  localities  (see  table  6).  Listed 
species  are  California  roach  (RCH),  Sacramento  sucker  (SKR),  Sacramento 
squawfish  (SO),  and  Rainbow  trout  (RD.  Asterisks  (**)  identify  assemblages 
co-occurring  at  frequencies  significantly  higher  than  expected  by  chance, 
based  on  a  critical  X^e^,T  a-ooos  ~  7.879,  adjusted  for  1 1  combinations  (see 
methods).  Probabilities  (p)  are  those  associated  with  calculated  values. 


Species 
assemblage 

Frequencies 

Probability 

Observed 

Expected 

RCH/RT/SKR/SQ 

2 

1.20 

0.077 

0.75<p<0.90 

RCH/SKR/SQ 

9 

3.15 

9.068** 

0.003 

RCH/RT/SQ 

2 

2.67 

0.011 

0.90<p<0.95 

RCH/RT/SKR 

2 

2.89 

0.053 

0.75<p<0.90 

RT/SQ/SKR 

2 

2.04 

0.104 

0.50<p<0.75 

RCH/RT 

5 

6.45 

0.139 

0.60<p<0.75 

RCH/SKR 

10 

7.62 

0.463 

0.25<p<0.50 

RCH/SQ 

9 

7.03 

0.305 

0.50<p<075 

RT/SKR 

3 

4.93 

0.415 

0.50<p<0.75 

RT/SQ 

3 

4.55 

0.243 

0.50<p<0.75 

SKR/SQ 

11 

5.38 

4.959 

0.026 

150 


lows,  the  plants  comprising  emergent 
and  shoreline  vegetation  at  such 
sites,  typically  shade  a  substantial 
surface  area  of  water  with  a  dense 
matrix  at  or  near  water  level.  Califor- 
nia red-legged  frogs  appear  sensitive 
to  the  presence  of  such  a  vegetation 
structure  because  most  sites  from 
which  frogs  were  recorded  lacked 
significant  alteration  of  emergent  or 
shoreline  vegetation  (see  table  2). 
Moreover,  because  only  juvenile 
frogs  were  recorded  from  most  sites 
with  limited  shoreline  or  emergent 
vegetation,  a  minimum  amount  of 
such  vegetation  appears  to  be  needed 
for  survival  of  adults.  Parallel  argu- 
ments apply  to  water  depth.  Previ- 
ous authors  have  characterized  R.  a. 
draytonii  as  a  p)Ool-  or  pond-dwelling 
species  (Stebbins  1966, 1985;  Storer 
1925;  Zweifel  1955)  and  descriptions 
corresponding  to  that  characteriza- 
tion were  recorded  for  this  frog  at 
most  sites.  Yet,  we  found  that  using 
minimum  water  depth  was  a  more 
encompassing  habitat  descriptor  be- 


cause it  included  canals  and  stream 
sites  where  adult  frogs  were  de- 
scribed as  being  conrvmon  and  that 
had  the  minimum  water  depth  re- 
quirement, but  could  not  be  de- 
scribed as  either  ponds  or  stream 
pools.  Available  description  of  such 
sites  indicates  that  they  fit  the  defini- 
tion of  a  run  (Armour  et  al.  1983), 
although  data  upon  which  part  of  the 
definition  is  based  (the  rate  of  water 
flow)  are  lacking. 

We  believe  that  California  red- 
legged  frogs  occur  primarily  in 
streams  because  alternative  sites 
(ponds)  that  have  suitable  water 
depth  and  vegetation  characteristics 
were  historically  rare  outside  of 
stream  habitats  rather  than  because 
red-legged  frogs  are  somehow  pre- 
adapted  for  survival  in  streams.  His- 
torically, pond  habitats  below  1500  m 
in  the  Central  Valley  were  mostly 
vernal  pools,  a  habitat  too  shallow 
and  ephemeral  to  develop  the  mac- 
rovegetation  found  associated  with 
R.  a.  draytonii  (see  Holland  1973,  Jain 


Table  8.— Spearman  rank  correlation  between  the  numerical  <non-coded) 
abundance  of  the  vertebrate  macrofauna  and  the  abundance  (coded)  of 
R.  boylUas  recorded  by  Moyle  (1973).  Sample  size  is  based  on  the  total 
number  of  sites  where  either  R.  boy/// or  the  species  being  compared  was 
present.  A  single  asterisk  (*)  Indicates  introduced  species.  A  double  aster- 
isk (••)  Identifies  significant  correlations  at  an  _  (two-tailed)  =  0.002,  ad- 
justed for  24  comparisons  (1 1  below  and  13  In  table  5;  see  methods). 
Probability  (p)  is  the  probability  of  obtaining  the  calculated  Spearman  cor- 
relation coefficient  (r^.  Common  names  for  the  listed  species  are  In 
table  6. 

Critical 


size 

coefficient 

Probability 

S|:>ecles 

(n=) 

(r,=; 

(p=) 

Cafosfomus  occidenfalis 

71 

-0.404" 

<0.001 

-0.363 

Gambusia  affinis* 

62 

-0.835" 

<0.001 

-0.388 

Icfalurus  cafus* 

41 

-0.798** 

<0.001 

-0.473 

Lovinia  eydlicauda 

40 

-0.760" 

<0.001 

-0.479 

Lavinla  symm  e  trie  us 

55 

-0.316 

0.020 

-0.411 

Lepomis  cyonellus* 

88 

-0.742" 

<0.001 

-0.327 

Lepomis  macrochirus' 

59 

-0.827" 

<0,001 

-0.397 

Micropferus  dolomfeur 

35 

-0.538" 

0.001 

-0.510 

Mylopharodon  conocephalus 

38 

-0.607" 

<0.001 

-0.491 

PlychocheHus  grandis 

66 

-0.54r* 

<0.001 

-0.376 

Rana  cafesbeiana* 

90 

-0.800" 

<0.001 

-0.323 

Salmo  gairdneriJ 

44 

-0.425 

0.005 

-0.458 

1976).  Even  the  only  two  exceptions 
to  R.  a.  draytonii  not  occurring  in  ver- 
nal pools  support  this  hypothesis.  A 
large  vernal  pool  in  San  (Dbispo 
County,  California  is  known  to  have 
a  population  of  California  red-legged 
frogs  (D.  C.  Holland,  pers.  comm.). 
However,  this  vernal  pool  is  atypical 
because  it  possesses  significant  mac- 
rovegetation  and  water  depth.  These 
features  appear  to  be  present  because 
this  large  (ca.  20  ha)  pool  does  not 
dry  down  each  year.  The  second  ex- 
ception is  a  vernal  pool  in  coastal 
southern  California  in  which  two 
frogs  with  abnormal  numbers  of  legs 
were  found  (Cunningham  1955). 
Cunningham  thought  that  the  defects 
were  induced  by  exposure  to  high 
temperatures  during  early  develop- 
ment, a  condition  facilitated  by  the 
limited  vegetative  cover  that  was 
present.  His  speculation  may  be 
valid  if  California  red-legged  frog 
embryos  have  a  low  critical  thermal 
maximum  (Hayes  and  Jennings 
1986).  Storer  (1925)  thought  that  R.  a. 
draytonii  was  excluded  from  tempo- 
rary (vernal)  pools  because  its  larval 
period  is  relatively  long,  but  the 
more  likely  mechanism  is  that  frogs 
immigrating  to  such  pools  were  un- 
able to  establish  because  suitable 
habitat  was  lacking.  The  latter  hy- 
pothesis is  supported  because  Cali- 
fornia red-legged  frogs  are  not  re- 
corded from  the  many  vernal  pools 
that  hold  water  for  intervals  longer 
than  the  minimum  time  required  by 
R.  a.  draytonii  to  complete  metamor- 
phosis (10  weeks;  Hayes,  unpubl. 
data;  see  also  Jain  1976,  Zedler  1987). 

Rana  a.  draytonii  also  appears  to 
have  responded  to  the  creation  of 
habitat  with  the  appropriate  vegeta- 
tion and  water  depth  characteristics. 
A  significant  aspect  of  the  changes  in 
aquatic  habitats  that  have  occurred 
in  the  Central  Valley  below  1500  m  is 
an  increase  in  the  number  of  perma- 
nent ponds  (Moyle  1973).  Storer 
(1925)  reported  that  R.  a.  draytonii 
occurred  in  a  number  of  water  stor- 
age reservoirs  and  artificial  ponds, 
but  the  habitat  features  of  those  sites 


151 


were  not  described.  Thus,  it  was  of 
special  interest  to  find  that  no  signifi- 
cant difference  could  be  identified 
between  the  probability  of  extinction 
of  R.  a.  draytonii  at  substrate-altered 
sites  (mostly  small  impoundments) 
and  at  sites  lacking  such  alteration. 
Moyle  (1973)  concluded  that  the  de- 
cline of  R.  a.  draytonii  was  related  in 
part  to  human-induced  alteration, 
including  creation  of  impoundments. 
Our  data  suggest  that  human-in- 
duced alteration  creating  small  im- 
poundments cannot  be  related  di- 
rectly to  the  disappearance  of  Cali- 
fornia red-legged  frogs.  We  empha- 
size that  these  data  do  not  exclude 
the  alternative,  discussed  later, 
which  indicates  that  the  creation  of 
small  impoundments  is  likely  to  have 
an  indirect  negative  effect  on  R.  a. 
draytonii  by  facilitating  the  dispersal 
of  introduced  aquatic  predators. 

Besides  features  of  habitat  struc- 
ture associated  with  R.  a.  draytonii,  its 
isolation  from  one  or  more  aquatic 
macrofaunal  predators  is  the  other 
key  element  suggested  by  these  data. 
No  significant  variation  was  found  in 
the  features  of  habitat  structure  im- 
portant to  R.  a.  draytonii  between 
intermittent  and  perennial  aquatic 
sites,  so  differences  in  habitat  struc- 
ture cannot  explain  why  R.  a.  dray- 
tonii is  recorded  most  frequently 
from  intermittent  aquatic  sites.  We 
believe  that  California  red-legged 
frogs  were  recorded  most  frequently 
from  intermittent  sites  because  the 
likelihood  of  extinction  at  perennial 
sites  is  now  higher  than  at  intermit- 
tent sites  (see  table  4)  and  few  his- 
torical data  are  available  from  when 
frogs  were  often  found  at  perennial 
sites. 

California  red-legged  frogs  are 
now  extinct  from  all  sites  on  the  Cen- 
tral Valley  floor,  all  of  which  were 
perennial  and,  except  for  one,  were 
recorded  prior  to  1950.  We  believe 
that  the  disadvantage  associated 
with  perennial  sites  and  the  advan- 
tage associated  with  intermittent 
sites  is  the  degree  to  which  the  for- 
mer allow,  and  the  latter  restrict,  the 


access  of  aquatic  macrofaunal  preda- 
tors. 

The  remaining  variation  in  fea- 
tures of  R.  a.  draytonii  habitat  we 
have  identified  can  be  directly,  or 
indirectly,  linked  to  a  hypothesis  in- 
voking the  influence  of  one  or  more 
aquatic  macrofaunal  predators.  The 
significantly  lower  likelihood  of  ex- 
tinction at  sites  with  small  drainage 
areas  (table  4)  and  R.  a.  draytonii 
being  recorded  from  a  greater  num- 
ber of  localities  with  smaller  drain- 
age areas  (fig.  1)  and  lower  stream 
orders  (fig.  2),  are  probably  unrelated 
to  either  drainage  area  or  stream  or- 
der effects  per  se.  Rather,  they  are  a 
function  of  both  the  bias  against  re- 
cording historical  data  and  the  fact 
that  sites  with  smaller  drainages  or 
lower  stream  orders  have  a  higher 
probability  of  being  intermittent 
aquatic  habitats,  which  have  a  higher 
probability  of  excluding  aquatic 
predators.  Limited  co-occurrence 
with  aquatic  predators,  namely  bull- 
frogs and  predatory  fishes,  and  a  sig- 
nificantly higher  likelihood  of  extinc- 
tion at  sites  where  bullfrogs  were  re- 
corded (table  4)  may  indicate  a  nega- 
tive interaction  with  one  or  more  of 
these  species.  Rana  a.  draytonii  did 
not  co-occur  with  any  fish  species 
frequently.  It  co-occurred  most  often 
with  California  roach,  a  small,  om- 
nivorous native  fish  that  is  thought 
to  have  declined,  in  part,  due  to  pre- 
dation  by  introduced  fishes  (Moyle 
and  Nichols  1974,  Moyle  1976).  We 
did  not  detect  a  significantly  higher 
likelihood  of  extinction  at  sites  with 
introduced  fishes.  However,  the 
sample  was  too  small  to  partition  to 
permit  testing  individual  fish  species, 
the  level  at  which  we  believe  such  an 
effect  is  most  likely. 

While  we  are  reasonably  con- 
vinced that  the  greater  restriction  of 
R.  a.  draytonii  to  intermittent  aquatic 
habitats  is  an  effect  due  to  novel 
aquatic  predators,  we  emphasize  that 
these  data  cannot  identify  which  are 
the  aquatic  predators  producing  such 
an  effect.  The  inability  to  identify  the 
responsible  predators  is  complicated 


by  the  condition  of  limited  overlap 
between  each  potential  predator  and 
R.  a.  draytonii.  That  condition  pre- 
vents excluding  the  alternative  that 
different  habitat  requirements  rather 
than  any  predatory  interaction  may 
explain  the  limited  overlap  in  habitat 
use  between  each  putative  predator 
and  California  red-legged  frogs 
(compare  Moyle  1973  for  bullfrogs 
and  Moyle  and  Nichols  (1973)  for 
various  fishes,  but  especially  mosqui- 
tofish  and  green  sunfish;  see  also 
Hayes  and  Jennings  1986  for  a  dis- 
cussion). It  is  this  fact  and  the  appar- 
ent intolerance  of  R.  a.  draytonii  to 
unshaded  habitat  that  leads  us  to 
suggest  that  some  alteration  of  ripar- 
ian vegetation  may  be  necessary  to 
create  the  conditions  for  a  negative 
interaction. 


Foothill  Yellow-Legged  Frog 

Partly  shaded,  shallow  streams  and 
riffles  with  a  rocky  substrate  that  is 
at  least  cobble-sized  are  the  habitat 
features  that  appear  to  be  important 
to  foothill  yellow-legged  frogs.  Previ- 
ous authors  agree  that  R.  hoylii  oc- 
curs in  streams  (Moyle  1973;  Stebbins 
1966, 1985;  Storer  1925;  Zweifel 
1955),  but  variation  exists  in  the  fea- 
tures of  streams  associated  with 
these  frogs.  Of  environmental  vari- 
ables that  appear  important  to  R. 
boy  Hi,  the  percentage  of  stream  area 
in  riffles  is  the  only  one  we  were  able 
to  correlate  significantly,  albeit 
weakly,  with  its  abundance.  Moyle 
(1973)  obtained  a  similar  positive 
correlation  in  his  original  analysis  of 
the  same  data,  and  Stebbins  (1966, 
1985)  also  emphasized  riffles  as  one 
of  the  key  aspects  of  R.  hoylii  habitat. 
The  reason  for  the  weak  correlation 
we  found  is  uncertain,  but  one  or 
more  of  three  factors  probably  pro- 
duced that  result.  First,  as  intermit- 
tent streams  lose  surface  flow  during 
late  summer,  riffles  disappear,  and  R. 
hoylii  can  then  be  found  associated 
with  stream  pools  (Fitch  1938,  Slevin 
1928,  Storer  1925,  Zweifel  1955). 


152 


Moyle's  data  were  collected  in  late 
summer  and  10  of  the  29  stream  sites 
at  which  R.  boylii  was  recorded  were 
intermittent,  so  data  from  these  sites 
may  have  diluted  the  correlation. 
Second,  riffle  area  may  be  correlated 
with  the  abundance  of  R.  boylii  only 
above  or  below  certain  values  (see 
fig.  5).  Lastly,  R.  boylii  has  been  re- 
ported from  sites  with  little  or  no 
riffle  habitat  unrelated  to  seasonal 
patterns  (Fitch  1938,  Zweifel  1955). 

Apart  from  riffles,  our  reanalysis 
of  environmental  variables  differs 
from  that  of  Moyle  (1973),  who 
found  that  five  of  the  other  variables 
that  we  re-examined  were  either 
positively  (i.e.,  shading  and  boulder/ 
bedrock;  compare  table  1  in  Moyle 
[1973]  and  our  table  5)  or  negatively 
(i.e.,  rooted  vegetation  [=  our  aquatic 
vegetation],  pools,  man  modified  [= 
our  human  alteration])  significantly 
correlated  with  the  abundance  of  jR. 
boylii.  We  attribute  this  difference,  in 
part,  to  our  analysis  being  more  con- 
servative because  we  adjusted  a  for 
the  experimentwise  error  rate,  our 
analysis  was  not  restricted  to  locali- 
ties where  only  frogs  were  found, 
and  we  used  non-parametric  tests. 
Some  of  the  correlations  that  Moyle 
(1973)  observed  with  R.  boylii  abun- 
dance may  have  been  significant  due 
to  one  or  more  of  these  differences. 
We  must  emphasize,  however,  that 
several  of  the  variables  that  Moyle 
found  correlated  with  R.  boylii  abun- 
dance vary  differentially  in  their  oc- 
currence between  riffles  and  pools 
(e.g.,  boulder /bedrock;  see  Moyle 
[1973]  and  Moyle  and  Nichols 
[1973]).  Those  variables  are  also  sus- 
ceptible to  the  seasonal  correlation- 
altering  effects  discussed  for  the  riffle 
variable.  Thus,  a  conservative  analy- 
sis, like  ours,  is  less  likely  to  detect 
variables  related  to  frog  abundance 
within  such  a  data  set. 

Nevertheless,  variables  identified 
as  important  to  R.  boylii  need  not  be 
correlated  to  its  abundance.  Stream 
depth,  shading,  and  substrate  type 
may  represent  such  variables.  Out 
reanalysis  of  Moyle's  data  suggests 


that  sites  with  a  shallow  average 
stream  depth  are  somehow  advanta- 
geous (see  fig.  3).  Moyle  (1973)  found 
no  significant  correlation  between  the 
abundance  of  R.  boylii  and  stream 
depth,  and  he  did  not  discuss  stream 
depth  with  respect  to  foothill  yellow- 
legged  frogs  in  any  other  context. 
Zweifel  (1955)  noted  that  streams  in 
which  R.  boylii  occurred  were  seldom 
more  than  0.3  m  deep,  and  Fitch 
(1936),  Storer  (1925),  and  Wright  and 
Wright  (1949)  found  that  R.  boylii 
usually  lays  eggs  in  shallow  water. 
Still,  overall  importance  of  stream 
depth  to  R.  boylii  remains  unclear. 
Our  reanalysis  also  suggests  that 
some  advantage  is  linked  to  in- 
creased shade  up  to  some  intermedi- 
ate level  (see  fig.  4).  Zweifel  (1955) 
described  shading  in  typical  R.  boylii 
habitat  as  interrupted,  whereas 
Moyle  (1973)  reported  a  positive  cor- 
relation between  frog  abundance  and 
the  degree  of  shading. 

Some  workers  have  emphasized 
the  degree  of  openness  or  insolation 
in  R.  boylii  habitat,  rather  than  ad- 
dressing shading  (Fitch  1938;  Steb- 
bins  1966, 1985).  Nevertheless,  even 
the  latter  imply  that  some  shading  is 
present.  Fitch's  (1938)  suggestion  that 
yellow-legged  frogs  are  excluded  by 
dense  canopy  may  be  supported  by 
Moyle's  data  because  he  recorded  no 
R.  boylii  at  sites  with  >90%  shading 
(see  also  fig.  4).  Our  reanalysis  also 
suggests  that  some  advantage  is  as- 
sociated with  sites  possessing  at  least 
a  cobble-sized  substrate  (see  fig.  6). 
Although  workers  have  most  fre- 
quently emphasized  the  rocky  aspect 
of  R.  boylii  habitat  (Fitch  1936, 1938; 
Moyle  1973;  Stebbins  1966, 1985; 
Storer  1925),  substrate  descriptions 
of  that  habitat  are  probably  as  varied 
as  any  other  single  variable.  Moyle 
(1973)  identified  a  positive  correla- 
tion between  the  percentage  of 
stream  area  with  bedrock  and  boul- 
ders and  the  abundance  of  R.  boylii, 
yet  sites  with  gravely  (Gordon  1939), 
sandy  (Zweifel  1955),  or  muddy  sub- 
strates have  also  been  recorded 
(Fitch  1938,  Storer  1925).  Because 


Moyle's  data  do  not  provide  frog 
age,  we  could  not  determine  whether 
sites  having  a  substrate  that  was  less 
than  cobble-sized  were  simply  mar- 
ginal habitat  with  juvenile  R.  boylii 
(see  Zweifel  1955),  or  whether  they 
represented  real  variation  in  habitat 
used  by  established  populations. 

Fitch  (1938)  and  Zweifel  (1955)  re- 
ported on  a  few  sites  with  adult  frogs 
that  lacked  a  substrate  that  was 
cobble-sized  or  larger  and  appeared 
to  have  few  predators.  They  sug- 
gested that  yellow-legged  frogs  are 
rarely  recorded  from  such  sites  be- 
cause their  predators  may  access  the 
"atypical"  habitat  more  easily.  Nev- 
ertheless, data  on  the  aforementioned 
variables  reinforce  the  conclusion  al- 
ready arrived  at  with  R.  a.  draytonii: 
Existing  data  cannot  distinguish  hy- 
potheses explaining  the  differential 
occurrence  of  R.  boylii  among  habitat 
categories  due  to  mechanistic  or 
physiological  restriction  (i.e.,  "habi- 
tat preference")  from  hypotheses  in- 
voking habitat  restriction  because  of 
some  novel  predator  (Hayes  and  Jen- 
nings 1986).  The  data  for  R.  boylii  dif- 
fer from  that  of  R.  a.  draytonii  in  that 
we  cannot  confidently  reject  the  al- 
ternative that  no  restriction  is  occur- 
ring. For  example,  it  remains  unclear 
whether  earlier  reports  of  "atypical" 
habitat  use  by  R.  boylii  were  simply 
rare  occurrences,  or  whether  those 
instances  actually  reflect  a  general 
pattern  of  broader  habitat  use  in 
years  prior  to  when  Moyle  (1973)  ob- 
tained his  data,  indicating  that  habi- 
tat restriction  had  occurred. 


Management  Implications 

Both  R.  a.  draytonii  and  R.  boylii  need 
immediate  management  considera- 
tion if  many  remaining  populations 
are  to  survive  into  the  next  century. 
Rana  a.  draytonii  is  extinct  on  the 
floor  of  the  Central  Valley,  and  is 
probably  extinct  from  over  half  of  the 
drainage  systems  in  the  Central  Val- 
ley from  where  it  was  historically  re- 
corded. We  consider  many  of  the 


153 


remaining  populations  at  risk  since 
over  half  of  the  localities  are  within 
areas  projected  to  be  flooded  by  res- 
ervoirs proposed  for  the  Coast  Range 
slope  of  the  Central  Valley  (Wemette 
et  al.  1980;  C.  J.  Brown,  Jr.,  pers. 
comm.).  Populations  at  an  additional 
10  localities  are  at  an  unknown,  but 
probably  high  level  of  risk.  Although 
these  additional  localities  will  not  be 
flooded  by  the  proposed  reservoirs, 
flooding  will  isolate  the  frogs  present 
in  small  (<10  km^)  drainage  basins 
upstream  of  the  reservoirs.  We  lack 
data  on  how  isolation  in  very  small 
drainage  basins  may  increase  the 
probability  of  extinction  (see  Fritz 
1979),  but  the  only  four  localities  iso- 
lated by  reservoirs  for  which  data 
exist  now  lack  red-legged  frogs 
(Hayes,  unpubl.  data).  California 
red-legged  frogs  were  recorded  at 
each  of  the  latter  sites  up  to  20  years 
ago,  between  one  and  five  years  after 
flooding  of  the  adjacent  reservoir 
had  taken  place.  Comparable  data  on 
the  decline  of  R.  boylii  in  the  Central 
Valley  are  lacking,  but  observations 
by  experienced  workers  indicate  that 
R.  boylii  no  longer  occurs  at  many 
localities  in  the  Central  Valley  drain- 
age basin  where  it  was  historically 
recorded  (Moyle  1973;  R.  Hansen,  D. 
Holland,  S.  Sweet,  D.  Wake,  pers. 
comm.;  Jennings,  unpubl.  data). 

Modal  habitat  requirements  for 
both  frog  species  suggested  by  exist- 
ing data  should  be  given  special  at- 
tention in  any  management  attempt. 
Since  our  comments  here  are  based 
on  data  for  both  species  in  the  Cen- 
tral Valley  of  California,  attempts  to 
apply  the  management  recommenda- 
tions we  make  to  other  areas  within 
the  geographic  range  of  each  species 
should  be  done  cautiously.  We  can- 
not overemphasize  that  preservation 
of  what  appears  to  be  the  preferred 
(modal)  habitat  condition  for  either 
species  should  be  stressed  where  it  is 
ambiguous  whether  restriction  is  due 
either  to  the  negative  impact  of  the 
introduced  aquatic  macrofauna,  or  to 
intrinsic  mechanical  or  physiological 
limitations.  Preservation  of  non-mo- 


dal habitat  is  not  only  likely  to  incur 
a  greater  cost  to  ensure  frog  survival, 
but  more  importantly,  it  may  still  not 
allow  survival  if  the  worst-case  sce- 
nario (restriction  of  habitat  by  the 
introduced  aquatic  macrofauna)  is 
true. 

The  modal  habitat  features  of  R.  a. 
draytonii  and  R.  boylii  are  similar  in 
two  ways.  First,  the  aquatic  habitat 
of  each  has  some  shading.  Yet,  shad- 
ing associated  with  California  red- 
legged  frogs  differs  because  of  the 
apparently  crucial  aspect  of  having 
dense  vegetation  at  or  near  water 
level.  We  lack  details  on  just  how  the 
streams  Moyle  (1973)  sampled  were 
shaded,  but  knowledge  of  some  of 
the  species  providing  shade  suggests 
that  a  higher  overstory  was  typical. 
Rana  a.  draytonii  will  always  be  at 
greater  risk  than  R.  boylii  where  al- 
teration of  riparian  vegetation  is  a 
problem  simply  because  of  its  shade 
requirement;  even  altered  stream  en- 
vironments may  retain  some  shad- 
ing, but  a  lesser  probability  will  al- 
ways exist  that  the  shading  that  re- 
mains will  have  the  structure  needed 
by  R.  a.  draytonii.  Second,  each  spe- 
cies occurs  most  frequently  in  the  ab- 
sence of  any  aquatic  macrofauna, 
and  both  species  have  probably  expe- 
rienced some  habitat  restriction  due 
to  introduced  aquatic  predators. 
Only  one  small  native  minnow  co- 
occurs  at  over  one-third  the  sites 
where  each  frog  species  was  re- 
corded, and  even  that  species  was 
not  positively  correlated  with  frog 
abundance.  For  R.  a.  draytonii,  the 
data  are  reasonably  convincing  that 
restriction  has  occurred  away  from 
perennial  aquatic  sites.  For  R.  boylii, 
data  do  not  clearly  indicate  habitat 
restriction.  Still,  the  fact  that  R.  boylii 
was  found  at  fewer  intermittent  sites 
leads  us  to  believe  that  if  habitat  re- 
striction has  taken  place,  it  has  oc- 
curred away  from  intermittent 
aquatic  sites.  We  reason  that  since 
riffles  disappear  seasonally  in  inter- 
mittent streams,  such  streams  lack 
the  condition  found  in  perennial 
streams  that  may  be  an  advantage  if 


riffle  habitat  is  a  refuge,  i.e.,  that  per- 
ennial streams  have  riffle  habitat 
year-round. 

CXir  analysis  indicates  that  at- 
tempts at  management  of  these  two 
frogs  should  address  at  least  three 
other  habitat  variables:  water  depth, 
stream  morphology,  and  substrate 
type.  Rana  boylii  appears  to  require  a 
shallow  water  depth  of  <0.6  m, 
whereas  R.  a.  draytonii  seems  to  re- 
quire some  water  _0.7  m  deep.  Data 
on  stream  morphology  and  substrate 
type,  which  were  recorded  only  for 
R.  boylii,  suggest  that  both  of  a  per- 
centage of  riffle  area  and  at  least 
cobble-sized  substrate  of  greater  than 
40%  best  suit  this  species.  Parallel 
data  for  R.  a.  draytonii  are  lacking, 
but  since  data  on  other  habitat  para- 
meters measured  for  R.  a.  draytonii 
are  largely  "reciprocals"  of  the  corre- 
lates of  riffle  habitat  associated  with 
R.  boylii,  we  anticipate  that  some  re- 
lationship to  the  more  lentic  water 
stream  morphology  categories  (i.e., 
pools  and  runs)  and  their  associated 
finer  substrate  categories  (i.e.,  silt 
and  sand)  will  be  demonstrated  for 
R.  a.  draytonii. 

Experiments  may  ultimately  iden- 
tify the  introduced  aquatic  predators 
likely  responsible  for  the  declines  of 
these  frogs,  but  management  based 
on  current  knowledge  should  ad- 
dress no  less  than  the  worst-case  sce- 
nario; i.e.,  that  any  member  of  the 
introduced  aquatic  macrofauna  pres- 
ents a  risk  to  the  survival  of  popula- 
tions of  R.  a.  draytonii  and  R.  boylii. 
Thus,  the  sound  management  deci- 
sion is  to  implement  measures  that 
will  maximize  the  degree  of  isolation 
between  existing  populations  of  each 
frog  species  and  any  members  of  the 
introduced  aquatic  macrofauna.  Just 
how  isolation  should  be  maintained 
will  vary  depending  on  the  site  con- 
sidered, but  some  general  sugges- 
tions can  be  made.  First,  passive 
measures  promoting  isolation  are 
preferable  because  they  are  less 
costly  and  are  less  likely  to  affect 
non-target  species.  Simply  avoiding 
habitat  modification  where  the  mo- 


154 


dal  habitat  features  for  each  frog  spe- 
cies already  exist  is  a  passive  meas- 
ure that  will  provide  some  degree  of 
within-habitat  isolation  since  mem- 
bers of  the  introduced  aquatic 
macrofauna  show  little  overlap  in 
their  habitat  requirements  with  each 
frog.  Yet,  populations  of  either  frog 
species  currently  coexisting  in  a  habi- 
tat mosaic  with  members  of  the  in- 
troduced aquatic  macrofauna  may 
still  be  doomed.  This  possibility 
leads  us  to  suggest  that  most  efforts 
at  management  should  be  spent  on 
frog  populations  at  sites  that  cur- 
rently lack  introduced  aquatic  preda- 
tors. We  consider  protection  of  the 
entire  hydrographic  basins  of  drain- 
age systems  tributaries  (see  methods 
for  definition)  an  important  part  of 
such  management  attempts  because 
intrusion  by  introduced  aquatic 
predators  is  probably  most  easily 
controlled  if  the  only  natural  access 
route  is  via  upstream  movement.  To 
our  knowledge,  no  locality  within  the 
Central  Valley  drainage  area  having 
an  extant  California  red-legged  frog 
population  has  its  entire  hydro- 
graphic  basin  protected.  Moreover, 
only  two  California  red-legged  frog 
populations  within  this  area  occur  at 
sites  where  the  habitat  is  currently 
offered  some  protection.  Second,  iso- 
lation strategies  may  differ  depend- 
ing on  whether  proximate  popula- 
tions of  introduced  aquatic  predators 
are  bullfrogs  or  fishes  or  both.  Apart 
from  being  physically  transported, 
fishes  are  effectively  prevented  from 
moving  upstream  by  a  barrier  (see 
Hayes  and  Jennings  1986),  whereas 
bullfrogs,  capable  of  overland  move- 
ment under  wet  conditions  (Hayes 
and  Warner  1985),  are  less  likely  to 
be  barrier-limited.  We  indicated  ear- 
lier that  creation  of  small  impound- 
ments may  enhance  the  ability  of  R. 
a.  draytonii  to  establish  at  certain  sites 
through  the  creation  of  features 
found  in  its  habitat,  but  attention  to 
the  positioning  of  such  impound- 
ments is  an  equally  important  con- 
siderahon.  If  impoundments  are 
close  enough  that  bullfrogs  reach 


them  from  an  adjacent  source  popu- 
lation, such  sites  can  also  act  as  local 
refuges  at  which  new  bullfrog  popu- 
lations can  become  established,  and 
can  serve  as  new  focal  points  from 
which  to  disp)erse.  Moreover,  new 
impoundments  probably  favor  the 
establishment  of  bullfrogs  simply  be- 
cause their  unvegetated  condition 
more  closely  matches  the  habitat  re- 
corded for  bullfrogs  (Moyle  1973). 
These  arguments  simply  indicate  that 
particular  attention  should  be  given 
to  avoiding  the  creation  of  "step- 
ping-stone" pathways,  i.e.,  provision 
of  access  into  currently  isolated 
drainages  by  the  positioning  of  im- 
poundments that  permit  introduced 
predators,  like  bullfrogs,  to  encroach 
progressively  by  dispersal. 

The  limits  of  our  analysis  indicate 
that  significant  aspects  of  habitat 
variation  for  both  frog  species  re- 
main to  be  understood.  In  particular, 
an  understanding  is  needed  as  to 
how  key  variables  influence  repro- 
duction and  refuge  sites.  Although 
available  data  on  oviposition  pat- 
terns suggest  a  link  between  R.  a. 
draytonii  and  the  presence  of  emer- 
gent vegetation  (Hayes  and 
Miyamoto  1984),  and  R.  boylii  and  a 
rocky  substrate  (Fitch  1936, 1938; 
Storer  1925;  Zweifel  1955),  it  is  un- 
clear for  either  species  to  what  de- 
gree the  substrate  can  vary  before 
oviposition  may  be  prevented  and 
also  how  aspects  of  reproduction  be- 
sides oviposition  may  be  linked  to 
habitat  variation.  Perhaps  the  most 
crucial  gap  is  a  lack  of  understanding 
of  what  aspects  of  habitat  variation 
are  related  to  frog  refuge  sites,  in- 
cluding the  often  temix)rary  refuges 
used  as  an  escape  from  predators  as 
well  as  those  refuges  used  during  the 
season  of  inactivity.  The  former  type 
of  refuge  site  may  be  related  to  the 
deep-water  and  dense  vegetation 
habitat  associated  with  R.  a.  draytonii, 
and  the  riffle  habitat  associated  with 
R.  boylii,  but  what  aspects  of  those 
habitat  features  really  comprise  the 
refuge  and  to  what  degree  they  may 
vary  before  they  are  no  longer  a  ref- 


uge is  unknown.  A  understanding  of 
the  latter  is  pivotal  to  the  identifica- 
tion of  predator-induced  habitat  re- 
striction. Most  importantly,  an 
understanding  of  how  reproduction 
and  refuge  sites  are  related  to  habitat 
variation  for  these  two  frogs  is  essen- 
tial if  management  is  to  ever  be  re- 
fined to  a  level  where  habitat  vari- 
ables, either  individually  or  in  con- 
cert, may  be  manipulated.  Finally,  if 
habitat  manipulations  are  attempted, 
they  will  have  to  be  implemented 
with  caution  in  aquatic  systems 
where  both  R.  a.  draytonii  and  R. 
boylii  co-occur;  differences  in  habitat 
characteristics  between  each  species 
suggest  that  whatever  way  one  or 
more  of  several  habitat  variables  are 
manipulated,  they  will  probably  re- 
sult in  a  tradeoff  between  habitat 
losses  and  habitat  gains  for  R.  a.  dray- 
tonii versus  R.  boylii. 

In  summary,  habitat  analysis  for 
the  two  ranid  frogs,  R.  a.  draytonii 
and  R.  boylii,  indicates  that  each  spe- 
cies is  most  frequently  associated 
with  discemibly  different  aquatic 
habitats,  the  former  with  densely 
vegetated,  deep  water  and  the  latter 
with  rocky,  shallow-water  riffles  in 
streams.  The  species  are  similar  in 
that  they  infrequently  co-occur  with 
any  aquatic  vertebrates,  especially 
the  introduced  aquatic  macrofauna. 
Low  levels  of  co-occurrence  between 
frogs  and  the  introduced  aquatic 
macrofauna  have  two  confounded 
explanations:  1)  preferential  use  of 
different  habitats  between  the  intro- 
duced aquatic  macrofauna  and  frogs, 
and  2)  habitat  restriction  because 
frogs  and  their  life  stages  are  preyed 
upon  by  the  introduced  aquatic 
macrofauna.  However,  even  though 
it  is  presently  impossible  to  idenrify 
the  responsible  predator,  temporal 
data  strongly  suggest  that  R.  a.  dray- 
tonii has  been  restricted  by  some  in- 
troduced aquatic  predator  and  the 
same  possibility  cannot  be  excluded 
for  R.  boylii.  For  t>oth  species,  a  man- 
agement scheme  is  necessary  to  avert 
existing  trends  of  decline,  and  ulti- 
mately, extinction.  A  management 


155 


scheme  that  minimizes  the  risk  of  ex- 
tinction based  on  current  data  must 
address  the  worst-case  scenario 
among  the  ahernatives  imphcated  in 
limiting  frog  distributions.  To  ad- 
dress anything  less  increases  the  risk 
of  extinction  if  that  alternative  is 
true.  Since  that  alternative  is  habitat 
restriction  by  an  introduced  aquatic 
macrofauna,  management  should 
strive  to  isolate  both  frog  species 
from  the  introduced  aquatic  macro- 
fauna.  Moreover,  available  data  indi- 
cate that  preservation  of  modal  con- 
ditions for  habitat  variables  identi- 
fied as  associated  with  each  species  is 
a  suitable  interim  strategy,  since  it  is 
more  likely  to  promote  isolation.  Sig- 
nificant refinements  of  this  manage- 
ment scheme  will  require  a  thorough 
understanding  of  how  habitat  vari- 
ables associated  with  each  frog  spe- 
cies are  linked  to  their  refuge  re- 
quirements and  their  reproductive 
patterns. 

ACKNOWLEDGMENTS 

Special  thanks  go  to  Charles  J. 
Brown,  Jr.,  Peter  B.  Moyle,  and 
David  B.  Wake  for  allowing  us  to  use 
data  in  their  care. 

Sean  J.  Barry,  John  M.  Brode, 
Charles  W.  Brown,  Mark  L. 
Cay  wood,  Henry  E.  Childs,  Jr., 
Arthur  L.  Cohen,  Nathan  W.  Cohen, 
Lawrence  R.  Cory,  John  B.  Cowan, 
Robert  G.  Crippen,  Henry  S.  Fitch, 
William  J.  Hamilton,  Jr.,  George  H. 
Hanley,  George  E.  Hansen,  Robert 
W.  Hansen,  John  Hendrickson  (Woo- 
dleaf,  Calif.),  John  R.  Hendrickson 
(University  of  Arizona),  Daniel  C. 
Holland,  Samuel  B.  Horowitz,  Alex- 
ander K.  Johnson,  William  F. 
Johnson,  Donald  R.  Kirk,  J.  Ralph  Li- 
chtenfels.  Amy  R.  McCune,  Roy  W. 
McDiarmid,  Milton  D.  Miller,  Rich- 
ard R.  Montanucci,  Garth  I.  Murphy, 
Robert  T.  Orr,  Thomas  L.  Rodgers, 
Stephen  B.  Ruth,  Robert  C.  Stebbins, 
the  late  Ruth  R.  Storer,  Samuel  S. 
Sweet,  Richard  Terry,  Walter  Tor- 
doff,  Jens  V.  Vindum,  Conrad 


Yamamoto,  and  Richard  G.  Zweifel 
all  contributed  ancillary  data.  Addi- 
tionally, important  data  were  ex- 
tracted from  the  unpublished  field 
notes  or  voucher  specimens  collected 
by  the  following  workers  no  longer 
living:  Adrey  E.  Borell,  Harold  C. 
Bryant,  Charles  L.  Camp,  Joseph  S. 
Dixon,  Adolphus  L.  Heermann, 
Henry  W.  Henshaw,  Carl  L.  Hubbs, 
Lloyd  G.  Ingles,  Henry  C.  Kellers, 
William  N.  Lockington,  Donald  R. 
McLean,  Joseph  R.  Slevin,  Tracy  I. 
Storer,  John  Van  Denburgh,  and  Al- 
bert H.  Wright.  Phyllis  A.  Buck,  Peter 
B.  Moyle,  C.  Mindy  Nelson,  and 
Richard  G.  Zweifel  kindly  reviewed 
the  manuscript. 

LITERATURE  CITED 

Abdulali,  Humayun.  1985.  On  the 
export  of  frog  legs  from  India. 
Journal  of  the  Bombay  Natural 
History  Society  82:347-375. 

Armour,  C.  L.,  K.  P.  Bumham,  and 
W.  S.  Platts.  1983.  Field  methods 
and  statistical  analyses  for  moni- 
toring small  salmonid  streams.  200 
p.  United  States  Fish  and  Wildlife 
Service,  FWS/OBS-83/33. 

Bailey,  James  A.  1984.  Principles  of 
wildlife  management.  373  p.  John 
Wiley  and  Sons,  Inc.,  New  York. 

Baker,  RoUin  H.  1942.  The  bullfrog:  A 
Texas  wildlife  resource.  Texas 
Game,  Fish  and  Oyster  Commis- 
sion, Bulletin  (23):3-7. 

Brocke,  R.  H.  1979.  The  name  of 
nongame.  Wildlife  Society  Bulletin 
7:279-282. 

Bury,  R.  Bruce.  1975.  Conservation  of 
non-game  wildlife  in  California:  A 
model  programme.  Biological 
Conservation  7:199-210. 

Bury,  R.  Bruce,  Howard  W. 

Campbell,  and  Norman  J.  Scott,  Jr. 
1980a.  Role  and  importance  of 
nongame  wildlife.  197-207  p.  In 
Transactions  of  the  45th  North 
American  Wildlife  and  Natural 
Resources  Conference,  Wildlife 
Management  Institute,  Washing- 
ton, D.C. 


Bury,  R.  Bruce,  C.  Kenneth  Dodd,  Jr., 
and  Gary  M.  Fellers.  1980b.  Con- 
servation of  the  Amphibia  of  the 
United  States:  A  review.  Depart- 
ment of  the  Interior,  United  States 
Fish  and  Wildlife  Service,  Re- 
source Publication  (134):l-34. 

Bury,  R.  Bruce,  and  Jill  A.  Whelan. 
1984.  Ecology  and  management  of 
the  bullfrog.  Department  of  the 
Interior,  United  States  Fish  and 
Wildlife  Service,  Resource  Publica- 
tion (155):l-23. 

Chamberlain,  F.  M.  1898.  Notes  on 
the  edible  frogs  of  the  United 
States  and  their  artificial  propaga- 
tion. 249-261  p.  In  Report  of  the 
United  States  Commissioner  of 
Fish  and  Fisheries  for  the  Year 
Ending  June  30, 1897.  United 
States  Government  Printing  Of- 
fice, Washington,  D.C. 

Childs,  Henry  E.,  Jr.,  and  Walter  E. 
Howard.  1955.  The  vertebrate 
fauna  of  the  San  Joaquin  Experi- 
mental Range.  United  States  De- 
partment of  Agriculture,  Califor- 
nia Range  and  Experiment  Station, 
Miscellaneous  Paper  (19):l-20. 

Cowan,  J.  B.  1979.  Mammals,  reptiles 
and  amphibians  of  Gray  Lodge 
Wildlife  Area.  12  p.  Publication  of 
the  Department  of  Fish  and  Game, 
Resources  Agency,  State  of  Cali- 
fornia, Sacramento. 

Cresswell,  Richard.  1862.  Aristotle's 
history  of  animals  in  ten  books. 
326  p.  Henry  G.  Bohn,  York  Street, 
Covent  Garden,  London. 

Cunningham,  John  D.  1955.  Notes  on 
abnormal  Rjina  aurora  draytonii. 
Herpetologica  11:149. 

Dodd,  C.  Kenneth,  Jr.  1978.  Amphibi- 
ans and  reptiles:  the  declining  spe- 
cies. Water  Spectrum  10:24-32. 

Fitch,  Henry  S.  1936.  Amphibians 
and  reptiles  of  the  Rogue  River 
Basin,  Oregon.  American  Midland 
Naturalist  17:634-652. 

Fitch,  Henry  S.  1938.  Rana  hoylii  in 
Oregon.  Copeia  1938:148. 

Fitch,  Henry  S.  1949.  A  study  of 
snake  populations  in  central  Cali- 
fornia. American  Midland  Natu- 
ralist 41:513-579. 


156 


Fritz,  R.  S.  1979.  Consequences  of  in- 
sular population  structure:  distri- 
bution and  extinction  of  spruce 
grouse  (Camchites  canadensis) 
populations.  Oecologia  42:57-66. 

Gordon,  Kenneth.  1939.  The  amphib- 
ian and  reptiles  of  Oregon.  Oregon 
State  College  Monographs  (l):l-82. 

Grinnell,  Joseph,  Joseph  Dixon,  and 
Jean  M.  Linsdale.  1930.  Vertebrate 
natural  history  of  a  section  of 
northern  California  through  the 
Lassen  Peak  region.  University  of 
California  Publications  in  Zoology 
35:1-594. 

Grinnell,  Joseph,  and  Tracy  I.  Storer. 
1924.  Animal  life  in  the  Yosemite; 
an  account  of  the  mammals,  birds, 
reptiles,  and  amphibians  in  a 
cross-section  of  the  Sierra  Nevada. 
752  p.  University  of  California 
Press,  Berkeley. 

Hallowell,  Edward.  1854.  Descrip- 
tions of  new  reptiles  from  Califor- 
nia. Proceedings  of  the  Academy 
of  Natural  Sciences  of  Philadel- 
phia 7:91-97. 

Hallowell,  Edward.  1859.  Report  on 
the  reptiles  collected  on  the  Sur- 
vey. 27  p.  Pacific  Railroad  Reports 
10  (part  4,  Williamson's  Report), 
zoology  report  no.  1. 

Hayes,  Marc  P.,  and  Mark  R.  Jen- 
nings. 1986.  DecHne  of  ranid  frog 
species  in  western  North  America: 
Are  bullfrogs  (Rana  catesbeiana) 
responsible?  Journal  of  Herpetol- 
ogy  20:490-509. 

Hayes,  Marc  P.,  and  Dana  M.  Krem- 
pels.  1986.  Vocal  sac  variation 
among  frogs  of  the  genus  Rana 
from  western  North  America. 
Copeia  1986:927-936. 

Hayes,  Marc  P.,  and  Michael  M. 
Miyamoto.  1984.  Biochemical,  be- 
havioral and  body  size  differences 
between  Rana  aurora  aurora  and  R. 
a.  draytoni.  Copeia  1984:1018-1022. 

Hayes,  Marc  P.,  and  Joan  Warner. 
1985.  Life  history  notes:  Rana 
catesbeiana,  food.  SSAR  Herpeto- 
logical  Review  16:109. 

Hine,  Ruth  L.,  Betty  L.  Les,  and 
Bruce  F.  Hellmich.  1981.  Leopard 
frog  populations  and  mortality  in 


Wisconsin.  Department  of  Natural 
Resources,  Madison,  Wisconsin, 
Technical  Bulletin  (122):l-39. 

Holland,  Robert  F.  1978.  The  geo- 
graphic and  edaphic  distribution 
of  vernal  pools  in  the  Great  Valley, 
California.  California  Native  Plant 
Society  (4):1-12. 

Honegger,  Rene  E.  1981.  Threatened 
Amphibians  and  Reptiles  in  Eu- 
rope. 158  p.  Akademie  Verlagsges, 
Wiesbaden,  Germany. 

Husain,  K.  Z.,  and  M.  M.  Rahman. 
1978.  The  amphibian  fauna  of 
Bangladesh.  Bangladesh  Journal  of 
Zoology  6:157-158. 

Ingles,  Lloyd  G.  1932a.  Four  new 
species  of  Haematoloechus  (Trema- 
toda)  from  Rana  aurora  draytonii 
from  California.  Univ.  of  Calif. 
Publications  in  Zool.  37(7):189-201. 

Ingles,  Lloyd  G.  1932b.  Cephalogoni- 
mus  brevicirrus,  a  new  species  of 
Trematode  from  the  intestines  of 
Rarm  aurora  from  California.  Uni- 
versity of  California  Publications 
in  Zoology  37(8):203-210. 

Ingles,  Lloyd  G.  1933.  The  specificity 
of  frog  flukes.  Science  78:168. 

Ingles,  Lloyd  G.  1936.  Worm  para- 
sites of  California  amphibia. 
Transactions  of  the  American 
Microscopical  Society  55:73-92. 

Jain,  Subodh.  (editor)  1976.  Vernal 
pools:  their  ecology  and  conserva- 
tion. A  symposium  sponsored  by 
the  Institute  of  Ecology,  Univer- 
sity of  California,  Davis,  May  1 
and  2, 1976.  Institute  of  Ecology 
Publication  (9):l-93. 

Jennings,  Mark  R.,  and  Marc  P. 
Hayes.  1985.  Pre-1900  overharvest 
of  California  red-legged  frogs 
(Rana  aurora  draytonii):  The  in- 
ducement for  bullfrog  (Rana 
catesbeiana)  introduction.  Herpetol- 
ogica  41:94-103. 

Kellert,  Stephen  R.  1985.  Social  and 
perceptual  factors  in  endangered 
sp)ecies  management.  Journal  of 
Wildlife  Management  49:528-536. 

Leidy,  R.  A.  1984.  Distribution  and 
ecology  of  stream  fishes  of  the  San 
Francisco  Bay  drainage.  Hilgardia 
52(8):1-175. 


Leopold,  Aldo  S.  1931 .  A  history  of 
ideas  in  game  management.  Out- 
door America  (June):22-24,38- 
39,47. 

Leopold,  Aldo  S.  1933.  Game  man- 
agement. 481  p.  Charles  Scribner's 
Sons,  New  York. 

McAuliffe,  J.  R.  1978.  Biological  sur- 
vey and  management  of  sport- 
hunted  bullfrog  populations  in 
Nebraska.  78  p.  Nebraska  Game 
and  Parks  Commission,  Lincoln, 
Nebraska. 

Moyle,  Peter  B.  1973.  Effects  of  intro- 
duced bullfrogs,  Rana  catesbeiana, 
on  the  native  frogs  of  the  San 
Joaquin  Valley,  California.  Copeia 
1973:18-22. 

Moyle,  Peter  B.  1976.  Inland  fishes  of 
California.  405  p.  University  of 
California  Press,  Berkeley. 

Moyle,  Peter  B.,  and  Robert  D. 
Nichols.  1973.  Ecology  of  some 
native  and  introduced  fishes  of  the 
Sierra  Nevada  Foothills  in  Central 
California.  Copeia  1973:478-490. 

Moyle,  Peter  B.,  and  Robert  D. 
Nichols.  1974.  Decline  of  the  fish 
fauna  of  the  Sierra  Nevada  foot- 
hills. Central  California.  American 
Midland  Naturalist  92:72-83. 

Moyle,  Peter  B.,  Jerry  J.  Smith, 
Roberts  A.  Daniels,  Thomas  L. 
Taylor,  Donald  G.  Price,  and 
Donald  M.  Baltz.  1982.  Distribu- 
tion and  ecology  of  stream  fishes 
of  the  Sacramento-San  Joaquin 
drainage  system,  California.  Uni- 
versity of  California  Publications 
in  Zoology  115:1-256. 

Neill,  Wilfred  T.  1974.  Reptiles  and 
amphibians  in  the  service  of  man. 
248  p.  Biological  Sciences  Curricu- 
lum Study  Book,  Pegasus  Press, 
The  Bobbs-Merrill  Company,  Inc., 
Indianapolis,  Indiana. 

Pister,  E.  P.  1976.  A  rationale  for  the 
management  of  nongame  fish  and 
wildlife.  Fisheries  1:11-14. 

Rutter,  Cloudsley.  1908.  The  fishes  of 
the  Sacramento-San  Joaquin  Basin, 
with  a  study  of  their  distribution 
and  variation.  Bulletin  of  the 
United  States  Bureau  of  Fisheries 
27:103-152. 


157 


Slevin,  Joseph  R.  1928.  The  amphibi- 
ans of  western  North  America. 
Occasional  Papers  of  the  Califor- 
nia Academy  of  Sciences  (16):1- 
144. 

Sokal,  Robert  R.,  and  R.  James  Rohlf. 
1981.  Biometry.  687  p.  W.  H.  Free- 
man and  Co.,  San  Francisco. 

Stebbins,  Robert  C.  1966.  A  field 
guide  to  western  reptiles  and  am- 
phibians. 279  p.  Houghton  Mifflin 
Company,  Boston. 

Stebbins,  Robert  C.  1985.  A  field 
guide  to  western  reptiles  and  am- 
phibians. Second  edition,  revised. 
336  p.  Houghton  Mifflin  Com- 
pany, Boston. 

Storer,  Tracy  1. 1925.  A  synopsis  of 
the  Amphibia  of  California.  Uni- 
versity of  California  Publications 
in  Zoology  27:1-342. 

Storer,  Tracy  1. 1933.  Frogs  and  their 
commercial  use.  California  Fish 
and  Game  19:203-213. 

Strahler,  A.  N.  1957.  Quantitative 
analysis  of  watershed  geomor- 
phology.  Transactions  of  the 
American  Geophysicists  Union 
38:913-920. 

Sweet,  Samuel  S.  1983.  Mechanics  of 
a  natural  extinction  event:  Ram 
boylii  in  southern  California.  Ab- 
stract of  a  paper  presented  at  the 
26th  annual  meeting  of  the  Society 
for  the  Study  of  Amphibians  and 
Reptiles,  and  the  31st  annual  meet- 
ing of  the  Herpetologists'  League, 
University  of  Utah,  August  7-12, 
1983. 

Treanor,  Robert  R.  1975a.  Manage- 
ment of  the  bullfrog  (Ram  catesbe- 
iam)  resource  in  California.  Cali- 
fornia Department  of  Fish  and 
Game,  Inland  Fisheries  Adminis- 
trative Report  (75-l):l-30. 

Treanor,  Robert  R.  1975b.  Manage- 
ment of  the  bullfrog  resource  in 
California.  Cal-Neva  Wildlife 
Transactions  1975:81-92. 

Treanor,  Robert  R.,  and  Stephen  J. 
Nicola.  1972.  A  preliminary  study 
of  the  commercial  and  sporting 
utilization  of  the  bullfrog.  Ram 
catesbeiam  Shaw,  in  California. 
California  Department  of  Fish  and 


Game,  Inland  Fisheries  Adminis- 
trative Report  (72-4):l-23. 

United  States  Geological  Survey. 
1970a.  Surface  water  supply  of  the 
United  States  1961-1965.  Part  II. 
Pacific  slope  basin  in  California. 
Volume  3.  Southern  Central  Valley 
basins.  U.  S.  Geological  Survey, 
Water-Supply  Paper  (1930):1-155. 

United  States  Geological  Survey. 
1970b.  Surface  water  supply  of  the 
United  States  1961-1965.  Part  II. 
Pacific  slope  basin  in  California. 
Volume  4.  Northern  Central  Val- 
ley basin.  Uruted  States  Geological 
Survey,  Water-Supply  Paper 
(1931):1-111. 

Walker,  M.  V.  1946.  Reptiles  and 
amphibians  of  Yosemite  National 
Park.  Yosemite  Nature  Notes  25:1- 
48. 

Wemette,  F.  G.,  F.  A.  Hall,  Jr.,  C.  J. 
Brown,  Jr.,  C.  L.  Mayer,  and  N.  A. 
Villa.  1982.  Los  Vaqueros  project — 
fish  and  wildlife  impacts.  263  p. 
California  Department  of  Fish  and 
Game,  Status  Report. 

Williamson,  R.  S.  1855.  Report  of 
Lieutenant  R.  S.  Williamson.  43  p. 
In  Report  of  Exploration  in  Cali- 
fornia for  Railroad  Routes,  to  con- 
nect with  the  Routes  Near  the  35th 
and  32nd  Parallels  of  North 
Latitude;  In  Report  of  Explora- 
tions and  Surveys  to  Ascertain  the 
Most  Practicable  and  Economical 
Route  for  a  Railroad  from  the  Mis- 
sissippi River  to  the  Pacific  Ocean, 
Made  Under  the  Direction  of  the 
Secretary  of  War,  in  1853-54.  Vol- 
ume 5,  part  I.  Beverly  Tucker, 
Washington,  D.C.  (1856)  [Senate 
Executive  Document  No.  78, 33rd 
Congress,  2nd  Session]. 

Willis,  Yuell  L.,  Don  L.  Moyle,  and 
Thomas  S.  Baskett.  1956.  Emer- 
gence, breeding,  hibernation, 
movements  and  transformation  of 
the  bullfrog.  Ram  catesbeiam,  in 
Missouri.  Copeia  1956:30-41. 

Wright,  A.  H.  1920.  Frogs:  their  natu- 
ral history  and  utilization.  Depart- 
ment of  Commerce,  United  States 
Bureau  of  Fisheries  Document 
(888):l-44. 


Wright,  Albert  Hazen,  and  Anna  Al- 
len Wright.  1949.  Handbook  of 
frogs  and  toads  of  the  United 
States  and  Canada.  Third  edition. 
640  p.  Comstock  Publishing  Com- 
pany, Inc.,  Ithaca,  New  York. 

Zahl,  Paul  A.  1967.  In  quest  of  the 
world's  largest  frog.  National 
Geographic  Magazine  132:146-152. 

Zar,  Jerrold  H.  1974.  Biostatistical 
analysis.  Prentice-Hall,  Inc.  620  p. 
Englewood  Cliffs,  New  Jersey. 

Zedler,  Paul  H.  1987.  The  ecology  of 
southern  California  vernal  pools:  a 
community  profile.  Department  of 
the  Interior,  United  States  Fish 
and  Wildlife  Service,  Biological 
Report  85(7.11):1-136. 

Zweifel,  Richard  G.  1955.  Ecology, 
distribution,  and  systematics  of 
frogs  of  the  Ram  boylei  group. 
University  of  California  Publica- 
tions in  Zoology  54(4):207-292. 


158 


Integrating  Anuran 
Amphibian  Species  into 
Environmental  Assessment 
Programs^ 

Ronald  E.  Beiswenger^ 


Abstract.— Anurans  are  often  given  minimal 
attention  in  environmental  assessments  despite  their 
ecological  importance  and  potential  value  as 
indicator  species.  Habitat  and  guild-based  models 
must  be  adopted  to  include  all  life  cycle  stages  of 
anurans.  A  preliminary  habitat  suitability  model  for 
the  American  toad  shows  how  this  con  be 
accomplished. 


As  a  result  of  our  increased  under- 
standing of  the  roles  of  wildlife  spe- 
cies in  ecosystem  structure  and  func- 
tion, and  legal  requirements  to  de- 
velop holistic  approaches  to  environ- 
mental management,  it  has  become 
increasingly  common  to  include  all 
species  of  wildlife  in  resource  inven- 
tories and  monitoring  programs 
(Chalk  et  al.  1984).  However,  am- 
phibians are  often  ignored  or  given 
minimal  attention  in  such  programs, 
even  though  they  are  important 
wildlife  resources  and  should  be 
given  serious  consideration  in  man- 
agement evaluations  (Bury  and  Ra- 
phael 1983,  Bury  et  al.  1980,  Jones 
1986).  If  included  in  resource  evalu- 
ations at  all,  amphibians  are  usually 
lumped  with  reptiles  in  a  category 
called  herpetofauna  and  even  then 
are  often  only  represented  as  items  in 
a  species  list. 

This  is  unfortunate  because,  in 
addition  to  their  ecological  impor- 
tance, anurans  are  potentially  valu- 
able as  a  unique  form  of  indicator 
species  capable  of  integrating  envi- 
ronmental changes  occurring  in  both 
the  terrestrial  and  aquatic  phases  of 
their  habitats.  Furthermore,  because 
they  occupy  small  ponds  and  the 
shallow  margins  of  lakes,  anurans 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles  and  Small 
Mammals  in  North  America.  (Flagstaff.  AZ. 
July  19-21.  1988.) 

'^Ronald  E.  Beiswenger  is  Professor.  De- 
partment of  Geography  and  Recreation. 
The  University  of  Wyoming.  Laramie,  WY 
82071. 


are  likely  to  be  the  first  vertebrates  to 
come  in  contact  with  contaminated 
run-off  or  acidified  snowmelt.  This 
could  make  them  useful  as  elements 
of  an  early  warning  system  for  the 
detection  of  environmental  contami- 
nation. Campbell  (1976)  found  that 
the  boreal  toad,  Bufo  boreas,  would  be 
an  especially  effective  indicator  spe- 
cies for  monitoring  the  impact  of 
cloud  seeding  in  the  mountains  of 
Colorado.  It  is  also  significant  that 
many  anurans  require  specialized 
habitats  in  wetland  areas  and  ripar- 
ian zones,  and  could  serve  as  indica- 
tor species  for  the  overall  health  of 
these  areas  of  special  ecological  im- 
portance. 

Despite  their  potential  usefulness, 
there  are  several  reasons  why  am- 
phibians are  not  given  adequate  at- 


tention in  environmental  assess- 
ments. The  importance  of  amphibi- 
ans in  ecosystems  is  generally  unrec- 
ognized, particularly  by  the  general 
public  and  the  resource  managers 
who  must  respond  to  the  desires  of 
this  public  as  they  set  management 
priorities.  Also,  the  secretive  habits 
during  the  non-breeding  season,  and 
complex  life  cycles  of  amphibians 
make  them  relatively  difficult  to 
study.  Consequently,  the  natural  his- 
tory of  many  amphibian  species  is 
not  well  known.  Another  factor  is 
that  current  models  for  monitoring 
and  assessment  have  been  developed 
for  either  terrestrial  or  aquatic  spe- 
cies and  have  not  been  adapted  to 
species  with  divergent  life  cycle 
stages  which  depend  on  both  aquatic 
and  terrestrial  habitats  (table  1). 


Table  1  .—  Habitat  components  and  life  cycle  stages  of  anurans. 


Habitat 
component 


Eggs/Pre- 
feeding 
tadpoles 


Feeding  IVIetamorphosing 

tadpoles        tadpoles  Juveniles 


Adults 


Aquatic  Phase 

Spawning  sites  X 

Tadpole  habitat  X 

Aquatic/Terrestrial  Interface  Phase 

Tadpole  habitat 

Juvenile  habitat 

Terrestrial  Phase 

Summer  habitat 

Hibernation  sites 

Movement  corridors 

Interspersion  Factors 

Distribution  of  habitat  components 

Density  of  habitat  components 


X 


X 

X 
X 


X 

X 
X 
X 


X 
X 
X 

X  X 
X  X 


159 


Approaches  for  incorporating 
wildlife  into  resource  evaluations  in- 
clude inventories  of  relative  abun- 
dance and  species  richness,  develop- 
ment of  databases,  the  use  of  indica- 
tor species,  and  the  development  of 
species  diversity  indices  and  models 
using  guild  concepts.  However,  the 
application  of  these  approaches  to 
species  of  Amphibia  has  not  kept 
pace  with  applications  to  other  spe- 
cies of  vertebrates. 

The  primary  purpose  of  this  paper 
is  to  suggest  ways  to  use  single  spe- 
cies models,  and  models  which  use 
guilds  and  habitat  structure,  to  more 
effectively  integrate  anuran  amphibi- 
ans into  resource  assessments.  A 
single  species  model  for  the  Ameri- 
can toad,  stressing  the  importance  of 
tadpole  habitat,  is  presented  in  some 
detail. 


Models  for  Anurons 


Guilds  and  Habitat  Structure 

Guild-based  environmental  assess- 
ments are  especially  useful  from  an 
ecological  perspective,  although  they 
are  most  effective  when  used  in  com- 
bination with  other  methods  (Karr 
1987).  Unfortunately,  when  amphibi- 
ans are  included  in  guild-based  pro- 
grams they  are  usually  considered 
too  simplistically.  A  common  proce- 
dure is  to  categorize  them  according 
to  their  general  spawning  and  feed- 
ing habitat,  but  to  include  no  further 
detail  (e.g.  see  Thomas  et  al.  1979). 

The  habitat  models  developed  for 
Arizona  (Short  1984)  represent  a 
good  starting  point  for  producing 
effective  models  for  anurans.  In  these 
models  wildlife  guilds  are  used  to 
correlate  habitat  use  with  habitat 
structure  (layers)  by  associating  a 
species  with  a  particular  plant  com- 
munity (habitat  or  cover  type),  and 
then  with  a  habitat  layer.  Layers  of 
both  terrestrial  and  aquatic  habitat 
are  included. 


This  system  is  as  appropriate  for 
terrestrial  adult  anurans  as  it  is  for 
any  small,  terrestrial  vertebrate. 
However,  the  aquatic  phases  of  the 
model  require  further  development  if 
it  is  to  be  used  with  the  aquatic  larval 
stages  of  amphibians.  The  adaptive 
significance  of  the  tadpole  stage  has 
been  established  by  Wassersug  (1975) 
and  Wilbur  (1980),  and  it  is  clear  that 
the  habitat  requirements  of  larval 
anurans  should  be  an  important 
component  of  habitat  models.  The 
selection  of  a  spawning  site  that  will 
provide  high  quality  habitat  for  the 
tadpole  stage  is  likely  to  be  critical  to 
the  evolutionary  success  of  an  anu- 
ran species. 


Single  Species  Models 

Habitat  models  for  indicator  species 
have  been  developed  by  the  U.S.  Fish 
and  Wildlife  Service  (1981),  the  U.S. 
Forest  Service  (Berry  1986)  and  oth- 
ers (e.g.  Clawson  et  al.  1984)  for  use 
in  assessing  environmental  impacts 
and  in  making  management  deci- 
sions. A  comprehensive  habitat 
model  for  an  anuran  species  must 
encompass  spawning  sites,  tadpole 
habitat,  metamorphic  sites,  juvenile 
and  adult  feeding  habitat,  movement 
corridors  and  hibernation  sites.  For 
example,  a  model  developed  for  the 
bullfrog  (Rana  catesbeiana)  illustrates 
how  the  approach  can  be  applied  to 


Table  2.— Components  of  habitat  for  Bufo  americanus  (measurable  attrib- 
ute in  parentheses). 

Spawning  Habitat 

Shallow,  emphemeral  ponds  (depth  range) 
m^m::  Emergent  or  submergent  vegetation  (%  cover) 
■    Exposure  to  direct  sunlight  (%  of  area  shaded) 

Tadpole  Habitat 

Ponds  with  access  to  shallow  shoreline  areas  (<  10  cm)  and  to 

deeper  areas  (10-100  cm) 
Substrates  with  food  W-:W-r 

periphyton  (%  cover) 

bottom  areas  with  detritus  or  microorganisms  (%  cover) 
Microorganisms  suspended  in  water  column  (density) 
Exposure  to  direct  sunlight  (%  of  area  shaded) 

Metamorphic  Habitat 

Shallow  depth  gradient  at  shoreline  (<  1 0  cm) 
Exposure  to  direct  sunlight  (%  of  area  shaded) 
Moist  substrate  on  shore  (moisture  content) 
Vegetative  cover  on  shore  (%  cover) 

Juvenile  and  Adult  Habitat 

Availability  of  insect  and  other  invertebrate  prey  (prey  density) 
Access  to  moist  substrates  and  refugia  (moisture  content  and  refu- 
gia  density) 

Access  to  vegetative  cover  (distance  to  cover) 

Hibernation  Site 

Unoccupied  animal  burrows  (burrow  density) 

Friable  soils  (soil  texture) 

Root  zones  of  large  trees  (large  tree  density) 

Interspersion 

Movement  corridors  between  hibernation  and  spawning  sites  (distri- 
bution of  continuous  open  areas  with  adequate  cover) 

Distribution  and  density  of  potential  spawning  sites  within  the  home 
range  of  the  population  (density  of  spawning  sites) 


160 


an  anuran  species  that  is  primarily 
aquatic  (Graves  and  Anderson  1987). 
While  this  model  is  well  constructed, 
a  different  modeling  approach  would 
be  needed  for  anurans  with  terres- 
trial adult  stages.  A  limitation  of  the 
bullfrog  model  is  that  the  habitat  re- 
quirements of  the  tadpole  stage  are 
not  given  in  sufficient  detail.  This  is 
important  because  the  larval  stage 
(up  to  three  years  in  duration)  repre- 
sents a  significant  proportion  of  a 
bullfrog's  total  lifespan. 

A  different  array  of  habitat  com- 
ponents for  a  species  that  is  predomi- 

HABITAT  VARIABLES 


nantly  terrestrial  is  an  adult,  the 
American  toad  (Bufo  americanus)  is 
outlined  in  table  2.  This  outline  is 
based  on  extensive  field  studies  in 
Michigan  (Beiswenger  1975, 1977), 
field  observations  of  related  toad 
species  in  Oregon  and  Wyoming 
(Beiswenger  1978,  1981,  1986),  and 
information  found  in  the  literature. 

Including  the  terrestrial  features  of 
toad  habitat  in  assessments  does  not 
represent  a  particularly  difficult  chal- 
lenge because  these  features  can  be 
described  using  well-established  ap- 
proaches developed  for  other  small 

COMPONENTS 


Percent  of  water  area  1  m  or  less 
in  depth  (VI) 

Percent  cover  of  rooted  aquatic 
vegetation  (V2) 

Percent  of  shoreline  v/ith  shading 
riparian  vegetation  (V3) 

Percent  of  shoreline  v/ith  strip  of 
invegetated  shallow  water  (V4) 


Percent  of  shoreline  with  terrestrial 
vegetative  cover  or  ground  debris 
within  1  m  of  water  (V5) 

Percent  tree  canopy  closure  (V5) 

Percent  of  trees  that  are  deciduous 
species  (V7) 

Percent  herbaceous  canopy  cover  (V8) 

Number  of  burrows,  decaying  logs,  and 
debris  objects  larger  than  20  cm  in 
diameter  on  the  ground  (V9) 

Distance  along  a  protected  dispersal 
corridor  to  potential  spawning 
sites  (V10) 


_ Aquatic  cover/ 
reproduction 


Jerrestrial  cover/ 
hibernation 


Interspersion 


Figure  1  .—Relationships  of  habitat  variables  to  components  of  an  HSI  model  for  the  Ameri- 
can toad. 


vertebrates  that  live  on  and  below 
the  surface  of  the  ground.  However, 
tadpole  habitat  is  also  important  and 
must  be  incorporated  into  habitat  as- 
sessment procedures.  This  is  some- 
what more  challenging  because  less 
is  known  about  tadpole  ecology  and 
techniques  for  describing  tadpole 
habitat  are  not  well  developed. 

A  Habitat  Model  for  the  American 
Toad 

A  preliminary  version  of  a  habitat 
suitability  model  for  the  American 
toad  is  described  here  to  show  how 
the  requirements  of  all  life  cycle 
stages  could  be  incorporated  into 
such  a  model  (figs.  1  and  2).  The 
model  includes  10  variables  and  is 
based  primarily  on  the  author's  expe- 
rience and  a  partial  literature  review. 
Consequently,  the  model  should  be 
refined  through  a  more  extensive 
analysis  of  the  literature  and  a  peer 
review  process  before  it  is  field 
tested. 

The  habitat  requirements  of 
spawning  adults  and  tadpoles  are 
included  in  the  aquatic  cover/repro- 
ductive component  of  the  model.  The 
quality  of  spawning  sites  selected  by 
American  toads  is  influenced  by 
structural  features  such  as  depth  gra- 
dients and  vegetation.  Adult  toads 
typically  lay  their  eggs  in  shallow, 
unshaded,  vegetated  areas  (variables 
2  and  3),  distributing  them  in  strands 
on  the  vegetation.  At  first  the  newly 
hatched  tadpoles  do  not  feed,  but 
remain  at  the  site  where  the  eggs 
were  laid. 

Older  tadpoles  are  active  swim- 
mers and  display  a  variety  of  feeding 
modes  that  arc  influenced  to  a  large 
measure  by  structural  features  of  the 
habitat  (e.g.  aquatic  vegetation  and 
depth  gradients)  (variables  1,  2,  and 
4).  Wassersug  (1975)  has  shown  that 
tadpoles  are  essentially  non-discrimi- 
nant suspension  feeders,  although 
they  use  a  variety  of  means  for  ob- 
taining food.  Tadpoles  of  the  Ameri- 
can toad  most  commonly  graze 


161 


t  0.5- 


100!? 


PERCEN"0-  WATER 
AREA  I  n  OR  LE55 
IN  DEPTH 
(Variable  1) 


m% 


PERCENT  SHORELINE 
WITH  SHADING  RIPARIAN 
VEGETATION 
(variaole  3) 


PERCFHT  f.OVFR  OF 
ROOTED  ACUATIC 
VEGETATION 
(Variable  2) 


TO-i 


05- 


l(50« 


PERCENT  SHORELINE 
WITH  30-50  cn  WIDE 
STRIP  OF  UNVFGFTATFD 

SHALLOW  WATER 
;iOcn  OR  LESS  DEEP) 
(Variable  4) 


PERCENT  SHORELINE 
WITH  VEGETATIVE 
COVER  OB  GROUND 
DEBRIS  WITHIN  1  m 
0-  WATER 
(Variable  5) 


PERCENT  OF  TREES 
THAT  ARE  DEC  DUOUS 
SPECIES 
(Variable  7) 


100^ 


PERCENT  TREE 
CANOPY  CLOSURE 
(Variable  6) 


0.5- 


PERCENT  HERBACEOUS 
CANOPY  COVER 
(V.¥lable  a) 


150  m 


300  n 


NUMBER  OF  BURROWS, 
DECAYING  LOGS,  AND 
DEORISODjECTS 
LARGER  THAN  20  cm 
INDiAMETEROM 
THE  GROUND 
(Va-.able 


DISTANCE  ALONG  A 
PROTECTED D  3PER5AL 
CORRIDOR  TO  POTENTIAL 
SPAWNING  SITES 
(VariablelO) 


Figure  2.— The  assumed  relationships 
among  habitat  variables  and  suitability 
index  values  for  the  American  toad. 


periphyton  from  emergent  or  sub- 
mergent  vegetation,  or  scrape  micro- 
organisms and  detritus  from  the 
pond  bottom  and  other  substrates. 
However,  when  blooms  of  sus- 
pended algae  are  present,  the  tad- 
poles become  midwater  filter  feed- 
ers. They  also  feed  on  organic  mate- 
rial supported  by  the  surface  film  of 
the  pond.  At  other  times,  the  tad- 
poles are  facultatively  cannibalistic 
or  coprophagic.  The  particular  feed- 
ing mode  employed  is  usually  influ- 
enced by  a  combination  of  factors 
including  the  type  of  food  available, 
depth  and  temperature  gradients, 
vegetation  structure  and  the  degree 
of  social  behavior  exhibited  by  the 
tadpoles  (Beiswenger  1975).  Most  of 
the  time  toad  tadpoles  feed  from 
substrates  provided  by  the  structural 
features  of  their  environment.  Diaz- 
Paniagua  (1987)  also  found  structural 
features  of  aquatic  vegetation  to  be 
important  in  the  distribution  of  the 
tadpoles  of  five  anuran  species  in 
Spain. 

Habitat  use  by  tadpoles  is  strongly 
influenced  by  temperature,  which  in 
the  shallow  ponds  they  occupy  is 


highly  correlated  with  depth  and  so- 
lar radiation  (variables  1,  3,  and  4). 
For  example,  in  northern  Michigan 
ponds  were  early  summer  tempera- 
tures varied  greatly  over  the  diel  pe- 
riod, toad  tadpoles  consistently  se- 
lected the  warmest  available  water  in 
thermally  stratified  ponds 
(Beiswenger  1977).  Thus,  they  occu- 
pied the  deepest  areas  of  the  pond 
(greater  than  50  cm  in  depth)  at 
night,  avoiding  the  shallow  pond 
margin  where  temperatures  were  5.5 
C  cooler.  During  the  day  tadpoles 
moved  to  shallow  areas  near  shore 
which  were  9  C  warmer  than  the 
deeper  areas  of  the  pond.  During 
those  times  when  there  was  no  ther- 
mal stratification  (e.g.  cloudy  days), 
or  later  in  the  summer  when  pond 
temperatures  were  uniformly  high, 
the  tadpoles  used  all  parts  of  the 
pond  (Beiswenger  1977).  These  ob- 
servations indicate  that  tadpole  habi- 
tat quality  is  partly  determined  by 
thermal  stratification  associated  with 
depth  gradients  and  exposure  to  di- 
rect sunlight. 

Habitat  quality  for  mctamorphic 
tadpoles  is  strongly  influenced  by 


their  vulnerability  to  predation  (vari- 
ables 4  and  5).  As  Arnold  and  Was- 
sersug  (1978,  p.  1019)  expressed  it, 
"the  transforming  anuran  is  neither  a 
good  larva  nor  a  good  frog."  The  lar- 
vae develop  forelimbs  which  impede 
swimming,  the  tail  remnant  on  the 
newly  emergent  juvenile  interferes 
with  its  jumping  ability.  Conse- 
quently, the  availability  of  structural 
features  such  as  hiding  cover  and 
moist  substrates  is  important  for  the 
successful  emergence  and  dispersal 
of  metamorphosing  tadpoles. 

Habitat  quality  for  juvenile  and 
adult  toads  is  determined  by  factors 
generally  associated  with  deciduous 
or  mixed  coniferous/deciduous  for- 
ests. These  factors  include  moderate 
temperature  regimes,  invertebrate 
prey  density,  protected  microhabitats 
with  moist  substrates,  vegetative 
cover,  and  access  to  hibernation  sites. 
Some  of  the  variables  used  as  surro- 
gate measures  of  substrate  moisture 
and  other  forest  floor  conditions  in 
the  HSI  model  for  the  red-spotted 
newt  (Sousa  1985)  were  adapted  for 
the  American  toad  model  (variables 
6,  7,  and  8).  Juvenile  and  adult  toads 


162 


also  need  moist  cover  during  hot  dry- 
periods  and  for  winter  hibernacula. 
These  can  be  provided  by  soils  which 
are  suitable  for  burrowing,  existing 
small  mammal  burrow  systems,  or 
decaying  logs  and  other  debris  ob- 
jects on  the  ground  (variable  9). 

The  American  toad  model  in- 
cludes interspersion  as  a  habitat-re- 
lated factor.  Movement  corridors 
interconnecting  spawning  areas, 
summer  habitat  and  hibernation  sites 
are  an  im|X)rtant  component  of  juve- 
nile and  adult  habitat  (variable  10). 
Brode  and  Bury  (1984)  have  pointed 
out  (cited  in  Ohmart  and  Anderson 
1986),  that  such  corridors  are  impor- 
tant for  dispersal  and  genetic  conti- 
nuity, and  anurans  use  riparian 
zones  as  travel  lanes.  Habitat  frag- 
mentation by  road  construction 
(Rittschof  1975),  or  other  forms  of 
habitat  destruction  can  disrupt  these 
travel  lanes  and  prevent  anurans 
from  reaching  spawning  ponds  or 
hibernation  sites. 

Attention  must  also  be  paid  to 
other  aspects  of  interspersion.  For 
example,  the  reproductive  success  of 
toads  depends  on  the  continuing 
availability  of  shallow  water  habitats. 
Ponds  with  optimum  spawning  con- 
ditions in  a  given  year  may  be  dry  in 
years  with  low  precipitation,  or  too 
deep  in  years  when  flooding  pre- 
vails. At  the  same  time,  changing  wa- 
ter levels  may  result  in  the  availabil- 
ity of  new  spawning  sites,  apparently 
in  response  to  this  kind  of  variation, 
some  species  of  toads  do  not  use  the 
same  spawning  site  every  year 
(Kelleher  and  Tester  1969)  and  in 
some  years  may  not  breed  at  all.  Be- 
cause of  variation  like  this,  it  is  im- 
portant to  describe  the  distribution  of 
habitat  components,  such  as  spawn- 
ing sites  and  movement  corridors,  in 
a  broad  geographic  area  and  over  a 
range  of  environmental  conditions. 

Relationships  among  the  habitat 
variables  and  habitat  components  are 
expressed  by  equations  in  HSI  mod- 
els. A  value  for  the  aquatic  cover/ 
reproduction  (SIA)  component  is  ob- 
tained by  combining  the  suitability 


index  values  for  variables  1  through 
4,  as  shown  in  the  following  equa- 
tion. 

SIA  =  SIVl  X  SIV2  x(SI\/3+SIV4) 

2 

This  assumes  that  the  suitability  of 
aquatic  habitats  is  primarily  deter- 
mined by  the  presence  of  water 
depths  ranging  from  less  than  10  cm 
to  1  m,  rooted  aquatic  vegetation  to 
provide  cover  and  substrates  for 
food,  and  shallow,  unshaded  shore- 
line areas. 

It  is  assumed  that  terrestrial  habi- 
tat suitability  (SIT)  is  determined  by 
the  availability  of  cover  with  moist 
substrates,  invertebrate  prey  and  hi- 
bernation sites.  The  following  equa- 
tion shows  how  these  habitat  values 
could  be  evaluated  using  variable  5 
to  assess  cover  for  metamorphic 
stages,  6,  7,  and  8  as  surrogate  meas- 
ures of  substrate  moisture,  and  vari- 
able 9  for  the  availability  of  hibernac- 
ula. 

SIT=  (SIV5+SIV6+SIV74-SIV9) 
4 

Overall  habitat  suitability  (HSI)  is 
determined  by  combining  the  suita- 
bility values  for  the  aquatic  (SIA)  and 
terrestrial  (SIT)  habitat  components 
with  the  suitability  value  for  inter- 
spersion (SII)  as  shown  in  the  follow- 
ing equation. 

HSI  =  (SIA  x  SIT  x  511)^^3 

This  form  is  used  because  a  value  of 
zero  for  the  suitability  index  for  any 
one  of  the  three  components  indi- 
cates a  lack  of  habitat  to  maintain  vi- 
able populations  of  American  toads. 

Once  it  has  been  fully  developed, 
a  habitat  model  for  the  American 
toad  could  be  used  to  assess  the  ef- 
fects of  such  activities  as  road  build- 
ing, housing  construction,  environ- 
mental pollution,  landfill  operations, 
clearing  of  deciduous  forests,  drain- 
ing or  dredging  of  ponds  and  wet- 
lands, intensive  recreational  use  of 
wetlands,  floodplains  and  the  shore- 
line areas  of  lakes,  and  large  changes 
in  water  level  by  removing  or  intro- 
ducing water. 


Habitat  Models  and  Endangered 
Species  Protection 

The  Wyoming  toad  (Bufo  hemiophrys 
baxteri)  has  recently  been  listed  as 
endangered  by  the  U.S.  Fish  and 
Wildlife  Service  (Baxter  et  al.  1982). 
As  of  June  1988,  there  was  only  one 
small  breeding  population  known  to 
exist.  There  are  no  habitat  models 
available  for  this  subspecies  and 
there  have  been  few  studies  of  its 
natural  history.  This  is  unfortunate 
because  there  is  an  urgent  need  to 
begin  a  recovery  program.  Informa- 
tion about  the  related  Manitoba  toad 
(Bufo  hemiophrys)  which  has  been 
more  extensively  studied  could  be 
used  to  infer  habitat  relationships, 
but  this  is  obviously  not  as  valid  as 
studying  the  Wyoming  toad  directly. 
This  situation  illustrates  why  it  is 
important  to  intensify  our  efforts  to 
develop  databases  and  habitat  mod- 
els for  all  species  before  they  reach 
the  point  of  becoming  endangered.  It 
also  exemplifies  the  role  a  habitat 
model  can  play  in  identifying  infor- 
mation gaps  and  focusing  research 
efforts. 


Discussion 

Resource  assessments  require  the 
development  of  models  for  the  quan- 
titative assessment  of  habitat  suitabil- 
ity. It  is  essential  that  such  models  be 
developed  in  combination  with  com- 
prehensive databases.  A  long  range 
goal  should  be  to  develop  databases 
with  efficient  retrieval  systems  so 
that  it  is  possible  to  access  all  of  the 
site-specific  natural  history  informa- 
tion available  in  the  literature,  and  in 
the  files  of  researchers  and  resource 
managers.  The  databases  should  also 
be  constructed  so  that  information 
gaps  and  priority  areas  for  research 
can  be  identified. 

This  paper  has  emphasized  pro- 
ducing habitat  models  for  individual 
species  as  if  these  species  exist  in  iso- 
lation. Hutto  et  al.  (1987)  have  criti- 
cized the  overemphasis  on  species 


163 


approaches  in  conservation  pro- 
grams as  too  narrow  and  they  point 
out  that  we  must  not  lose  sight  of  the 
higher  order  patterns  and  processes 
which  occur  among  interacting  spe- 
cies. They  suggest  supplementing  the 
species  approach  with  approaches 
that  consider  such  things  as  land- 
scape patterns  that  maintain  ecosys- 
tem level  processes,  the  use  of  geo- 
graphic information  systems,  and 
other  land-based  approaches. 

Studies  emphasizing  the  role  of 
anurans  in  ecosystems  should  result 
in  a  better  understanding  of  ecologi- 
cal process  occurring  at  the  terres- 
trial-aquatic interface,  and  could  also 
contribute  to  more  effective  manage- 
ment of  species  which  depend  on 
these  edge  habitats  and  ecotones. 

Literature  Cited 

Arnold,  Stevan  J.  and  Richard  J.  Was- 
sersug.  1978.  Differential  preda- 
tion  on  metamorphic  anurans  by 
garter  snakes  (Thamnophis):  social 
behavior  as  a  possible  defense. 
Ecology  59: 1014-1022. 

Baxter,  George  T.,  Mark  K. 

Stromberg  and  C.  Kenneth  Dodd. 
1982.  The  status  of  the  Wyoming 
toad  (Bufo  hemiophrys  baxteri).  En- 
vironmental Conservation  9:  348, 
338. 

Beiswenger,  Ronald  E.  1975.  Struc- 
ture and  function  in  aggregations 
of  tadpoles  of  the  American  toad, 
Bufo  americanus.  Herpetologica  31: 
222-233. 

Beiswenger,  Ronald  E.  1977.  Diel  pat- 
terns of  aggregative  behavior  in 
tadpoles  of  Bufo  americanus,  in  re- 
lation to  light  and  temperature. 
Ecology  58:  98-108. 

Beiswenger,  Ronald  E.  1978.  Re- 
sponses of  Bufo  tadpoles  to  labora- 
tory gradients  of  temperature. 
Journal  of  Herpetology  12: 499- 
504. 

Beiswenger,  Ronald  E.  1981.  Preda- 
tion  by  gray  jays  on  aggregating 
tadpoles  of  the  boreal  toad  (Bufo 
boreas).  Copeia  1981:  459-460. 


Beiswenger,  Ronald  E.  1986.  An  en- 
dangered species,  the  Wyoming 
toad,  Bufo  hemiophrys  baxteri — the 
importance  of  an  early  warning 
system.  Biological  Conservation 
37:  59-71. 

Berry,  K.H.  1986.  Introduction:  de- 
velopment, testing,  and  applica- 
tion of  wildlife-habitat  models. 
Pages  3-4.  In  J.  Verner,  M.L.  Mor- 
rison, and  C.J.  Ralph,  eds.  Wildlife 
2000:  modeling  habitat  relation- 
ships of  terrestrial  vertebrates. 
Univ.  Wise.  Press,  Madison.  470  p. 

Brode,  J.M.  and  R.B.  Bury.  1984.  The 
importance  of  riparian  systems  to 
amphibians  and  reptiles.  Pages  30- 
36.  In  Warner,  R.E.  and  K.M.  Hen- 
drix,  eds.  California  riparian  sys- 
tems: ecology,  conservation,  and 
productive  management.  Univ. 
California  Press,  Berkeley.  1035  p. 

Bury,  R.  Bruce,  Howard  W.  Camp- 
bell and  Norman  J.  Scott,  Jr.  1980. 
Role  and  importance  of  nongame 
wildlife.  Transactions  North 
American  Wildlife  and  Natural 
Resources  Conference  45: 197-207. 

Bury,  R.  Bruce  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles.  Pages 
416-419.  In  John  F.  Bell  and  Toby 
Atterbury,  eds.  Proceedings  of  the 
international  conference.  Renew- 
able resource  inventories  for 
monitoring  changes  and  trends. 
Oregon  State  University,  Corval- 
lis.  737  p. 

Campbell,  James  B.  1976.  Environ- 
mental controls  on  boreal  toad 
populations  in  the  San  Juan  Moun- 
tains. Pages  289-295.  In  Harold  W. 
Steinhoff  and  Jack  D.  Ives  (eds). 
Ecological  impacts  of  snowpack 
augmentation  in  the  San  Juan 
Mountains,  Colorado.  Final  report 
San  Juan  ecology  project.  Colo- 
rado State  University  Publ.,  Fort 
Collins.  480  p. 

Chalk,  David  E.,  Stephen  A.  Miller 
and  Thomas  W.  Hoekstra.  1984. 
Multispecies  inventories:  integrat- 
ing information  on  wildlife  re- 
sources. Wildlife  Society  Bulletin 
12:  357-364. 


Clawson,  Mary  E.,  Thomas  S.  Bas- 
kett,  and  Michael  J.  Armbruster. 
1984.  An  approach  to  habitat  mod- 
eling for  herpetofauna.  Wildlife 
Society  Bulletin  12:  61-69. 

Diaz-Paniagua,  Carmen.  1987.  Tad- 
pole distribution  in  relation  to 
vegetal  heterogeneity  in  tempo- 
rary ponds.  Herpetological  Journal 
1: 167-169. 

Graves,  Brent  M.  and  Stanley  H.  An- 
derson. 1987.  Habitat  suitability 
index  models:  bullfrog.  U.S.  Fish 
Wildl.  Serv.  Biol.  Rep.  82  (10: 138). 
22  p. 

Hutto,  Richard  L.,  Susan  Reel  and 
Peter  B.  Landres.  1987.  A  critical 
evaluation  of  the  species  approach 
to  biological  conservation.  Endan- 
gered Species  UPDATE  4(12):  1-4. 

Jones,  K.  Bruce.  1986.  Amphibians 
and  reptiles.  Pages  267-290.  In  Co- 
operrider,  A.Y.,  R.J.  Boyd  and 
H.R.  Stuart  eds.  Inventory  and 
monitoring  of  wildlife  habitat.  U.S. 
Department  of  the  Interior,  Bureau 
of  Land  Management,  Service 
Center.  Denver,  Co.  xviii,  858  p. 

Karr,  James  R.  1987.  Biological  moni- 
toring and  environmental  assess- 
ment: a  conceptual  framework. 
Environ.  Manage.  11:  249-256. 

Kelleher,  Kelvin  E.  and  John  R. 
Tester.  1969.  Homing  and  survival 
in  the  Manitoba  toad,  Bufo  hemio- 
phrys, in  Minnesota.  Ecology  50: 
1040-1048. 

Ohmart,  Robert  D.  and  Bertin  W. 
Anderson.  1986.  Riparian  habitat 
Pages  169-199.  In  Cooperrider, 
A.Y.,  R.J.  Boyd  and  H.R.  Stuart, 
eds.  Inventory  and  monitoring  of 
wildlife  habitat.  U.S.  Dept.  Inter., 
Bur.  Land  Manage.  Service  Center. 
Denver,  Co.  xviii,  858  p. 

Rittschof,  Daniel.  1975.  Some  aspects 
of  the  natural  history  and  ecology 
of  the  leopard  frog,  Rana  pipiens. 
Ph.D.  Dissertation,  University  of 
Michigan,  Ann  Arbor. 

Short,  H.L.  1984.  Habitat  suitability 
index  models:  The  Arizona  guild 
and  layers  of  habitat  models.  U.S. 
Fish  Wildl.  Serv.  FWS/OBS-82/ 
10.70.  37  p. 


164 


Sousa,  P.J.  1985.  Habitat  suitability 
index  models:  Red-spotted  newt. 
U.S.  Fish  Wildl.  Serv.  Biol.  Rep. 
82(10.111).  18  p. 

Thomas,  Jack  Ward  (ed).  1979.  Wild- 
life habitats  in  managed  forests. 
The  Blue  Mountains  of  Oregon 
and  Washington.  Agricultural 
Handbook  No.  553.  USDA  Forest 
Service,  Washington,  D.C.  512  p. 

U.S.  Fish  and  Wildlife  Service.  1981. 
Standards  for  the  development  of 
habitat  suitability  index  models. 
ESM103.  U.S.  Depart.  Int.  Fish 
Wildl.  Serv.,  Division  of  Ecological 
Services.  Washington,  D.C.  68  p. 
and  appendices. 

Wassersug,  Richard  J.  1975.  The 
adaptive  significance  of  the  tad- 
pole stage  with  comments  on  the 
maintenance  of  complex  life  cycles 
in  anurans.  American  Zoologist 
15:  405-417. 

Wilbur,  Henry  M.  1980.  Complex  life 
cycles.  Annual  Review  of  Ecology 
and  Systematics  11:  67-93. 


165 


Preliminary  Report  on  Effect 
of  Bullfrogs  on  Wetland 
Herpetofaunas  in 
Southeastern  Arizona^ 

Cecil  R.  Schwalbe  and  Philip  C.  Rosen^ 


Abstract.— Ranid  frogs  (Rona  cofesbeiano,  R. 
chiricohuensis,  and  R.  yavopaiensis),  garter  sr^akes 
(Thamnophis  eques,  T.  marcianus)  and  Sonoran 
mud  turtles  (Kinosfernon  sonoriense)  were  surveyed 
in  soutt^eastern  Arizona.  Distribution  of  the 
introduced  bullfrog  (Rano  cofesbeiano)  was 
negatively  correlated  witt^  distributions  of  the  two 
leopard  frogs  and  garter  snakes,  The  hypothesis  that 
bullfrog  predotion  caused  decline  of  a  native 
wetland  herpetofouno  is  supported  by  data  on 
bullfrog  diet,  on  garter  snake,  leopard  frog  and  mud 
turtle  population  structure,  and  natural  history 
observations  on  the  snakes.  An  experimental 
removal  of  bullfrogs  has  been  initiated  at  the  San 
Bernardino  National  Wildlife  Refuge. 


The  bullfrog  (Ram  catesbeiam)  is 
North  America's  largest  frog  and  one 
of  the  most  widely  distributed  anu- 
rans  on  the  continent.  Occurring 
naturally  from  Florida  to  Nova  Scotia 
and  west  into  central  Texas,  Okla- 
homa, and  Kansas,  the  bullfrog  has 
been  introduced  widely  into  perma- 
nent waters  throughout  the  West 
(Bury  and  Whelan  1984,  Stebbins 
1985,  Wright  and  Wright  1949) 
Known  to  be  voracious,  opportunis- 
tic predators,  they  have  been  impli- 
cated in  declines  of  native  anuran 
populations  (Bury  and  Luckenbach 

1976,  Bury  et  al.  1980,  Conant  1975, 

1977,  Jameson  1956,  Moyle  1973, 
Nussbaum  et  al.  1983,  Vitt  and 
Ohmart  1978  and  others).  Much  less 
is  known  about  their  impacts  on 
other  vertebrate  classes. 

A  recent  investigation  of  factors 
producing  decline  of  Mexican  garter 
snakes  (Thamnophis  eques)  in  Arizona 
(Rosen  and  Schwalbe  1988)  sug- 
gested that  predation  by  introduced 
bullfrogs  (see  fig.  1)  is  a  present  and 

'Paper  presented  of  the  Symposium,  The 
Monagemenf  of  Amphibians,  Repfiles,  and 
Small  Mammals  in  North  America.  July  18- 
22.  1988.  Flagstaff.  Arizona. 

^Cecil  R.  Schwalbe  is  Nongame  Herpe- 
tologist  and  Philip  C.  Rosen  is  Contract  Bi- 
ologist. Arizona  Game  and  Fish  Depart- 
ment. 2222  West  Greenway  Road.  Phoenix. 
Arizona  65023-4399.  Rosen's  present  ad- 
dress is  Department  of  Ecology  and  Evolu- 
tionary Biology.  University  of  Arizona. 
Tucson.  Arizona  85721. 


serious  impact  on  some  of  the  few 
remaining  snake  populations.  Obser- 
vations during  the  garter  snake  sur- 
vey suggested  a  similar  effect  on 
leopard  frogs  (Rana  yavapaiensis,  R. 
chiricahuensis). 

Recently,  Hayes  and  Jennings 
(1986)  questioned  the  importance  of 
bullfrog  predation  in  declines  of 
western  North  American  ranid  frogs. 
They  include  predation  by  bullfrogs 
as  one  of  three  major  hypotheses  to 
explain  decline  of  ranid  frogs  in  Cali- 
fornia, but  suggest  that  predation  by 
introduced  fish  has  had  greater  im- 


pact on  native  frogs.  Hayes  and  Jen- 
nings (1986)  indicate  further  that 
their  hypotheses  need  to  be  tested  to 
determine  actual  causal  factors  in 
population  declines.  In  this  paper  we 
present  distributional  and  natural 
historical  data  implicating  bullfrogs 
in  population  declines  of  native  wet- 
land reptiles  and  amphibians  in 
southeastern  Arizona.  We  then  de- 
scribe an  experimental  program  of 
bullfrog  removal  we  have  initiated  to 
test  the  direct  and  indirect  effects  of 
this  introduced  predator  on  wetland 
herpetofaunas. 


166 


Methods  and  Materials 

We  report  on  two  phases  of  our 
work.  The  first  phase  involves  exten- 
sive surveys,  principally  for  garter 
snakes.  The  second  focuses  on  inten- 
sive surveying  and  experimental 
manipulation  at  one  locality  that  is 
heavily  infested  with  bullfrogs. 

Extensive  Phase 

We  sampled  over  80  localities 
throughout  much  of  central  and 
southern  Arizona  during  1985-1987, 
searching  appropriate  aquatic  and 
semi-aquatic  habitats  (Rosen  and 
Schwalbe  1988).  Methods  and  results 
are  briefly  summarized  here.  Lotic 
habitats  were  surveyed  for  2-6  mile 
reaches  on  foot.  Lentic  habitats  were 
also  examined  on  foot,  in  their  en- 
tirety in  most  cases.  During  these 


surveys,  attempts  were  made  to  cap- 
ture, measure,  mark  and  release  all 
garter  snakes  seen.  Detailed  observa- 
tions were  made  on  distribution  and 
abundance  of  other  biota  on  the  sites 
sampled,  with  special  attention  to 
anurans,  turtles  and  other  snakes. 
Intensive  mark-recapture  studies 
were  conducted  at  four  sites  using 
trapping  methods  described  below. 

Intensive  Phase 

San  Bernardino  National  Wildlife 
Refuge  (SBNWR),  one  of  four  sites 
where  mark-recapture  procedures 
were  initiated  during  the  extensive 
phase  of  our  work,  was  selected  for 
ongoing  observation  and  experimen- 
tation. Beginning  in  September  1986, 
we  visited  the  refuge  in  September 
and  May  of  each  year,  marking 
snakes,  observing  herpetofaunal  dis- 


SAN  BEffmom  NWR 


Mexico 


Figure  2.— Diagrammatic  map  of  San  Bernardino  National  Wildlife  Refuge.  Stippled  line  indi- 
cates boundary  between  upland  Chihuahuan  desertscrub  and  riparian  scrub  and  wood- 
land vegetation  types. 


tributions  and  abundances,  and  ex- 
perimentally removing  bullfrogs. 

Intensive  Site  Description 

SBNWR  (fig.  2)  consists  of  984  ha  in 
the  San  Bernardino  Valley  on  the 
Mexican  border  in  Cochise  County, 
Arizona.  Elevations  range  from  1134 
to  1183  m.  Higher,  rocky  slopes  and 
mesas  supporting  Chihuahuan  de- 
sertscrub and  lower  terraces  grading 
into  desert  grassland  comprise  al- 
most two-thirds  of  the  refuge. 

The  heart  of  the  refuge  is  a  low- 
land supporting  dense  mesquite 
(Prosopis  velutina)  bosques  and 
sacaton  (Sporobolus)  grasslands  inter- 
spersed with  four  spring-fed  ponds 
and  seven  additional  springs.  In  the 
center  of  this  low  ground  is  deeply 
incised  Black  Draw,  headwater  of  the 
Rio  Yaqui,  which  normally  arises  at  a 
natural  spring  about  halfway  be- 
tween the  Mexican  border  and  north 
boundary  of  the  refuge.  Large,  iso- 
lated, living  and  dead  cottonwood 
trees  (Populus  fremontii)  occur  near 
almost  all  aquatic  habitats.  Broad 
swamplikc  ciencgas  with  little  open 
water  occur  at  the  artesian  wells  that 
do  not  supply  ponds. 

Vegetation  in  Black  Draw  varies 
from  rank  herbaceous  plants  and  tall 
grass  in  the  northern  one-half, 
through  open  riparian  thicket  and 
cat-tail  (Typha  domingensis)  stands, 
into  almost  impenetrable  thickets  of 
sapling  cottonwood  and  willow 
(Salix  gooddingi)  throughout  the 
lower  1.2  km  to  the  border.  Cienega 
pools  are  cold  and  reach  a  depth  of 
about  2  meters. 

North  Pond,  focal  point  for  the 
experimental  removal  of  bullfrogs, 
contains  0.1  ha  of  open  water  sur- 
rounded by  earthen  levees.  Artesian 
well  flow  is  piped  into  the  pond  and 
into  a  small  marshy  area  north  of  the 
pond.  North  and  west  banks  are 
lined  with  mesquite.  South  and  west 
banks  are  open  or  overgrown  with 
herbaceous  vegetation.  Cat-tail  is 
spreading  rapidly  around  the  pond 


167 


margin  from  foci  in  northeast  and 
southwest  corners.  Open  water  is 
largely  choked  with  submergent 
macrophytes. 

The  wetland  herpetofauna  of  the 
refuge  includes  bullfrogs  (Rana 
catesbeiana),  lowland  leopard  frogs 
(R.  yavapaiensis),  Mexican  garter 
snakes  (Thamnophis  eques),  checkered 
garter  snakes  (T.  marcianus) ,  and 
Sonoran  mud  turtles  (Kinosternon 
sonoriense). 


Intensive  Field  Procedures 

Garter  snakes  were  collected  by  hand 
at  all  times  of  day  and  night,  and 
with  minnow  traps  connected  by 
aquatic  drift  fences  (see  Rosen  and 
Schwalbe  1988  for  details).  Four  drift 
fences,  each  with  a  trap  at  each  end, 
were  set  in  North  Pond  during  each 
visit  to  the  refuge.  Two  drift  fences 
with  traps  were  set  in  Twin  Pond  in 
August  1985  and  August-September 
1986.  Twin  Pond  was  drained  during 
summer  1987  and  remains  dry. 

The  following  data  were  recorded 
for  each  snake  captured:  date,  loca- 
tion, sex,  snout-vent  length  (SVL), 
tail  length,  total  weight,  presence/ 
absence  and  number  of  food  items, 
and  injuries.  Females  were  palped  to 
determine  presence /absence  and 
number  of  developing  young.  For 
hand-caught  snakes  we  recorded  ac- 
tivity at  time  of  first  sighting,  mi- 
crohabitat,  time,  and  cloacal  and  am- 
bient temperatures.  Each  individual 
was  uniquely  marked  by  clipping 
subcaudal  scales. 

Bullfrogs  were  collected  mostly 
with  four-pronged  spears  at  night  by 
using  head  lamps  to  find  and  blind 
them.  Additionally,  many  were  col- 
lected in  turtle  hoop  nets,  which 
were  set  along  seine  nets  rigged  as 
aquatic  drift  fences.  Some  hoop  nets 
were  baited  to  capture  turtles,  and 
these  captured  bullfrogs,  as  well.  A 
few  were  collected  by  hand  and  with 
air  guns  and  light  arms.  Initial  col- 
lecting efforts  were  focused  on  larger 


(>100  mm  SVL)  bullfrogs.  Every 
aquatic  habitat  on  the  refuge  was 
checked  for  frogs  by  listening  for 
their  calls  and  searching  visually  at 
night.  Captured  bullfrogs  were  kept 
on  ice  overnight  and  the  following 
data  were  recorded  the  next  day: 
capture  location  and  date,  sex,  snout- 
vent  length,  total  weight.  Most  were 
dissected  to  determine  stomach  con- 
tents and  reproductive  condition. 


Results 

Distribution  and  Natural  History 

Leopard  frogs  are  significantly  less 
common  where  bullfrogs  abound 
(table  1:  Spearman  rank  correlation 
r^=-0.434,  p<0.025,  Rosner  1982,  Sokal 
and  Rohlf  1981).  SBNVVR  is  the  only 
site  where  we  found  both  bullfrogs 


and  leopard  frogs.  Among  the  sites 
shown  in  table  1,  introduced,  non- 
native  predatory  fish  were  found  in 
abundance  only  at  Bog  Hole  and 
Babocomari  Cienega,  where  ranid 
frogs  were  absent.  Historical  records 
indicate  that  leopard  frogs  once  were 
abundant  in  two  areas  now  support- 
ing dense  bullfrog  populations,  Ari- 
vaca  Creek  (Wright  and  Wright  1949) 
and  SBNWR  (Lanning  1981,  Lowe 
personal  communication). 

Mexican  garter  snakes  also  are  sig- 
nificantly less  abundant  in  the  pres- 
ence of  bullfrogs  (table  1:  Spearman 
rank  correlation  r^=-0.420,  p<0.03).  At 
the  Potrero  Canyon  locality,  Mexican 
garter  snakes  were  known  as  late  as 
1970  (Rosen  and  Schwalbe  1988),  but 
we  found  only  checkered  garter 
snakes  (N  =  24)  during  19854987. 

At  SBNWR,  all  museum  records  of 
Thamnophis  prior  to  1970  (N=7)  were 


Table  1. —Distribution  and  abundance  of  ranid  frogs  and  garter  snal<es  in 
wetlands  of  southeastern  Arizona,  based  upon  field  work  during  1985-1988. 
0=absent,  l=rare,  2=common,  3=very  abundant;  P=pond,  C=cienega, 
M=marsti;  NWR=National  Wildlife  Refuge;  SB=San  Bernardino.  Leopard  frogs 
may  be  either  Rana  chiricahuensis  or  R.  yavapaiensis. 


Locality 

Ranid  abundance  Garter  snake  abundance 

Bull- 
frog 

Leopard 
frog 

Checkered 

Mexican 

San  Bernardino  NWR 

P 

3 

0 

1 

1 

San  Bernardino  NWR 

C 

3 

1 

1 

1 

Upper  SB  valley 

P 

0 

2 

3 

0 

Leslie  Creek 

C 

0 

3 

0 

0 

Lewis  Springs 

C 

0 

3 

0 

2 

San  Pedro  River 

2 

0 

2 

1 

San  Pedro  gravel  pit 

P 

3 

0 

0 

0 

Ramsey  Canyon 

P 

0 

3 

0 

0 

Parker  Canyon  Lake 

3 

0 

0 

1 

Sharp  Spring 

C 

1-2 

0 

0 

1 

Bog  Hole 

P 

0 

0 

0 

0 

Bog  Hole 

c 

0 

0 

0 

3 

Research  Ranch 

p 

0 

2 

0 

3 

Research  Ranch 

c 

0 

1 

0 

3 

Elgin  Cienega 

c 

0 

2 

0 

2 

Babocomari  River 

p 

0 

0 

0 

1 

Babocomari  River 

c 

0 

0 

0 

1 

Cienega  Creek 

c 

1 

0 

0 

2 

Potrero  Canyon 

M 

3 

0 

3 

0 

Potrero  Canyon 

C 

3 

0 

2 

0 

Sonoita  Creek 

M 

3 

0 

0 

0 

Sonoita  Creek 

C 

3 

0 

0 

0 

168 


eques,  while  all  subsequent  (N=5) 
were  marcianus.  Thamnophis  eques 
comprised  57%  of  the  garter  snakes 
seen  on  the  refuge  during  1985-1988 
(table  2).  On  the  refuge,  the  popula- 
tion of  Mexican  garter  snakes  was 
heavily  dominated  by  large  adults,  in 
significant  contrast  to  populations  in 
areas  lacking  bullfrogs,  where  year- 
lings and  small  adults  predominate 
(fig.  3,  Mann-Whitney  U  Test, 
p<0.001).  At  SBNWR,  most  Mexican 
garter  snakes  (61.9%)  had  damaged 
tails  which  bled  between  the  ventral 
scales  when  handled  (fig.  4),  suggest- 
ing unsuccessful  predation  attempts 
by  bullfrogs.  This  type  of  injury  was 
not  seen  at  any  other  locality. 

At  SBNWR  we  found  Sonoran 
mud  turtles  (Kinosternon  sonoriense) 
to  be  unexpectedly  rare.  Only  four 


turtles  were  captured  in  29  trap- 
nights  on  the  refuge,  a  rate  of  0.14 
captures  per  trap-night.  Elsewhere  in 
Arizona,  917  trap  nights  produced 
2,092  captures  at  the  17  other  locali- 
ties we  have  sampled  (Rosen  unpub- 
lished data,  Rosen  1987).  The  mean 
trap  success  for  those  17  localities 
was  4.32  +  0.23  captures  per  trap- 
night  (range  0.20-12.23).  For  the  five 
habitats  in  southeastern  Arizona 
which  were  comparable  to  the  ref- 
uge, and  where  at  least  20  trap-nights 
were  registered,  mean  trap  success 
was  5.42  +  1.03  captures  per  trap- 
night  (1.23-12.23).  Quitobaquito 
Pond,  with  0.20  captures  per  trap- 
night  was  the  only  area  in  Arizona 
with  trapping  success  approaching 
the  low  level  obtained  at  the  refuge. 
The  Quitobaquito  population  is 


Table  2.— Records  of  all  garter  snakes  captured  on  the  San  Bernardino  Na- 
tional Wildlife  Refuge,  Arizona,  1985-1988. 


Sampling 

Number 

Number 

Snakes 

period 

Mexican 

checkered 

captured 

garter  snakes 

garter  snakes 

per  day 

16-18  Aug  85 

3 

2 

1.67 

23-27  May  86 

4 

3 

1.40 

30  Aug- 1  Sep  86 

3 

1 

1.33 

23-25  May  87 

5 

0 

1.67 

5-7  Sep  87 

4 

6 

3.33 

29-30  May  88 

1 

3 

2,00 

Total 

20 

15 

1.84 

15- 


10-1 

w 
o 

K 

III 

K 


u.  O 


K 

U 

■  SJ 

3 
Z 


0- 


I     I     I     I   "  I    lll  I  1 1  I  ll  I 

200  400  600  800 

SNOUT-VENT  LENGTH  (mm) 

Figure  3.— Size-frequency  histograms  of 
Mexican  garter  snakes  in  1985  and  1986 
(rTX>clified  fronn  Rosen  and  Schwalbe  1988). 
Upper  histogrann  represents  snakes  fronn 
populations  where  bullfrogs  were  scarce  or 
absent.  Lower  histogrann  represents  San 
Bernardino  National  Wildlife  Refuge  sample. 


Figure  4.— Bullfrog  damage  to  tail  of  large 
Mexican  garter  snake,  San  Bernardino  Na- 
tional Wildlife  Refuge,  Cochise  County,  Ari- 
zona, 1986. 


known  to  have  been  markedly  re- 
duced by  human  activities  (Rosen 
1986). 

Including  captures  obtained  by  all 
methods,  only  six  Sonoran  mud 
turtles  have  been  found  by  us  on  the 
refuge.  All  were  large  adults,  and, 
according  to  growth  ring  analysis 
(see  Rosen  1987),  all  were  born  prior 
to  1981.  In  all  other  populations,  ju- 
veniles comprised  over  207o  of  the 
sample  (Rosen,  unpublished  data). 


Bullfrog  Diet 

Stomach  contents  confirmed  the  op- 
portunistic feeding  behavior  of  bull- 
frogs (table  3).  Invertebrates  consti- 
tuted the  majority  of  food  items,  with 
the  snail,  Planorbella  tenuis,  and  in- 
sects of  the  orders  Coleoptera, 
Diptera,  Hemiptera,  Hymenoptera, 
Odonata  and  Orthoptera  commonly 
eaten.  Arthropods  consumed  in- 
cluded adults  and  larvae  of  terres- 
trial, aquatic  and  flying  forms. 

Vertebrates  were  found  in  14.6 
percent  of  the  stomachs  that  con- 
tained some  food.  The  most  com- 
monly consumed  vertebrates  were 
other  frogs,  including  bullfrogs.  At 
least  two  species  of  native  fishes, 
both  endangered,  were  eaten,  the 
Yaqui  chub  (Gila  purpurea)  and  the 
Yaqui  topminnow  (Poeciliopsis  oc- 
cidentalis  sonoriensis).  Mammal  prey 
included  Peromyscus,  a  Sigmodon  and 
other  as  yet  unidentified  small  ro- 
dents. The  two  reptile  food  items 
were  a  neonate  checkered  garter 
snake  in  a  frog  from  House  Pond  and 
a  spiny  lizard  (genus  Sceloporus).  Not 
shown  in  table  3  was  a  nestling  bird, 
thought  to  be  a  red-winged 
blackbird,  Agelaius  tricolor,  found  in 
the  stomach  of  a  subadult  bullfrog 
(100  mm  SVL). 


Bullfrog  Density 

Using  the  numbers  of  bullfrogs  re- 
moved from  North  Pond  (table  4), 
we  can  estimate  density  and  bio- 


169 


mass.  After  removing  74  adult  bull- 
frogs in  spring  1987,  we  estimated  5 
adults  remained.  Including  the  small 
area  of  marsh  north  of  the  levee, 
there  was  0.11  ha  of  habitat  for  this 
population,  giving  a  minimum  den- 
sity estimate  of  718  adults/ha.  Mean 
weight  for  all  frogs  removed  in  the 
spring  1987  census  at  North  Pond 
was  217.1  g,  yielding  a  total  biomass 
of  23.7  kg,  or  215.5  kg/ha.  Excluded 
from  this  biomass  estimate  were  re- 
maining adults,  and  numerous  juve- 
niles that  were  not  hunted.  These  es- 
timates are  conservative  since  we 
had  already  removed  51  adults  and 
23  juveniles  during  fall  1986,  before 
we  had  determined  the  most  effec- 
tive means  of  removing  the  frogs. 

The  fall  1987  census  at  North  Pond 
reflects  thorough  removal  the  previ- 
ous spring,  with  only  about  10  frogs 
either  maturing  into  adults  or  immi- 
grating between  May  24  and  Septem- 
ber 5, 1987.  We  estimated  that  4-6 
adults  remained  in  North  Pond  at  the 
end  of  our  1987  collecting.  Because  of 
extremely  cool,  windy  weather  dur- 
ing the  spring  1988  trip,  we  were  un- 
able to  collect  bullfrogs  effectively 
during  the  last  night  and  left  an  esti- 
mated 15-20  adults. 

A  total  of  552  bullfrogs  has  been 
removed  from  SBNWR  as  of  June 
1988  (tables  4-5),  including  358  of 
adult  size,  from  a  total  area  of  2.4  ha 
of  open  water.  We  estimate  that  take 
to  represent  55-80%  of  the  adult  bull- 
frogs on  the  refuge  at  that  time. 

Preliminary  Experimer^tal  Results 


Leopard  frogs  bred  successfully  at 
the  spring  source  in  central  Black 
Draw  in  early  1987,  a  time  of  unusu- 
ally good  rainfall.  This  area  was  vir- 
tually devoid  of  bullfrogs  because  it 
is  open  enough  for  predators  and  re- 
source managers  to  kill  all  or  almost 
all  adults.  In  May  1987,  leopard  frog 
tadpoles  and  juveniles  were  moder- 
ately abundant  from  the  spring  to  the 
northernmost  reach  of  cienega- 
stream  and  dense  sapling  thicket. 


where  they  were  replaced  by  bull- 
frogs. The  first  confirmation  of  leop- 
ard frogs  in  North  Pond  was  five 
found  in  bullfrog  stomachs  in  May 
1987.  No  noticeable  further  increase 
in  leopard  frog  numbers  or  distribu- 
tion was  observed  in  May  1988. 

The  first  juvenile  Mexican  garter 
snake  on  the  refuge  during  this  study 
was  recorded  in  fall  1987.  The  cap- 
ture rate  of  garter  snakes  on  the  ref- 
uge doubled  between  May  and  Sep- 
tember 1987  following  bullfrog  re- 
moval (table  2).  Extremely  cold. 


windy  weather  on  the  May  1988  trip 
greatly  depressed  reptile  activity. 
Thus,  the  2.0  garter  snakes  captured 
per  day  (table  2)  may  reflect  a  de- 
crease in  activity  rather  than  a  de- 
crease in  the  numbers  of  garter 
snakes  on  the  refuge. 

Discussion 

Distributional  and  natural  historical 
data  from  southeastern  Arizona  pro- 
vide prima  facie  evidence  that  bull- 


Table  3  -Stomach  contents  of  aduit  (>120  mm  snout-vent  length)  bullfrogs. 
San  Bernardino  National  Wildlife  Refuge.  Arizona. 

Sampling  date 


Prey  type 


30  Aug- 
1  Sep  86 


5-6 
Sep  .87 


22-24 
May  87 


29-30 
May  88 


Total 


Amphibians 
Bullfrogs 
Tadpoles 
Juveniles 
Leopard  frogs 

Juveniles 
Unknown  anurans 
Fishes 

Yaqui  chub 
Yaqui  topminnovv 
Unidentified 
I  Mammals 
Reptiles 
Invertebrates 
Detritus 

Empty  stomachs 
Total  food  items 
No.  frogs  dissected 


r 


sidered  to  be  adults.    


Sampling 
period 


Adult 
males 


Fall  1986 
Spring  1987 
Fall  1987 
Spring  1988 

^  Totals  


33 
43 
14 
17 

107 


Adult 
females 

Total 
juveniles 

Total 
removed 

18 
31 
1 

15 

23 
35 
13 
48 

74 
109 
28 
80 

65 

119 

291 

170 


frogs  play  a  causative  role4n  popula- 
tion decline  and  disappearance  of 
native  wetland  amphibians  and  rep- 
tiles (table  1;  Results).  For  Mexican 
garter  snakes,  this  evidence  is  bol- 
stered by  data  on  population  struc- 
ture (fig.  3)  and  by  observations  of 
injuries  caused  by  bullfrogs  (fig.  4; 
Rosen  and  Schwalbe  1988). 

That  bullfrogs  are  predatory  gen- 
eralists  has  been  thoroughly  docu- 
mented (see  extensive  review  of  bull- 
frog foods  in  Bury  and  Whelan  1984). 
In  Arizona  alone,  bullfrogs  have  con- 
sumed such  vertebrate  prey  as  a 
nestling  bird,  young  muskrat  (On- 
datra zibethicus),  cotton  rat  (Sigmo- 
don),  softshell  turtle  (Trionyx  spinif- 
erus),  spiny  lizard  (Sceloporus), 
kingsnake  (Lampropeltis  getulus),  sev- 
eral species  of  fish  and  frogs,  garter 
snakes,  even  a  rattlesnake  (Crotalus 
atrox)  (fig.  1,  table  3;  Clarkson  and 
deVos  1986). 

To  our  knowledge,  in  southeastern 
Arizona,  the  only  place  where  bull- 
frogs abound  and  where  leopard 
frogs  and  Mexican  garter  snakes  also 
still  occur,  albeit  rarely,  is  SBNVVR. 
We  beheve  the  native  species  persist 
there  because  the  extent  and  diver- 
sity of  aquatic  habitats  is  greater  than 
elsewhere  in  the  region.  Specifically, 
the  relatively  sparse  vegetation  and 
absence  of  deep  pools  at  the  spring 
source  area  in  central  Black  Draw  has 
remained  largely  free  of  adult  bull- 
frogs. This  is  where  leopard  frogs 


have  bred  and  where  the  smallest 
Mexican  garter  snakes  have  been 
found. 

We  believe  the  reason  only  five 
leopard  frogs  and  one  garter  snake 
were  found  in  bullfrog  stomachs  is 
due  to  already  severe  reduction  of 
leopard  frog  and  garter  snake  popu- 
lations. The  same  reasoning  may  ap- 
ply to  the  absence  of  hatchling  Sono- 
ran  mud  turtles  in  bullfrog  stomachs. 

The  bullfrog  density  at  North 
Pond  (SBNWR)  was  quite  high  for 
Arizona  populations,  although  not 
necessarily  high  for  other  parts  of  its 
range  (Currie  and  Bellis  1969).  Such  a 
density  is  equalled  and  possibly  ex- 
ceeded at  Arivaca,  Pima  County,  Ari- 
zona, where  both  leopard  frogs  and 
Mexican  garter  snakes  have  been  ex- 
tirpated or  become  extremely  rare 
(Rosen  and  Schwalbe  1988).  Concen- 
trations of  bullfrogs  similar  to  that  in 
lower  Black  Draw  have  only  been 
seen  in  comparable  habitat  in  por- 
tions of  one  cienega  in  the  San  Ra- 
phael grasslands  of  Santa  Cruz 
County.  Abundances  comparable  to 
those  in  House  Pond  occur  at  a 
gravel  mine  south  of  Arizona  High- 
way 90  on  the  San  Pedro  River,  Co- 
chise County;  at  Page  Springs, 
Yavapai  Count}';  and  possibly  at 
Parker  Canyon  Lake,  Cochise  and 
Santa  Cruz  counties  and  Potrero 
Canyon  marsh,  eight  kilometers 
north  of  Nogales,  Santa  Cruz 
County. 


r 


Table  5.— Bullfrog  removals  from  aquatic  habitats  other  than  North  Pond, 
San  Bernardino  National  Wildlife  Refuge,  Arizona.  Individuals  >  120  mm 
snout-vent  length  are  considered  adults. 


Adult 

Adult 

Juveniles 

Total 

Locality 

Date 

males 

females 

removed 

Twin  Pond 

Fall  86 

2 

2 

1 

5 

Tula  Pond 

Spring  87 

2 

3 

3 

8 

House  Pond 

Spring  87 

35 

42 

34 

111 

Black  Draw 

Spring  87 

32 

25 

15 

72 

Tuie  Pond 

Spring  88 

0 

0 

3 

3 

House  Pond 

Spring  88 

9 

10 

11 

30 

Black  Draw 

Spring  88 

6 

18 

8 

32 

Totals 

86 

100 

75 

261 

At  Potrero  Canyon  marsh,  Mexi- 
can garter  snakes  have  disappeared 
and  checkered  garter  snakes  are 
abundant.  In  the  preceding  three  lo- 
calities, checkered  garter  snakes  are 
absent,  and  Mexican  garter  snakes 
persist  in  low  numbers.  Both  garter 
snakes  occur  along  the  San  Pedro 
River  but  neither  utilize  the  gravel 
pit  pond  (Rosen  and  Schwalbe  1988, 
Rosen  personal  observations). 

Natural  cienega-streams,  includ- 
ing Turkey  and  O'Donnell  Creeks, 
where  bullfrogs  are  absent,  and 
Cienega  Creek,  where  they  are  rare, 
have  high  densities  of  Mexican  garter 
snakes  and  include  many  juveniles 
and  young  adults.  One  spring  fed 
pond  north  of  Canelo  Hills,  which  is 
structurally  and  vegetatively  similar 
to  North  Pond,  contained  about  95 
Mexican  garter  snakes  at  a  density 
near  1055  individuals/ha,  and 
yielded  an  average  of  5.4  snakes  per 
trapping  day  (Rosen  and  Schwalbe 
1988).  In  contrast,  only  seven  garter 
snakes  have  been  trapped  on  SB- 
NWR in  fifteen  days  of  similar  trap- 
ping. 

Central  Black  Draw  would  ordi- 
narily be  regarded  as  relatively  poor 
habitat  for  Mexican  garter  snakes, 
because  the  vegetative  cover  is  too 
thin,  particularly  at  the  water's  edge. 
The  abundance  of  Mexican  garter 
snakes  there  and  the  regular  occur- 
rence of  checkered  garter  snakes  at 
North  Pond  display  an  inversion  of 
the  usual  habitat  preferences  of  the 
two  species  in  Arizona.  In  competi- 
tion, in  a  broad  sense,  with  Mexican 
garter  snakes,  checkered  garter 
snakes  may  be  favored  by  the  pres- 
ence of  bullfrogs  because  they  are 
less  aquatic  and  hence  less  affected 
by  the  increased  predation  pressure. 

Hayes  and  Jennings  (1986)  argued 
that  predation  by  introduced  bull- 
frogs was  not  a  compelling  hypothe- 
sis to  explain  population  declines  of 
native  ranid  frogs  in  western  North 
America.  They  suggest  that  preda- 
tion by  introduced  fish,  mainly  cen- 
trarchids,  is  a  more  promising  hy- 
pothesis. In  southeastern  Arizona  we 


171 


found  that  bullfrogs  have  invaded  a 
greater  variety  of  wetland  environ- 
ments than  exotic  predatory  fish, 
and,  in  some  instances,  have 
achieved  population  densities  suffi- 
cient to  impact  the  native  herpe- 
tofauna.  While  we  do  suspect  that 
introduced  fish  impact  native  wet- 
land herpetofaunas  in  Arizona  (see 
Rosen  and  Schwalbe  1988),  our  data 
for  the  southeastern  portion  of  the 
state  compellingly  incriminate  the 
bullfrog. 

Our  approach  is  to  attempt  to 
manage  or  eliminate  bullfrogs  from 
selected  areas.  It  is  principally  in- 
tended to  develop  practical  manage- 
ment techniques  for  controlling  bull- 
frogs, but  should  also  provide  an  ex- 
perimental test  of  the  bullfrog  preda- 
tion  hypothesis. 

Effective  January  1, 1988,  the  Ari- 
zona Game  and  Fish  Commission 
opened  the  season  year  round  and 
set  an  unlimited  bag  and  possession 
limit  on  dead  bullfrogs  statewide  ex- 
cept for  La  Paz,  Mohave,  and  Yuma 
counties  (Arizona  Game  and  Fish 
Commission  1988).  The  stipulation  of 
unlimited  possession  of  dead  frogs 
was  to  decrease  the  likelihood  of  ac- 
cidental or  intentional  release  of  bull- 
frogs into  new  habitats.  The  new 
regulations  will  make  it  easier  for 
agencies,  organizations  and  individu- 
als to  put  pressure  on  bullfrog  popu- 
lations in  specific  areas  in  favor  of 
native  species. 

No  data  exist  to  show  impacts  of 
bullfrogs  on  native  species  in  the 
three  western  counties,  so  they  have 
retained  a  July  1  to  November  30  sea- 
son with  a  bag  and  possession  limit 
of  12  per  day  or  in  possession  live  or 
dead.  Because  Arizona's  amphibian 
and  reptile  regulations  are  reviewed 
annually,  new  data  can  be  incorpo- 
rated into  management  decisions. 

Conclusions 

There  is  evidence  that  bullfrogs  have 
negatively  impacted  populations  of 
native  amphibians  and  reptiles  in 


Arizona.  Although  some  of  the 
trends  are  encouraging,  preliminary 
data  from  bullfrog  removal  exp>eri- 
ments  are  inconclusive  as  to  whether 
or  not  bullfrog  control  measures  may 
augment  recruitment  in  lowland 
leopard  frogs,  Mexican  garter  snakes 
or  Sonoran  mud  turtles.  More  inten- 
sive efforts  will  be  required  to  elimi- 
nate bullfrogs  from  even  local  habi- 
tats when  such  habitats  are  structur- 
ally complex. 

Acknowledgments 

We  are  thankful  to  the  U.S.  Fish  and 
Wildlife  Service  Office  of  Endan- 
gered Species  in  Albuquerque  for 
funding  parts  of  this  study.  We  thank 
the  U.S.  Fish  and  Wildlife  Service 
and  Refuge  Manager  Ben  Robertson 
in  particular  for  permission  to  con- 
duct this  research  on  the  refuge,  and 
C.H.  Lowe  for  his  information  on  the 
history  of  the  herpetofauna  on  the 
refuge  and  elsewhere  in  the  South- 
west. 

For  enthusiastic  assistance  in  the 
field  we  gratefully  acknowledge  the 
following:  Ron  Armstrong,  Randy 
Babb,  Howard  Berna,  Andrew, 
Cindy  and  Ted  Cordery,  Mary 
Gilbert,  Rich  Glinski,  Dean  and  Gar- 
rett Hendrickson,  Julia  Hoffman, 
Terry  Johnson,  Charlie  Painter,  Bruce 
Palmer,  David  Parizek,  David 
Propst,  Cathy  Schmidt,  Adam  and 
Ethan  Schwalbe,  Berney  Swinburne, 
Ross  Timmons  and  Sabra  Tonn. 


Literature  Cited 

Arizona  Game  and  Fish  Commission. 
1988.  Commission  Order  41:  Am- 
phibians. Arizona  Game  and  Fish 
Department  Publication.  1  p. 

Bury,  R.  Bruce,  C.  Kenneth  Dodd,  Jr., 
and  Gary  M.  Fellers.  1980.  Conser- 
vation of  the  amphibia  of  the 
United  States:  a  review.  U.S.  De- 
partment of  the  Interior,  Fish  and 
Wildlife  Service,  Resource  Publica- 
tion 134, 34  p.  Washington,  D.C. 


Bury,  R.  Bruce,  and  Roger  A.  Luck- 
enbach.  1976.  Introduced  amphibi- 
ans and  reptiles  in  California.  Bio- 
logical Conservation  10:1-14. 

Bury,  R.  Bruce,  and  Jill  A.  Whelan. 
1984.  Ecology  and  management  of 
the  bullfrog.  U.S.  Department  of 
the  Interior,  Fish  and  Wildlife 
Service,  Resource  Publication  155, 
23  p.  Washington,  D.C. 

Clarkson,  Robert  W.,  and  James  C. 
deVos,  Jr.  1986.  The  bullfrog.  Ram 
catesbeiam  Shaw,  in  the  Lower 
Colorado  River,  Arizona-Califor- 
nia. Journal  of  Herpetology  20:42- 
49. 

Conant,  Roger.  1975.  A  field  guide  to 
reptiles  and  amphibians  of  eastern 
and  central  North  America.  429  p. 
Houghton  Mifflin  Company,  Bos- 
ton. 

Conant,  Roger.  1977.  Semiaquatic 
reptiles  and  amphibians  of  the 
Chihuahuan  Desert  and  their  rela- 
tionships to  drainage  patterns  of 
the  region,  p.  455-491.  In  Transac- 
tions of  the  symposium  on  the  bio- 
logical resources  of  the  Chihua- 
huan Desert  region.  United  States 
and  Mexico.  R.H.  Wauer  and  D.H. 
Riskind,  editors.  U.S.  National 
Park  Service,  Transaction  Proceed- 
ings Series  3, 658  p. 

Currie,  W.,  and  E.  D.  Bellis.  1969. 
Home  range  and  movements  of 
the  bullfrog.  Ram  catesbeiam 
Shaw,  in  an  Ontario  pond.  Copeia 
1969:688-692. 

Dowe,  Brian  John.  1979.  The  effect  of 
time  of  oviposition  and  microenvi- 
ronment  on  growth  of  larval 
bullfrogs  (Ram  catesbeiam)  in  Ari- 
zona. Unpubl.  masters  thesis,  55  p. 
Arizona  State  University,  Tempe. 

Fritts,  Thomas  H.,  Randy  D.  Jen- 
nings, and  Norman  J.  Scott,  Jr. 
1984.  A  review  of  the  leopard 
frogs  of  New  Mexico.  Unpubl.  re- 
port to  New  Mexico  Department 
of  Game  and  Fish  (Santa  Fe,  New 
Mexico).  157  p. 

Hayes,  Marc  P.,  and  Mark  R.  Jen- 
nings. 1986.  Decline  of  ranid  frog 
species  in  western  North  America: 
Are  bullfrogs  (Ram  catesbeiam) 


172 


ogy  20:490-509. 

Hendrickson,  Dean  A.,  and  W.L. 
Minckley.  1984.  Cienegas — vanish- 
ing climax  communities  of  the 
American  Southwest.  Desert 
Plants  6:130-175. 

Hulse,  Arthur  C.  1974.  An  autecol- 
ogical  study  of  Kinosternon  sonori- 
ense  Lcconte  (Chelonia:  Kinoster- 
nidae).  Unpubl.  doctoral  disserta- 
tion, 105  p.  Arizona  State  Univer- 
sity, Tempe. 

Jameson,  David  L.  1956.  Growth,  dis- 
persal and  survival  of  the  Pacific 
treefrog.  Copeia  1956:25-29. 

Lanning,  Dirk  V.  1981.  The  verte- 
brates of  San  Bernardino  Ranch, 
Cochise  County,  Arizona.  Unpubl. 
report  from  the  Arizona  Natural 
Heritage  Program  (Tucson,  Ari- 
zona) to  The  Nature  Conservancy 
and  the  U.S.  Fish  and  Wildlife 
Service,  Albuquerque,  New  Mex- 
ico. 88  p. 

Moyle,  Peter  B.  1973.  Effects  of  intro- 
duced bullfrogs,  Rana  catesbeiana, 
on  the  native  frogs  of  the  San 
Joaquin  Valley,  California.  Copeia 
1973:18-22. 

Nussbaum,  Ronald  A.,  Edmund  D. 
Brodie,  Jr.,  and  Robert  M.  Storm. 
1983.  Amphibians  and  reptiles  of 
the  Pacific  Northwest.  332  p. 
Northwest  Naturalist  Books,  The 
University  of  Idaho  Press, 
Moscow. 

Rosen,  Philip  C.  1986.  Population  de- 
cline of  Sonoran  mud  turtles  at 
Quitobaquito  Springs.  Final  Re- 
port to  National  Park  Service,  Co- 
operative Park  Service  Research 
Unit,  Tucson,  Arizona.  24  p. 

Rosen,  Philip  C.  1987.  Variation  of 
female  reproduction  among  popu- 
lations of  Sonoran  mud  turtles 
(Kinosternon  sonoriense)  in  Arizona. 
M.S.  thesis,  Arizona  State  Univer- 
sity, Tempe. 

Rosen,  Philip  C,  and  Cecil  R.  Sch- 
walbe.  1988.  Status  of  the  Mexican 
and  narrow-headed  garter  snakes 
(Thamnophis  eques  megalops  and 
Thamnophis  rufipunctatus)  in  Ari- 
zona. Unpubl.  report  from  Ari- 
zona Game  and  Fish  Dept.  (Phoe- 


nix, Arizona)  to  U.S.  Fish  and 
Wildlife  Service,  Albuquerque, 
New  Mexico.  69  p. 

Rosner,  Bernard.  1982.  Fundamentals 
of  Biostatistics.  547  p.  Duxbury 
Press,  Boston. 

Sokal,  Robert  R.  and  F.  James  Rohlf. 
1981.  Biometry,  the  Principles  and 
Practice  of  Statistics  in  Biological 
Research.  859  p.  W.H.  Freeman 
Co.,  San  Francisco. 

Stebbins,  Robert  C.  1985.  A  field 
guide  to  western  reptiles  and  am- 
phibians. 336  p.  Houghton  Mifflin 
Company,  Boston. 

Vitt,  Laurie  J.,  and  Robert  D. 

Ohmart.  1978.  Herpetofauna  of  the 
lower  Colorado  River:  Davis  Dam 
to  the  Mexican  border.  Proceed- 
ings of  the  Western  Foundation  of 
Vertebrate  Zoology  2:35-72. 

Wright,  Albert  Hazen,  and  Anna  Al- 
len Wright.  1949.  Handbook  of 
frogs  and  toads  of  the  United 
States  and  Canada.  640  p.  Com- 
stock  Publishing  Company,  New 
York. 


173 


Developing  Management 
Guidelines  for  Snapping 
Turtles^ 

Ronald  J.  Brooks,^  David  A.  Galbraith,^  E. 
Graham  Nanceklvell/  and  Christine  A. 
Bishop^ 


Abstract.— We  examined  demographic  features 
of  2  Ontario  populations  of  snapping  turtles 
(Chelydra  serpentina)  io  provide  an  empirical  basis 
for  developing  management  guidelines.  The 
northern  population  matured  later  (18-20  yr)  than 
did  the  southern  populations  (<10  yr),  and  displayed 
an  older  age  distribution.  Long-lived,  "bet-hedging" 
species  have  low  annual  reproductive  success  and 
are  unusually  susceptible  to  exploitation.  A 
preliminary  life  table  is  presented  for  the  northern 
population.  Our  results  indicate  that  the  northern 
population  cannot  sustain  even  minimal  levels  of 
exploitation  by  humans  without  undergoing  a 
decline  in  numbers. 


In  general,  turtles  have  not  been  a 
major  concern  of  wildlife  managers 
in  North  America,  and  in  many  juris- 
dictions they  are  given  little  or  no 
protection.  They  are  perceived  to 
have  limited  ecological,  commercial, 
aesthetic  or  recreational  value,  and 
because  they  are  usually  cryptic  and 
slow  moving  they  are  uninteresting 
to  most  people.  Partly  for  these  rea- 
sons, there  have  been  remarkably 
few  studies  of  their  life  history  and 
ecology.  In  addition,  their  great  lon- 
gevity makes  them  difficult  to  study, 
except  on  a  long-term  basis.  Never- 
theless, turtles  are,  or  should  be,  of 
interest  to  wildlife  managers  for  at 
least  three  major  reasons. 

First,  they  are  major  components 
of  a  variety  of  both  terrestrial  and 
aquatic  ecosystems  and  therefore 
play  significant,  though  often  unrec- 
ognized roles  as  carnivores,  herbi- 
vores and  scavengers.  In  both 
aquatic  and  terrestrial  habitats,  the 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff ,  AZ,  July  19-21,  1988.) 

^Professor,  Department  of  Zoology,  Uni- 
versity of  Guelph,  Guelph,  Ontario,  Can- 
ada. N1G2W1. 

^Graduate  Student,  Biology  Depart- 
ment, Queen's  University,  Kingston,  Ontario, 
Canada.  K7L  3N6. 

"Wildlife  Technician,  Department  of  Zo- 
ology. University  of  Guelph,  Guelph,  On- 
tario, Canada.  N1G2W1. 

^Graduate  Student,  Biology  Depart- 
ment, York  University,  Toronto,  Ontario, 
Canada. 


Standing-crop  biomass  of  turtles  is 
generally  much  higher  than  that  of 
any  other  reptile  (Iverson  1982).  In 
aquatic  systems,  turtle  biomass  often 
exceeds  that  of  sympatric  endoth- 
erms  by  an  order  of  magnitude  and 
is  similar  to  levels  reported  for  fishes 
(Iverson  1982).  Similarly,  annual  pro- 
duction of  turtles  is  comparable  to 
that  reported  in  most  other  verte- 
brates, although  well  below  levels 
found  in  some  fishes  (Iverson  1982). 
Many  turtles  that  are  especially  long- 
lived  may  have  low  annual  produc- 
tivity. This  low  productivity  may  be 
overestimated  because  of  the  high 
standing-crop  biomass  of  turtles. 
Their  life  history  is  markedly  differ- 
ent from  those  of  the  birds  and  mam- 
mals that  typically  occupy  the  atten- 
tion of  wildlife  managers.  As  such, 
these  species  represent  special  prob- 
lems in  conservation  and  manage- 
ment. Therefore,  turtles  should  be  of 
interest  to  managers,  because  they 
are  important  components  of  a  vari- 
ety of  ecological  communities  and 
because  in  many  cases  their  longevity 
and  low  annual  production  relative 
to  standing  crop,  characteristic  of  a 
"bet-hedger"  (Obbard  1983)  is  a  life- 
history  strategy  that  may  be  highly 
susceptible  to  exploitation  or  to  other 
sources  of  mortality  of  adult  animals 
such  as  unsuitable  overwintering 
conditions  or  heavily  polluted  wa- 
ters. 

Secondly,  managers  should  have 
an  interest  in  turtles  because  many 
species  are  harvested  for  commercial 


profit,  usually  as  food  or  for  the  pet 
trade  (Bergmann  1983,  Congdon  et 
al.  1987,  Lovisek  1982).  There  is  evi- 
dence of  marked,  recent  declines  in 
harvests  of  most  turtle  species,  but 
this  evidence  is  difficult  to  quantify 
because  estimates  of  total  stocks  do 
not  exist  for  any  turtle  species.  For 
snapping  turtles,  the  annual  commer- 
cial catch  in  Minnesota  was  esti- 
mated at  36000-40800  kg  or  approxi- 
mately 6000-6800  average-sized 
adults  (Helwig  and  Hora  1983).  In 
southern  Ontario,  Lovisek  (1982)  esti- 
mated the  annual  catch  of  C.  ser- 
pentina to  be  30000-50000  kg  or  5000- 
8300  adults.  There  is  evidence  from 
trappers  (J.  Bullard  pers.  comm.)  that 
numbers  of  this  species  are  a  fraction 
of  former  numbers  over  much  of 
their  southern  range  in  Ontario,  but 
again  no  quantitative  estimates  exist. 
At  present,  therefore,  it  is  necessary 
to  measure  the  impact  of  harvesting 
turtles  on  a  local  basis  (Hogg  1975). 

Thirdly,  snapping  turtles  may  be 
of  interest  to  managers  because  they 
are  often  regarded  as  pests  or  as  a 
danger  to  human  swimmers,  or  as 
destructive  predators  of  waterfowl 
and  game  fish  (Hammer  1969,  Kiviat 
1980,  Pell  1941). 

In  this  paper,  we  review  the  biol- 
ogy of  snapping  turtles  in  relation  to 
these  three  areas  of  potential  impor- 
tance for  wildlife  managers.  We  pres- 
ent demographic  characteristics  of  2 
populations  in  Ontario,  and  in  addi- 
tion, we  develop  a  life  table  for  the 
more  northern  population  of  snap- 


174 


ping  turtles  which  will  allow  us  to 
predict  the  impact  of  different  levels 
of  harvesting  pressure  on  this  popu- 
lation. 


Snapping  Turtles  in  Aquatic 
Ecosystems 

Regulation  of  Population  Density 

There  is  at  present  little  understand- 
ing of  what  factors  regulate  popula- 
tions of  any  turtle  species,  but  it  is 
known  that  turtles  may  reach  very 
high  densities  and  high  biomass  den- 
sities (Galbraith  et  al.,  in  press;  Iver- 
son  1982).  It  seems  likely  that  pri- 
mary productivity  would  be  the  best 
predictor  of  variation  in  numbers  of 
turtles  in  a  habitat.  In  snapping 
turtles,  population  density  ranges 
from  1-75  adult  turtles  per  ha  (Gal- 
braith et  al.,  in  press).  Density  among 
populations  correlates  positively 
with  latitude  and  primary  produc- 
tion levels  and  negatively  with  the 
size  of  the  body  of  water  (Galbraith 
et  al.,  in  press),  although  data  are  too 
sparse  to  rely  heavily  on  these  corre- 
lations. Other  possible  factors  influ- 
encing density  are  predation  pres- 
sure, especially  on  nests  and  hatch- 
lings,  climatic  influences  on  egg  sur- 
vival and  embryo  development,  and 
availability  of  suitable  nesting  sites. 
Again,  the  role  of  these  factors  has 
not  been  studied. 


Annual  Energy  Budgets 

No  complete  energy  budget  has  been 
determined  for  any  turtle  population, 
although  some  efforts  have  been 
made  to  estimate  critical  components 
of  the  energy  budget  (Congdon  et  al. 
1982).  Almost  all  efforts  in  this  area 
have  concentrated  on  the  energy  con- 
tent and  cost  of  the  eggs  (Congdon 
and  Gibbons  1985,  Congdon  and 
Tinkle  1982,  Shine  1980)  and  on  the 
rates  of  digestion,  especially  in  rela- 
tion to  temperature  (Parmenter 
1981). 


Food-Web  Connections 

Snapping  turtles  are  widely  regarded 
as  voracious  predators,  but  most 
studies  of  their  diet  indicate  that 
plant  material  is  a  major  com|X)nent 
of  their  food  (Alexander  1943,  Ham- 
mer 1972,  Pell  1941).  Hammer  (1972) 
found  that  plants  made  up  the  major- 
ity of  the  diet  of  snapping  turtles  in  a 
North  Dakota  marsh.  In  Connecticut, 
fish  (mostly  nongame  species)  and 
aquatic  plants  were  of  equal  impor- 
tance and  birds  made  up  only  a  small 
fraction  of  the  diet  (Alexander  1943). 
In  Maine,  snapping  turtles  ate  signifi- 
cant numbers  of  ducklings  in  local 
areas  where  both  turtles  and  water- 
fowl were  common,  but  widespread 
control  of  turtles  was  not  recom- 
mended (Coulter  1957).  Lagler 
(1943),  working  in  Michigan,  con- 
cluded that  snapping  turtles  had 
minimal  impact  on  waterfowl  and 
pan  fish  and  subsisted  primarily  on 
plant  material  and  invertebrates.  In 
general  then,  snapping  turtle  preda- 
tion on  waterfowl  or  game  and  sport 
fish  poses  no  serious  problem  to 
these  valuable  species  except  perhaps 
in  local  situations  where  numbers  of 
turtles  may  be  very  high  and  the 
turtles  have  easy  access  to  young  wa- 
terfowl. 

Adult  snapping  turtles  are  largely 
immune  to  predation  other  than  by 
humans  over  most  of  their  range.  A 
wide  diversity  of  predators  prey  on 
snapping  turtle  eggs  (foxes,  skunks, 
raccoons)  and  hatchlings  (herons, 
large  fish),  and  mortality  is  very  high 
during  these  stages. 

Rationale  for  \he  Development  of 
Life  Tables 

The  demography  of  populations  of 
freshwater  turtles  under  exploitation 
has  not  been  extensively  studied. 
Some  reports  have  cited  large  catches 
being  removed  from  specific  loca- 
tions with  apparently  little  impact  on 
remaining  numbers  in  the  short  term 
(e.g.  Hogg  1975)  but  no  study  has 


followed  an  exploited  population  in 
detail  for  any  length  of  time.  It  is  nec- 
essary, therefore,  to  infer  the  effect  of 
harvesting  on  populations  using 
demographic  parameters  of  unex- 
ploited  populations  under  long-term 
study.  This  paper  describes  2  snap- 
ping turtle  populations  in  Ontario, 
Canada  and  presents  a  life  table  for 
one  of  these  populations. 

Study  Areas 

Lake  Sasajewun,  Algonquin 
Provincial  Park,  Ontario 

The  Ontario  Ministry  of  Natural  Re- 
sources Wildlife  Research  Area 
(W.R.A.  45'35'  N,  78'30'W,  mean  an- 
nual temperature  4.4 'C),  is  located  in 
the  central  area  of  Algonquin  Provin- 
cial Park,  in  a  region  of  mixed  forest 
last  logged  in  the  1930s.  The  snap- 
ping turtles  inhabiting  the  lakes  and 
streams  running  through  the  W.R.A. 
have  been  studied  since  1972.  Each 
year,  adult  female  turtles  are  cap- 
tured after  nesting  and  both  males 
and  females  are  captured  using 
baited  hoop  traps.  Of  the  approxi- 
mately 185  tagged  snapping  turtles  in 
the  watershed  of  the  North  Mada- 
waska  River,  about  100  are  recap- 
tured each  year.  Approximately  70 
nests  of  known  females  are  located 
each  year. 

Snapping  turtles  are  the  largest 
aquatic  vertebrate  in  the  W.R.A., 
with  the  exception  of  beavers  (Castor 
canadensis)  and  occasional  river  otters 
(Lutra  canadensis).  The  only  other  spe- 
cies of  turtle  in  this  watershed  is  the 
midland  painted  turtle  ( Chrysemys 
picta  marginata),  present  in  very  small 
numbers  (<  10).  The  density  of  the 
W.R.A.  snapping  turtle  population  is 
approximately  1.5  adults/ha  in  lakes 
(Galbraith  et  al.,  in  press).  The  study 
area  and  the  snapping  turtle  popula- 
tion have  been  described  extensively 
elsewhere  (Galbraith  and  Brooks 
1987;  Galbraith  et  al.  1987,  in  press; 
Obbard  1983). 


175 


Royal  Botanical  Gardens, 
Hamilton,  Ontario 

The  Royal  Botanical  Gardens  (R.B.G.) 
consist  of  approximately  700  ha  of 
woodlands  and  waterways  within 
the  metropolitan  Hamilton  area 
(43'17'N,  79'53'W;  mean  annual  tem- 
perature 9.8'C).  This  study  area  and 
the  snapping  turtle  population  in  the 
R.B.G.  have  been  described  previ- 
ously (Galbraith  et  al.,  in  press).  We 
have  captured,  tagged,  and  released 
adult  and  juvenile  snapping  turtles  in 
this  watershed  since  1984.  In  addi- 
tion to  snapping  turtles,  map  turtles 
(Malaclemys  geographica)  and  painted 
turtles  are  common  aquatic  cheloni- 
ans  in  this  system.  The  painted  turtle 
is  at  least  as  common  as  the  snapping 
turtle. 

The  turtles  inhabit  a  highly  pro- 
ductive, eutrophic  waterway  which 
is  artificially  enriched  by  effluent 
from  a  sewage  treatment  plant.  West 
Pond  (9.8  ha),  where  our  trapping 
has  taken  place,  also  connects  with 
heavily-polluted  Hamilton  Harbour. 
Despite  the  contaminants,  this  popu- 
lation exhibits  one  of  the  highest  den- 
sities yet  reported  for  this  species, 
approximately  60-70  adults/ ha  (Gal- 
braith et  al.,  in  press). 

Methods  and  Results 
Life  Tables 

Two  approaches  are  commonly  taken 
in  preparing  life  tables.  Static  or  ver- 
tical life  tables  are  prepared  by  deriv- 
ing mortality  rates  from  the  observed 
population  age  structure.  Cohort- 
specific,  or  horizontal  life  tables  are 
prepared  by  following  a  specific  co- 
hort and  observing  age-specific  mor- 
tality rates  throughout  life  (Deevy 
1947).  At  present,  only  static  life 
tables  can  be  prepared  for  snapping 
turtle  populations,  because  individ- 
ual cohorts  cannot  be  followed  effec- 
tively in  these  animals  which  may 
have  a  maximum  longevity  of  over  a 
century  (Galbraith  and  Brooks  1987). 


Therefore,  we  will  only  consider 
static  life  tables. 


Life-Table  Parameters  for 
Algonquin  Park  (W.R.A.) 

Snapping  turtles  experience  large 
fluctuations  in  annual  reproductive 
success  (Obbard  1983).  In  the  W.R.A. 
population,  for  example,  most  years 
do  not  produce  any  emergent  hatch- 
lings  (R.J.  Brooks,  unpubl.  data) 
whereas  occasional  years  may  pro- 
duce large  numbers  of  hatchlings. 
This  highly  stochastic  survivorship 
throws  some  doubt  on  the  utility  of 
static  life  tables,  because  age  curves 
could  be  highly  biased  by  errors  due 
to  irregular  recruitment.  Therefore, 
we  will  use  an  average  mark-recap- 
ture survivorship  rate  (Galbraith  and 
Brooks  1987)  for  all  adult  females  for 
the  construction  of  the  life  table. 

Several  critical  pieces  of  informa- 
tion have  never  been  obtained  for 
any  snapping  turtle  population.  For 
example,  no  estimate  of  survivorship 
of  hatchlings  or  juveniles  has  ever 
been  published.  A  crude  estimate  of 
this  rate  can  be  obtained  by  assuming 
that  the  number  of  turtles  recruited 
per  year  into  the  population  is  fairly 
represented  by  the  average  recruit- 
ment rate,  and  that  the  number  of 
eggs  being  produced  per  year  has  not 
varied  greatly  between  the  years 
when  recruits  were  initially  pro- 
duced (i.e.  as  eggs)  and  the  present 
time.  In  the  W.R.A.  population,  on 
average,  one  new  nesting  female  is 
captured  per  year  on  nesting  sites 
used  by  approximately  85  other  fe- 
males. The  mean  clutch  size  of  34 
eggs  once  per  year  gives  an  annual 
egg  production  of  2890  eggs.  Assum- 
ing half  these  eggs  produce  females, 
the  net  survivorship  across  all  age 
classes  (including  eggs)  until  age  at 
first  nesting  (approximately  19  yr, 
(Galbraith  1986))  is  therefore  1/1445 
(0.000692). 

In  the  W.R.A.  population,  Obbard 
(1983)  observed  a  mean  rate  of  emer- 
gence of  hatchlings  from  eggs  of 


0.0635,  averaged  over  142  nests  in  5 
yr.  Taking  this  into  account,  in  addi- 
tion to  the  adult  recruitment  rate  of 
one  mature  female  per  year,  the 
probability  of  mortality  between 
hatching  and  maturity  for  females  in 
this  population  is  99.17%.  Average 
annual  juvenile  survivorship  from 
this  estimate  is  therefore  0.7541  from 
hatching  to  19  yr  (table  1). 

High  rates  of  statistical  errors 
within  age  estimates  of  individual 
turtles  (Galbraith  1986)  make  docu- 
mentation of  horizontal  rates  of  age- 
specific  changes  in  fecundity  unreli- 
able, and  therefore  we  have  con- 
structed our  life  table  using  mean 
clutch  size  for  all  age  classes.  Net  fe- 
cundity, however,  is  a  function  of 
both  clutch  size  and  clutch  fre- 
quency. Obbard  (1983)  estimated  that 
72.1%  of  adult  females,  on  average, 
lay  a  clutch  each  year  in  this  popula- 
tion. Mean  annual  egg  production  is 
therefore  24.514  eggs  per  female 
(mean  clutch  size  is  34  eggs).  For  the 
purposes  of  a  life  table,  the  female 
turtles  are  considered  as  producing 
only  female  offspring.  It  is  also  neces- 
sary, therefore,  to  consider  the  effects 
of  biases  in  hatchling  sex  ratios. 
Snapping  turtles  experience  environ- 
mental sex  determination,  whereby 
incubation  temperature  during  the 
middle  third  of  the  incubation  period 
determines  offspring  sex  (Yntema 
1976).  Between  1981  and  1985,  the 
mean  hatchling  sex  ratio  of  naturally 
incubated  nests  in  the  W.R.A.  was 
66%  female  (R.J.  Brooks,  unpubl. 
data).  Therefore,  each  female  turUe, 
on  average,  produces  16.18  female- 
destined  embryos  per  nesting  season. 
Although  snapping  turtles  are  long- 
lived,  the  life  table  for  female  snap- 
ping turtles  in  the  W.R.A.  suggests 
that  they  do  not  reproduce  enough  to 
sustain  the  population  (table  1). 

Life-Table  Parameters  for  the 
Royal  Botanical  Gardens  (R.B.G.) 

Although  data  are  inadequate  to  con- 
struct a  meaningful  life  table  for 


176 


snapping  turtles  from  the  R.B.G., 
some  population  parameters  are 
known.  For  example,  females  in  the 
very  large  snapping  turtle  population 
in  the  R.B.G.  appear  to  nest  for  the 
first  time  at  10  yr  of  age  (RJ.  Brooks, 
unpubl.  data),  and  the  mean  clutch 
size  in  the  R.B.G.  population  between 


1985  and  1987  was  45  eggs.  The  rate 
of  mortality  in  this  population  is 
likely  higher  than  in  the  Algonquin 
population,  because  numerous  dead 
turtles  are  found  each  year  (C.A. 
Bishop,  unpubl.  data).  Essential  but 
currently  unavailable  information 
from  the  R.B.G.  population  includes 


Table  l.—Ufe  table  for  female  snapping  turtles  In  Algonquin  Park  (W.R.A.), 
Ontario,  Canada. 


Year 


class 

a  ' 

X 

1  = 

X 

^x 

X 

ml, 

X  X 

Zm.l  * 

X  X 

n 

u 

1007  A 

1  nnn 

1 

1 

.uooo 

9 

01  '^zL'^ 

0470 

O 

oo.ooo 

.UOO  1 

*4 

CI  o^n 

0979 

C. 

'^0  17ft 

090^^ 

.  /  0*4  1 

A 

O 

90  '^i4A 

OT^"^ 
.u  I 

7^41 

7 

99  9ft9 

01 17 

o 
o 

iA  on  A 

.uuoo 

7c; /ll 

0 
y 

19  A7'^ 

OOAA 

i^Ay 
,  /  0*4 1 

in 

0  ^^^7 

y .  oo  / 

00 '^O 

7'^dl 
.  /  0*4  1 

1  1 

1  9nft 
/  .zuo 

00  "^A 

l^A^ 
.  1  1 

009 A 

7^41 

1  o 

A  noo 

0091 

.  /C^  t 

15 

3.091 

.0316 

.7541 

16 

2.331 

,0012 

.7541 

17 

1.758 

.0009 

.7541 

18 

1.326 

.0007 

.7541 

19 

1.000 

.000524 

.9660 

16.18 

.00848 

0.00848 

20 

.000506 

.9660 

16.18 

.00819 

0.0167 

21 

.000489 

.9660 

16.18 

.00791 

0.0246 

22 

.000472 

.9660 

16.18 

.00764 

0.0322 

23 

.000456 

.9660 

16.18 

.00738 

0.0396 

24 

.000441 

.9660 

16.18 

.00714 

0.0468 

25 

.000426 

.9660 

16.18 

.00689 

0.0536 

30 

.000358 

.9660 

16.18 

.03107 

0.0847 

35 

.000301 

.9660 

16.18 

.02615 

0.1109 

40 

.000253 

.9660 

16.18 

.02199 

0.1329 

50 

.000179 

.9660 

16.18 

.03404 

0.1633 

60 

.000127 

.9660 

16.18 

.02409 

0.1873 

70 

.000090 

.9660 

16.18 

.01705 

0.1990 

80 

.000064 

.9660 

16,18 

.01209 

0.2111 

90 

.000045 

.9660 

16,18 

.00853 

0.2196 

100 

.000032 

.9660 

16.18 

.00730 

0.2269 

'         =  numbers  of  individuals. 

ax 

-  probability  of  survival  from  year  class  0  to  year  class  x. 
^n,       =  probability  of  survival  from  year  class  x  to  year  class  x+ 1 . 

=  net  fecundity  at  year  class  x  (female-destined  embryos  produced). 

*Z//n^  =  sum  of  all  reproduction  from  year  class  0  to  year  class  x.  equals  Ro.  total 
lifetime  reproduction,  when  xisat  its  maximum. 


long-term  estimates  of  emergence 
rates  of  hatchlings  or  of  adult  survi- 
vorship, annual  nesting  frequency, 
and  primary  sex  ratio. 


Life-Table  Implications  for 
Management  Guidelines 

Clearly,  exploitation  of  a  population 
similar  to  that  in  Algonquin  Park 
would  quickly  reduce  numbers  be- 
low any  chance  of  recovery  by  repro- 
duction within  that  population.  In 
formulating  our  life  table  for  the 
W.R.A.,  we  have  had  to  make  several 
assumptions.  The  most  important 
concerns  our  estimate  of  the  rate  of 
survival  of  hatchlings  and  juveniles. 

A  comparison  between  the  2 
populations  indicates  that  the  advan- 
tages in  the  R.B.G.  population  of  hav- 
ing a  larger  clutch  size  than  the  more 
northern  population  and  being  able 
to  initiate  nesting  almost  10  yr  before 
the  W.R.A.  population  may  be  tem- 
p>ered  by  overestimating  adult  survi- 
vorship in  the  R.B.G.  population. 
Consequently,  lifetime  reproduction 
may  not  be  as  high  as  one  might  pre- 
dict. These  comparisons  must  be  im- 
proved by  direct  observation  of  sur- 
vival in  the  critical  juvenile  years, 
and  by  following  individuals  of 
known  age  throughout  life,  in  a  vari- 
ety of  populations. 

Considerable  variation  in  popula- 
tion characteristics  exists  between 
these  2  populations  located  about  280 
km  apart.  Trapping  guidelines  appli- 
cable to  the  R.B.G.  population  may 
not  be  suitable  to  the  population  in 
the  W.R.A.  Regardless,  neither  could 
likely  tolerate  harvests  of  more  than 
10%  of  the  adult  population. 


Management  Practices  to 
Increase  Yields  of  Snapping 
Turtles 

It  is  evident  that  unregulated  har- 
vesting of  adult  snapping  turtles  will 
rapidly  decrease  population  sizes, 
because  adult  turtles  are  normally 


177 


subject  to  very  low  rates  of  mortality 
(Galbraith  and  Brooks  1987).  Two 
strategies  are  possible  to  increase 
harvestable  numbers  of  turtles. 

First,  practical  experience  with  sea 
turtle  farming  has  shown  that  large 
numbers  of  eggs  can  be  incubated 
under  artificial  or  protected  condi- 
tions (Mrosovsky  and  Yntema  1980), 
although  care  must  be  taken  to  incu- 
bate the  eggs  at  a  selection  of  tem- 
peratures which  will  produce  a  bal- 
anced sex  ratio.  Similar  propagation 
of  snapping  turtles  should  result  in 
increased  numbers  of  juveniles  in 
populations  where  adult  numbers 
are  not  density-dependent. 

Secondly,  enrichment  of  the  envi- 
ronment could  provide  faster  growth 
rates  for  these  poikilotherms.  In- 
creases in  available  protein  will 
probably  result  in  an  increase  in 
growth  rates  of  individuals  and  in- 
creases in  adult  carrying  capacities 
(MacCulloch  and  Secoy  1983). 

Organochloride  Contaminants 
and  Hunnan  Consumption 

Long-lived  bottom-dwellers  can  ac- 
cumulate high  levels  of  environ- 
mental toxins,  and  snapping  turtles 
have  been  found  to  carry  very  high 
loads  of  PCBs  of  various  forms 
(Bryan  et  al.  1987a).  Several  studies 
have  considered  the  way  in  which 
PCBs  accumulate  and  in  which  tis- 
sues, and  snapping  turtles  are  now 
being  employed  as  biomonitors  for 
organochlorides  in  some  studies 
(C.A.  Bishop  et  al.,  unpubl.  data). 

Bryan  et  al.  (1987)  demonstrated 
that  local  levels  of  pollutants  mark- 
edly affected  the  levels  of  organo- 
chloride toxins  in  snapping  turtle  tis- 
sues. Tissue-specific  accumulation  of 
PCBs  is  not  random  in  snapping 
turtles,  but  is  a  function  of  lipopro- 
tein content  of  the  tissue  and  the  high 
lipoprotein  solubility  of  the  toxins. 
Especially  high  concentrations  (as 
high  as  1600  ppm  PCB  in  turtles  from 
polluted  locations)  are  found  in  fat 
bodies,  brain,  and  testes.  However, 


Bryan  et  al.  (1987)  indicated  that 
toxic  PCB  congeners  did  not  remain 
in  the  large  fat  reserves  of  female 
turtles,  as  some  had  suggested,  but 
were  passed  on  in  bulk  to  the  egg 
yolks. 

It  is  necessary,  therefore,  to  test 
tissue  or  egg  samples  to  ensure  that 
turtles  being  harvested  for  human 
consumption  are  not  loaded  to  a  dan- 
gerous degree  with  organochloride 
contaminants. 


Management  of  Snapping  Turtles 
as  Predators 

Several  studies  have  considered  the 
impact  of  snapping  turtles  on  water- 
fowl populations  (Alexander  1943, 
Hammer  1972,  Lagler  1943).  Highly- 
productive  bodies  of  water  present 
ideal  habitat  for  waterfowl  and  for 
turtles. 

Destroying  turtle  nesting  locations 
may  not  reduce  local  populations  of 
snapping  turtles,  because  females 
may  migrate  several  kilometers  be- 
tween their  usual  home  range  and 
their  nesting  sites  (Obbard  1977).  In 
addition,  such  habitat  interference 
will  remove  nesting  opportunities  for 
other  turtle  species. 

Reduction  in  numbers  of  adult 
snapping  turtles  through  trapping 
will  rapidly  deplete  isolated  popula- 
tions and  should  reduce  risks  to  prey 
species.  However,  if  turtles  can  emi- 
grate into  the  management  area,  then 
the  expected  long-term  effect  of  cull- 
ing adults  will  not  be  realized  be- 
cause the  population  can  increase 
from  these  new  adult  immigrants. 

Acknowledgments 

We  are  grateful  to  the  Ontario  Minis- 
try of  Natural  Resources  and  to  D. 
Strickland  for  their  support  and  for 
granting  permission  to  conduct  re- 
search in  Algonquin  Provincial  Park, 
and  to  the  Royal  Botanical  Gardens, 
Hamilton,  Ontario,  for  permission  to 
work  in  Cootes  Paradise.  We  thank 


C.  Bell,  M.L.  Bobyn,  M.  Fruetel,  J. 
Hughes,  K.  Lampman,  J.  Lay  field, 
and  S.  Plourde  for  field  and  technical 
assistance,  and  K.  Kovacs  and  S.  In- 
nes  for  computational  help.  This 
study  has  been  supported  by  Natural 
Sciences  and  Engineering  Research 
Council  Canada  Grant  A5990,  and  an 
Ontario  Ministry  of  Natural  Re- 
sources Renewable  Resource  Pro- 
gram Grant  to  Ronald  J.  Brooks. 

Literature  Cited 

Alexander,  Maurice  M.  1943.  Food 
habits  of  the  snapping  turtle  in 
Connecticut.  Journal  of  Wildlife 
Management  7:278-282. 

Bergmann,  Brian.  1983.  Carapace  ca- 
pers: uncontrolled  trapping  threat- 
ens Ontario's  snapping  turtles. 
Equinox  2(5):127. 

Bryan,  A.  M.,  P.  G.  Olafsson,  and  W. 
B.  Stone.  1987.  The  disposition  of 
low  and  high  environmental  con- 
centrations of  PCB's  in  snapping 
turtle  tissues.  Bulletin  of  Environ- 
mental Contamination  and  Toxi- 
cology 38:1000-1005. 

Congdon,  Justin  D.,  Gary  L.  Breiten- 
bach,  Richard  C.  van  Loben  Sels, 
and  Donald  W.  Tinkle.  1987.  Re- 
production and  nesting  ecology  of 
snapping  turtles  (Chelydra  ser- 
pentina) in  southeastern  Michigan. 
Herpetologica  43:39-54. 

Congdon,  Justin  D.,  A.  E.  Dunham, 
and  Donald  W.  Tinkle.  1982.  En- 
ergy budgets  and  life  histories  of 
reptiles,  p.  233-271.  In  C.  Gans  and 
F.  H.  Rough,  editors.  Biology  of 
the  Reptilia.  13.  Physiology  D- 
Physiological  Ecology.  Academic 
Press.  New  York. 

Congdon,  Justin  D.,  and  J.  Whitfield 
Gibbons.  1985.  Egg  components 
and  reproductive  characteristics  of 
turtles:  relationships  to  body  size. 
Herpetologica  41: 194-204. 

Congdon,  Justin  D.,  and  Donald  W. 
Tinkle.  1982.  Reproductive  ener- 
getics of  the  painted  turtle 
(Chrysemys  picta).  Herpetologica 
38:228-237. 


178 


Coulter,  Malcolm  W.  1957.  Predation 
by  snapping  turtles  upon  aquatic 
birds  in  marine  marshes.  Journal 
of  Wildlife  Manage.  21(1):17-21. 

Deevy,  Edward  S.,  Jr.  1947.  Life 
tables  for  natural  populations  of 
animals.  Quarterly  Review  of  Biol- 
ogy 22:283-314. 

Galbraith,  David  A.  1986.  Age  esti- 
mates, survivorship,  growth,  and 
maturity  of  female  Chelydra  ser- 
pentina Linnaeus  in  Algonquin 
Provincial  Park,  Ontario.  M.Sc. 
Thesis,  University  of  Guelph, 
Guelph,  Ontario. 

Galbraith,  David  A.,  and  Ronald  J. 
Brooks.  1987.  Survivorship  of 
adult  females  in  a  northern  popu- 
lation of  common  snapping 
turtles,  Chelydra  serpentina.  Cana- 
dian Journal  of  Zoology  65:1581- 
1586. 

Galbraith,  David  A.,  Ronald  J. 

Brooks,  and  Martyn  E.  Obbard.  In 
Press.  Female  age  and  body  size  at 
the  onset  of  reproduction  in  a 
northern  population  of  snapping 
turtles  (Chelydra  serpentina).  Cana- 
dian Journal  of  Zoology. 

Galbraith,  David  A.,  Mark  W.  Chan- 
dler, and  Ronald  J.  Brooks.  1987. 
The  fine  structure  of  home  ranges 
of  male  Chelydra  serpentina:  are 
male  snapping  turtles  territorial? 
Canadian  Journal  of  Zoology 
65:2623-2629. 

Hammer,  Donald  A.  1969.  Parame- 
ters of  a  marsh  snapping  turtle 
population:  Lacreek  Refuge,  South 
Dakota.  Journal  of  Wildlife  Man- 
agement 33:  995-1005. 

Hammer,  Donald  A.  1972.  Ecological 
relations  of  waterfowl  and  snap- 
ping turtle  populations.  Ph.D.  Dis- 
sertation, Utah  State  University, 
Salt  Lake  City,  Utah. 

Helwig,  D.  D.,  and  M.  E.  Hora.  1983. 
Polychlorinated  biphenyl,  mercury 
and  cadmium  concentrations  in 
Minnesota  snapping  turtles.  Bulle- 
tin of  Environmental  Contamina- 
tion and  Toxicology  30:186-190. 

Hogg,  D.  M.  1975.  The  snapping 
turtles  of  Wye  marsh.  Ontario  Fish 
and  Wildlife  Review  14(2):16-20. 


Iverson,  John  B.  1977.  Reproduction 
in  freshwater  and  terrestrial 
turtles  of  north  Rorida.  Herpetol- 
ogica  33:205-212. 

Iverson,  John  B.  1982.  Biomass  in 
turtle  populations:  a  neglected 
subject.  Oecologia  55:69-76. 

Kiviat,  Erik.  1980.  A  Hudson  River 
tidemarsh  snapping  turtle  popula- 
tion. Transactions  of  the  Northeast 
SecHon,  Wildlife  Society  37:158- 
168. 

Lagler,  Karl  F.  1943.  Food  habits  and 
economic  relations  of  the  turtles  of 
Michigan  with  special  reference  to 
fish  management.  29:257-312. 

Lovisek,  Jim.  1982.  An  investigation 
of  the  harvesting  of  turtles  in  On- 
tario. Nongame  Program  Wildlife 
Branch,  Ontario  Ministry  of  Natu- 
ral Resources. 

MacCulloch,  Ross  D.,  and  D.  M. 
Secoy.  1983.  Demography,  growth, 
and  food  of  western  painted 
turtles,  Chrysemys  picta  bellii 
(Gray),  from  southern  Sas- 
katchewan. Canadian  Journal  of 
Zoology  61:1449-1509. 

Mrosovsky,  N.,  and  C.  L.  Yntema. 
1980.  Temperature  dependence  of 
sexual  differentiation  in  sea 
turtles:  implications  for  conserva- 
tion practices.  Biological  Conser- 
vation 18:271-280. 

Obbard,  Martyn  E.  1977.  A  radio- 
telemetry  and  tagging  study  of 
activity  in  the  common  snapping 
turtle,  Chelydra  serpentina.  M.Sc. 
Thesis,  University  of  Guelph, 
Guelph,  Ontario. 

Obbard,  Martyn  E.  1983.  Population 
ecology  of  the  common  snapping 
turtle,  Chelydra  serpentina,  in  north- 
central  Ontario.  Ph.D.  dissertation. 
University  of  Guelph,  Guelph,  On- 
tario. 

Parmenter,  Robert  R.  1981.  Digestive 
turnover  rates  in  freshwater 
turtles:  the  influence  of  tempera- 
ture and  body  size.  Comparative 
Biochemistry  and  Physiology 
70(A):235-238. 

Pell,  Stuyvesant  Morris.  1941.  Notes 
on  the  habits  of  the  common  snap- 
ping turtle,  Chelydra  serpentina 


(Linn.),  in  central  New  York.  M.Sc. 
Thesis,  Cornell  University,  Ithaca, 
New  York. 

Shine,  Richard.  1980.  "Costs"  of  re- 
production in  reptiles.  Oecologia 
46:92-100. 

Yntema,  C.  L.  1976.  Effects  of  incuba- 
tion temperatures  on  sexual  differ- 
entiation in  the  turtle,  Chelydra  ser- 
pentina. Journal  of  Morphology 
150:453-462. 


179 


Spatial  Distribution  of  Desert 
Tortoises  (Gopherus  agassizii) 
at  Twentynine  Palms, 
California:  Implications  for 
Relocations^ 

Ronald  J.  Baxter^ 


Abstract.— The  spatial  distribution  of  desert 
tortoises  in  relation  to  plant  communities  was 
compared  against  randomness.  Tortoise  captures  (n 
=  120)  and  tortoise  burrows  (n  =  160)  exhibited  non- 
random  distributions  across  a  1.29  square  kilometer 
study  plot  at  Twentynine  Palms,  California.  Results 
imply  high  diversity  plant  ecotones  and 
communities,  and  possibly  soil  characteristics  are 
important  in  determining  tortoise  densities.  Non- 
randomness  in  tortoise  populations  dictates  that 
relocation  sites  must  include  specific  vegetational, 
topographic  and  edaphic  habitats  used  by  the 
parental  populations. 


The  desert  tortoise  (Gopherus  agas- 
sizii) is  a  species  whose  future  is  un- 
certain. Increased  use  of  the  deserts 
by  man  (Luckenbach  1982)  has  led  to 
the  point  where  the  tortoise  was  offi- 
cially listed  as  "threatened"  in  the 
state  of  Utah  (Dodd  1980).  The  U.S. 
Fish  and  Wildlife  Service  stated  in 
1985  that  ''...listing  [of  the  desert  tor- 
toise as  a  threatened  or  endangered 
species]  is  warranted  but  precluded 
by  other  pending  proposals  of  higher 
priority"  (Federal  Register.  50(234): 
49868-49870, 1985). 

In  California,  the  desert  tortoise  is 
the  official  state  reptile,  and  is  fully 
protected  under  law.  The  tortoise  is 
also  protected  in  Arizona  and  Ne- 
vada. 

As  part  of  a  larger  population 
study  (Stewart  and  Baxter  1987)  at 
the  Twentynine  Palnns  Marine  Corps 
Air  Ground  Combat  Center 
(MCAGCC),  the  spatial  distributions 
of  tortoise  captures  and  burrows 
were  analyzed  and  compared  against 
randomly  generated  distributions. 
Questions  asked  were:  (1)  Are  tor- 
toise captures  and  burrows  ran- 
domly located  across  the  landscape 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff, AZ,  July  19-21,  1988). 

'Ronald  J.  Baxter  received  his  master's 
degree  in  biology  for  working  on  the  desert 
tortoise  while  at  California  State  Polytech- 
nic University,  Pomona.  He  is  currently  com- 
pleting his  doctorate  at  the  Department  of 
Biological  Sciences.  Northern  Arizona  Uni- 
versity. Flagstaff.  AZ,  8601 1-5640. 


and /or  are  they  associated  with  cer- 
tain habitat  types  or  site  characteris- 
tics, and  if  so,  (2)  what  implications 
do  these  distributions  have  for  future 
management  decisions? 

Mettiods 

Twentynine  Palms  MCAGCC  is  lo- 
cated approximately  5  kilometers 
north  of  Twentynine  Palms,  San  Ber- 
nardino County,  California,  in  the 
southwestern  extreme  of  the  Mojave 
Desert.  All  fieldwork  was  performed 
in  the  Sand  Hill  Training  Area  which 
is  in  the  southwest  corner  of  the 
MCAGCC.  Elevations  ranged  from 
865  meters  atop  Sand  Hill  to  about 
730  meters  in  the  bottom  of  Surprise 
Springs  wash.  Data  were  collected 
Monday  through  Friday,  14  April 
through  18  July,  1986. 

Systematic  searching  methods  for 
tortoises  and  tortoise  burrows  were  a 
derivation  of  procedures  described 
by  Berry  (1984).  A  1.29  square  kilo- 
meter permanent  study  plot  was  es- 
tablished, with  its  approximate  cen- 
ter being  the  NE  1  /4,  SW  1  /4,  NE  1  / 
4,  of  S7,  T2N,  R7E  (San  Bernardino 
Base  Meridian)  This  site  offered  a 
wide  variety  of  habitats  including 
washes,  sandy  basins,  rolling  hills 
and  alluvial  bajadas.  The  plot  was  di- 
vided into  64  equal  sized  "grids"  of 
142  meters  on  a  side,  with  grid  cor- 
ners marked  by  posts.  Grids  were 
searched  in  parallel  belts  until  the 
entire  plot  had  been  searched  twice; 


once  with  the  belts  running  north- 
south,  and  once  with  the  belts  run- 
ning east-west.  The  plot  was  also 
randomly  searched. 

When  an  active  tortoise  was  en- 
countered, it  was  marked,  weighed, 
sexed,  measured  and  photographed. 
Each  tortoise  was  assigned  a  unique 
number,  and  marginal  scutes  were 
notched  with  a  small  triangular  file 
for  relatively  permanent  identifica- 
tion. The  precise  location  of  the  cap- 
ture was  noted  by  its  distance  (meas- 
ured by  rangefinder)  and  compass 
aspect  to  the  nearest  grid  post.  Data 
collected  at  each  capture  site  in- 
cluded plant  community,  tempera- 
tures at  the  ground,  1  centimeter, 
and  1  meter,  cloud  cover,  wind 
speed  and  direction,  closest  burrow, 
closest  plant,  and  any  unusual  be- 
havior. 

Precise  location  of  tortoise  bur- 
rows were  similarly  determined  by 
rangefinder  and  compass.  Data  col- 
lected at  each  burrow  included  plant 
community,  distance  and  identifica- 
tion of  nearest  ecotone,  distance  to 
nearest  wash,  distance  to  nearest  Hi- 
laria  rigida,  slope  aspect  and  steep- 
ness, opening  compass  aspect  and 
position,  length,  depth,  and  tunnel 
characteristics.  In  this  study  area,  it 
was  difficult  to  determine  if  a  bur- 
row high  on  a  slope  above  a  wash 
was  part  of  the  wash  "system." 
Therefore,  it  was  arbitrarily  decided 
to  include  burrows  in  the  wash  plant 
community  only  if  they  were  actually 
found  the  wash  bed. 


180 


Six  visually  identified  plant  com- 
munities (Latr/  Amdu,  Hiri/ Amdu, 
Mixed,  Wash,  Sparse  Wash  and 
Meadow)  were  mapped  within  the 
study  plot,  and  seven  15-meter  line 
transects  (total  of  105  meters)  were 
measured  which  included  bare 
ground  as  a  species.  Transects  were 


i:latr/amdu  5:  meadow 
2:mixed  •'.burrow 


3:SPARSE  WASH  ff-mSH  NORTH 

4:HIRI/AMDU  Soo' 

METERS 


Figure  1  .—Approximate  distribution  of  piant 
communities  and  tortoise  burrows  across 
the  study  piot.  See  text  for  expianation  of 
piant  community  names. 


randomly  located  in  each  of  these  six 
communities.  Standard  transect  sta- 
tistics (density,  coverage,  frequency, 
relative  density,  relative  coverage, 
relative  frequency  and  importance 
values;  Brower  and  Zar  1984)  were 
computed  for  each  community. 
Simpson's  diversity  indices  (Simpson 
1949)  were  computed  and  compared 
with  Student's  t-tests  (Keefe  and 
Bergerson  1977).  Available  annuals 
as  well  as  perennials  were  used  to 
give  the  best  estimate  possible  for 
diversity.  In  addition,  seven  soil 
samples  were  taken  in  each  commu- 
nity, and  analyzed  for  soil  separates 
(Brower  and  Zar  1984)  and  soil  cal- 
cium (Hach  1983).  Finally,  nine 
''sand  scats"  were  collected  during 
the  field  work  and  tested  for  calcium. 

A  random  model  for  capture  and 
burrow  locations  was  formed  by 
combining  a  number  of  statistical 
tests.  First  a  master  map  of  the  plot 
was  constructed  from  actual  field 
data  at  a  scale  of  1:2000.  All  capture 
positions,  burrows  and  plant  com- 
munity boundaries  were  plotted  on 
this  map  and  checked  against  aerial 
photographs.  The  area  covered  by 
each  plant  community  was  then  de- 
termined by  the  use  of  a  planimeter. 
An  X-Y  scale  ranging  from  0  to  8  was 
plotted  on  the  sides  of  the  map,  and 
a  list  of  328  random  numbers  was 
generated  by  computer.  These  num- 
bers were  paired,  and  the  pairs  be- 


came the  X-Y  coordinates  of  random 
positions  against  which  observed 
capture  and  burrow  locations  were 
compared.  Distances  to  the  nearest 
wash  and  ecotone  were  determined 
for  these  random  locations  by  meas- 
uring them  on  the  map,  and  com- 
pared against  observed  by  Student's 
t-tests  (Zar  1974).  Observed  capture 
distances  were  sometimes  combined 
with  previous  data  recorded  in  this 
area  (Baxter  and  Stewart  1986). 

A  lack  of  habitat  preference  may 
be  suggested  if  burrows  and  captures 
were  found  in  the  same  relative 
abundance  as  the  plant  communities. 
In  addition,  if  the  expected  plant 
abundance  distribution  differed  sig- 
nificantly from  random  an  extrapola- 
tion of  observed  distributional  char- 
acteristics could  be  accomplished.  An 
assumption  of  this  test  was  that  a 
distribution  of  randomly  generated 
locations  (with  randomness  con- 
firmed) produced  a  random  fre- 
quency distribution.  Expected  fre- 
quencies for  burrows  and  captures 
were  generated  by  multiplying  the 
total  number  of  actual  burrows  or 
captures  by  the  percent  of  the  plot 
encompassed  by  each  plant  commu- 
nity. These  values  were  compared  by 
a  goodness-of-fit  chi-square  test  (Zar 
1974).  In  addition,  the  number  of 
burrows  or  captures  per  grid  were 
compared  against  expected  values  as 
derived  from  the  Poisson  distribu- 
tion by  a  goodness-of-fit  test. 

Results 

Plant  Communities  and  Soils 

Vegetation  analyses  revealed  six  dis- 
tinct plant  assemblages  (table  1;  fig. 
1).  Plant  community  distributions 
generally  reflected  the  relief  of  the 
plot.  The  higher,  more  well-drained 
hills  were  dominated  by  an  associa- 
tion of  Larrea  tridentata  and  Ambrosia 
dumosa,  which  encompassed  plot 
area  the  most  ("Latr/ Amdu";  table 
1)  and  exhibited  relatively  high  plant 
diversity. 


r 


Table  1  .—Summary  of  plant  community  data  from  plant  transects  (total 
transect  length  =  105  meters). 


Simpson's 

Percent 

Plant 

No.  of 

No.  of 

diversity 

of  plot 

community^ 

species 

individuals 

index 

area 

Sparse  Wash 

16 

733 

0.6069 

5.6 

Hiri/Amdu 

8 

501 

0.6841 

4.0 

Mixed 

12 

349 

0,7247 

37.2 

Latr/ Amdu 

11 

306 

0.7688 

50.6 

Wash 

25 

292 

0.7914 

0.2 

Meadow 

15 

662 

0.7497 

1.7 

Bore  Areas 

0.7 

°See  text  for  explanation  of  community  names. 


181 


Found  on  37.2%  of  the  plot  area 
was  the  "n\ixed"  community  that 
generally  occupied  intermediate  ar- 
eas between  the  Latr/ Amdu  and  ei- 
ther washes  or  areas  of  high  Hilaria 
rigida  density.  It  was  characterized 
by  the  association  of  L.  tridentata,  A. 
dumosa  and  H.  rigida,  and  was  found 
most  often  on  the  slopes  above,  and 
narrow  linings  next  to  washes.  The 
edge,  or  ecotone,  of  this  community 
with  the  Latr/  Amdu  community  is 
extensively  discussed  below. 

A  highly  diverse  plant  community 
was  found  in  the  washes  (table  1;  ap- 
pendix 1).  Such  areas  not  only  con- 
tained these  perennial  species,  but 
also  a  significant  number  of  other 
species  found  only  in  this  commu- 
nity, giving  it  the  highest  species 
richness  of  any  community. 

Small  uplifts  within  wash  channels 
seemed  to  support  a  more  open  type 
of  wash  vegetation,  "sparse  wash." 
Such  areas  had  many  species  com- 
mon to  the  washes  (appendix  1 ),  yet 
much  of  this  community  was  essen- 
tially pure  stands  of  the  opportunis- 
tic grass,  Schismus  barbatus. 

A  community  ("Hiri/Amdu") 
consisting  primarily  of  H.  rigida  and 
A.  dumosa  was  located  in  upland  ba- 
sins where  L.  tridentata  was  not 
found.  Such  areas  were  low  in  habit 
and  diversity,  and  very  sandy. 

Finally,  near  the  south  boundary 
of  the  plot,  a  small  "meadow"  of 
mostly  Bailey  a  multiradiata  was 
found.  Since  no  tortoises  or  tortoise 
burrows  were  found  there,  it  was 
eliminated  from  further  analyses. 

Bare  ground,  when  treated  as  a 
species  in  transect  analyses,  had 
overriding  importance  values  and 
dominance  in  all  communities  (ap- 
pendix 1).  This  is  often  the  case  in 
desert  environments.  Likewise  im- 
portance values  of  S.  barbatus  were 
extremely  high  in  all  communities, 
pointing  to  the  generally  disturbed 
nature  of  the  site.  Comparisons  of 
Simpson  indices  for  the  communities 
revealed  significant  differences  (p  < 
0.05)  in  diversity  for  all  communities 
except  two.  The  Latr/ Amdu  and 


wash  communities  were  not  signifi- 
cantly different  (p  >  0.50)  in  their  di- 
versity. 

Soils  were  found  to  be  somewhat 
similar  in  constituency  (table  2),  each 


being  composed  to  a  large  degree  of 
sand.  Soil  calcium  levels  (table  3) 
were  shown  to  differ  significantly. 
No  detectable  calcium  was  found  in 
any  of  the  sand  scats  tested. 


Tabie  2.— Summary  of  percent  soli  separcstes  for  plant  communities. 


Plant 

Silt 

community" 

Sand 

(%) 

Clay 

Classification 

Sparse  mixed 

87 

2 

11 

loamy  sand 

Hiri/Amdu 

90 

8 

2 

sand 

Mixed 

85 

12 

3 

loamy  sand 

Latr/ Amdu 

70 

20 

10 

sandy  loam 

Wash 

81 

3 

16 

sandy  loam 

Meadow 

63 

3 

33 

sandy  day  loam 

°See  text  for  explanation  of  community  names. 


r 


Table  3.— Summary  of  soil  calcium  levels  and  their  significance. 

Plant 

Mean  soil  calcium 

Significantly 

community*' 

Cmeq/100  mg  soli) 

different  from 

Latr/Amdu 

6.43 

Hiri/Amdu 

<0.006 

Wash 

<0.05 

Mixed 

NS 

Hiri/Amdu 

3.00 

Wash 

N$ 

Mixed 

<0.05 

Mixed 

1.48 

Wash 

NS 

Wash 

0.57 

°Se0  text  for  explanation  of  community  names. 

^2'Sampte  t-test;  corrected  for  type  t  errors:  N$  =  not  significant. 

Table  4.— Distributions  and  significance  of  tortoise  burrows  per  grid  (Pols- 
son,  n  =  64). 


°P>0.9d 
^P>Q.7S 


J 


Number  of 

Number  of  grids 

Expected  values 

burrows/grid 

random 

observed 

random" 

observed^ 

0 

6 

6 

4.95 

5.52 

1 

11 

17 

12.67 

13.53 

2 

15 

11 

16.21 

16.58 

3 

15 

14 

13.83 

13.54 

4 

10 

8 

8.85 

8.29 

5  or  more 

7 

8 

4.67 

6.30 

182 


Table  5.— Summary  of  frequency  of  tortoise  burrows  compared  to  plant 
community  abundance. 


Plant  community*' 

Sparse 
Wasti 

HIrl/ 
Amdu 

Mixed 

Latr/ 
AmHij 

Wast! 

Other 

%  of  Diot 

5  6 

A  n 

37.2 

50.6 

0.2 

2.4 

Random'^ 

observed 
expected 
Observed^ 

14 
9.2 

6 

6.6 

55 
61.0 

76 
83.0 

8 

0.3 

5 

3.9 

observed 
expected 

11 
8.8 

2 

6.3 

68 
58.8 

75 
80.0 

1 

0.3 

1 

3.8 

°See  text  for  explanation  of  community  names. 

''P<0.00hn=164. 

^P>0.25,n=  158. 

  _J 


Table  6.— Comparison  of  distance  to  washes  between  observed  and  ran- 
dom tortoise  burrows. 


Plant  Mean  distance  (m  (SEM)) 


community*" 

random 

observed 

t 

freedom 

P 

All 

96.83 

101.21 

0.424 

318 

>0.50 

communities 

(6.80) 

(7.83) 

Mixed 

79.54 

68.66 

0.821 

120 

>0.50 

(9.75) 

(8.90) 

Latr/ Amdu 

132.05 

145.40 

0.845 

148 

>0.50 

(10.07) 

(12.17) 

°See  text  for  explanation  of  community  names. 


J 


r 


Table  7.— Comparison  of  distances  to  ecotone  between  observed  and  ran- 
dom tortoise  burrows. 


Plant 

community*' 

Mean  distance  (m  fSFMD 
random  observed 

t 

Degrees  of 
freedom 

P 

All 

96.83 

101.21 

0.424 

318 

>0.50 

Latr/ Amdu 

33.63 

15.21 

5.360 

137 

<  0.0005 

(2.65) 

(1.80) 

Mixed 

38.33 

12.18 

3.650 

65 

<  0.0005 

(9.09) 

(1.60) 

Combined 

34.05 

13.99 

6.493 

203 

<  0.0005 

(3.00) 

(1.26) 

""See  text  for  explarxjtion  of  community  names. 

  / 


Tortoise  Burrov^s 

A  total  of  164  tortoise  burrows  was 
found  on  the  study  plot  (fig.  1).  Sev- 
enty-five percent  were  found  under 
bushes,  14%  with  the  opening  under 
a  bush  but  the  tunnel  proceeding  into 
an  open  area,  8%  with  entrances  in 
the  open  but  the  tunnels  proceeding 
under  a  bush,  and  3%  entirely  in  an 
open  area.  Thus,  almost  all  of  bur- 
rows (97%)  were  associated  with 
shrubs.  Of  these,  71%  were  associ- 
ated with  L.  tridentata,  13%  each  with 
H.  rigida  and  A.  dumosa,  and  another 
3%  with  other  species. 

Neither  the  distribution  of  ob- 
served or  random  burrows  differed 
significantly  from  the  Poisson  ex- 
pected frequencies  (table  4).  Like- 
wise, when  the  distribution  of  ob- 
served burrows  was  compared 
against  the  distribution  of  random 
burrows,  no  significant  difference 
was  found  (chi-square  =  2.224;  DF  = 
5;  p  >  0.50).  Thus,  when  the  entire 
plot  area  is  considered,  tortoise  bur- 
rows exhibited  a  random  pattern 
across  the  landscape.  However,  this 
was  a  relatively  large  scale  test  of 
burrows  per  arbitrary  unit  area,  and 
says  nothing  about  the  pattern  of  tor- 
toise burrows  in  relation  to  plant 
communities. 

The  abundances  of  tortoise  bur- 
rows (both  observed  and  random)  in 
each  plant  community  were  com- 
pared against  expected  frequencies 
generated  by  the  abundances  of  the 
plant  communities  (table  5).  Burrows 
were  sparse  in  the  Hiri/Amdu  and 
wash  communities.  Observed  bur- 
row frequency  distribution  did  not 
differ  significantly  (p  >  0.25)  from  the 
expected  frequency  distribution.  The 
observed  frequency  distribution  dif- 
fered significantly  from  the  random 
distribution  (chi-square  =  11.74;  DF  = 
5;  p  <  0.05),  as  did  the  expected  dis- 
tribution (chi  square  =  158.9;  DF  =  5; 
p<  0.001). 

Mean  observed  burrow  distance 
to  the  closest  wash  was  compared  to 
the  mean  distance  from  the  ran- 
domly located  burrows  (table  6). 


183 


Comparisons  for  the  sparse  wash 
and  Hiri/Amdu  communities  were 
not  done  because  they  would  be  bio- 
logically meaningless  or  had  too  low 
a  sample  size,  respectively.  For  all 
burrows,  and  for  burrows  found  in 
either  the  Latr/ Amdu  or  mixed  com- 
munities, no  significant  differences 
between  random  and  observed  wash 
distances  were  detected.  Thus,  ob- 
served tortoise  burrows  were  not  lo- 
cated closer  to  washes  than  a  set  of 
random  points  predicted.  However, 
examination  of  the  spatial  pattern 
(fig.  1)  reveals  a  lack  of  burrows  deep 
within  Latr/Amdu  and  Hiri/Amdu 
areas  which  were  furthest  away  from 
any  possible  wash  influence. 

Past  observations  seemed  to  indi- 
cate a  correlation  between  burrow 
location  and  the  presence  of  the  edge 
of  the  H.  rigida  distribution  (Baxter 
and  Stewart  1986).  The  approximate 
distribution  of  observed  burrows  to 
this  edge  may  be  seen  in  figure  1 . 
Mean  edge  (ecotone)  distance  of  ob- 
served burrows  was  compared  to 
that  of  random  sites  (table  7).  Highly 
significant  differences  in  ecotone  dis- 
tances were  found  in  both  communi- 
ties, and  also  when  combined.  Thus, 
burrows  were  found  closer  to  the 
ecotone  than  a  set  of  random  points. 


Tortoise  Captures 

Similar  analyses  were  performed  for 
tortoise  capture  sites.  There  were  a 
total  of  120  tortoise  captures  and  re- 
captures of  41  individual  tortoises. 
The  observed  captures  per  grid, 
along  with  the  randomly  located  cap- 
ture frequencies  (same  points  used 
for  random  burrow  sites)  were  com- 
pared against  expected  values  de- 
rived from  the  Poisson  distribution 
(table  8).  Observed  capture  sites 
showed  a  statistically  significant  de- 
parture from  Poisson  expected  fre- 
quencies by  the  goodness-of-fit  test 
(p  <  0.05). 

Frequencies  of  capture  sites  in 
each  plant  community  were  com- 
pared against  expected  values  gener- 


ated by  community  abundance  (table 
9).  Observed  distributions  for  both 
all  captures,  and  for  captures  of  ac- 
tive tortoises  (those  found  outside  of 
burrows)  differed  significantly  from 
expected.  These  two  observed  distri- 
butions did  not  differ  from  each 
other  (chi-square  =  0.5385;  DF  =  5;  p 
>  0.99),  yet  differed  significantly 
from  the  randomly  generated  distri- 
bution (chi-square  =  18.957  and 
19.556,  respectively;  DF  =  5;  p  < 
0.005).  Thus,  tortoise  captures  were 
not  found  across  the  plot  in  a  ran- 
dom fashion  as  would  be  predicted 


by  a  set  of  randomly  generated 
points.  Habitat  preference  for  washes 
was  seemingly  indicated,  as  was  a 
lack  of  preference  for  Hiri/Amdu 
areas.  These  results  also  gave  further 
support  to  the  non-randomness  ex- 
hibited in  the  Poisson  analyses. 

To  further  examine  this  apparent 
non-random  distribution  of  capture 
locations,  the  mean  observed  capture 
distance  to  washes  was  compared  to 
that  of  the  randomly  located  sites 
(table  10).  When  all  capture  sites,  or 
captures  within  the  mixed  commu- 
nity were  considered,  a  significant 


Table  8.—Distrlbutlons  and  significance  of  tortoise  captures  per  grid  (Pois- 
son, n=  64). 


Number  of 
captures/grid 


Number  of  grids 
random  observed 


Expected  values 
random"  observed^ 


0 
1 
2 
3 

4  or  more 


6 
VI 
15 
15 
17 


19 
17 
15 
4 
9 


4.95 
12.67 
16.21 
13.83 
13.59 


9.97 
18.54 
17.23 
10.68 

5.54 


''P>0.90 
'>P<  0,026 


Table  9.— Summary  of  frequt^ncy  of  captures  compared  to  plant  commu- 
nity abundance. 


Plant  community*" 


Sparse 

HIrl/ 

IViixed 

Latr/ 

Wash 

Other 

Wash 

Amdu 

Amdu 

%  of  plot 

5.6 

4.0 

37.2 

50.6 

0.2 

2.4 

Random'^ 

observed  (n=164) 

14 

6 

55 

76 

8 

5 

expected 

9.2 

6.6 

61.0 

83.0 

0.3 

3.9 

Observed  (oll)^ 

observed  (n=120) 

14 

1 

33 

48 

23 

1 

expected 

6.7 

4.8 

44.6 

60.7 

0.3 

2.9 

Observed  (active)^ 

observed  (n=81) 

9 

1 

20 

32 

18 

1 

expected 

4.5 

3.3 

30.1 

41.0 

0.2 

1.9 

°See  text  for  explanation  of  community  names. 
""Pk  0.001 


184 


difference  between  random  and  ob- 
served locations  was  demonstrated. 
However,  mean  distance  to  washes 
within  Latr/ Amdu  sites  was  not  sig- 
nificantly different  from  the  random 
set  of  points,  possibly  because  the 
Latr/ Amdu  communities  were  gen- 
erally located  further  away  from 
washes,  as  well  as  the  high  variation 
in  observed  Latr/Amdu  distances. 
These  results,  along  with  the  results 
of  the  community  analysis  above, 
seemed  to  indicate  a  high  degree  of 
tortoise  activity  near  the  washes. 

Distances  to  the  edge  of  the  H. 
rigida  were  compared  between  ran- 
domly generated  and  observed  cap- 
ture locations  (table  11).  Highly  sig- 
nificant differences  in  mean  distances 
were  demonstrated  for  both  the 
Latr/Amdu  community,  and  for  cap- 
tures found  in  the  mixed  and  Latr/ 
Amdu  communities  combined.  Cap- 


tures within  the  mixed  community 
alone  were  not  significantly  different 
from  randomly  generated  locations. 
It  seems  then  that  captures,  like  bur- 
rows, were  generally  not  found  far 
within  Latr/Amdu  areas,  but  tended 
to  be  near  its  edge  with  the  H.  rigida 
distribution  (i.e.  the  mixed  commu- 
nity). Because  there  was  no  differ- 
ence within  the  nnixed  community 
alone,  differences  from  random  for 
captures  within  the  mixed  and  Latr/ 
Amdu  communities  combined  were 
probably  significant  due  to  the 
higher  number  of  observations 
within  the  Latr/Amdu  community 
biasing  the  sample.  Thus,  it  seems 
that  tortoises  tended  to  stay  either 
near  the  washes,  the  mixed  commu- 
nity, or  its  ecotone  with  the  Latr/ 
Amdu  community,  and  generally 
were  not  going  far  within  the  Latr/ 
Amdu  community. 


Table  10.— Comparison  of  distance  to  washes  between  observed  and  ran- 
dom capture  locations. 


Plant 

community*' 

Mean  distance  rSfM)) 

t 

Degrees  of 
freedom 

random 

observed 

P 

All 

96,83 

71.86 

2.189 

258 

<0.05 

communities 

(6.80) 

(9.39) 

Mixed 

79.54 

44.14 

2.081 

73 

<0.05 

(9.75) 

(10.61) 

Latr/Amdu 

132.05 

133.66 

0.917 

117 

>  0.50 

(10.07) 

(15.12) 

''See  text  for  explanation  of  community  names. 


r   N 

Table  1 1  .—Comparison  of  distances  to  ecotone  between  observed  and 
random  capture  locations. 


Plant 

community*' 

Mean  distance  (m  (SEM)) 
random  observed 

Degrees  of 
t         freedom  P 

Latr/Amdu 

32.33 

18.59 

3.389 

114 

<  0.001 

(9.09) 

(3.05) 

Mixed 

38.63 

21.05 

1.595 

42 

>0.10 

(2.65) 

(4.65) 

Combined 

34.05 

13.99 

3.485 

157 

<  0.001 

(3.00) 

(2.53) 

=See  text  for  explanation  of  community  names. 


Discussion 

Since  the  establishment  in  1975  of  the 
Desert  Tortoise  Council,  the  amount 
of  literature  published  on  the  desert 
tortoise  has  been  considerable. 
Oddly  enough,  only  a  few  papers 
may  be  found  that  attempt  to  say 
what  exactly  makes  good  tortoise 
habitat. 

A  paper  by  Schwartzmann  and 
Ohmart  (1978)  quantified  the  fre- 
quency of  use  by  tortoises  in  a  num- 
ber of  "habitat  types."  Their  study 
took  place  in  the  Picacho  Mountains 
of  Arizona's  Sonoran  Desert,  where 
tortoises  are  known  to  frequent 
rocky  hillsides  and  are  absent  from 
valley  bottoms  (Fritts  1985).  Habitat 
preferences  are  just  the  opposite  in 
the  Mojave  Desert,  and  thus  their  re- 
sults may  not  be  applicable.  Like- 
wise, Walchuck  and  Devos  (1982) 
studied  tortoise  habitat,  but  this  was 
also  in  the  Sonoran  Desert  of  Ari- 
zona. 

In  a  draft  report,  Weinstein  et  al. 
(1986)  performed  several  multivari- 
ate analyses  on  the  large  Bureau  of 
Land  Management  tortoise  database. 
Several  attempts  were  made  to  corre- 
late abundance  with  habitat  charac- 
teristics. Not  only  were  many  of 
these  characteristics  derived  from  the 
extrapolation  of  large  scale  map  data, 
but  the  best  fit  analysis  was  found  by 
designating  "corrected  sign"  of  the 
transects  (the  dependent  variable; 
not  actual  population  numbers)  into 
arbitrary  categories.  Indeed,  one  of 
the  authors  (Berry  and  Nicholson 
1984)  has  shown  that  roughly  one- 
third  of  population  estimates  (7  out 
of  20  and  4  out  of  6)  based  on  sign 
transects  did  not  agree  with  intensive 
plot  censuses.  Also,  Turner  et  al. 
(1982)  stated  that  sign  transects 
"...cannot  provide  the  accuracy  and 
precision  needed..."  In  addition, 
Fritts  (1985)  stated  that  such 
transects  are  "...subject  to  error." 
Thus  the  accuracy  of  sign  transects 
are  open  to  serious  debate,  and  al- 
though the  discriminant  analysis 
showed  some  promise  as  a  method 


185 


for  accessing  regional  abundances, 
the  nature  of  the  analysis  and  the 
underlying  assumptions  of  both  the 
data  acquisition  and  techniques  leave 
much  to  be  desired. 

When  viewed  from  the  larger 
scale  of  regional  or  even  plot  area, 
these  data  seem  to  indicate  that  bur- 
rows were  found  in  a  random  fash- 
ion when  predicted  by  burrows  per 
unit  area.  However,  different  results 
may  have  been  obtained  by  changing 
the  size  and  shape  of  the  grids.  For 
example,  32  larger  rectangular  grids 
may  very  well  have  produced  differ- 
ent results  than  the  64  smaller  square 
grids  used  in  this  study.  In  addition, 
such  an  analysis  said  nothing  about 
distributions  in  relation  to  habitat 
characteristics.  Therefore,  such  a  test 
should  be  used  as  a  starting  point 
and /or  support  for  other  tests,  and 
locally  is  of  limited  use  by  itself  for 
describing  ecologically  meaningful 
patterns  which  may  exist. 

With  closer  examination,  these 
data  also  indicate  that  burrow  loca- 
tions were  assembled  in  a  pattern 
similar  to  the  non-random  distribu- 
tion of  plant  communities.  Within- 
community  examinations  revealed 
patterns  of  burrow  site  utilization, 
and  such  patterns  were  strongly  non- 
random.  At  Sand  Hill  then,  while  a 
majority  of  burrows  were  not  found 
in  washes,  they  were  often  found 
within  easy  walking  distance  to  a 
wash.  Very  often,  burrows  were  on 
slopes  high  above  washes,  and  possi- 
bly within  its  area  of  influence.  They 
were  not  found  far  within  either  the 
Latr/  Amdu  or  Hiri/ Amdu  commu- 
nities, but  were  tied  strongly  to  the 
edge  of  these  communities  with  the 
mixed  community. 

Washes  are  sometimes  cited  as 
being  of  great  importance  to  tortoise 
populations  (Burge  1978,  Hohman 
1977,  Lowe  1964).  However,  results 
of  this  study  indicated  that  tortoise 
burrows  were  not  significantly  closer 
to  washes  than  a  set  of  randomly  se- 
lected sites.  Burge  (1978)  found  207 
(26%)  of  783  burrows  and  pallets 
were  associated  with  washes.  Of 


these,  56  (27%)  were  actually  within 
a  wash  bed.  However,  Burge  appar- 
ently eliminated  some  burrows  from 
the  analysis  due  to  their  physical 
characteristics.  The  discrepancy  may 
be  due  to  the  definition  used.  In  this 
study,  wash  burrows  were  defined 
as  such,  only  if  they  were  actually 
within  the  sandy  wash  bottoms.  In 
this  way,  burrows  which  were  on 
wash  banks,  were  counted  as  being 
in  the  plant  community  of  the  bank. 
Burrows  located  on  wash  banks,  and 
even  further  away,  may  have  been 
associated  with  the  wash,  and  a  re- 
classification of  these  burrows  may 
show  washes  to  have  a  more  impor- 
tant influence  in  burrow  analyses. 
Examinations  of  the  actual  burrow 
distribution  (fig.  1)  seemed  to  indi- 
cate that  they  were  mostly  absent 
from  areas  highly  isolated  from  wash 
influence. 

The  significance  of  capture  loca- 
tions in  relation  to  the  washes  also 
seemed  to  refute  the  burrow /wash 
results.  Washes  clearly  supported  a 
disproportionate  amount  of  activity 
in  relation  to  their  abundance  on  the 
plot.  Preliminary  investigations  of 
tortoise  communities  near  Kramer 
Junction,  San  Bernardino  County, 
have  also  shown  tortoises  are  proba- 
bly localizing  their  activities  in  the 
vicinity  of  washes  (Baxter,  unpub. 
data). 

Several  things  may  explain  the 
disproportionate  amount  of  captures 
in  the  washes.  Greater  visibility  of 
tortoises  in  the  washes  may  be  a  fac- 
tor. Utilization  of  highly  diverse 
plant  resources  there  may  also  con- 
tribute to  the  localization  of  activity. 
Finally,  washes  may  simply  serve  a 
natural  highways  for  tortoise  move- 
ments. For  instance,  several  relocated 
tortoises  at  Kramer  Junction  abruptly 
turned  and  followed  trails  and 
washes  upon  their  release  (Baxter, 
unpub.  data).  Regardless,  these  data 
seem  to  support  washes  as  an  impor- 
tant habitat  characteristic  for  tor- 
toises at  Sand  Hill.  If  this  population 
is  representative  of  other  Mojave 
populations,  the  importance  of 


washes  in  p>otential  relocation  sites 
will  be  highly  significant  in  assuring 
the  best  chance  of  survival  for  the 
relocatees.  Further,  impacts  to 
washes  may  have  highly  significant 
impacts  on  a  population  if  it  is  local- 
izing its  activities  there. 

These  data  support  the  impor- 
tance of  large  woody  shrubs  (i.e.,  L. 
tridentata)  for  successful  burrow  con- 
struction at  this  site.  Similar  results 
have  been  reported  by  Burge  (1978) 
who  found  72%  of  "cover  sites"  as- 
sociated with  shrubs.  Berry  and 
Turner  (1984)  found  75%  of  juvenile 
burrows  associated  with  bushes. 
Support  for  the  burrow  roofs  and 
added  protection  from  predators  are 
likely  reasons  for  this  association. 
Regardless,  the  absence  of  L.  triden- 
tata from  the  Hiri/Amdu  community 
is  probably  a  major  reason  for  the 
tortoises  not  utilizing  those  areas. 
Unsuccessful  burrow  construction  by 
virtue  of  the  sandier  soils  is  another 
possibility.  This  latter  assumption  is 
supported  by  the  Weinstein  et  al. 
(1986)  analysis  which  showed  "soil 
diggibility"  as  a  highly  significant 
regression  variable. 

However,  the  lack  of  burrows 
deep  within  Latr/ Amdu  communi- 
ties is  not  explained  by  the  spatial 
abundance  of  L.  tridentata.  The  high 
frequency  of  burrows  and  captures 
point  out  that  something  is  being 
sought  there  by  the  tortoises.  Yet, 
deep  ventures  within  these  areas  ap- 
parently do  not  provide  resources 
that  are  unavailable  at  their  edges. 
Perhaps  the  higher  levels  of  soil  cal- 
cium found  there  are  being  utilized. 
Tortoises  must  support  a  massive, 
ossified  shell,  as  well  as  lay  eggs,  and 
calcium  may  be  a  very  important  nu- 
trient. Tortoises  have  been  observed 
eating  dirt  (geophagy)  and  then  pro- 
ducing "sand  scats,"  and  calcium 
levels  have  been  hypothesized  as  an 
explanation  for  this  behavior  (Sokol 
1971).  The  lack  of  calcium  in  the  sand 
scats  tested  seems  to  support  this 
hypothesis. 

In  contrast,  such  deep  ventures 
would  take  the  tortoises  away  from 


186 


the  distribution  of  H.  rigida,  and  the 
frequented  and  diverse  washes.  Al- 
though detailed  scat  analyses  were 
not  performed,  field  examination  of 
hundreds  of  scats  seemed  to  suggest 
that  H.  rigida  is  a  significant  dietary 
component.  Turner  and  Berry  (1986) 
found  H.  rigida  as  a  part  of  the  diet  of 
tortoises  near  Goffs,  California. 

It  would  seem  then  that  tortoises 
in  this  area  are  exhibiting  some  char- 
acteristics similar  to  "edge"  species. 
That  is,  tortoise  activity  is  centered 
on  the  two  communities  with  the 
highest  vegetational  diversity  that 
border  extensive  areas  of  H.  rigida. 
Since  burrows  are  closely  associated 
with  L.  tridentata,  they  in  turn  are 
found  primarily  along  the  only 
highly  diverse  ecotone  of  the  H. 
rigida  distribution  where  L.  tridentata 
importance  is  the  highest.  This  im- 
portance of  H.  rigida  and  L.  tridentata 
is  further  shown  in  appendix  1 .  The 
two  communities  where  tortoises 
were  not  found  (i.e.,  deep  Latr/ 
Amdu  and  Hiri/Amdu)  each  lack 
one  of  these  species.  The  assumption 
that  they  are  focusing  on  high  diver- 
sity areas  is  further  supported  by 
Weinstein  et  al.  (1968)  which  shows 
"food  availability"  as  the  single  most 
significant  regression  variable.  Fi- 
nally, Speake  (1986)  reports  that  for 
the  gopher  tortoise  (G.  polyphemus), 
"Edge  habitats  or  ecotonal  areas  ap- 
pear important  to  tortoises.  In  each 
habitat  type  except  oldfields  tortoises 
tended  to  cluster  near  the  edges.  In 
general,  the  more  edge  availability  in 
a  given  habitat,  the  higher  the  tor- 
toise density." 

In  summary,  tortoises  utilized  the 
environment  at  Sand  Hill  in  a  mostly 
non-random  fashion.  Tortoise  cap- 
tures were  spread  out  between  two 
communities  of  highly  diverse  re- 
sources, with  clustering  occurring  at 
either  edge.  Tortoises  frequented 
washes  and  the  ecotonal  edge  of  the 
Latr/Amdu  community,  with  many 
found  in  the  intermediate  mixed 
community.  Tortoises  were  not 
found  deep  within  Latr/Amdu  or 
Hiri/Amdu  areas.  Burrows  were 


found  close  to  the  ecotone  of  the 
mixed  and  Latr/Amdu  connmunities. 
Burrows  were  not  found  closer  to 
washes  than  randomly  located  bur- 
rows, although  this  point  is  far  from 
clear.  Burrows  were  located  close  to 
the  one  highly  diverse  edge  of  tor- 
toise activity  area  where  the  impor- 
tance of  L.  tridentata  and  soil  calcium 
were  the  greatest,  and  were  not 
found  in  Hiri/Amdu  areas  where  L. 
tridentata  was  absent,  and  soils  were 
the  most  unconsolidated. 

Non-randomness  in  tortoise  popu- 
lations is  especially  important  for  the 
management  considerations  of  relo- 
cation. Clearly,  despite  the  best  ef- 
forts of  concerned  managers,  the  use 
of  the  deserts  will  continue  to  in- 
crease and  the  frequency  of  tortoise 
relocations  will  also  undoubtedly  in- 
crease. If  tortoise  distributions  are 
random,  relocation  management  es- 
sentially becomes  a  search  for  safe 
relocation  sites  roughly  similar  to  the 
"parental"  area.  No  special  consid- 
erations of  unique  habitat  types  are 
required.  If  on  the  other  hand  they 
are  not,  then  the  relocation  site(s) 
must  include  such  high-use  habitats 
as  those  found  in  the  parental  site.  In 
addition,  severe  disturbance  of  such 
favored  habitats  will  in  turn  have  se- 
vere impacts  on  the  populations,  par- 
ticularly if  small. 

This  study  indicates  that  the  non- 
randomness  exhibited  by  the  Sand 
Hill  tortoises  is  probably  a  function 
of  the  non-randomness  of  highly  di- 
verse plant  assemblages  and  edaphic 
characteristics.  Thus,  the  presence  of 
diverse  land  forms  and  their  associ- 
ated plant  communities  and  diverse 
edges  within  future  relocation  sites 
should  be  of  significant  importance 
to  the  manager.  Areas  which  "look 
good"  to  the  relocation  manager  may 
not  supply  the  needed  resources  for 
the  relocatees.  These  data  are  in  need 
of  further  support  however.  If  such 
patterns  are  exhibited  in  other  popu- 
lations, biologists  and  managers  may 
use  such  techniques  to  successfully 
determine  possible  habitat  require- 
ments, and  help  insure  the  survival 

187 


of  one  of  the  Mojave's  most  enig- 
matic species. 

Acknowledgments 

The  author  wishes  to  express  sincere 
thanks  to  Dr.  Glenn  R.  Stewart  of  Cal 
Poly,  Pomona  for  physical  help  and 
moral  support  during  the  fieldwork, 
and  for  his  abiding  friendship.  Many 
thanks  also  to  the  entire  staff  at  the 
MCAGCC  for  logistical  support.  Fi- 
nally, thanks  to  K.  Berry,  D.  Speake 
and  R.  Szaro  for  their  constructive 
reviews  of  this  manuscript. 

This  work  was  supported  by 
United  State  Navy  contract 
N6247484RPOOV48,  which  was  ad- 
ministered by  the  Cal  Poly  Kellogg 
Unit  Foundation.  Additional  equip- 
ment support  was  supplied  by 
graduate  research  funds  of  Cal  Poly, 
and  monies  received  from  the  Chuck 
Bayless  and  Tim  Brown  memorial 
scholarship  funds.  Travel  funds  were 
supplied  by  Sigma  Xi,  The  Scientific 
Research  Society. 

Literature  Cited 

Baxter,  Ronald  J.,  and  Glenn  R.  Ste- 
wart. 1986.  Report  of  the  continu- 
ing fieldwork  on  the  desert 
tortoise  (Gopherus  agassizii)  at  the 
Twentynine  Palms  marine  corps 
base.  Proceedings  of  the  sympo- 
sium, [Palmdale,  Calif.,  March, 
1986].  The  Desert  Tortoise  Coun- 
cil, Long  Beach,  Calif  [in  press]. 

Berry,  Kristin  H.  1984.  A  description 
and  comparison  of  field  methods 
used  in  studying  and  censusing 
desert  tortoises.  Appendix  II.  In 
The  status  of  the  desert  tortoise 
(Gopherus  agassizii)  in  the  United 
States.  Kristin  Berry,  editor.  Re- 
port to  the  U.S.  Fish  and  Wildlife 
Service,  Sacramento,  Calif.  Order 
No.  11310-0083-81. 

Berry,  KrisHn  H.,  and  Lori  L. 

Nicholson.  1984.  The  distribution 
and  density  of  desert  tortoise 
populations  in  California  in  the 


1970's.  In  Kristin  Berry,  editor. 
The  status  of  the  desert  tortoise 
(Gopherus  agassizii)  in  the  United 
States.  Report  to  the  U.S.  Fish  and 
Wildlife  Service,  Sacramento,  CA. 
Order  No.  11310-0083-81. 

Berry,  Kristin  H.,  and  Frederick  B. 
Turner.  1984.  Notes  on  the  behav- 
ior and  habitat  preferences  of  ju- 
venile desert  tortoises  (Gopherus 
agassizii)  in  California.  Proceed- 
ings of  the  symposium,  [Lake  Ha- 
vasu  City,  Ariz.,  March,  1984].  The 
Desert  Tortoise  Council,  Long 
Beach,  CA. 

Brower,  James  E.  and  Jerrold  Zar. 
1984.  Field  and  laboratory  meth- 
ods for  general  ecology.  William 
E.  Brown,  publishers.  Dubuque, 
Iowa. 

Burge,  Betty.  1978.  Physical  charac- 
teristics and  patterns  of  utilization 
of  cover  sites  used  by  Gopherus 
agassizii  in  southern  Nevada.  Pro- 
ceedings of  the  symposium,  [Las 
Vegas,  Nevada,  1978].  The  Desert 
Tortoise  Council,  Long  Beach,  CA. 

Dodd,  C.  Kenneth.  1980.  Endangered 
and  threatened  wildlife  and 
plants:  listing  as  threatened  with 
critical  habitat  for  the  Beaver  Dam 
slope  population  of  desert  tortoise 
in  Utah.  Federal  Register  45(163): 
55654-55666. 

Fritts,  Thomas  H.  1985.  Ecology  and 
conservation  of  North  American 
Tortoises  (genus  Gopherus).  II. 
Evaluation  of  tortoise  abundance 
based  on  tortoise  sign  detected  in 
field  surveys.  Prepared  for:  U.S. 
Fish  and  Wildlife  Service,  Denver 
Wildlife  Research  Center,  Univer- 
sity of  New  Mexico,  Albuquerque. 
NM. 

Hach  Company.  1983.  Soil  calcium 
and  magnesium  test  kit  (model 
14855)  instruction  manual.  Hach 
Company,  Inc.,  Loveland,  CO. 

Hohman,  Judy  P.  1977.  Preliminary 
investigations  of  the  desert  tor- 
toise on  the  Beaver  Dam  slope  in 
Arizona.  Proceedings  of  the  sym- 
posium, [Las  Vegas,  Nev.,  March 
1977].  The  Desert  Tortoise  Coun- 
cil, Long  Beach,  CA. 


Keefe,  T.  J.  and  E.  Bergerson.  1977.  A 
simple  diversity  index  based  on 
the  theory  of  runs.  Water  Re- 
sources 11:689-691. 

Lowe,  Charles.  1964.  The  vertebrates 
of  Arizona.  The  University  of  Ari- 
zona press,  Tucson,  Ariz. 

Luckenbach,  Roger  A.  1982.  Ecology 
and  management  of  the  desert 
tortoise  (Gopherus  agassizii)  in 
California.  In  R.  Bruce  Bury,  edi- 
tor. North  American  tortoises: 


conservation  and  ecology.  U.S. 
Fish  and  Wildlife  Service,  Wildlife 
research  report  No.  12,  Washing- 
ton, D.C. 

Schwartzmann,  James  L.,  and  Robert 
D.  Ohmart.  1976.  Quantitative 
vegetational  data  of  desert  tortoise 
(Gopherus  agassizii)  habitat  in  the 
lower  Sonoran  Desert.  Proceed- 
ings of  the  symposium,  [Las  Ve- 
gas, Nev.  March,  1978]  The  Desert 
Tortoise  Council,  Long  Beach,  CA. 


Appendix  1 

Summary  of  Importance  Values^  From  Plant  Transect  Data. 


Plant  Community^ 


Species 

1 

2 

3 

4 

5 

Dare  vjrouna 

1  Oft  0 

1  IJ.O 

1 1  n 

1  lU.O 

114  7 
1 14./ 

yD.D 

Dcnismus  oaroatus 

Q 

Rl  ft 

AQ  1 
^y.  1 

OA  ft 
Zo.o 

40  A 

Larrea  tridentata 

16.5 

21.6 

36.7 

6.7 

Ambrosia  dumosa 

lie 

lie: 
1 1.0 

07  Q 

Z/.y 

0'2  ft 
ZO.O 

A  1 

riuaria  rtgiaa 

0  Q 

l^Q  Q 

Oy.y 

07 
L/ .o 

A  7 
O./ 

Erodtum  texanum 

11.9 

6.8 

10.2 

26.4 

19.3 

Malacothrix  spp. 

7  A 

OQ  1 

1C  o 

oo.z 

1  n  c 

Eriogonum  spp. 

/.U 

o  c 
Z.o 

1  "3  1 
lO.l 

o.o 

Hymenoclea  salsola 

111 
11.1 

O  7 
Z./ 

OA  A 

Zo.4 

/iTnsincicui  spp. 

Q  ft 
y.o 

0  0 

z.z 

Oenothera  deltoides 

6.4 

4.8 

13.6 

Baileya  multiradiata 

2.0 

6.8 

6.5 

Abronia  villosa 

5.6 

2.2 

2.1 

Bromus  rubens 

2.7 

2.2 

Langloisia  Matthervsii 

4.9 

2.6 

Langloisia  Palmeri 

2.3 

Oryzopsis  hymenoides 

2.5 

Eriophyllum  Wallacei 

2.0 

Menodora  spinescens 

2.9 

5.7 

Lesquerella  Palmeri 

2.3 

3.1 

Salazaria  mexicana 

3.0 

Dalea  Fremontii 

10.6 

Cucurbita  foetidissima 

2.3 

Euphorbia  polycarpa 

2.0 

Isomeris  arborea 

3.4 

Prunus  fasiculata 

6.7 

Spheralcea  ambigua 

2.6 

Salvia  columbariae 

4.0 

Phacelia  spp. 

2.2 

Petalonyx  Thurberi 

2.6 

Unknown  composite  #1 

2.9 

2.3 

7.3 

Unknown  composite  #2 

2.3 

'Importance  value  =  relative  density  +  rel.  domin.  +  rel.  freq. 

'Plant  community:  See  text  for  description  of  community  names:  Meadow  and  bare 
areas  not  listed:  I  =  Sparse  Wash):  2  =  Hiri/Amdu:  3  =  Mixed:  4  =  Latr/Amdu:  5  =  Washi. 


188 


Simpson,  E.  H.  1949.  Measurement  of 
diversity.  Nature  163:466-467. 

Sokol,  O.  M.  1971.  Lithophagy  and 
geophagy  in  reptiles.  Journal  of 
Herpetology  5:69-71. 

Speake,  Daniel  W.  1986.  Gopher  tor- 
toise density  in  various  south  Ala- 
bama habitats.  Alabama  Coopera- 
tive Fish  and  Wildlife  Research 
Unit,  Research  Information  Bulle- 
tin 86-105, 1  p.  Auburn,  Alabama. 

Stewart,  Glenn  R.,  and  Ronald  J. 
Baxter.  1987.  Final  report  and 
habitat  management  plan  for  the 
desert  tortoise  (Gopherus  agassizii) 
in  the  West  and  Sand  Hill  training 
areas  of  the  Twentynine  Palms 
MCAGCC.  Report  prepared  for 
the  U.S.  Department  of  the  Navy, 
San  Bruno,  Calif.  Contract  number 
N6247484RPOOV48. 

Turner,  Frederick  B.,  and  Kristin  H. 
Berry.  1986.  Population  ecology  of 
the  desert  tortoise  at  Goffs,  CaH- 
fornia,  in  1985.  University  of  Cali- 
fornia, Los  Angeles  publication 
number  12-1544. 

Turner,  Frederick  B.,  and  Carl 
Thelander,  Daniel  Pearson,  and 
Betty  Burge.  1982.  An  evaluation 
of  the  transect  technique  for  esti- 
mating desert  tortoise  density  at  a 
prospective  power  plant  site  in 
Ivanpah  Valley,  California.  Pro- 
ceedings of  the  symposium,  [Las 
Vegas,  Nev.,  March,  1982].  The 
Desert  Tortoise  Council,  Long 
Beach,  CA. 

Walchuck,  Sandra  L.,  and  James  C. 
devos,  Jr.  1982.  An  inventory  of 
desert  tortoise  populations  near 
Tucson,  Arizona.  Proceedings  of 
the  symposium,  [Las  Vegas,  Nev., 
March,  1982].  The  Desert  Tortoise 
Council,  Long  Beach,  CA. 

Weinstein,  Michael  and  Frederick  B. 
Turner  and  Kristin  H.  Berry.  1986. 
An  analysis  of  habitat  relation- 
ships of  the  desert  tortoise  in  Cali- 
fornia. Draft  report  prepared  for 
Southern  California  Edison  Com- 
pany, Los  Angeles,  Calif. 

Zar,  Jerrold  H.  Biostatistical  analysis. 
1974.  Prentice-Hall,  Inc.,  Engle- 
wood  Cliffs,  NJ. 


Changes  in  a  Desert  Tortoise 
(Gopherus  agassizii) 
Population  After  a  Period  of 
High  IVIortality^ 

David  J.  Germano^  and  Michele  A.  Joyner^ 


Abstract.— An  apparent  high  rate  of  mortality  for 
desert  tortoises  at  the  Piute  Valley  in  southern 
Nevada  between  1979  and  1983  significantly 
decreased  mean  carapace  length  and  average 
age  of  the  population  by  1983.  but  not  density.  By 
1987,  average  size  and  age  of  the  population  had 
increased  and  density  remained  stable. 


Chelonians,  as  a  group,  are  charac- 
terized by  high  rates  of  adult  sur- 
vival, delayed  maturity,  and  low 
rates  of  juvenile  survival  (Wilbur  and 
Morin  1988).  Many  chelonians  live  a 
long  time  after  reaching  adulthood 
(Gibbons  1987),  potentially  leading  to 
a  long  period  of  reproduction  offset- 
ting low  juvenile  survival  (Wilbur 
and  Morin  1988).  The  desert  tortoise 
(Gopherus  agassizii)  (fig.  1)  is  an  her- 
bivorous chelonian  of  the  desert 
Southwest  that  exhibits  these  popula- 
tion traits  (Berry  1986,  Luckenbach 
1982,  Osorio  and  Bury  1982,  Turner 
et  al.  1984, 1986).  In  1983,  a  large 
number  of  desert  tortoise  skeletons 
were  collected  from  a  study  plot  lo- 
cated in  southern  Nevada  and  deaths 
were  believed  to  have  occurred  since 
the  initial  census  in  1979  (unpub- 
lished report,  C.  Mortimore  and  P. 
Schneider,  Nevada  Department  of 
Wildlife,  Las  Vegas,  NV).  It  was  re- 
ported that  since  1979,  mean  cara- 
pace length  of  the  population  de- 
creased, sex  ratio  had  become  male 
biased,  and  that  population  density 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff. AZ,  July  19-21.  1988). 

'David  J.  Germane  is  a  doctoral  candi- 
date. Museum  of  Southwestern  Biology. 
Department  of  Biology.  University  of  New 
Mexjco,  Albuquerque  87131 

^Michele  A.  Joyner  is  an  undergraduate. 
Museum  of  Southwestern  Biology.  Depart- 
ment of  Biology.  University  of  New  Mexico. 
Albuquerque  87131 


decreased,  and  that  these  changes  oc- 
curred because  long-term  grazing  of 
this  plot  by  cattle  weakened  tortoises 
to  such  a  degree  that  decreased  for- 
age production  resulting  from  below- 
average  rainfall  in  1981  killed  many 
individuals  (unpublished  report,  C. 
Mortimore  and  P.  Schneider,  Nevada 
Department  of  Wildlife,  Las  Vegas). 

We  recensused  this  population  in 
1987  in  order  to  determine  changes 
that  might  have  taken  place  since 
1983  in  age  distribution,  size  distri- 
bution, sex  ratios,  and  population 
density  in  order  to  address  the  fol- 
lowing questions:  Of  what  signifi- 
cance are  such  periods  of  high  mor- 


tality to  the  p)opulations'  probability 
of  survival?  How  do  desert  tortoise 
populations  respond  to  high  rates  of 
mortality?  Are  changes  in  population 
demographics  long-lasting?  Can  we 
predict  future  changes  in  desert  tor- 
toise populations?  We  also  reassess 
possible  causes  of  the  high  rate  of 
mortality  between  1979  and  1983. 

METHODS 

Study  Area 

The  2.59  km^  plot  is  located  in  the 
Piute  Valley  of  southern  Nevada  in 


190 


Figure  3.— Creosote  bush  and  white  bursage  are  the  most  conspicuous  plants  of  much  of  the 
study  plot  (top)  with  Mojave  yucca  abundant  In  the  northwestern  portion  (bottom).  Other 
abundant  plants  at  this  site  are  California  buckwheat  (Eriogonum  fasiculatum),  rayless 
goldenhead  (Acamtopappus  sphaerocephalus),  Opuntia  spp,,  bush  muhly  (Muhlenbergia 
porteri),  gig  galleta  (Hilaria  riglda),  six-week  fescue  (Festuca  octoflora),  filaree  (Erodium 
cicutarium),  desert  dandelion  (Malacothrix  glabrata),  and  Chaenactis  spp. 


the  eastern  Mojave  desert  (fig.  2). 
Vegetation  is  Mojave  desert  scrub 
dominated  by  creosote  bush  (Larrea 
tridentata)  and  white  bursage  (Ambro- 
sia dumosa)  over  the  southeastern  2/3 
of  the  plot  grading  into  an  area  with 
an  overstory  of  Mojave  desert  yucca 
(Yucca  schidigera)  in  the  northwestern 
third  (fig.  3). 

Field  Methods 

The  population  was  censused  be- 
tween April  and  June  1979  by  the  Bu- 
reau of  Land  Management  (unpub- 
lished report,  A.  Karl,  BLM,  Las  Ve- 
gas, NV)  and  again  between  April 
and  June  1983  by  the  Nevada  Depart- 
ment of  Wildlife  (unpublished  re- 
port, C.  Mortimore  and  P.  Schneider, 
Nevada  Department  of  Wildlife,  Las 
Vegas,  NV).  Each  tortoise  encoun- 
tered was  measured,  weighed, 
marked,  its  sexed  determined,  and 
its  location,  behavior  and  general 


Figure  2.— The  location  of  the  desert  tortoise 
permanent  study  plot  (PSP)  in  the  Piute  Val- 
ley of  southern  Nevada.  The  dashed  and 
dotted  lines  show  major  washes. 


condition  noted.  Shells  were  col- 
lected and  are  catalogued  in  the  Mu- 
seum of  Southwestern  Biology,  Uni- 
versity of  New  Mexico,  Albuquer- 
que. 


We  recensused  the  plot  13-27  May 
and  18-25  August  1987.  We  collected 
similar  data  on  tortoises,  but  in- 
cluded making  casts  of  the  second 
costal  scute  using  dental  casting  ma- 


191 


terial  (Galbraith  and  Brooks  1987). 
Measurements  of  growth  rings  from 
the  impressions  on  the  casts  were 
taken. 

Growth  rings  of  desert  tortoises 
have  been  found  to  be  valuable  for 
determining  age  and  growth  histo- 
ries of  many  individuals  (Germano 
1988).  Shells  were  collected  and  de- 
posited in  the  Museum  of  Southwest- 
ern Biology. 

Data  Analysis 


Density 

Densities  in  1979  and  1983  were  de- 
termined by  the  investigators  who 
conducted  the  censuses  using  the 
Schnabel  estimator.  This  method  in- 
volves making  periodic  estimates  of 
density  during  the  census  based  on 
the  number  of  marked  and  un- 
marked animals  found  (Tanner  1978). 
Because  of  immigration  into  the  plot, 
we  reestimated  density  for  1983  us- 
ing the  Jolly-Seber  estimator  (Tanner 
1978),  which  does  not  assume  a 
closed  population. 

As  a  first  approximation  of  den- 
sity for  1987,  we  used  a  simple  mark- 
recapture  estimator  with  May  as  the 
period  of  marking  animals  and  Au- 
gust as  the  recapture  period.  Only  1  / 
2  the  plot  was  recensused  in  August 
because  of  time  constraints.  Density 
was  computed  for  this  half  of  the 
plot. 

Carapace  Lengtti  Distributions 

Carapace  lengths  (CD  of  individuals 
were  plotted  and  mean  CLs  com- 
puted for  live  tortoises  and  remains 
for  each  census  year.  Mean  CLs  of 
the  total  population,  tortoises  >180 
mm  CL,  and  tortoises  <180  mm  CL 
were  compared  among  years  using 
anova  with  comparisons  among 
means  using  Scheffe's  multiple  com- 
parisons test. 


Age  Distributions 

Ages  of  individuals  were  plotted  for 
live  tortoises  and  remains  and  mean 
ages  compared  in  a  manner  similar 
to  CLs.  Ages  of  skeletons  and  1987 
live  tortoises  were  determined  for 
most  individuals  using  scute  annuli, 
a  technique  that  is  accurate  up  to  20- 
25  years  (Germano  1988).  Several  in- 
dividuals were  considered  to  be 
older  than  the  number  of  easily  seen 
annuli  based  on  non-growth  since 
last  capture,  or  scute  edge  beveling, 
which  indicates  continued  slow 
growth.  These  individuals  were  cate- 
gorized as  >25  years  old. 

Ages  were  estimated  for  live  tor- 
toises found  in  1979  and  1983  using 
an  age-CL  regression  (Age  =  0.106 
CL  -  3.82).  The  number  of  scute  an- 
nuli is  well  correlated  with  CL  (r^  = 
0.908,  n  =  150),  although  the  relation- 
ship is  less  accurate  in  larger  indi- 
viduals. We  corrected  for  the  pres- 
ence of  older  individuals  in  our  esti- 
mates by  assigning  a  portion  of 
adults  of  various  sizes  to  the  >25  age 
category  based  on  the  percentage  of 
adults  that  were  into  this  category 
from  the  1987  live  and  1983  and  1987 
shell  groups. 

Mortality  Rates 

Age-specific  mortality  rates  were  de- 
termined for  1979-1983  and  1983- 
1987  using  the  equation  q^  =  (k  [f  J)/ 
g^,  where  q^  is  the  mortality  rate  per 
year  for  age  x,  k  is  the  per  capita 
mortality  rate  of  the  population,  f^  is 
the  proportion  of  animals  age  x  that 
are  known  to  have  died  in  the  past 
year,  and  g^  is  the  proportion  of  ani- 
mals of  age  X  in  the  preceding  live 
population  (Fryxell  1986).  In  order  to 
compare  mortality  rates  to  age  distri- 
butions, we  determined  mortality 
rates  for  age  groups  0-14  years,  15-27 
years,  and  >25  years.  The  per  capita 
mortality  rate  was  divided  by  4  to 
obtain  the  yearly  mortality  rate  for 
each  time  period. 


Sex  Ratios 

Sex  ratios  were  compared  among  live 
tortoises  and  shells.  Sex  was  assigned 
to  tortoises  >180  mm  CL  based  on 
secondary  sex  characteristics  or,  in 
some  instances,  for  males  >170  mm 
CL  when  plastron  concavity  was  ob- 
vious. Sex  can  be  determined  reliably 
in  desert  tortoises  based  on  shell 
characters  after  180  mm  CL  (unpub- 
lished report,  F.  Turner  and  K.  Berry, 
Southern  California  Edison  Co.,  CA) 
and  female  tortoises  in  this  part  of 
the  Mojave  desert  reproduce  at  189 
mm  CL  (Turner  et  al.  1986),  indicat- 
ing that  sexual  maturity  probably  oc- 
curs between  180-190  mm  CL.  Ratios 
were  tested  for  deviation  from  a  1:1 
sex  ratio  with  Chi-square  analysis  (p 
<  0.05). 


CL/Weigtit  Regressions 

Carapace  length  to  weight  regres- 
sions were  constructed  for  1979  and 
1987  tortoises  based  on  the  logarith- 
mic transformation  of  both  variables. 
Data  for  1983  were  not  available. 
Slopes  were  tested  against  0  and 
against  each  other  using  f-tests  (Sokal 
and  Rohlf  1981). 


Growtti  Rate  Comparisons 

Individual  growth  was  compared 
among  1987  live  tortoises  and  shell 
groups  in  two  ways.  Growth  rings 
were  compared  among  groups  using 
mean  annual  widths  (AW)  and  mean 
percent  growth  for  rings  1-24  (See 
Germano  1988  for  a  description  of 
growth  ring  measurements).  Percent 
growth  for  a  ring  is  AW/ estimated 
CL  for  the  preceding  year.  CLs  were 
estimated  using  the  length  of  growth 
rings  from  the  second  costal  scute, 
which  are  highly  correlated  to  CL  (r^ 
=  0.96,  n  =  174).  Growth  estimates 
based  on  annuli  have  been  found  to 
accurately  reflect  carapace  growth  in 
gopher  tortoises  (Landers  et  al.  1982) 
and  desert  tortoises  (Germano  In 


192 


Press).  Means  of  these  variables  for 
each  ring  were  compared  among 
groups  using  the  nonparametric 


Wilcoxon  sign  test.  We  also  com- 
pared the  mean  AW  and  mean  per- 
cent growth  of  the  last  two  growth 
rings  for  the  shells  found  in  1983  to 
the  mean  AW  and  mean  percent 
growth  of  the  1980  and  1981  growth 
rings  from  live  tortoises  found  in 
1987  using  f-tests. 


'rill  i'ri"i"l"rT"("i'T'i"r"i"ri"i'TiTi  i  i  i  i  i — 
I   t   s   7   I   II  II  B  17  II  n  u  tt  n  n  >za 


AGE  (yeors) 


1   I   7   *   II  n  II  IT  If  ai  M  tt  IT  II  >ii 

AGE   (y  e  a  r  t ) 


Figure  4.— Population  size  distributions  for 
iive  desert  tortoises  from  \he  Piute  Valley 
permanent  study  plot.  Mean  carapace 
lengths  and  somple  sizes  are  given  in  table 
1. 


Figure  5.— Population  size  distrit>utions  for 
desert  tortoises  found  dead  in  1983  and 
1987  from  the  Piute  Valley  permanent  study 
plot.  Mean  carapace  lengths  and  sample 
sizes  are  given  in  table  1 . 


Table  1  .—Mean  carapace  lengths  (mm)  of  tortoises  from  the  Piute  Valley 
permanent  study  plot.  Standard  deviation  and  sample  size  are  given  be- 
low \he  mean. 


All 

%  of 

%of 

Group 

tortoises 

>180  mm  CL 

total 

<180  mm  CL 

total 

1979  live 

186.8 

217.1 

58 

144.5 

42 

(44.0,  84) 

(21.0,49) 

(30.8,  35) 

1983  live 

148.2 

211.8 

37 

110.8 

63 

(59.6,81) 

(24.9,  30) 

(38.3,51) 

1987  live 

181.1 

213,8 

60 

125.8 

40 

(46.6,  48) 

(20.0, 29) 

(37.8,  19) 

1983  shells 

197.6 

212.9 

78 

106.4 

22 

(93.3,  108) 

(22.6,  84) 

(39.0,  24) 

1987  shells 

165.4 

216.3 

49 

117.2 

51 

(58.1,37) 

(19.4,  18) 

(36.9,  19) 

Climate  Analysis 

Climate  was  analyzed  using  weather 
information  from  Searchlight,  Ne- 
vada. Data  were  compared  for  3  time 
periods;  1970-June  1979,  July  1979- 
1982,  and  July  1979-July  1987.  Means 
and  variances  of  rainfall,  both  annual 
and  winter,  were  compared  among 
time  periods.  Mean  monthly  tem- 
peratures were  compared  among 
time  periods  and  temperatures  below 
freezing  were  analyzed  for  duration 
and  relation  to  unusually  warm  win- 
ter daily  highs. 

RESULTS 
Density 

Tortoise  density  was  estimated  to  be 
50/km2  in  1979  and  72/km2  in  1983 
by  the  authors  of  these  censuses. 
Eighty-four  and  81  tortoises  were 
found  in  1979  and  1983,  respectively. 
We  reestimated  the  1983  density  to 
be  44  tortoises/km^.  We  estimated 
the  density  in  1987  to  be  59  tortoises/ 
km^  (95%  confidence  intervals,  19- 
173).  We  found  48  tortoises  in  1987, 
33  in  May  and  19  on  the  southern 
half  of  the  plot  in  August,  of  which  4 
had  been  marked  in  May. 

Carapace  Length  Distributions 

Distributions  of  CLs  of  live  tortoise 
populations  varied  significantly  for 
each  census  (fig.  4).  Mean  CL  was 
significantly  smaller  in  1983  than  in 
either  1979  (p<.05)  or  1987  (p<.05). 
Mean  CLs  in  1979  and  1987  were  not 
significantly  different,  however 
(p>.05,  table  1).  No  significant  differ- 
ences were  found  among  mean  CLs 
for  adults  (>180  mm  CL).  Adults 
comprised  58%  of  the  1979  popula- 
tion, 37%  of  the  1983  population,  and 
60%  of  the  1987  population.  The  ' 
mean  CL  of  non-adults  (<180  mm 
CL)  was  significantly  smaller  in  1983 
than  1979  (p<.05),  but  was  not  sig- 
nificantly different  than  1987  (p>.05. 


193 


table  1).  The  mean  CL  of  non-adults 
was  not  significantly  different  be- 
tween 1979  and  1987  (p>.05). 

Remains  of  37  tortoises  were 
found  in  1987  compared  to  109  found 
in  1983  (fig.  5).  Ten  shells  were  found 
in  1979.  CLs  of  remains  were  not  sig- 
nificantly different  (p>.05),  although 
mean  CL  in  1983  was  considerably 
larger  than  for  1987  (table  1).  Mean 
CLs  of  adult  remains  in  1983  and 
1987  were  similar,  as  were  non-adult 
CLs,  but  adults  comprised  78%  of  the 
1983  collection  and  only  49%  of  the 
1987  collection.  The  mean  CL  of  re- 
mains from  1983  was  not  signifi- 
cantly different  from  the  mean  CL  of 
live  tortoises  in  1979  or  1987,  but  was 
significantly  larger  than  live  tortoises 
in  1983  (p<.05).  Mean  CL  of  remains 
from  1987  was  not  significantly  dif- 
ferent than  any  live  tortoise  means. 


Age  Distributions 

Ages  of  tortoises  varied  significantly 
among  years  (table  2).  Changes  in 
age  distributions  of  live  tortoises 
were  similar  to  the  changes  seen  for 
CLs  (fig.  6).  The  estimated  mean  age 
for  1979  was  significantly  older  than 
1983  (p<.05)  but  not  1987  (p>.05). 
Mean  age  for  1987  was  not  signifi- 
cantly different  than  1983  (p>.05),  but 
non-adults  were  significantly  older 
(p<.05).  Mean  age  of  1983  remains 
was  significantly  older  than  1983  live 
tortoises  (p<.05),  but  was  not  signifi- 
cantly different  than  1987  live  tor- 
toises or  remains  (p>.05,  fig.  7). 


Mortality  Rotes 

Death  rates  for  1983-1987  were  lower 
than  for  1979-1983.  Per  capita  mortal- 
ity rate  (k)  for  1979-1983  was  0.21/ 
year  (N  =  130)  and  was  0.08/year  for 
1983-1987  (N  =  115).  Mortality  rates 
dropped  for  all  age  classes  after  1983. 
For  1979-1983  mortality  rates  were 
0.145/year  for  0-14  year  olds,  0.247/ 
year  for  15-25  year  olds,  and  0.195/ 
year  for  tortoises  >25  years.  For  1983- 


1987  mortality  rates  were  0.061 /year 
for  0-14  year  olds,  0.093/year  for  15- 
25  year  olds,  and  0.103  for  tortoises 
>25  years.  Mortality  rates  for  all 
adults  (15-25  years  and  >25  years)  for 
1979-1983  was  0.240/year  and  for 
1983-1987  was  0.103/year. 


Sex  Ratios 

Sex  ratios  of  live  tortoises  show  an 
increasing  proportion  of  males  (table 
3),  although  only  1987  showed  a  sig- 
nificantly biased  sex  ratio.  When  the 
1987  sex  ratio  was  analyzed  by  size, 
92%  of  tortoises  >220  mm  CL  were 
males,  whereas  only  53%  of  tortoises 
180-219  mm  CL  were  males  (table  3). 
When  analyzed  by  age,  63%  of 
tortoises  >20  years  were  males,  but 
71%  of  tortoises  of  known  sex  be- 
tween 13-19  years  were  males,  a  sig- 
nificantly higher  proportion  than  fe- 
males. The  sex  ratios  of  dead  tor- 
toises were  not  significantly  different 
than  1:1  (table  3). 


CL/Weigtit  Regressions 

The  regressions  of  weight  against  CL 
had  significant  slopes  for  1979  and 


CARAPACE       LENGTH  (mm) 

Figure  6.— Population  age  distributions  for 
live  desert  tortoises  from  the  Piute  Valley 
perrrKinent  study  plot.  The  1979  and  1983 
age  distributions  are  estimates  based  on  a 
carapace  length  to  annulus  number  re- 
gression. A  proportion  of  adults  were 
placed  in  the  >25  age  category  based  on 
the  proportion  of  adults  In  this  category 
from  the  age  distributions  for  which  ages 
were  assigned  by  annuli  counts.  The  1987 
age  distribution  Is  based  on  annuli  counts. 


r 


Table  2.--Mean  ages  of  tortoises  from  the  Piute  Valley  permanent  study 
plot  in  southern  Nevada.  Standard  deviation  and  sample  size  are  given 
below  the  mean.  Ages  for  1 979  and  1 983  are  estimates  based  on  cara- 
pace length  (see  Methods). 

Ages  (years) 


Group 

0-27 

0-14 

15-27 

^>25 

1979  live 

16.6 

10.9 

19.5 

(5.1,72) 

(3.4, 24) 

(2.9,  48) 

(12) 

1983  live 

12.1 

7.5 

18.8 

(6.6.74) 

(3.7,41) 

(3.0, 30) 

(7) 

1987  live 

14.1 

11.3 

17.0 

(3.8, 43) 

(3.2, 22) 

(2.2,21) 

(5) 

1983  shells 

17.0 

7.8 

19.9 

(6.2, 94) 

(3.6, 22) 

(3.3,  72) 

(14) 

1987  sl-»ells 

14.0 

8.4 

19.3 

(62.,  31) 

(2.6, 15) 

(3.5,  16) 

(6) 

'Mean  age  cannot  be  determined. 


194 


1987  (fig.  8).  The  regression  equation 
for  1979  is  gram  weight  =  0.000317 
CL2.924     ^  0.952,  n  =  73)  and  for 
1987  is  gram  weight  =  0.000505 
(2L2.826  (jj  ^  0.969,  n  =  53).  Regression 
slopes  were  not  significantly  differ- 
ent from  each  other  (p>.10). 


Growth  Rate  Comparisons 

No  significant  differences  were 
found  in  a  ring  by  ring  comparison  of 
growth  between  1987  live  tortoises 
and  1983  remains  for  either  annual 
widths  (AW)  or  percent  growth. 


When  1980  and  1981  rings  were  com- 
pared, no  significant  difference  ex- 
isted between  the  mean  AW  for  the 
last  two  rings  of  1983  mortalities  (X  = 
1.98mm,  n  =  72)  and  the  1980  and 
1981  rings  for  1987  live  tortoises  (X  = 
1.92nmi,  n  =  79;  p>.10). 


CARAPAC  E       LENGTH  (mm) 


Figure  7.— Population  age  distributions  for 
desert  tortoises  found  dead  in  1 983  and 
1987  from  the  Piute  Valley  pernnanent  study 
plot.  Both  the  1983  and  the  1987  age  distri- 
butions are  based  on  counts  of  annuli. 


9 


4.0- 
l.t- 
1.0- 
IJ 

t.o- 
1,1 

1.0 


I  •  T  1 


I    I  \ — r— I — l—l — I — I — r— I — I   I  I — r— I — I — I — I — r 


M 

1.0 

u 

l.« 
I.I- 

1.0- 

•.I 


HIT 


I  T  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I — I — r 
to      M     no     HO     NO      NO     WO     tot    tW    140  tM 

CARAPACE       LENGTH  (mn) 


Figure  8.— Regressions  of  carapace  length 
to  weight  for  desert  tortoises  found  in  1979 
and  1987.  Slopes  of  both  regressions  are 
significantly  different  from  0  but  not  from 
each  other. 


Table  3.— Numbers  of  males  to  females  for  desert  tortoises  from  the  Piute 
Valley  permanent  study  plot.  Significant  departures  from  a  1 :1  sex  ratio 
were  determined  by  Chl-square  analysis.  The  1987  live  totals  were  sub- 
categorized  by  size  and  age. 


Year 

Males 

Females 

Ratio 

1979 

live 

24 

30 

0.88 

1 

0.667 

shells 

4 

3 

1.33 

1 

0.001 

1983 

live 

22 

11 

2 

1 

3.667 

shells 

35 

41 

0.85 

1 

0.474 

1987 

live  (total) 

20 

9 

2.22 

1 

M.172 

size:  180-219  mm  CL 

9 

8 

1.13 

1 

0.059 

>220mm  CL 

11 

1 

11 

1 

^8.330 

age:  13-19  years 

15 

6 

2.5 

1 

^3.857 

>20  years 

5 

3 

1.67 

1 

0.500 

shells 

11 

6 

1.83:1 

1.471 

'Significanf  departure  from  1:1  ratio  (p<.05). 


Climate  Analysis 

Average  precipitation  were  higher 
between  July  1979  and  July  1987  than 
the  previous  10  years  (table  4).  The 
highest  average  precipitation  was 
recorded  between  July  1979  and  De- 
cember 1982.  Winter  rainfall  (Octo- 
ber-March) followed  the  same  pat- 
tern, with  both  1979-1987  and  1979- 
1982  averages  higher  than  1970-1979 
(table  4).  The  period  1970-1979  was  a 
drought  period  with  average  rainfall 
7%  below  the  long-term  average  of 
183.8  mm  and  7  of  the  10  years  were 
well  below  average  (table  4).  When 
1978  and  1979  are  excluded,  average 
precipitation  drops  to  129.3  mm,  30% 
below  the  long-term  average.  July 
1979-December  1982  averaged  40% 
higher  rainfall  than  the  long-term  av- 
erage with  only  1981  experiencing 
below-average  rainfall.  Mean 
monthly  high  and  low  temperatures 
were  similar  among  time  periods.  No 
extended  periods  of  freezing  tem- 
peratures were  found  for  daily  read- 
ings between  1979  and  1983. 


DISCUSSION 

Population  Parameters 

The  desert  tortoise  p>opulation  in  the 
Piute  Valley  study  plot  experienced  a 
high  rate  of  mortality,  particularly  of 
adults,  between  July  1979  and  1983. 
Related  to  this  event  was  a  signifi- 
cant decrease  in  the  size  and  age  dis- 
tributions of  the  population  in  1983, 
although  both  were  returning  to  1979 
dimensions  by  1987.  The  lower  mean 
age  in  1983  is  probably  a  result  of  in- 
creased survival  of  hatchlings  and 
increased  immigration.  The  increased 


195 


survival  of  hatchlings,  as  shown  by 
the  significant  increase  of  tortoises  in 
the  1-4  age  group  in  1983,  may  be 
due  to  more  favorable  conditions  be- 
cause of  lower  densities  just  after  the 
high  rate  of  mortality,  or  to  optimal 
climatic  and  habitat  conditions. 

It  is  possible  that  the  greater  num- 
bers of  smaller  tortoises  found  in 
1983  could  have  resulted  from  better 
search  effort  for  these  sizes  (Berry 
and  Turner  1984),  but  we  censused 
the  plot  carefully  in  1987,  specifically 
looking  for  small  tortoises,  yet  we 
found  relatively  few.  While  we  do 
not  doubt  that  young  are  missed  be- 
cause of  their  inconspicuousness,  we 
believe  that  the  changes  in  size  and 
age  distributions  between  1979  and 
1987  reflect  actual  population 
changes. 

The  size  and  estimated  age  distri- 
butions for  1983  indicate  that  a  sig- 
nificant number  of  smaller  and 
younger  tortoises  came  into  the  plot 
between  1979  and  1983.  Judging  by 
the  male-dominated  sex  ratio  after 
1979,  immigration  largely  has  been 
by  young  males.  The  biased  sex  ra- 
tios are  not  due  to  higher  adult  male 
survival  since  equal  proportions  of 
males  and  females  died.  Most  of  the 
males  in  the  present  population  are 
fairly  young,  although  they  are  large. 
Male  turtles  are  known  to  disperse 
greater  distances  than  females  (Gib- 
bons 1986). 

Although  many  turtle  populations 
have  biased  sex  ratios,  evolutionary 
theory  indicates  that  these  ratios 
should  be  under  selective  pressure  to 
be  relatively  even,  in  most  instances 
(Fisher  1930,  Trivers  1972).  However, 
desert  tortoise  age  to  maturity  is  ca. 
15  years  (Germano  In  Press,  Woo- 
dbury and  Hardy  1948),  therefore  a 
reproductive  solution  mediated  by 
selection  would  require  hundreds  of 
years. 

Censuses  in  other  parts  of  this  val- 
ley in  1983  indicate  that  this  high  rate 
of  mortality  was  confined  to  this  plot 
and  areas  close  by  (unpublished  re- 
port, C.  Mortimore  and  P.  Schneider, 
Nevada  Department  of  Wildlife,  Las 


Vegas,  NV).  Differences  in  sex  ratios 
at  this  plot  may  be  more  a  reflection 
of  higher  male  movement  rates  com- 
pared to  females  and  not  to  a  real 
difference  in  numbers  of  males  and 
females  in  the  population  as  a  whole. 
Over  time  the  sex  ratios  may  change 
by  movement  of  females  into  the  plot 
from  outside. 


Density  may  have  decreased 
slightly  since  1979,  but  it  does  not  ap- 
pear to  have  changed  significantly 
over  the  8  year  period,  although  we 
recognize  the  imprecision  of  these 
density  estimates.  The  number  of  tor- 
toises found  has  decreased  in  each 
census,  but  investigators  and  time 
periods  in  the  field  have  varied,  ren- 


Table  4.— Annual  and  winter  precipitation  (mm)  for  1970-1987  and  for  3 
time  periods  from  the  Searchlight,  Nevada  NOAA  Station.  Winter  precipita- 
tion is  defined  by  the  months  October-March.  Means  and  standard  devia- 
tions are  given  for  the  3  time  periods.  Precipitation  for  1987  only  Includes 
the  months  of  January-July. 


Year 


1970 
1971 
1972 
1973 
1974 
1975 
1976 
1977 
1978 
1979 
1980 
1981 
1982 
1983 
1984 
1985 
1986 
1987 


Time  period 


Annual  Winter 
total  total 


Jan.  1970-  July  1979-  July  1979 
June  1 979      Dec.  1 982     July  1 987 


127.76 
68.83 
136,65 
114.81 
184.40 
132.08 
161.80 
107.70 
473.71 
256.54 
313.44 
162.81 
366.10 
376.68 
300.48 
149.35 
166.88 
73.66 


30,73 

17.02 
179.02 

54,36 
100,08 

82.79 

52.58 
183.90 
249.43 
260.10 

67.06 
101.09 
216.15 

61.47 
191.52 

91.69 
126.24 


Annual  precipitation 


170.9 
(113.0) 


281.2 
(86.4) 


265.5 
(148.8) 


Winter  precipitation 


104.4 
(33.0) 


139.4 
(73.7) 


161.0 
(71.7) 


196 


dering  this  comparison  unreliable. 
We  believe  that  the  lower  number  of 
live  tortoises  found  in  1987  is  due  to 
inexperienced  field  personnel  and 
the  shorter  duration  of  time  in  the 
field.  The  most  valid  of  these  density 
estimates  is  the  Jolly-Seber  estimate 
of  44  tortoises /km^,  because  more 
assumptions  are  met  with  this  tech- 
nique. Unfortunately,  estimates  can- 
not be  made  for  the  first  or  last  cen- 
sus with  this  technique.  Density  esti- 
mates, though,  are  similar  in  magni- 
tude and  we  believe  this  indicates 
that  density  has  remained  relatively 
stable  since  1979.  The  population 
must  have  experienced  a  decline  af- 
ter 1979  but  we  believe  that  increased 
survival  of  young  and  immigration 
from  adjacent  non-affected  areas  has 
quickly  returned  the  density  to  1979 
levels. 


Mortality  Factors 

Causes  of  the  high  rate  of  mortality 
have  not  been  demonstrated.  The 
hypothesis  that  long-term  grazing 
confounded  by  a  drought  in  1981 
was  the  cause  of  the  high  number  of 
tortoise  deaths  is  not  supported  by 
growth  analysis  of  annuli,  CL/ 
weight  data,  or  climate  data.  Growth 
did  not  differ  significantly  between 
those  that  died  before  1983  and  those 
that  survived  to  1987.  In  addition, 
the  weight  to  size  regressions  for 
1979  and  1987  were  the  same  and 
both  were  almost  identical  to  the  re- 
gression for  tortoises  from  an  un- 
grazed  plot  in  Nevada  (Medica  et  al. 
1975).  As  for  a  drought  in  1981,  aver- 
age rainfall  was  only  9%  below  the 
long-term  average  (up  to  1987)  and 
was  actually  at  the  average,  up  to 
1981,  given  the  drought  in  the  1970s. 
Preceding  1981  were  3  years  of  ex- 
ceptionally high  rainfall.  In  contrast, 
rainfall  in  1977  was  41%  below  aver- 
age and  followed  many  drought 
years  (table  4). 

Desert  tortoises  are  known  to 
store  water  (Nagy  and  Medica  1986) 
and  may  be  able  to  store  fat.  It  seems 


doubtful  that  one  average  year  of 
rainfall  after  3  very  good  years  could 
cause  starvation  or  lethal  dehydra- 
tion. The  2  years  preceding  our  cen- 
sus in  1987  were  below  average  in 
precipitation,  yet  mortality  rates 
dropped.  The  period  1970-1977  was  a 
drought,  yet  only  10  shells  were 
found  in  1979.  If  these  low  rainfall 
years  didn't  produce  a  high  rate  of 
mortality  that  could  be  detected  in 
1979,  it  is  hard  to  imagine  that  one 
average  year  after  3  good  years 
would  result  in  excess  mortality.  Es- 
timates of  yearly  adult  death  rates 
from  1972-1982  for  a  population  only 
42  km  south  of  this  site  was  1.2%,  in 
an  area  that  has  been  grazed  by  live- 
stock for  100  years  (Berry  and 
Nicholson  1984a). 

Other  possible  causes  for  this  mor- 
tality could  have  been  disease,  pre- 
dation,  or  flooding.  Diseases  are 
known  to  affect  other  turtle  species 
in  the  wild  (Jacobson  1980a,b),  but  no 
evidence  exists  for  disease  as  a  fac- 
tor. Many  of  the  shells  show  signs  of 
chewing  by  carnivores,  although 
whether  this  indicates  predation  or 
scavenging  cannot  be  determined. 
Flooding  occurred  in  or  near  the  plot 
in  1980  and  1982  (unpublished  re- 
port, J.  Jamrog  and  R.  Stager,  BLM, 
Las  Vegas,  NV).  The  plot  is  dissected 
by  numerous  washes  that  are  most 
prevalent  in  this  part  of  the  valley 
(fig.  2). 

The  exact  cause  of  the  high  rate  of 
mortality  may  never  be  known.  Star- 
vation, disease,  flooding,  and  preda- 
tion may  have  all  had  an  effect.  No 
singular  explanation  is  supported  by 
the  data.  Whatever  the  causative 
agent,  the  population  appears  to  be 
returning  to  a  density  and  popula- 
tion structure  as  occurred  before  the 
f)eriod  of  high  mortality. 

Management  Implications 

As  a  long-lived  reptile,  the  desert  tor- 
toise is  more  vulnerable  to  fluctua- 
tions in  adult  mortality  than  to  simi- 
lar fluctuations  in  younger  age 


groups.  Many  desert  tortoise  popula- 
tions consist  of  adult  segments  that 
usually  have  yearly  survivorship 
rates  of  95-98%  (Berry  and  Nicholson 
1984b).  High  adult  survivorship  is 
often  coupled  with  low  juvenile  sur- 
vivorship (Wilbur  and  Morin  1988) 
and  part  of  the  concern  for  tortoise 
populations  is  that  they  may  not 
have  the  ability  to  withstand  distur- 
bance because  of  low  juvenile  survi- 
vorship. Female  desert  tortoises  in 
the  eastern  Mojave  desert  have  the 
ability  to  lay  2-3  clutches  in  a  season 
(Turner  at  al.  1986).  The  significant 
increase  in  1983  of  tortoises  1-4  yr  of 
age  suggests  more  hatchlings  have 
survived  between  1979-1983  than 
previously.  As  with  any  other  popu- 
lation parameter,  juvenile  survivor- 
ship can  vary,  and  this  may  lead  to 
periodic  additions  of  greater  num- 
bers of  young  surviving  to  adult  age. 

It  appears  that  desert  tortoises 
have  the  ability  to  recover  from  dis- 
turbance in  some  instances.  This  ap- 
pears to  be  what  is  happening  at  the 
Piute  plot.  Increased  juvenile  survi- 
vorship and  immigration  are  holding 
the  population  density  stable  and  the 
age  and  size  distributions  are  return- 
ing to  1979  dimensions.  This  kind  of 
recovery  may  not  occur  if  a  distur- 
bance is  prolonged  or  is  widespread. 
Those  managing  desert  tortoises 
must  be  aware  of  the  dynamics  of 
each  population,  but  it  is  apparent 
that  tortoise  populations  can  recover 
from  short-term  high  mortality. 

ACKNOWLEDGMENTS 

We  thank  T.  Fritts  and  the  National 
Ecology  Research  Center  of  the  U.S. 
Fish  and  Wildlife  Service  for  provid- 
ing support  during  data  collection 
and  analyses.  We  also  thank  R.  Wil- 
ingham,  J.  Talbert,  and  C.  Isbell  for 
assistance  with  the  May  census.  R. 
Haley  and  B.  Turner  of  the  Nevada 
Department  of  Wildlife  provided  re- 
|X)rts  and  shells  for  this  site.  T.  Fritts, 
M.  Molles,  N.  Scott,  H.  Snell,  K.  Sev- 
erson,  and  2  anonymous  reviewers 


197 


read  drafts  of  this  manuscript  and 
greatly  improved  its  content,  but  any 
errors  or  omissions  are  our  own. 


LITERATURE  CITED 

Berry,  Kristin  H.  1986.  Desert  tortoise 
(Gopherus  agassizii)  research  in 
Cahfornia,  1976-1985.  Herpetol- 
ogica  42:62-67. 

Berry,  Kristin  H.,  and  Lori  L. 
Nicholson.  1984a.  Attributes  of 
populations  at  twenty-seven  sites 
in  California,  p.  154-241.  In  The 
Status  of  the  Desert  Tortoise  (Go- 
pherus agassizii)  in  the  United 
States,  Kristin  H.  Berry,  editor. 
Desert  Tortoise  Council  Report  to 
U.S.  Fish  and  Wildlife  Service, 
Sacramento,  California.  Order  No. 
11310-0083-81. 

Berry,  Kristin  H.,  and  Lori  L. 

Nicholson.  1984b.  A  summary  of 
human  activities  and  their  impacts 
on  desert  tortoise  populations  and 
habitat  in  California,  p.  61-117.  In 
The  Status  of  the  Desert  Tortoise 
(Gopherus  agassizii)  in  the  United 
States,  Kristin  H.  Berry,  editor. 
Desert  Tortoise  Council  Report  to 
U.S.  Fish  and  Wildlife  Service, 
Sacramento,  California.  Order  No. 
11310-0083-81. 

Berry,  Kristin  H.,  and  Frederick  B. 
Turner.  1984.  Notes  on  the  behav- 
ior and  habitat  preference  of  juve- 
nile desert  tortoises  (Gopherus 
agassizii)  in  California,  p.  111-130. 
In  Proceedings  of  the  1984  Sympo- 
sium of  the  Desert  Tortoise  Coun- 
cil, Mary  W.  Trotter,  editor.  Long 
Beach,  California. 

Fisher,  Ronald  A.  1930.  The  genetical 
theory  of  Natural  Selection.  Ox- 
ford Univ.  Press,  London.  272  p. 

Fryxell,  J.  M.  1986.  Age-specific  mor- 
tality: an  alternative  approach. 
Ecology  67:1687-1672. 

Galbraith,  David  A.,  and  Ronald  J. 
Brooks.  1987.  Photographs  and 
dental  casts  as  permanent  records 
for  age  estimates  and  growth 
studies  of  turtles.  Herpetological 
Review  18:69-71. 


Germano,  David  J.  1988.  Age  and 
growth  histories  of  desert  tortoises 
using  scute  annuli.  Copeia  1988: [In 
Press]. 

Germano,  David  J.  In  Press.  Prelimi- 
nary comparison  of  growth  in  des- 
ert tortoises  and  gopher  tortoises. 
In  Proceedings  of  the  1987  Sympo- 
sium of  the  Gopher  Tortoise 
Council,  Richard  Franz  and  C. 
Kenneth  Dodd,  Jr.,  editors. 

Gibbons,  J.  Whitfield.  1986.  Move- 
ment patterns  among  turtle  popu- 
lations: applicability  to  manage- 
ment of  the  desert  tortoise.  Herpe- 
tologica  42:104-113. 

Gibbons,  J.  Whitfield.  1987.  Why  do 
turtles  live  so  long?  Bioscience 
37:262-269. 

Jacobson,  Elliott  R.  1980a.  Viral 
agents  and  viral  diseases  of  rep- 
tiles, p.  197-202.  In  Reproductive 
Biology  and  Diseases  of  Captive 
Reptiles,  James  B.  Murphy  and 
Joseph  T.  Collins,  editors.  Society 
for  the  Study  of  Amphibians  and 
Reptiles,  Contribution  to  Herpe- 
tology  Number  1. 

Jacobson,  Elliott  R.  1980b.  Mycotic 
diseases  of  reptiles,  p.  235-241.  In 
Reproductive  Biology  and  Dis- 
eases of  Captive  Reptiles,  James  B. 
Murphy  and  Joseph  T.  Collins, 
editors.  Society  for  the  Study  of 
Amphibians  and  Reptiles,  Contri- 
butions to  Herpetology  Number  1. 

Landers,  J.  Larry,  W.  Alan  McRae, 
and  James  A.  Garner.  1982. 
Growth  and  maturity  of  the  go- 
pher tortoise  in  southwestern 
Georgia.  Bulletin  of  the  Rorida 
State  Museum,  Biological  Sciences, 
27:81-110. 

Luckenbach,  Roger  A.  1982.  Ecology 
and  management  of  the  desert  tor- 
toise (Gopherus  agassizii)  in  Califor- 
nia, p.  1-37.  In  North  American 
Tortoises:  Conservation  and  Ecol- 
ogy, R.  Bruce  Bury  editor.  U.S. 
Fish  and  Wildlife  Service,  Wildlife 
Research  Report  12.  Washington, 
D.C 

Medica,  Philip  A.,  R.  Bruce  Bury,  and 
Frederick  B.  Turner.  1975.  Growth 
of  the  desert  tortoise  (Gopherus 


agassizi)  in  Nevada.  Copeia 
1975:639-643. 

Nagy,  Kenneth  A.,  and  Philip  A. 
Medica.  1986.  Physiological  ecol- 
ogy of  desert  tortoises  in  southern 
Nevada.  Herpetologica  42:73-92. 

Osorio,  Sandalio  Reyes,  and  R.  Bruce 
Bury.  1982.  Ecology  and  status  of 
the  desert  tortoise  (Gopherus  agas- 
sizii) on  Tiburon  Island,  Sonora.  p. 
39-55.  In  North  American  Tor- 
toises: Conservation  and  Ecology, 
R.  Bruce  Bury,  editor.  U.S.  Fish 
and  Wildlife  Service,  Wildlife  Re- 
search Ref)ort  12.  Washington, 
D.C. 

Sokal,  Robert  F.,  and  F.  James  Rohlf. 
1981.  Biometry.  Second  edition. 
W.  H.  Freeman  and  Company, 
New  York.  859  p. 

Tanner,  James  T.  1978.  Guide  to  the 
study  of  animal  populations.  Uni- 
versity of  Tennessee  Press,  Kn- 
oxville.  186  p. 

Trivers,  Robert  L.  1972.  Parental  in- 
vestment and  sexual  selection,  p. 
136-179.  In  Sexual  Selection  and 
the  Descentof  Man— 1871-1971, 
Bernard  Cambell  editor.  Aldine 
Publishing  Co.,  Chicago. 

Turner,  Frederick  B.,  Page  Hayden, 
Betty  L.  Burge,  and  Jan  B.  Rober- 
son.  1986.  Egg  production  by  the 
desert  tortoise  (Gopherus  agassizii) 
in  California.  Herpetologica  42:93- 
104. 

Turner,  Frederick  B.,  Philip  A. 

Medica,  and  Craig  L.  Lyons.  1984. 
Reproduction  and  survival  of  the 
desert  tortoise  (Gopherus  agassizii) 
in  Ivanpah  Valley,  California. 
Copeia  1984:811-820. 

Wilbur,  Henry  M.,  and  Peter  J. 

Morin.  1988.  Life  history  evolution 
in  turtles,  p.  387-439.  In  Biology  of 
the  Reptilia,  Volume  16,  Carl  Gans 
and  Raymond  B.  Huey,  editors. 
Alan  R.  Liss,  Inc.,  New  York. 

Woodbury,  Angus  M.,  and  Ross 
Hardy.  1948.  Studies  of  the  desert 
tortoise,  Gopherus  agassizii.  Eco- 
logical Monographs  18:145-200. 


198 


A  Survey  Method  for 
Measuring  Gopher  Tortoise 
Density  and  Habitat 
Distribution^ 

Daniel  M.  Spillers  and  Dan  W.  Speake^ 


The  only  tortoise  to  occur  in  the 
southeast,  the  gopher  tortoise  (Go- 
pherus  polyphemus)  (fig.  1),  is  limited 
to  six  states.  Of  these  six  states,  legal 
protection  is  offered  by  South  Caro- 
lina, Mississippi,  Georgia,  Florida 
and  Alabama;  Louisiana  does  not  re- 
strict the  harvest  on  gopher  tortoises 
at  present.  The  gopher  tortoise  is 
now  federally  listed  as  threatened  in 
the  portion  of  its  range  west  of  the 
Tombigbee  river  in  Alabama. 

During  the  past  several  years,  an 
apparent  decline  of  gopher  tortoise 
p)opulations  has  been  noted.  Boze- 
man  (1971)  and  Wharton  (1978) 
noted  the  rapid  loss  and  alteration  of 
sand  ridge  habitat,  the  habitat  in 
which  most  gopher  tortoise  popula- 
tions occur,  and  argued  for  the  pres- 
ervation of  these  habitats  not  only  for 
gopher  tortoises  but  also  for  other 
aspects  of  their  ecological  signifi- 

'A  contribution  of  the  Alabama  Coop- 
erative Rshi  and  Wildlife  Research  Unit:  Au- 
burn University  Agricultural  Experiment  Sta- 
tion and  Department  of  Zoology  and  Wild- 
life Science,  Game  and  Rsh  Division  of  the 
Alabama  Department  of  Conservation  and 
Natural  Resources,  the  U.S.  Fish  and  Wildlife 
Service  and  the  Wildlife  Management  Insti- 
tute cooperating.  Presented  at  the  Sympo- 
sium on  Management  of  Amphibians,  Rep- 
tiles, and  Small  Mammals  in  North  America, 
July  19,  1988. 

'Spillers  is  a  research  technician  and 
Speake  is  assistant  unit  leader/wildlife  with 
the  Alabama  Cooperative  Fish  and  Wildlife 
Research  Unit.  Auburn  University,  Alabama 
36849-5414. 


Abstract.— An  underground  closed-circuit 
television  camera  and  Landsat  satellite  imagery 
were  utilized  in  a  2-year  study  to  examine  status  of 
the  gopher  tortoise  in  southern  Alabama.  Use  of  this 
camera  resulted  in  a  complete  count  of  gopher 
tortoises  in  the  sample  transects.  The  transects  were 
located  precisely  on  standard  topographic  maps 
and  on  Landsat  images.  An  estimation  was  then 
made  of  the  amount  of  each  habitat  type  in 
southern  Alabama  based  on  light  reflectance  of  the 
vegetation  and  soil  type  of  the  sample  transects. 
Density  measurements  were  then  expanded  to 
estimate  tortoise  numbers  for  the  entire  area.  This 
method  is  effective  for  estimating  gopher  tortoise 
numbers  and  for  determining  quantity  and  location 
of  gopher  tortoise  habitat. 


cance.  Auffenberg  and  Franz  (1982) 
documented  a  decline  of  gopher  tor- 
toise populations  on  specific  sites  in 
the  Southeast.  Landers  et  al.  (1980) 
found  that  gopher  tortoises  have 
such  a  low  reproductive  rate  that 
human  exploitation  of  tortoises  can 
drastically  reduce  local  populations. 
Landers  and  Speake  (1980)  showed 
that  population  densities  of  gopher 
tortoises  can  fluctuate  widely  in  re- 
sponse to  habitat  manipulation  or 
neglect.  Other  conceivable  reasons 
for  this  apparent  decline  were  noted 
by  Diemer  (1986). 

Sand  ridge  habitat  is  not  only  im- 
portant for  gopher  tortoises,  but  also 
for  many  other  animals  that  use  go- 
pher tortoise  burrows  for  nesting, 
feeding,  or  escape  cover.  Three  sub- 
species of  the  crawfish-gopher  frog 
complex  that  are  closely  associated 
with  gopher  tortoise  burrows  are  the 
dusky  gopher  frog  (Ram  areolata 
sevosa),  the  Rorida  gopher  frog,  (R.  a. 
aesypus),  and  the  Carolina  gopher 
frog  (R.a.  capito).  The  threatened 
eastern  indigo  snake  (Drymarchon 
corais  couperi)  is  dependent  on  tor- 
toise burrows  for  winter  cover  in  the 
northern  part  of  its  range  (Speake  et 
al.,  1978;  Landers  and  Speake,  1980; 
Diemer  and  Speake,  1981).  Several 
species  of  mammals  and  birds  use 
gopher  tortoise  burrows,  most  often 
as  escape  cover.  Several  authors  have 
noted  the  diversity  of  animal  life 
(both  vertebrate  and  invertebrate) 


Figure  1  .—A  gopher  tortoise  from  southern 
Alabama. 


inhabiting  tortoise  burrows  and  the 
dependence  of  some  species  on  tor- 
toise burrows  for  survival  (Allen  and 
Neill,  1951;  Hubbard,  1894;  Hutt, 
1967;  Landers  and  Speake,  1980; 
Speake  et  al.,  1978;  Woodruff,  1982). 

In  view  of  the  apparent  decline  of 
gopher  tortoise  populations,  it  is  im- 
p)ortant  to  be  able  to  accurately  meas- 
ure tortoise  density  in  an  area  and  to 
determine  quantity  and  distribution 
of  suitable  tortoise  habitat.  Tortoise 
density  has  been  previously  esti- 
mated by  means  of  a  correction  fac- 
tor applied  to  counts  of  burrows 
(Auffenberg  and  Franz,  1982),  dig- 
ging of  burrows,  and  use  of  listening 
devices.  Previous  methods  do  not 
ensure  accurate  determination  of  tor- 
toise density  without  burrow  de- 
struction and  prohibitive  labor.  De- 
termination of  quantity  and  location 
of  tortoise  habitat  is  becoming  neces- 


199 


sary  due  to  rapid  changes  in  land  use 
and  increasing  relocation  and  re- 
stocking efforts  (Diemer,  1984;  Lan- 
ders, 1981). 

The  objectives  of  this  study  were 
to  develop  and  employ  a  method  to: 
(1)  accurately  measure  gopher  tor- 
toise density  and  (2)  locate  and  quan- 
tify tortoise  habitat  in  a  24-county 
area  of  southern  Alabama. 

We  are  indebted  to  James  Altiere, 
Eugene  Carver,  Kevin  Dodd,  Lane 
Knight,  Sonny  Mitchell,  Claud 
Searcy,  and  William  Sermons,  who 
assisted  in  collecting  field  data.  We 
are  especially  indebted  to  Walter 
Stephenson,  Chief  of  the  Resource 
Development  Section  of  the  State 
Planning  Division,  Department  of 
Economic  and  Community  Affairs, 
State  of  Alabama  for  his  help  and  co- 
operation in  giving  us  access  to  the 
Landsat  remote  sensing  system.  Ap- 
preciation is  extended  to  Joe  Exum, 
Raymond  Metzler,  and  Nick  Wiley 
for  their  assistance  in  experimental 
design  and  data  analysis.  Special  ap- 
preciation is  extended  to  Dr.  Charles 
Williams  of  the  Research  and  Data 
Analysis  Department,  Auburn  Uni- 
versity, for  his  advice  and  aid  with 
statistical  design  and  analysis.  The 
project  was  funded  by  a  grant  from 
the  U.S.  Fish  and  Wildlife  Service 
and  by  the  Alabama  Cooperative 
Fish  and  Wildlife  Research  Unit. 


Methods 

Study  Area  Determination  and 
Questionnaires 

Our  study  area  was  determined  by 
the  reported  historical  range  of  the 
gopher  tortoise  in  Alabama  (Mount, 
1978;  Auffenberg  and  Franz,  1982). 
This  included  24  counties  in  the 
coastal  plain  of  Alabama  (excluding 
the  counties  west  of  the  Tombigbee 
river  which  were  surveyed  by  other 
researchers).  Questionnaires  were 
sent  out  to  wildlife  biologists,  conser- 
vation officers,  herpetologists, 
county  agents,  soil  conservation 


agents  and  other  people  who  were 
likely  to  have  knowledge  of  gopher 
tortoise  populations  in  our  24-county 
study  area.  These  questionnaires 
asked  for  locations  of  areas  that  sup- 
ported or  had  supported  tortoise 
populations,  and  names  of  landown- 
ers or  other  persons  who  might  have 
additional  knowledge  of  tortoise 
populations.  A  map  was  included 
with  each  questionnaire  so  that  loca- 
tions could  be  marked.  A  total  of  132 
questionnaires  was  mailed  out  and 
58%  were  returned. 

Soil  conservation  offices  were  vis- 
ited in  each  surveyed  county  and  fur- 
ther inquiries  were  made  concerning 
tortoise  population  occurrence  and 
habitat  availability.  Areas  in  each 
county  that  had  soils  with  sand  to  a 
depth  of  at  least  1  m  and  that  pref- 
erably contained  a  variety  of  habitat 
types  were  delineated  on  maps. 
These  areas  were  considered  poten- 
tial tortoise  habitat  (Garner  and  Lan- 
ders, 1981;  Landers,  1981;  Landers 
and  Garner,  1981)  and  were  used  to 
sample  tortoise  densities. 

After  evaluation  of  the  informa- 
tion from  the  questionnaires,  per- 
sonal interviews,  and  discussion  with 
soil  conservation  agents,  the  24- 
county  study  area  was  divided  into 
three  classes  (fig.  2).  Class  I  counties 
(n=14)  contained  widely  distributed 
gopher  tortoise  {X)pulations  and 
habitat.  Class  II  counties  (n=4)  con- 
tained relict  or  disjunct  populations 
and  scattered,  spotty  habitat.  Class 
III  counties  (n=6)  were  those  in 
which  no  tortoise  populations  could 
be  found. 


Sampling  Sctieme 

In  Class  I  counties,  regions  deline- 
ated by  the  soil  conservation  agents 
(sandy  soil  >  1  m)  were  located  on 
1:24,000  scale  topographic  maps. 
Within  these  areas,  a  reference  point 
for  initiation  of  sampling  was  chosen 
from  the  map  which  had  a  variety  of 
habitat  types  (at  least  2)  within  a  1 
km  radius  of  the  reference  point. 


These  points  were  chosen  before  vis- 
iting the  site.  Where  necessary,  per- 
mission was  obtained  for  sampling 
on  private  property. 

Upon  arrival  at  the  location  as 
many  of  the  following  habitat  types 
were  located  as  possible:  unburned 
pine/ scrub  oak,  burned  pine/ scrub 
oak,  planted  pines,  clearcuts,  old- 
fields,  agricultural  fields,  pasture, 
and  corresponding  edges  for  each 
type.  The  example  of  each  habitat 
type  nearest  to  the  reference  point 
was  then  sampled. 

Belt  transects  measuring  265  x  15 
m  (0.4  ha)  were  systematically  lo- 
cated within  the  habitat  types  avail- 
able; edge  transects  were  centered  on 
and  followed  the  edge.  If  there  were 
open  burrows  in  the  transect,  the 
burrows  were  examined  using  the 
MUTVIC  (Miniature  Underground 
Television  Inspection  Camera) 
(Speake  and  Altiere,  1983).  This  de- 
vice enabled  us  to  insert  a  closed- 


Figure  2.— Distribution  of  the  gophier  tortoise 
in  24  counties  of  Alabama. 


200 


circuit  television  camera  to  the  bot- 
toms of  the  burrows  and  determine  if 
they  were  occupied  (figs.  3-5).  Bur- 
row width  measurements  were  made 
with  calipers  inserted  approximately 
70  cm  into  the  burrow.  Data  gathered 
for  each  transect  included  habitat 


type,  number  of  open  burrows,  num- 
ber of  active  burrows  (burrows  with 
sign  of  recent  tortoise  use),  number 
of  tortoises,  and  width  of  burrows. 

In  Class  II  counties  we  searched 
each  area  where  tortoise  populations 
had  been  reported  or  where  gopher 


Figure  3.— Closed-circuit  television  camera  with  protective  glass  globe. 


Figure  4.— Crew  inserting  closed-circuit  television  camera  into  gopher  tortoise  burrow. 


tortoise  habitat  (sandy  soil  >  1  m)  ex- 
isted. Observations  were  made  of  the 
total  number  of  burrows,  and  total 
number  of  active  burrows.  Since 
these  counties  lay  along  the  northern 
border  of  the  gopher  tortoise's  range 
in  Alabama,  tortoise  populations 
were  scattered  and  did  not  occur  as 
uniformly  in  specific  habitat  types  as 
those  populations  in  Class  I  counties. 
Therefore  we  did  not  sample  here 
but  instead  used  a  correction  factor 
similar  to  the  one  described  by 
Auffenberg  and  Franz  (1982).  The 
correction  factor  (0.67  tortoises/ac- 
tive burrow)  was  obtained  from  our 
sampling  of  Class  I  counties  by  di- 
viding the  total  number  of  tortoises 
by  the  total  number  of  active  bur- 
rows. The  estimated  total  number  of 
tortoises  for  Class  II  counties  was 
very  low  (56),  and  did  not  signifi- 
cantly affect  our  population  estimate. 

Landsat  Satellite  Imagery 

Having  measured  tortoise  density  on 
sample  areas  of  the  habitat  types, 
Landsat  digital  satellite  imagery  was 
used  to  obtain  an  estimate  of  the  area 
of  each  habitat  type  in  Class  I  coun- 
ties. Characteristics  and  usage  of  this 
remote  sensing  technique  are  de- 
scribed by  Anderson,  Wentz  and 
Treadwell  (1980),  Brabander  and 
Barclay  (1977),  Diemer  and  Speake 
(1983),  Graham  et  al.  (1981),  Taranik 
(1978a),  and  Taranik  (1978b).  The 
system  we  used  makes  a  scan  of  the 
earth  every  eighteen  days  from  a 
geosynchronous  orbit.  The  multis- 
pectral  scanner  operates  in  seven  dif- 
ferent wavelengths  of  light — four  vis- 
ible and  three  infrared.  We  used  near 
infrared  because  it  showed  vegeta- 
tion characteristics  more  clearly.  By 
making  several  passes,  the  scanner 
senses  light  reflectance  based  on  0.1 
ha  pixels.  Each  0.1  ha  of  the  earth's 
surface  is  assigned  1  of  256  gray  val- 
ues based  on  its  reflectance.  Using 
these  gray  values  we  separated  the 
following  habitat  types  based  on 
their  spectral  signature:  unburned 


201 


pine/scrub  oak,  burned  pine/ scrub 
oak,  planted  pine,  old-field,  agricul- 
tural fields,  pasture  and  composite 
edge. 

Before  sampling,  we  used  ground- 
truthing  to  determine  if  it  was  fea- 
sible to  attempt  to  classify  each  habi- 
tat type  using  Landsat  imagery.  On 
70-0.4  ha  sample  plots  in  Baldwin 
County  (10  plots  in  each  habitat 
type),  each  plot  was  correctly  classi- 
fied. Clearcuts  were  not  included  be- 
cause they  were  a  rapidly  changing 
transient  stage  (1-2  years)  leading  to 
planted  pine  habitat,  and  as  such 
could  not  be  identified  on  Landsat 
images  accurately  due  to  their  rapid 
vegetational  change.  Habitat  was 
considered  planted  pine  if  pine  was  a 
prominent  understory  or  midstory 
component  (at  least  0.3  m  tall).  Indi- 
vidual edge  types  were  combined 
because  edge  transects  had  similar 
vegetation  characteristics  and  thus  a 
similar  spectral  signature.  Combined 
edge  habitat  was  identifiable. 

NASA  software  used  with  Land- 
sat imagery  includes  a  program  for 
referencing  Landsat  digital  data  to 
any  scale  map.  We  referenced  our 
data  to  standard  1:24,000  topo- 
graphic maps  using  known  control 
points.  This  enabled  us  to  use  Uni- 
versal Trans  Mercator  coordinates  to 
locate  each  transect  on  the  Landsat 
image  and  obtain  the  correct  gray 
value  for  each  transect.  We  then  as- 
signed a  range  of  gray  values  to  each 
habitat  type  based  on  the  reflectance 
of  the  sample  transects.  The  accuracy 
of  the  habitat  classifications  was 
checked  throughout  this  process. 

A  polygon  was  then  constructed 
enclosing  all  the  Class  I  counties,  and 
areas  of  each  gray  value  within  this 
polygon  were  measured.  From  these 
measurements  we  determined  the 
total  area  for  each  habitat  type  in 
Class  I  counties. 


Data  Analysis 

We  had  two  concerns  relative  to  data 
analysis:  (1)  to  derive  a  population 


estimate  based  on  mean  tortoise  den- 
sity f>er  hectare  multiplied  by  the  es- 
timated area  of  the  respective  habitat 
type,  and  (2)  to  identify  and  locate 
gopher  tortoise  habitat. 

In  order  to  obtain  a  population 
estimate  we  multiplied  the  mean 
density  of  gopher  tortoises  per  hec- 
tare in  a  specific  habitat  type  by  the 
total  area  of  that  habitat  type  in  Class 
I  counties.  An  allowance  was  made 
for  standard  error  of  the  mean.  The 
habitat  totals  were  then  summed  to 
give  a  final  population  estimate  of 
the  Class  I  counties. 

In  addition  to  these  concerns  we 
examined  age  class  structure.  Lan- 
ders et  al.  (1982)  noted  that  gopher 
tortoises  pass  through  two  general 
life-history  stages  before  they  reach 
sexual  maturity.  The  juvenile  stage 
lasts  until  the  carapace  is  approxi- 
mately 100-120  mm.  During  the  juve- 
nile stage,  the  shells  are  very  soft  and 
carapacial  scutes  usually  have  dis- 
tinct yellow  centers.  This  stage  usu- 


ally lasts  until  about  5  years  of  age. 
Juvenile  coloration  fades  and  the 
shells  begin  to  harden  during  the 
subadult  stage  which  generally  lasts 
from  5  to  21  years  of  age.  Carapace 
lengths  range  from  about  120-220 
mm.  At  sexual  maturity,  body  vol- 
ume has  drastically  increased  and 
sexual  dimorphism  is  apparent.  This 
occurs  at  approximately  21  years  of 
age  and  a  carapace  length  of  230  mm. 
Alford  (1980)  established  a  mathe- 
matical relationship  between  the 
widths  of  gopher  tortoise  burrows 
and  the  carapace  lengths  of  their  oc- 
cupants in  northern  Florida  (this  rela- 
tionship has  not  been  thoroughly 
tested  in  other  states).  Using  Alford's 
equation  log^^y  =  0.879  log^^x  +  0.149, 
where  y  is  carapace  length  and  x  is 
burrow  width,  we  used  our  burrow 
width  measurements  of  occupied 
burrows  to  divide  tortoise  popula- 
tions into  juvenile,  subadult,  and 
adult  age  classes.  We  considered  age 
class  structure  to  be  an  important  cri- 


Figure  5.— Closed-circuit  television  monitor  displaying  picture  of  a  gopher  tortoise  inside  a 
burrow. 


202 


 ^  .  

Table  1  .—Summary  of  sample  variables  and  derived  estimates  for  Class  I  counties  from  339-0>4  ha  transects  in  south- 
em  Alabama,  1984-1985. 


Standard    Area  Population 
Habitat  n  Habitat  totals  Mean  densitles/ha       error*"      (ha)  estimate 

Open     Active  Open  Active 

burrows  burrows  Tortoises  burrows  burrows  Tortoises 


Old-field 

21 

23 

17 

13 

2.72 

2.00 

1.53 

0.47 

35,822 

207,808  ± 

63,836 

Planted  Pine 

17 

7 

6 

5 

1.01 

0.87 

0.72 

0.35 

99,855 

71,896  ± 

34,949 

Burned  Pine/ 

Scrub  Oak 

34 

36 

13 

9 

2.62 

0.94 

0.64 

0.27 

209,108 

133,829  ± 

56,459 

Edge 

129 

85 

54 

34 

1.63 

1,04 

0.64 

0.15 

102,408 

65,541  ± 

15,361 

Pasture 

46 

1 

1 

1 

0.05 

0.05 

0.05 

0.05 

61,225 

3,061 

±  3,061 

Agriculture 

31 

0 

0 

0 

0.00 

0.00 

0.00 

0,00 

210,386 

0 

Unburned  Pine/ 

Scrub  Oak 

10 

0 

0 

0 

0.00 

0.00 

0.00 

0.00 

133,004 

0 

Clearcuts 

51 

1 

1 

0 

0.05 

0.05 

0.00 

0.00 

Totals 

339 

153 

92 

62 

951,808 

482,135  ± 

173,666 

"Standard  error  o  f  the  tortoise  mean  dertsity/ha. 


teria  along  with  density  in  evaluating 
tortoise  population  viability.  Re- 
search has  not  yet  revealed  an  opti- 
mum age  class  structure.  Intuitively, 
in  a  long-lived  animal  such  as  the 
gopher  tortoise,  the  age  class  struc- 
ture of  a  healthy  population  would 
be  skewed  toward  the  adult  class. 
The  presence  of  juvenile  and  sexually 
mature  adult  tortoises  does  definitely 
indicate  recent  reproduction. 

Results 

Gopher  Tortoise  Densities  and 
Habitat  Areas 

Tortoise  densities  and  habitat  areas 
were  measured  in  Class  I  counties. 
These  results  are  summarized  in 
table  1,  which  includes  sampling 
variables  by  habitat  type  along  with 
estimates  derived  from  sampling. 


Age  Class  Structure 

Five  percent  of  the  sampled  popula- 
tion (n=100  tortoises)  were  juvenile 
tortoises,  48%  were  subadult,  and 
47%  were  adults.  This  structure 


shows  that  there  has  been  recent  re- 
production, and  that  there  is  a  large 
segment  of  breeding  size  adults  pres- 
ent. This  suggests  that  the  potential 
for  successful  population  mainte- 
nance over  the  estimated  951,808  ha 
area  of  tortoise  habitat  in  Class  I 
counties  is  good. 

Discussion 

Using  the  referenced  Landsat  data 
and  knowing  the  range  of  gray  val- 
ues for  each  habitat  type,  we  were 
able  to  examine  any  area  in  Class  I 
counties  and  determine  the  size  and 
quantity  of  gopher  tortoise  habitat 
units.  Using  a  plotter,  figures  can  be 
made  of  all  the  0.1  ha  pixels  that  cor- 
respond to  a  given  habitat  type  and 
then  the  figure  can  be  overlaid  on  a 
map.  For  our  purposes  we  only 
needed  the  area  of  each  habitat  type 
in  Class  I  counties. 

This  technique  has  two  distinct 
sources  of  error.  First  is  the  variation 
of  the  gopher  tortoise  densities 
within  habitat  types.  These  variations 
are  inherent  in  sampling  biological 
populations.  In  this  study  the  vari- 
ance was  fairly  low.  Increased 
sample  size  would  likely  lower  this 


 J 

error.  The  second  source  of  error  is 
in  estimating  total  areas  of  the  habi- 
tat types  over  a  large  region.  Al- 
though in  our  preliminary  ground- 
truthing,  Landsat  imagery  correctly 
classified  all  our  habitat  types  (ex- 
cluding clearcuts  and  individual 
edge  types),  we  suspect  that  when 
this  technique  is  applied  to  a  large 
diverse  region  some  areas  will  be 
misclassified.  Ground-tru thing 
should  be  done  after  the  classifica- 
tion to  determine  what  percentage 
has  been  misclassified,  which  would 
allow  the  researcher  to  make  allow- 
ances for  this  error  in  final  computa- 
tions. 


Conclusions 

We  found  this  technique  to  be  useful 
for  measuring  tortoise  density  and 
for  determining  quantity  and  loca- 
tion of  tortoise  habitat.  The  error  in 
this  technique  seems  to  b>e  less  than 
that  for  techniques  used  for  census- 
ing  most  other  animals.  Although  it 
is  difficult  to  estimate  numbers  of 
animals  over  a  large  area,  it  is  helpful 
to  be  able  to  accurately  measure  den- 
sity in  small  areas  and  then  extrapo- 
late this  density  on  the  basis  of  a 


203 


quantitative  measurement  of  a  desig- 
nated area.  This  method  should  be 
especially  valuable  for  surveys  of 
animals  that  are  habitat  specific. 

Literature  Cited 

Alford,  Ross  A.  1980.  Population 
structure  of  Gopherus  polyphemus 
in  northern  Florida.  Journal  of 
Herpetology.  14:177-182. 

Allen,  Ross,  and  Wilfred.  T.  Neill. 
1951.  The  gopher.  Florida  Wildlife. 
5(7):10, 32. 

Anderson,  William  H.,  W.  Alan 
Wentz,  B.  Dean  Tread  well.  1980.  A 
guide  to  remote  sensing  for  wild- 
life biologists.  Pages  291-303.  In  S. 
D.  Schemnitz  (ed.).  Wildlife  man- 
agement techniques  manual.  The 
Wildlife  Society,  Washington,  D.C. 
viii  +  686  p. 

Auffenberg,  Walter  and  Richard 
Franz.  1982.  The  status  and  distri- 
bution of  Gopherus  polyphemus.  p. 
95-126.  In  R.B.  Bury  (ed)  North 
American  tortoises:  Conservation 
and  ecology.  U.S.  Fish  and  Wild- 
life Service  Research  Rep.  No.  12. 

Bozeman,  John.  R.  1971.  A  sociologic 
and  geographic  study  of  the  sand 
ridge  vegetation  in  the  Coastal 
Plain  of  Georgia.  Ph.D.  disserta- 
tion. University  of  North  Carolina, 
Chapel  Hill.  244  p. 

Brabander,  Jerry  J.  and  John  S. 
Barclay.  1977.  A  practical  applica- 
tion of  satellite  imagery  to  wildlife 
habitat  evaluation.  Proceedings  of 
the  Annual  Conference  of  South- 
eastern Fish  and  Wildlife  Agen- 
cies. 31:300-306. 

Diemer,  Joan  E.  1984.  Tortoise  reloca- 
tion in  Florida:  solution  or  prob- 
lem? Proceedings  of  the  9th  An- 
nual Symposium  of  the  Desert 
Tortoise  Council,  p.  131-135. 

Diemer,  Joan  E.  1986.  The  ecology 
and  management  of  the  gopher 
tortoise  in  the  southeastern  United 
States.  Herpetologica  42(1):125- 
133. 

Diemer,  Joan  E.  and  Dan  W.  Speake. 
1981.  The  status  of  the  eastern  in- 


digo snake  in  Georgia.  In  Ron  R. 
Odom  and  J.  W.  Guthrie  (eds.). 
Proceedings  of  the  Nongame  and 
Endangered  Wildlife  Symposium. 
Ga.  Dept.  Nat.  Resources,  Game  & 
Fish. Tech.  Bull.  WL5.  p.  52-61. 

Diemer,  Joan  E.  and  Dan  W.  Speake. 
1983.  The  distribution  of  the  east- 
ern indigo  snake  (Drymarchon  co- 
rais  couperi)  in  Georgia.  Journal  of 
Herpetology  17:256-264. 

Garner,  James  A.  and  J.  Larry  Lan- 
ders. 1981.  Foods  and  habitat  of 
the  gopher  tortoise  in  southwest- 
em  Georgia.  Proceedings  of  the 
Annual  Conference  of  the  South- 
eastern Association  of  Fish  and 
Wildlife  Agencies.  35:120-133. 

Graham,  Marcellus  H.,  Ronnie  W. 
Pearson,  Benjamin  R.  Seyfarth, 
Bobby  G.  Junkin,  Maria  T.  Kalcic. 
1980.  ELAS,  earth  resources  labo- 
ratory applications  software.  Re- 
port No.  183  of  the  NASA,  Na- 
tional Space  Technologies  Labora- 
tories, Earth  Resources  Labora- 
tory. 

Hubbard,  Henry.  G.  1894.  The  insect 
guests  of  the  Florida  land  tortoise. 
Insect  Life  6:302-315. 

Hutt,  Art.  1967.  The  gopher  tortoise, 
a  versatile  vegetarian.  Florida 
Wildlife.  21(7):20-24. 

Landers,  J.  Larry.  1981.  Techniques 
for  restocking  gopher  tortoise 
populations.  In  The  Gopher  Tor- 
toise: Distribution,  Ecology,  and 
Effects  of  Forest  Management.  Fi- 
nal Report  to  Georgia  Department 
of  Natural  Resources. 

Landers,  J.  Larry  and  James  A.  Gar- 
ner. 1981.  Status  and  distribution 
of  the  gopher  tortoise  in  Georgia, 
p.  45-51.  In  Ron  R.  Odom  and  J. 
W.  Guthrie  (eds.).  Proceedings  of 
the  Nongame  and  Endangered 
Wildlife  Symposium.  Ga.  Dept. 
Nat.  Resources,  Game  &  Fish 
Tech.  Bull.  WL5. 

Landers,  J.  Larry,  James  A.  Garner, 
and  W.  Alan  McRae.  1980.  Repro- 
duction of  gopher  tortoises  (Go- 
pherus polyphemus)  in  southwest- 
em  Georgia.  Herpetologica 
36(4):353-361. 


Landers,  J.  Larry  and  Dan  W. 

Sf)eake.  1980.  Management  needs 
of  sandhill  reptiles  in  southern 
Georgia.  Proceedings  of  the  An- 
nual Conference  of  the  Southeast- 
em  Association  of  Fish  and  Wild- 
life Agencies.  34:515-529. 

Mount,  Robert  H.  1975.  The  Reptiles 
and  Amphibians  of  Alabama.  Ala- 
bama Agricultural  Experiment 
Station,  Auburn.  347  p. 

Speake,  Dan  W.  and  James  A.  Alti- 
ere.  1983.  A  device  for  viewing 
and  filming  the  contents  of  tree 
cavities,  p.  185-187.  In  Davis,  J.  W., 
Gregory,  A.  F.  and  R.  A.  Ocken- 
tield  (Tech.  Coords.)  1983  Snag 
Habitat  Management:  proceedings 
of  the  symposium.  USDA  Forest 
Service  General  Technical  Report 
RM-99.  226  pages.  Rocky  Moun- 
tain Forest  and  Range  Experiment 
Station,  Ft.  Collins,  CO. 

Speake,  Dan  W.,  Joe  A.  McGlincy, 
Thaggard  A.  Colvin.  1978.  Ecology 
and  management  of  the  eastern 
indigo  snake  in  Georgia:  a  prog- 
ress report.  Pages  64-73.  In  Ron  R. 
Odom  and  Larry  Landers  (eds.). 
Proceedings  of  the  Rare  and  En- 
dangered Wildlife  Symposium. 
Ga.  Dept.  Nat.  Resources,  Game  & 
Fish  Tech.  Bull.  WL4. 

Taranik,  James  V.  1978a.  Characteris- 
tics of  the  Landsat  Multispectral 
Data  System:  U.S.  Geological  Sur- 
vey Open-file  Report  78-187,  76  p., 
Sioux  Falls,  South  Dakota. 

Taranik,  James  V.  1978b.  Principles  of 
computer  processing  of  Landsat 
data  for  geologic  applications:  U.S. 
Geological  Survey  Open-file  Re- 
port 78-117,  50  p.,  Sioux  Falls, 
South  Dakota. 

Wharton,  Charles  H.  1978.  The  natu- 
ral environments  of  Georgia. 
Georgia  Dept.  Nat.  Res.  277  p. 

Woodruff,  Robert  E.  1982.  Arthro- 
pods of  gopher  burrows,  p.  24-48. 
In  R.  Franz  and  R.  J.  Bryant  (Eds.), 
The  Gopher  Tortoise  and  Its 
Sandhill  Habitat.  Proceedings  of 
the  3rd  Annual  Meeting,  Gopher 
Tortoise  Council. 


204 


Evaluation  and  Review  of 
Field  Techniques  Used  to 
Study  and  Manage  Gopher 
Tortoises^ 

Russell  L.  Burke^  and  James  Cox^ 


Abstract. —This  paper  reviews  methods  used  to 
census  gopher  tortoises  as  well  as  techniques  for 
demographic,  reproduction,  and  movement 
studies.  We  also  evaluate  a  refinement  for  line 
transect  estimates  of  gopher  tortoise  abundance.  In 
situations  where  dense  vegetation  structure  may 
hinder  abilities  to  locate  burrows  along  transects, 
Fourier  series  estimators  of  abundance  con  be  used 
to  overcome  the  problem,  However,  our  results 
indicate  that  many  transects  may  be  needed  to 
provide  precise  estimates  of  gopher  tortoise 
abundance  over  large  areas.  The  collection  of 
vegetation  data  along  transects  may  also  be  helpfu 
in  evaluating  habitat  preference  in  this  species. 


Introduction 

Of  the  approximately  107  genera  and 
267  species  of  North  American  rep- 
tiles, two  species  of  tortoises  have 
received  a  relatively  large  amount  of 
scientific  attention.  Organizations 
dedicated  to  the  conservation  and 
protection  of  the  gopher  tortoise  (Go- 
pherus  polyphemus)  (The  Gopher  Tor- 
toise Council)  and  the  desert  tortoise 
(G.  agassizi)  (The  Desert  Tortoise 
Council)  attest  to  heightened  levels 
of  amateur  and  scientific  interest  in 
these  species.  Past  bibliographies 
(Diemer  1981,  Douglass  1975, 
Douglass  1977,  Hohman  et  al.  1980) 
together  record  over  775  different 
publications  concerning  the  genus, 
and  more  have  been  published  since 
then.  Compared  to  most  other  reptile 
species,  an  exceptional  diversity  of 
techniques  has  been  employed,  and 
many  field  methods  have  been  devel- 
oped and  used  to  study  their  status 
and  biology. 

The  gopher  tortoise  is  a  large  ter- 
restrial turtle  (15-37  cm  carapace 
length,  3.6-5.0  kg)  that  exhibits  low 
rates  of  juvenile  recruitment,  extreme 

'Paper  presented  at  syrDposium,  Man- 
agement of  Amphibiarts.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff. AZ,  July  19-21  1988.) 

'Researct)  Associate.  Tall  Timbers  Re- 
search! Station.  Route  1.  Box  678.  Tallahas- 
see. Florida.  32312. 

'Biologist.  Nongame  Wildlife  Program. 
Florida  Game  and  Fresh  Water  Fish  Com- 
mission. 620  S.  Meridian  Street.  Tallahassee. 
Rorida.  32399-1600. 


adult  longevity,  and  persistent  use  of 
a  small  number  of  burrows,  often  in 
a  loose  aggregation  of  10  to  15  indi- 
viduals. As  a  result,  tortoises  display 
a  social  system  that  involves  indi- 
viduals who  may  have  interacted 
regularly  for  decades  (Douglass  1976, 
Landers  et  al.  1980,  McRae  et  al. 
1980).  Tortoises  were  once  a  common 
feature  of  the  upland  habitats  of  the 
southeastern  coastal  plain  (Auffen- 
berg  and  Franz  1982),  but  the  species 
is  now  less  common  and  appears  on 
several  state  and  federal  lists  of  rare 
or  endangered  species  (Lohoefener 
and  Lohmeier  1984,  Wood  1987).  The 
principal  forces  driving  these  popu- 
lation declines  are  rapid  urbaniza- 
tion, certain  forest  management  prac- 
tices, and  human  predation  (Diemer 
1986). 

Gopher  tortoise  burrows  are  im- 
portant to  a  large  wildlife  commu- 
nity, and  332  other  species  have  been 
documented  to  use  tortoise  burrows 
at  least  occasionally  (Jackson  and 
Milstrey  in  press).  Included  among 
the  several  rare  species  that  rely 
heavily  on  tortoise  burrows  are  the 
Florida  mouse  (Podomys  floridanus), 
Florida  and  dusky  crawfish  frogs 
(Ram  areolata  aesopus  and  R.  areolata 
sevosa),  sand  skink  (Neoseps  reynoldsi), 
Florida  pine  snake  (Pituophis  melano- 
leucus  mugitus),  and  eastern  indigo 
snake  (Drymarchon  corais  couperi). 

In  this  paper  we  review  tech- 
niques used  in  field  research  on  the 
gopher  tortoise  community.  We  also 
discuss  future  areas  of  research  and 


analyze  the  use  of  Fourier  series  esti- 
mators (Burnham  et  al.  1980)  in  line 
transect  censusing  techniques.  In 
doing  so  we  suggest  appropriate 
methods  for  future  work,  standard- 
ize some  techniques,  bring  some 
lesser  known  techniques  to  the  fore, 
and  suggest  refinements  to  com- 
monly used  methods. 

Estinnating  Population  Size 

Burrow  Count  Transects 

Burrow-count  transects  are  currently 
the  most  widely  used  method  for  es- 
timating the  size  of  local  gopher  tor- 
toise populations,  though  some  tor- 
toise populations  do  not  dig  burrows 
(Auffenberg  1969),  while  others  may 
use  seven  or  more  burrows  per  indi- 
vidual (McRae  et  al.  1980).  Burrows 
are  particularly  amenable  to  transect 
analysis  since  they  are  stationary  and 
generally  visible  in  many  of  the  open 
areas  occupied  by  gopher  tortoises. 
Transects  also  require  little  equip- 
ment, can  be  used  to  cover  relatively 
large  areas  in  a  short  time,  and  can 
be  used  to  estimate  abundance  over  a 
large  area  using  random  or  stratified- 
random  sampling  procedures.  A  con- 
version factor  (Auffenberg  and  Franz 
1982)  is  used  to  relate  the  number  of 
different  tortoise  burrows  to  the 
number  of  gopher  tortoises  in  an 
area. 

The  dimensions  of  reported 
transects  ranges  from  100  to  250  m  in 


205 


length  to  7  to  10  m  in  width  (Auffen- 
berg  and  Franz  1982,  Cox  et  al.  1987, 
Lohoefener  and  Lohmeier  unpub. 
rep.).  Lohoefener  (in  press)  points 
out  that  strip  transect  burrow  counts 
assume  that  all  burrows  are  detected 
within  a  strip.  Breininger  et  al.  (in 
press),  however,  expressed  concern 
that  dense  vegetation  could  make 
strip-transect  estimates  unreliable 
unless  the  transects  were  narrow. 
The  thick  oak  scrub  (Quercus  spp.) 
vegetation  common  on  many  of  their 
study  sites,  for  example,  would  have 
prohibited  surveyors  from  seeing 
burrows  more  than  a  few  meters 
from  transect  lines. 

A  possible  method  of  correcting 
this  problem  (Cox  et  al.  1987, 
Lohoefener  in  press)  is  to  take  {per- 
pendicular distance  measures  from 
transect  lines  to  observed  gopher  tor- 
toise burrows.  Perpendicular  dis- 
tances can  be  used  in  Fourier  series 
density  estimators  (or  other  estima- 
tors) (Bumham  et  al.  1981)  to  account 
for  differences  in  the  detectability  of 
burrows  due  to  vegetation  or  the  size 
of  the  burrow. 

To  look  at  this  problem  in  more 
depth,  we  compared  strip  transects 
and  line  transects  by  establishing  12 
transects  (250  m  by  20  m)  in  each  of 
three  areas  containing  gopher  tor- 
toise populations.  The  areas  selected 
had  noticeable  differences  in  vegeta- 
tive structure.  The  first  site  was  a 
mixed  longleaf  pine  (Pinus  palustris), 
turkey  oak  (Quercus  laevis)  habitat  on 
a  private  ranch;  the  second  site  was 
an  early  successional  sand  pine  scrub 
(P.  clausa)  forest  on  private  timber 
lands;  and  the  third  site  was  a  ma- 
ture longleaf  pine  forest  in  the 
Apalachicola  National  Forest.  The 
starting  points  and  directions  of 
transects  within  these  areas  were 
randomly  selected. 

Perpendicular  distances  from  bur- 
rows to  transect  lines  were  measured 
to  the  nearest  0.25  m,  and  only  bur- 
rows detected  from  the  transect  line 
were  recorded  (i.e.,  burrows  located 
while  measuring  perpendicular  dis- 
tances to  burrows  seen  from  the 


transect  line  were  ignored).  Burrow 
densities  for  each  of  the  three  areas 
were  estimated  directly  using  the 
number  recorded  on  transects  and 
Fourier  series  estimators  obtained 
from  perpendicular  distance  data 
(table  1)".  Fourier  series  estimators 
were  calculated  using  the 
TRANSECT  program  developed  by 
Laake  et  al.  (1979)  and  are  presented 
in  table  1  for  the  three  sites. 

Vegetation  structure  appeared  to 
influence  the  estimate  of  burrow  den- 
sity on  the  early  successional  site 
(Site  2),  but  the  Fourier  series  esti- 
mate of  density  was  no  different  than 
the  estimate  provided  by  direct  com- 
putations on  the  other  sites.  The 
early  successional  site  had  a  very 
thick  shrub  component  that  made  it 
difficult  to  locate  burrows  several 
meters  from  the  transect  line.  Ten 
meters  was  probably  too  wide  a 
transect  width  in  this  particular  set- 
ting. The  direct  computation  of  bur- 
row density  from  transect  data  on 
Site  2  is  only  half  the  density  estimate 
developed  by  the  Fourier  series  esti- 
mate. 

The  level  of  variation  observed 
among  transects  (whether  they  be 
strip  or  line  transects)  within  a  site 
can  be  used  to  estimate  the  number 
of  additional  transects  needed  to  at- 
tain a  higher  level  of  accuracy  for  the 
estimate  of  density  (Bumham  et  al. 
1981).  To  increase  the  precision  of 
our  estimates  by  10%,  for  example, 
an  additional  24  transects  would  be 
needed  for  Site  1, 40  for  Site  2,  and  78 
for  Site  3.  Such  an  analysis  can  help 
determine  whether  additional  sur- 
veys are  needed,  given  the  level  of 
accuracy  desired.  For  some  ques- 
tions, levels  of  accuracy  of  20-30% 
may  be  acceptable. 

Detecting  small  burrows  of  juve- 
nile tortoises  in  transect  sampling 
can  be  particularly  difficult  even  in 
fairly  open  habitats  (Douglass  1978). 
This  problem  weakens  the  reliability 
of  transect  data  in  estimating  the 
abundance  of  juveniles.  Fourier  se- 
ries estimators  again  could  be  used, 
in  conjunction  with  an  estimate  of 


burrow  size,  to  gauge  detectability  of 
small  burrows,  but  extremely  large 
samples  are  probably  needed  to  ob- 
tain an  accurate  detectability  func- 
tion and  estimate  of  abundance  for 
smaller  tortoises. 


Point-Center  Burrow  Counts 

Tortoises  often  form  small  colonies  of 
aggregated  burrows  (McRae  et  al. 
1980),  and  H.  Mushinsky  and  E. 
McCoy  (Pers.  comm..  University  of 
South  Florida,  Tampa,  Florida)  use  a 
point-center  method  (Cottam  and 
Curtis  1956)  to  estimate  the  size  of 
tortoise  colonies.  The  approximate 
center  of  the  aggregation  of  burrows 
is  estimated,  and  the  center  point  of 
the  census  station  is  placed  there. 
The  distance  from  the  center  point  to 
several  tortoise  burrows  is  deter- 
mined, and  a  burrow  density  esti- 
mate is  derived  using  standard 
point-center  calculations  (Cottam 
and  Curtis  1956).  If  the  abundance  of 
tortoises  over  a  large  area  is  desired, 
all  aggregations  should  be  located. 

Ottier  Indirect  Estimates  of  Density 

In  some  situations  (e.g.,  intensive  col- 
ony analysis  or  preparation  for  popu- 
lation relocation),  complete  burrow 
counts  are  needed.  We  have  used 
teams  of  6  to  12  inexf)erienced  field 
assistants,  spaced  at  arm's  length,  to 


Table  1  .—Mean  burrow  density  esti- 
mates (burrows  per  ha)  and  stan- 
dard deviations  calculated  from 
transect  data  using  Fourier  series 
estimators  (D)  and  direct  computa- 
tions. Data  were  collected  at  three 
sites  in  north  Florida. 


Location 

Fourier  series 

Direct 

estimator  <D) 

computations 

Site  1 

5.3  +  0.957 

5.5  +  0.932 

Site  2  7.9 ±0.464  3.3  +  1.351 
Site  3      3.8  +  0799    3.8  +  0.873 


206 


traverse  an  area  and  search  inten- 
sively for  burrows.  Later  searches  by 
a  more  exf)erienced  researcher  did 
not  reveal  any  previously  undiscov- 
ered burrows,  except  for  a  few  cryp- 
tic hatchling  burrows. 

Trained  dogs  and  aerial  searches 
by  helicopter  (Humphrey  et  al.  1986) 
have  also  been  used  to  locate  gopher 
tortoise  burrows.  Gopher  tortoises 
often  defecate  in  or  near  their  bur- 
rows, and  a  motivated  dog  can  de- 
tect and  locate  the  resulting  olfactory 
source.  Scats  and  carcasses  are  also 
important  field  sign  used  as  indices 
of  desert  tortoise  populations  (Berry 
and  Nicholson  1984,  Woodman  and 
Berry  1984). 

Regularly  used  burrows  often 
have  several  well-defined  trails  lead- 
ing to  foraging  areas  and  other  bur- 
rows (Ernst  and  Barbour  1972).  We 
have  used  these  trails  to  find  bur- 
rows hidden  in  extremely  dense 
vegetation. 


Activity  Patterns  and  Correction 
Factors  for  Burrow  Counts 

Although  estimates  of  gopher  tor- 
toise burrow  abundance  are  rela- 
tively easy  to  collect,  calculating  the 


number  of  tortoises  associated  with 
those  burrows  can  be  difficult.  It 
seems  logical  that  the  number  of  tor- 
toise burrows  would  be  positively 
correlated  with  the  number  of  go- 
pher tortoises  in  an  area,  but  the  pre- 
cise nature  of  this  relationship  is 
poorly  understood.  Complicating 
factors  include  the  level  of  human 
disturbance,  soil  type,  and  factors 
that  influence  gopher  tortoise  activity 
patterns  (e.g.,  time  of  day,  season, 
and  weather  conditions). 

Most  researchers  have  used  a  cor- 
rection factor  of  0.614  times  the  num- 
ber of  "active"  and  "inactive"  bur- 
rows to  estimate  tortoises  abundance 
from  burrow  counts.  This  conversion 
factor  is  based  on  information  pre- 
sented in  Auffenberg  and  Franz 
(1982)  that  was  derived  from  long- 
term  data  on  the  occupation  rates  of 
122  burrows.  Burrow  activity  was 
defined  by  Auffenberg  and  Franz 
(1982)  in  the  following  manner: 

active  (burrow)  if  the  soil  of 
the  burrow  had  been  recently 
disturbed  by  the  tortoise,  inac- 
tive if  the  soil  were  undis- 
turbed but  the  burrow  ap- 
peared to  be  maintained,  and 
old  if  the  mouth  had  been 


Table  2.-- Examples  of  reported  correction  factors. 


Tortois©s/active  TortoIses/InactlveTortolses/actlve+ 

burrow  burrow  inoctiv©  Source 


•11%" 

49/103(48%)" 

33/103(32%)" 

67/124(54%) 

43/44(98%) 

4/19(21%) 

35/174(20%) 

« 

'61.5% 
9/19(47%) 
7/10(70%) 
•66,0% 


0/30(0%) 
3/16(19%) 
0/25  (0%) 
0/144(0%) 


0/225  (0%) 
0/47(0%) 


n=122,614% 


67/154(44%) 
45/60  (75%) 
4/44  (9%) 
35/318(11%) 
127/411  (31%) 

» 

9/244  (4%) 
7/57  (12%) 

10/89(11%)"* 


Auffenberg  and  Franz  (1982) 
Breininger  et  al.  (in  press) 


Burke  (pers.obs.) 
Doonan(1986) 

Fucigna  and  Nickerson  (in  press) 

Linley(1986) 
Lohoefener(1982) 
Speake  (1983) 

Spiders  and  Speake  (1986) 
Stout  et  al.  (in  press) 


'Not  reported  or  additional  details  not  reported. 
"Includes  'maybe  active'  activity  classification. 
'"Unknown  number  of  tortoises  hiad  been  harvested  prior  to  survey. 


washed  in  or  covered  with 
debris  (1982:96)  (italics  ours). 

Little  exp)erience  is  needed  to  learn 
to  make  these  distinctions,  but  differ- 
ent investigators'  classifications  may 
vary,  increasing  the  imprecision  of 
tortoise  abundance  estimates.  The 
precision  is  also  affected  by  the  activ- 
ity level  of  tortoises.  During  warm 
periods  tortoises  may  move  among 
several  burrows  during  a  day;  dur- 
ing cooler  periods  a  tortoise  may  stay 
in  a  burrow  for  several  weeks. 

R.  Stratton  (Pers.  comm.)  suggests 
that  it  is  possible  to  determine 
whether  a  burrow  is  occupied  (i.e., 
active)  by  the  direction  of  foot  tracks 
on  the  burrow  apron.  Stratton  was 
able  to  identify  correctly  14  of  15  oc- 
cupied burrows  using  this  technique, 
but  he  incorrectly  identified  19  unoc- 
cupied burrows  as  being  occupied. 

I.  J.  Stout  (Pers.  connm..  University 
of  Central  Florida,  Orlando  Florida) 
has  successfully  used  a  "sewer 
snake"  to  determine  if  a  burrow  is 
occupied.  When  extended  to  the  end 
of  the  burrow,  the  sound  of  the  end 
of  the  wire  tapping  a  tortoise  shell  is 
distinctive.       Other  methods  in- 
clude "feeling"  for  tortoises  using 
long  PVC  pipes  (Pers.  comm.,  J.  Di- 
emer,  Florida  Game  and  Fresh  Water 
Fish  Connmission  Wildlife  Research 
Laboratory,  Gainesville,  Florida)  and 
listening  for  tortoises  using  either  a 
flexible  garden  hose  (Pers.  comm., 
D.B.  Means,  Coastal  Plains  Institute, 
Tallahassee,  Florida)  or  an  electronic 
"ear"  to  amplify  breathing  sounds 
(Pers.  comm.,  D.  W.  Speake,  Ala- 
bama Cooperative  Research  Unit, 
Auburn,  Alabama). 

Several  small  twigs  stuck  verti- 
cally into  the  soil  at  the  burrow 
mouth  can  also  be  used  to  determine 
if  a  burrow  is  occupied  (Hallinan 
1923,  Beiinger  et  al.  in  press).  If  prop- 
erly spaced,  one  or  more  twigs  will 
be  knocked  over  the  next  time  a  tor- 
toise passes.  Direction  of  travel  can 
be  determined  by  uniquely  marking 
the  top  of  each  twig  (or  using  a  "Y" 
shaped  stick)  and  noting  which  di- 


207 


rection  the  twig  falls.  The  twigs  can 
be  resurveyed  1-3  days  after  place- 
ment. 

Some  recent  studies  involving  to- 
tal colony  capture  (Doonan  1986, 
Stout  et  al.  in  press,  Fucigna  and 
Nickerson  in  press,  Linley  1986,  R.L. 
Burke  unpublished  data),  using  a 
miniature  underground  television 
camera  (Burke  p>ers.  obs.,  Breininger 
et  al.  in  press,  Spillers  and  Speake 
1986)  or  other  techniques  have  pro- 
vided reliable  determinations  of  the 
number  of  tortoises  per  burrow. 
These  studies  (table  2)  have  reported 
a  wide  variation  in  the  appropriate 
correction  factor,  from  4%  of  active 
and  inactive  burrows  (Speake  1983?) 
to  75%  (Doonan  1986). 

Breininger  et  al.  (in  press)  suggest 
that  an  appropriate  correction  factor 
must  be  determined  on  a  case-by- 
case  basis.  They  recommended  that 
at  least  20  active  and  inactive  bur- 
rows be  surveyed  by  other  methods 
(e.g.,  by  camera  techniques,  trapping, 
or  by  stick  placement  at  the  mouth  of 
the  burrow)  to  establish  an  accurate 
correction  factor  for  a  site. 


Capture  Techniques 

Gopher  tortoises  spend  most  of  their 
time  in  burrows  (McRae  et  al.  1980), 
which  makes  it  difficult  to  observe  or 
capture  animals  above  ground.  It  is 
not  known  how  much  time  gopher 
tortoises  spend  in  above  ground  ac- 
tivities, but  the  congener  desert  tor- 
toise is  inactive  for  about  98%  of  its 
life  (Nagy  and  Medica  1986). 

Once  inhabited  burrows  are  lo- 
cated, tortoises  may  be  captured  and 
counted  directly  by  any  of  several 
methods.  The  methods  vary  in  terms 
of  time  and  resource  expenditures 
required  and  the  degree  to  which 
habitat  conditions  are  disturbed. 


Trapping 

Many  researchers  use  a  version  of 
bucket  trapping  similar  to  that  origi- 


nally reported  by  Agassiz  (1857). 
This  fairly  non-disruptive  technique 
involves  burying  a  smooth  sided 
plastic  bucket  (usually  a  five-gallon 
size)  immediately  in  front  of  the  bur- 
row, and  covering  the  trap  loosely 
with  a  cloth  or  a  sheet  of  heavy  pa- 
per. The  trap  is  then  disguised  with  a 
thin  layer  of  soil. 

Drainage  holes  may  be  drilled  in 
the  bottom  and  sides  to  prevent  ac- 
cumulation of  rainwater,  which  can 
drown  a  captured  tortoise.  However, 
in  extremely  hydric  soils,  traps 
should  not  have  holes  because  water 
entering  from  the  ground  can  cause 
the  same  problem. 

In  general,  traplines  should  be 
closed  down  during  periods  of  heavy 
rains.  Traps  should  be  checked  at 
least  daily,  and  during  very  hot 
weather  there  is  a  risk  of  overheating 
and  killing  captured  animals  (Burke 
1987,  Taylor  1982).  It  may  help  to 
shade  exposed  traps.  Smaller  cans 
and  containers  may  be  used  for  cap- 
turing juvenile  and  subadult  tor- 
toises. 

Bucket  trapping  is  labor  intensive, 
but  once  traps  are  in  place  they  are 
easy  to  monitor.  Up  to  forty  traps 
may  be  installed  by  an  experienced 
person  per  day,  and  over  100  traps 
can  be  checked  and  reset  if  necessary 
per  person  per  day.  We  found  that 
over  90%  of  bucket-trapped  tortoises 
were  captured  in  the  first  21  days, 
suggesting  that  three  to  four  weeks  is 
required  to  capture  nearly  all  tor- 
toises. 

These  results  are  very  similar  to 
the  results  obtained  by  J.  Diemer 
(Pers.  comm.,  Florida  Game  and 
Fresh  Water  Fish  Commission  Wild- 
life Research  Laboratory,  Gainesville, 
Florida).  An  absence  of  signs  of 
above-ground  activity  after  place- 
ment of  traps  helps  to  indicate 
whether  all  occupied  burrows  in  the 
area  have  been  located  and  trapped. 

Martin  and  Layne  (1987)  placed 
standard  live  mammal  traps  at  the 
entrance  of  the  burrow  to  capture 
tortoises.  Snares  have  also  been  used 
by  Novotny  (1986)  and  ourselves 


with  some  success.  They  may  be  set 
so  as  to  catch  the  leg  of  the  tortoise 
and  therefore  limit  possible  injury, 
though  Taylor  (1982)  describes  the 
use  of  snares  to  kill  pest  tortoises. 
Although  snares  are  inexpensive  and 
easy  to  set,  they  are  easily  evaded 
and  may  occasionally  injure  a  noosed 
animal. 

Auffenberg  (in  Plummer  1979)  and 
Recht  (1981)  described  using  me- 
chanical and  electronic  burrow-ex- 
cluding devices  to  force  tortoises  to 
remain  above  ground  after  leaving 
their  burrows.  Recht  (1981)  pointed 
out  that,  if  such  a  mechanism  was 
equipped  with  transmitting  appara- 
tus, the  tortoise  could  be  captured 
immediately. 

Deception 

"Handbobbing''  (Burke  1987,  Linley 
1986)  may  entice  tortoises  to  emerge 
from  burrows,  apparently  by  eliciting 
a  territorial  response.  This  technique 
involves  bobbing  a  clenched  fist  in 
short,  jerky  motions  at  the  mouth  of 
the  burrow,  which  is  similar  to  the 
head  bobbing  that  tortoise  engage  in 
as  part  of  social  interactions  (Auffen- 
berg 1969).  Once  a  territorial  re- 
sponse is  initiated,  tortoises  will  at- 
tempt to  push  the  intruding  hand 
from  the  burrow  and  can  be  maneu- 
vered into  a  position  to  be  extracted. 
Success  may  be  enhanced  by  striking 
the  ground  several  times  before 
handbobbing  and  by  tossing  a  small 
amount  of  soil  down  the  burrow. 
Mirrors  can  also  elicit  a  territorial 
response  (Legler  and  Webb  1961). 

A  somewhat  similar  technique, 
''tapping,"  has  been  used  to  capture 
desert  tortoises  (Medica  et  al.  1986). 
Tapping  involves  lightly  rapping  on 
the  tortoise's  shell  with  a  long  stick. 
This  procedure  would  be  difficult  to 
employ  successfully  where  burrows 
are  long  and  curved.  We  have  used 
sewer  snakes  to  probe  for  tortoises  at 
the  end  of  their  burrows,  but  we 
have  not  elicited  a  response  by  shell 
tapping. 


208 


Burrow  Excavation  and  Pulling 

Digging  up  the  entire  burrow  with  a 
backhoe  or  hand  shovel  is  both  time 
consuming  and  destructive.  At  one 
South  Florida  site,  it  took  an  experi- 
enced backhoe  operator  2.5  hours  to 
excavate  one  burrow  that  was  over 
11m  long  and  6  m  deep.  Most  bur- 
rows are  excavated  in  less  than  45 
minutes  using  a  backhoe,  which  com- 
pares favorably  to  the  approximately 
30  days  of  bucket  trapping  required 
to  remove  all  tortoises  from  an  area 
(Diemer  et  al.  in  press). 

When  excavating  a  burrow,  a 
sewer  snake  or  garden  hose  should 
be  extended  to  the  end  of  the  burrow 
to  keep  track  of  the  tunnel  path.  The 
entire  process  is  complicated  by 
loose,  sandy  soils  at  some  sites,  and 
it  is  difficult  to  retain  burrow  struc- 
ture and  avoid  potentially  dangerous 
cave-ins.  The  difficulty  of  the  process 
may  be  reduced  by  using  an  elec- 
tronic device  to  locate  the  burrow 
end  before  digging  (see  Wolcott 
1981).  Small  commensal  species  are 
likely  to  be  buried  when  a  burrow  is 
excavated  mechanically,  but  excava- 
tion by  hand  is  extremely  labor-in- 
tensive (Ernst  and  Barbour  1972). 

Taylor  (1982)  describes  the  history 
of  a  pulling  "hook"  first  reported  by 
Fisher  (1917).  It  is  the  only  simple, 
quick,  and  moderately  reliable 
method  for  capturing  tortoises,  used 
principally  by  tortoise  hunters.  Pull- 
ing requires  the  use  of  a  long  flexible 
rod  attached  to  a  short  stout  piece  of 
bent  wire.  The  apparatus  is  fed  into 
the  burrow,  maneuvered  behind  the 
tortoise,  and  wedged  between  the 
rear  of  the  plastron  and  the  flared 
carapace.  Success  rate  is  influenced 
by  a  puller's  skill  and  by  the  length 
and  curvature  of  the  burrow.  In  re- 
gions that  have  been  heavily 
"pulled"  in  the  past,  remaining  tor- 
toises are  most  often  found  in  wind- 
ing burrows  that  are  particularly  dif- 
ficult to  pull  (R.  Stratton,  Pers. 
comm.).  Taylor  (1982)  gives  details 
on  the  procedure,  as  well  as  statistics 
on  the  damage  to  captured  tortoises. 


Techniques  for  Studying  Tortoise 
Demography  and  Reproduction 

Estimates  of  Population  Structure 
Using  Burrow  Width 

Alford  (1980)  and  Martin  and  Layne 
(1987)  have  demonstrated  that  a 
simple  mathematical  relationship  ex- 
ists between  the  width  of  a  burrow 
and  the  size  of  the  resident  tortoise. 
Thus,  on  the  basis  of  a  burrow  cen- 
sus, burrow  widths,  and  a  reliable 
correction  factor,  it  is  possible  to  esti- 
mate population  size  and  evaluate 
demographic  structure  (Alford  1980, 
Sauer  and  Slade  1987).  The  relation- 
ship between  burrow  width  and  size 
of  occupant  may  be  slightly  biased, 
however,  since  small  tortoises  can 
occupy  large  burrows  but  the  ob- 
verse is  impossible. 

Marking  Techniques  and 
Determining  Sex  and  Age 

Marking  tortoise  shells  is  an  easy 
way  to  follow  the  fate  of  individuals 
over  long  periods  of  time.  Tech- 
niques for  marking  marginal  scutes 
of  turtles  have  been  reviewed  by 
Femer  (1979)  and  Plummer  (1979). 

Based  on  variation  in  the  shell  di- 
mensions of  183  adult  tortoises  of 
known  sex,  McRae  et  al.  (1981)  devel- 
oped a  discriminate  equation  that 
can  be  used  to  determine  accurately 
the  sex  of  adult  tortoises  from  north 
Florida  and  south  Georgia.  The  ap- 
plicability of  the  technique  to  tor- 
toises from  other  areas,  and  to 
smaller  size  classes,  is  untested 
(Wester  1986). 

Graham  (1979)  reviews  four  age- 
determination  techniques:  mark/re- 
capture, records  of  captive  speci- 
mens, exannination  of  long  bone  sec- 
tions, and  scute  ring  counts.  Of  these, 
only  scute  ring  counts  have  been  re- 
ported for  gopher  tortoises.  W. 
Auffenberg  (Pers.  comm.,  Florida 
State  Museum,  Gainesville,  Rorida) 
suggested  that  a  p>encil  rubbing  of 
the  plastron  was  an  accurate  way 


both  to  record  true  scute  rings  and  to 
avoid  counting  false  rings.  This  has 
been  confirmed  by  L.  Landers  (un- 
pub.  data.  Tall  Timbers  Research  Sta- 
tion, Tallahassee,  Florida).  Addi- 
tional methods  of  counting  and  re- 
cording scute  rings  are  given  by  Gal- 
braith  and  Brooks  (1987). 

Landers  et  al.  (1982)  demonstrated 
that,  in  southern  Georgia,  age  can  be 
accurately  estimated  by  carefully 
counting  plastron  scute  rings.  Ger- 
mano  and  Fritts  (in  press)  used 
mark/ recapture  data  to  show  a  high 
correlation  between  age  and  scute 
ring  counts  of  17  known-age  desert 
tortoises  (less  than  25  years  old)  from 
Nevada.  They  propose  microscopic 
examination  of  thin  scute  sections 
can  help  determine  age  of  older  tor- 
toises. However,  Berry  (in  press) 
presents  data  from  190  desert  tor- 
toises from  11  study  sites  in  which 
scute  rings  were  not  annual.  Ring 
deposition  varied  from  0  to  3  rings 
per  year.  Berry  and  Woodman  (1984) 
discuss  the  use  of  shell  wear  classes 
for  age  determination  of  adult  desert 
tortoises. 


Studies  of  Tortoise  Reproduction 

Indirect  indications  of  reproductive 
activity  include  swelling  of  the  sub- 
dentary  glands  and  recent  evidence 
of  gravidity.  Auffenberg  (1966)  and 
Rose  (1970)  suggested  that  the  sub- 
dentary  glands  produce  pheromones 
important  to  courtship  and  mating 
behavior,  and  Landers  et  al.  (1980) 
used  the  swollen  condition  of  these 
glands  in  some  captured  tortoises  as 
an  index  to  sexual  activity. 

Although  the  clutch  size  of  gravid 
tortoises  can  be  determined  by  radi- 
ography (Turner  et  al.  1986),  field 
methods  are  limited  to  palpation  and 
weight  loss.  T.  Linley  (Pers.  comm.) 
uses  palpation  to  estimate  clutch 
sizes  for  gravid  females  with  well 
calcified  eggs.  Turner  et  al.  (1986) 
also  regularly  weighed  transmittered 
desert  tortoises  and  used  sudden 
weight  loss  to  indicate  oviposition. 


209 


Given  the  fairly  predictable  nature 
of  tortoise  nest  location  (Hallinan 
1923),  it  is  surprising  that  so  few  field 
data  have  been  collected  on  nest  pre- 
dation,  nest  microclimate,  sex  of  off- 
spring, time  of  emergence,  etc. 
Auffenberg  and  Iverson  (1979)  in 
north  Florida,  and  Landers  et  al. 
(1980)  in  south  Georgia,  provide  esti- 
mates of  predation  rates  and  nest 
viability,  but  more  information  is 
needed  to  construct  accurate  esti- 
mates of  nesting  success  over  time, 
one  of  the  more  critical  portions  of 
tortoise  life  cycles  (Diemer  1984). 
Marshall  (1987)  and  Douglass  and 
Winegamer  (1977)  also  report  pre- 
liminary studies  on  nest  predation 
using  sign  at  a  small  number  of  regu- 
larly visited  nests. 

Camera  traps  may  be  particularly 
useful  in  egg  predation  studies,  al- 
lowing precise  identification  of  tim- 
ing and  predator.  R.L.  Burke  and  M. 
Noss  (pers.  obs.)  attempted  to  detect 
soil  disturbance  due  to  egg  laying  by 
burying  a  layer  of  colored  gravel  in 
46  burrow  mounds  before  oviposi- 
tion  season.  No  activity  was  de- 
tected, however.  Careful  use  of  an 
egg  probe  (Hallinan  1923)  may  facili- 
tate rapid  searching  of  large  numbers 
of  burrow  mounds  for  egg  clutches. 

Movement  Studies 

In  addition  to  studies  employing  di- 
rect observation  and  capture-recap- 
ture techniques  (e.g.,  Auffenberg  and 
Iverson  1979,  Douglass  and  Layne 
1978,  McRae  et  al.  1980,  Landers  et 
al.  1980),  various  remote  sensing  de- 
vices have  been  used  to  monitor  tor- 
toise movements. 

String  trailers  (see  Ferner  1979  and 
Plummer  1979)  have  been  used  for 
daily  movement  and  path  length 
studies  (Pers.  comm.,  W.  Auffenberg, 
Florida  State  Museum,  Gainesville, 
Fl.,  McRae  et  al.  1980).  Tortoises  too 
small  for  radio  transmitters  may  be 
tracked  using  a  metal  detector  to  lo- 
cate small  pieces  of  different  metals 
attached  to  their  shells. 


Radio  telemetry  (Legler  1979)  of 
gopher  tortoises  has  been  used  by 
Burke  (1987),  Fucigna  and  Nickerson 
(in  press),  McRae  et  al.  (1980),  Stout 
et  al.  (in  press),  J.  Diemer  (unpub- 
lished data,  Florida  Game  and  Fresh 
Water  Fish  Connmission  Wildlife  Re- 
search Laboratory,  Gainesville,  Flor- 
ida) and  others.  Radios  are  attached 
to  anterior  of  the  carapace  on  females 
(to  avoid  interference  with  copula- 
tion) and  either  the  anterior  or  poste- 
rior of  males.  Dental  acrylic  is  typi- 
cally used  to  fix  the  transmitter  on 
the  shell,  and  the  entire  device  is  cov- 
ered in  silicone  sealant  for  additional 
protection.  Other  researchers  (e.g.. 
Stout  et  al.  in  press)  have  used  ma- 
chine screws  or  wire  to  attach  the  ra- 
dio to  the  shell.  Antennae  are  usually 
glued  along  the  shell  or  left  dragging. 

Auffenberg  and  Iverson  (1979) 
used  a  series  of  microswitches  and 
sensors  buried  along,  and  extending 
into,  numerous  tortoise  burrows  to 
correlate  inner-burrow  movements 
with  microhabitat  environmental 
conditions. 


Commensal  Studies 

General  methods  for  trapping  reptile 
and  amphibian  species  are  reviewed 
by  Campbell  and  Christman  (1982) 
and  Vogt  and  Hine  (1982).  Crawfish 
frogs  may  be  seen  at  night  sitting  in 
the  mouth  of  the  burrow  (Hallinan 
1923),  and  are  sometimes  captured  in 
bucket  traps,  small  mammal  traps, 
and  funnel  traps  set  for  other  species 
(Franz  1986).  General  marking  tech- 
niques for  reptiles  and  amphibians 
are  reviewed  by  Ferner  (1979). 

Day  et  al.  (1980)  give  a  general  re- 
view of  capture  and  marking  tech- 
niques for  mammals,  birds  and  rep- 
tiles, and  Mengak  and  Guynn  (1987) 
compare  different  trapping  methods 
for  small  mammals  and  herpe- 
tofauna.  Eisenberg  (1983)  describes 
successful  placement  of  traps  for 
Florida  mice.  As  described  above, 
digging  up  the  burrow  by  hand  is  the 
only  known  way  reliably  to  capture 


all  burrow  commensals,  especially 
invertebrates.  W.  Auffenberg  (Pers. 
comm.,  Florida  State  Museum, 
Gainesville,  Rorida)  and  Milstrey 
(1986)  have  used  vacuum  systems  to 
sample  invertebrates  in  burrows. 
Milstrey  (1986)  and  Woodruff  and 
Klein  (in  prep.)  also  describe  various 
small,  baited  pitfall  traps  for  captur- 
ing invertebrates.  Butler  et  al.  (1984) 
describes  a  C02  trap  that  is  useful 
for  collecting  ticks  and  fleas. 

Vegetation  Analysis 

A  small  number  of  researchers  has 
attempted  to  characterize  gopher  tor- 
toise habitat  using  quantitative  meth- 
ods. Breininger  et  al.  (in  press), 
Marshall  (1987),  and  Wester  (1986) 
related  gopher  tortoise  densities  to 
vegetation  structure,  while  Auffen- 
berg and  Iverson  (1979)  analyzed  the 
relationship  between  tortoise  densi- 
ties and  a  single  vegetative  compo- 
nent, herbaceous  ground  cover. 
C^antitative  vegetation  sampling  has 
become  a  standard  element  in  survey 
techniques  used  for  other  groups 
(e.g.,  breeding  bird  censuses,  James 
and  Shugart  1970),  and  these  tech- 
niques should  be  more  widely  ap- 
plied to  tortoise  research. 

We  collected  vegetation  data  at  50 
m  points  as  part  of  the  transect  study 
described  above.  Percent  canopy 
cover  (trees  >  5  m),  percent  shrub 
cover,  percent  ground  cover,  percent 
wiregrass  (Aristida  stricta)  cover,  and 
the  relative  percent  of  deciduous 
trees  to  coniferous  trees  were  meas- 
ured using  methods  described  in  Cox 
et  al.  (1987).  These  five  variables 
were  selected  based  on  published 
information  about  gopher  tortoise 
habitat  preferences  (Campbell  and 
Christman  1982,  Diemer  1986),  but 
several  other  variables  could  also  be 
considered. 

A  principal  components  analysis 
was  performed  on  the  vegetation 
data  using  a  "varimax"  rotation  pro- 
cedure (Wilkinson  1980).  The  density 
(per  ha)  of  active  and  inactive  gopher 


210 


tortoise  burrows  along  each  of  the  32 
transect  segments  was  then  plotted 
against  the  transect's  vegetation 
score  on  the  first  principal  compo- 
nent axis.  This  procedure  helps 
gauge  the  degree  to  which  variation 
in  tortoise  density  along  transects 

Table  3.— Factor  loadings  for  6 
habitat  variables  measured  along 
transects.  Weightings  and  contrasts 
were  derived  from  a  "varlmax" 
principal  component  (PC)  analysis 
(Wilkinson  1983). 


Variable 

PC  1 

PC  2 

Canopy  cover 

0.809 

-0.278 

Shrub  cover 

-0.896 

0.171 

Ground  cover 

-0.832 

0.044 

Deciduous/conifer- 

ous overstory 

0.090 

0.900 

Percent  wiregrass 

0.607 

0.550 

Percent  variance 

explained  by  axis 

50.5% 

24.4% 

relates  to  variation  in  vegetation 
structure.  The  average  values  for 
vegetative  samples  recorded  along 
transects  was  used  to  compute  prin- 
cipal component  scores.  Too  few 
samples  were  collected  to  produce  a 
very  precise  evaluation  between  bur- 
row density  and  vegetation  struc- 
ture, so  the  effort  should  be  consid- 
ered only  as  an  example  of  the  appli- 
cation of  vegetation  data  collected 
along  transects. 

Principal  component  analysis  of 
vegetation  data  accurately  projected 
the  differences  we  casually  observed 
among  sites.  The  first  principal  com- 
ponent axis  explained  50.5%  of  the 
variation  among  samples  and  largely 
contrasted  decreasing  canopy  cover 
and  wiregrass  percentages  with  in- 
creasing shrub  and  ground  cover 
(table  3).  High  positive  scores  along 
this  axis  indicate  decreasing  {percent- 
ages of  canopy  cover  and  wiregrass, 
increasing  amounts  of  shrub  cover 


< 
X 

LU 

a. 


00 

o 
o: 

D 

m 

o 
> 

00 
Z 
LU 
Q 


15.0" 


10.0" 


5.0 


0.0 


0  (§) 


0 


CD 


0 


0©        Op.  0 

0   0    0  0 


0 


-1.6 


-0.2  1.2 
PC  1 


Figure  1  .—Gopher  tortoise  burrow  density  estirrvates  plotted  along  first  principal  connponent 
axis.  High  positive  scores  along  PCI  have  low  canopy  cover  and  relatively  high  levels  of 
herbaceous  ground  cover  and  shrubs. 


and  ground  cover,  and  increasing 
ratios  of  deciduous  to  coniferous 
trees.  The  second  principal  compo- 
nent axis  explained  an  additional 
24.4%  of  the  sample  variance  and  is 
weighted  by  decreasing  amounts  of 
wiregrass  cover  and  the  ratio  of  de- 
ciduous to  coniferous  trees  (table  3). 

A  plot  of  burrow  densities  against 
the  first  principal  component  shows 
a  general  trend  of  increasing  burrow 
density  with  decreasing  principal 
comp)onent  value  (fig.  1).  Areas  with 
greater  burrow  densities  generally 
had  a  lower  percentage  of  canopy 
cover,  but  higher  percentages  of 
shrub  and  ground  cover,  than  areas 
with  lower  densities.  The  regression 
line  drawn  through  the  points  has  an 
adjusted  r^  of  0.37  (p<0.05). 

Future  Directions 

Burrow-count  transects  are  efficient 
for  estimating  burrow  density,  but 
they  may  not  produce  sufficiently 
accurate  estimates  of  gopher  tortoise 
densities.  The  relationship  between 
burrow  density  and  tortoise  density 
is  poorly  understood,  and  studies 
analyzing  the  relationship  between 
burrow  occupancy  and  burrow  activ- 
ity class  are  needed  to  strengthen 
abundance  estimates.  Whether 
transects  are  appropriate  will  depend 
on  the  questions  being  addressed. 

The  combined  effects  of  variation 
in  occupancy  rates  and  variation  in 
burrow  counts  among  transects  may 
easily  produce  estimates  of  tortoise 
abundance  that  span  an  order  of 
magnitude.  For  example,  a  95-confi- 
dence  interval  for  the  density  of  ac- 
tive and  inactive  burrows  on  our  sec- 
ond study  area  (using  the  Fourier  se- 
ries estimate  from  table  1)  is  3.326- 
12.55  burrows  per  ha.  If  the  occu- 
pancy rate  of  20  active  and  inactive 
burrows  was  followed  for  a  week  on 
this  site  and  determined  to  be  0.60 
+0.20  for  any  one  day,  then  a  95-con- 
fidence  interval  for  the  estimated 
density  of  tortoises  on  the  site  could 
range  from  0.69  to  12.4  tortoise  per 


211 


ha.  Clearly  this  is  too  large  a  range 
for  some,  if  not  most,  ecological 
questions.  Many  more  transects  and 
more  precise  occupancy  rates  would 
be  needed  to  correct  these  problems. 

Fourier  series  estimators  should 
be  used  when  transects  are  con- 
ducted in  areas  with  a  dense  shrub 
component.  Some  strip-transect  esti- 
mates of  gopher  tortoise  densities  in 
thick,  scrubby  areas  may  have  under- 
estimated density.  Indeed,  Breininger 
et  al.  (in  press)  found  high  tortoise 
densities  on  areas  with  thick  shrub 
levels  that  traditionally  might  not 
have  been  considered  appropriate 
gopher  tortoise  habitat. 

Repeated  samples  of  burrow  activ- 
ity over  time  should  be  used  to  esti- 
mate site-specific  correction  factors, 
rather  than  rely  on  a  single  general- 
ized correction  factor.  This  can  be 
easily  done,  requiring  only  a  return 
visit  to  20  or  more  randomly  chosen 
burrows.  As  such  data  accumulate, 
they  may  lead  to  a  more  appropriate 
correction  factor. 

Additional  studies  of  the  commen- 
sal community  are  also  needed  since 
very  little  is  known  of  the  interac- 
tions that  occur  among  commensal 
species.  Certain  mutualistic  relation- 
ships may  be  critical  to  the  survival 
of  many  of  these  species  and  be  im- 
portant in  efforts  to  relocate  compo- 
nents of  the  burrow  community  (e.g., 
Diemer  et  al.  in  press).  Video  camera 
techniques  (Breininger  et  al.  in  press, 
Spillers  and  Speake  1986)  offer  a 
great  potential  for  investigating  bur- 
row ecology. 

Additional  studies  of  the  early  life 
cycles  of  gopher  tortoises  may  also 
be  worth  pursuing,  particularly  in 
terms  of  conducting  management  for 
this  species.  The  critical  survival  pe- 
riod in  the  gopher  tortoise  life  cycle 
occurs  during  the  first  few  years  of 
life  (Diemer  1984).  If  nesting  success 
and  hatchling  survival  can  be  effec- 
tively manipulated  through  manage- 
ment activities,  such  activities  would 
need  to  be  conducted  fairly  infre- 
quently to  enhance  population  size 
over  many  years. 


Acknowledgments 

The  authors  appreciate  the  sugges- 
tions of  W.  Auffenberg,  D.  Breinin- 
ger, R.  Franz,  L.  Landers,  J.  Layne,  H. 
Mushinsky,  I.  J.  Stout,  an  anonymous 
reviewer,  and  especially  K.  Berry  and 
J.  Diemer.  D.  Bentzein,  J.  Dudley, 
P.K.  Harpel-Burke,  T.  Linley,  M. 
Noss  and  R.  Stratton  provided  vital 
field  assistance  and  insightful  com- 
ments. J.  Layne,  B.  Woodruff  and  D. 
Wood  provided  in  press  manu- 
scripts. 

Literature  Cited 

Agassiz,  Louis.  1857.  Contributions 
to  the  natural  history  of  the  United 
States  of  America.  452  p.  Little, 
Brown,  and  Co. 

Alford,  Ross  A.  1980.  Population 
structure  of  Gopherus  polyphemus 
in  northern  Florida.  Journal  of 
Herpetology  14(2):177-182. 

Auffenberg,  Walter.  1966.  On  the 
courtship  of  Gopherus  polyphemus. 
Herpetologica  22(2):113-117. 

Auffenberg,  Walter.  1969.  Tortoise 
behavior  and  survival.  38  p.  Rand 
McNally,  Chicago. 

Auffenberg,  Walter,  and  Richard 
Franz.  1982.  The  status  and  distri- 
bution of  the  gopher  tortoise  (Go- 
pherus polyphemus).  p.  95-126.  In:  R. 
B.  Bury  (ed.).  North  American  tor- 
toises: Conservation  and  ecology. 
U.S.  Fish  and  Wildlife  Service, 
Wildlife  Research  Report  12, 
126  p.  Washington,  D.C. 

Auffenberg,  Walter,  and  John  B.  Iver- 
son.  1979.  Demography  of  terres- 
trial turtles,  p.  541-569.  In:  M.  Har- 
less  and  H.  Morlock  (eds.). 
Turtles:  Research  and  perspec- 
tives. 718  p.  John  Wiley  and  Sons, 
New  York. 

Berry,  Kristin,  in  press.  Using  growth 
rings  to  age  wild  juvenile  desert 
tortoises  (Gopherus  agassizii).  Pro- 
ceedings of  Desert  Tortoise  Coun- 
cil 1987  Symposium.  5319  Cerritos 
Avenue,  Long  Beach,  California 
90805. 


Berry,  Kristin  H.,  and  Lori  L.  Nichol- 
son. 1984.  The  distribution  and 
density  of  desert  tortoise  popula- 
tions in  California  in  the  1970's.  p. 
26-60.  In:  K.  H.  Berry,  (ed.).  The 
status  of  the  desert  tortoise  (Go- 
pherus agassizii)  in  the  United 
States.  Desert  Tortoise  Council  re- 
port to  the  U.S.  Fish  and  Wildlife 
Service,  Sacramento,  CA.  502  p.+ 
appendixes.  Order  No.  11310- 
0083-81. 

Berry,  Kristin,  and  A.  Peter 

Woodman.  1984.  Preliminary  in- 
vestigations of  shell  wear  in  deter- 
mining adult  age  groups  in  desert 
tortoises,  p.  A4/1-A4/54.  In:  K.  H. 
Berry,  (ed.).  The  status  of  the  des- 
ert tortoise  (Gopherus  agassizii)  in 
the  United  States.  Desert  Tortoise 
Council  report  to  the  U.S.  Fish  and 
Wildlife  Service,  Sacramento,  CA. 
502  p.+  appendixes.  Order  No. 
11310-0083-81. 

Breininger,  D.,  A.  Schmalzer,  D.  Ry- 
dene,  and  R.  Hinkle.  1986.  Burrow 
and  habitat  study  of  the  gopher 
tortoise  in  scrub  and  flatwood 
habitat  types.  Final  Project  Report, 
Project  No.  GFC-84-016, 313  p. 
Nongame  Wildlife  Program,  Flor- 
ida Game  and  Fresh  Water  Fish 
Commission,  Tallahassee,  Florida. 

Burke,  Russell  L.  1987.  An  experi- 
mental relocation  and  reintroduc- 
tion  of  a  gopher  tortoise  popula- 
tion. M.S.  thesis.  University  of 
Rorida,  Gainesville,  FL.  41  p. 

Burnham,  K.A.,  D.R.  Anderson,  and 
J.L.  Laake.  1980.  Estimations  of 
density  from  line-transect  sam- 
pling of  biological  communities. 
Wildlife  Monographs  72:1-202. 

Butler,  J.F.,  K.H.  Holscher,  O. 
Adeyeye,  and  E.P.  J.  Gibbs.  1984. 
Sampling  techniques  for  burrow 
dwelling  ticks  in  reference  to  po- 
tential African  swine  fever  virus 
vectors,  p.  1065-1074.  In:  D.A. 
Griffiths  and  C.E.  Bowman  (eds.). 
Acarology  IV,  Volume  2.  Ellis 
Horwood  Ltd.,  Chichester,  Eng- 
land. 1345  p. 

Campbell,  Howard  W.,  and  Steven  P. 
Christman.  1982.  Field  techniques 


212 


for  herpetofauna  community 
analysis,  p.  193-200.  In:  N.J.  Scott 
(ed.).  Herpetological  Communi- 
ties. U.S.  Fish  and  Wildlife  Service. 
Wildlife  Research  Report  13. 
Washington,  D.C.  239  p. 

Cottam,  G.,  and  J.T.  Curtis.  1956.  The 
use  of  distance  measures  in 
phytosociological  sampling.  Ecol- 
ogy. 37:451-460. 

Cox,  James.  In  press.  Demographic 
characteristics  of  small  gopher  tor- 
toise populations  and  their  pos- 
sible influence  on  relocation 
efforts.  In:  J.E.  Diemer,  D.R. 
Jackson,  J.L.  Landers,  J.N.  Layne 
and  D.A.  Wood.  Proceedings  of 
the  gopher  tortoise  relocation 
symposium.  Florida  Game  and 
Fresh  Water  Fish  Commission, 
Nongame  Wildlife  Program  Tech- 
nical Report.  Tallahassee,  FL. 

Cox,  James,  Douglas  Inkley,  and 
Randy  Kautz.  1987.  Ecology  and 
habitat  protection  needs  of  gopher 
tortoise  (Gopherus  polyphemus) 
populations  found  on  lands  slated 
for  large-scale  development  in 
Florida.  Florida  Game  and  Fresh 
Water  Fish  Commission, 
Nongame  Wildlife  Program  Tech- 
nical Report  No.  4.  Tallahassee, 
FL.  75  p. 

Day,  Gerald  I.,  Sanford  D.  Schem- 
nitz.  and  Richard  D.  Taber.  1980. 
Capturing  and  marking  wild  ani- 
mals, p.  61-88.  In:  S.D.  Schemnitz 
(ed.).  Wildlife  management  tech- 
niques manual.  686  p.  The  Wildlife 
Society.  Washington,  D.C. 

Diemer,  Joan  E.  1981.  Addendum  to 
the  bibliography  of  the  North 
American  land  tortoises:  Gopherus 
polyphemus.  p.  75-82.  In:  L.  Lo- 
hmeier  and  G.  Johnston  (eds.)  The 
future  of  gopher  tortoise  habitats. 
Proceedings  of  the  2nd  Annual 
Meeting  of  the  Gopher  Tortoise 
Council.  Florida  State  Museum, 
Gainesville,  Florida,  32611. 

Diemer,  Joan  E.  1986.  The  ecology 
and  management  of  the  gopher 
tortoise  in  the  southeastern  United 
States.  Herpetologica  42(1):  125- 
133. 


Diemer,  Joan,  Dale  Jackson,  J.  Law- 
rence Landers,  James  Layne,  and 
Don  Wood.  In  press.  Proceedings 
of  the  gopher  tortoise  relocation 
symposium.  Florida  Game  and 
Fresh  Water  Fish  Commission, 
Nongame  Wildlife  Program  Tech- 
nical Report.  Tallahassee,  FL. 

Doonan,  T.J.  1986.  A  demographic 
study  of  an  isolated  population  of 
the  gopher  tortoise,  Gopherus  pol- 
yphemus, and  an  assessment  of  a 
relocation  procedure  for  tortoises. 
M.S.  thesis.  University  of  Central 
Florida,  Orlando,  FL.  93  p. 

Douglass,  John  F.  1975.  Bibliography 
of  the  North  American  land  tor- 
toises (genus  Gopherus).  U.S.  De- 
partment of  the  Interior,  Special 
Science  Report — ^Wildlife  No.  190, 
60  p.  Washington,  D.C. 

Douglass,  John  F.  1976.  The  mating 
system  of  the  gopher  tortoise,  Go- 
pherus polyphemus,  in  southern 
Rorida.  M. A.  thesis.  University  of 
South  Florida,  Tampa,  FL.  79  p. 

Douglass,  John  F.  1977.  Supplement 
to  the  bibliography  of  the  North 
American  land  tortoises  (genus 
Gopherus).  Smithsonian  Herpeto- 
logical Information  Service,  No. 
39. 18  p. 

Douglass,  John  F.  1978.  Refugia  of 
juvenile  gopher  tortoises,  Gopherus 
polyphemus  (Reptilia,  Testudides, 
Testudinidae).  Journal  of  Herpe- 
tology  12(3):413-415. 

Douglass,  John  F.,  and  James  N. 
Layne.  1978.  Activity  and  ther- 
moregulation of  the  gopher 
tortoise  (Gopherus  polyphemus)  in 
southern  Florida.  Herpetologica 
34(4):359-374. 

Douglass,  John  F.,  and  Chester  E. 
Winegamer.  1977.  Predators  of 
eggs  and  young  of  the  gopher 
tortoise,  Gopherus  polyphemus 
(Reptilia,  Testudines,  Tes- 
tudinidae) in  southern  Florida. 
Journal  of  Herpetology  11(2):236- 
238. 

Eisenberg,  John  F.  1983.  Conserva- 
tion address:  The  gopher  tortoise 
as  a  keystone  species,  p.  1-4.  In: 
R.J.  Bryant  and  R.  Franz  (eds.). 


The  gopher  tortoise:  A  keystone 
species.  Proceedings  of  the  4th  an- 
nual meeting  of  the  Gopher  Tor- 
toise Council.  47  p.  Florida  State 
Museum,  Gainesville,  Florida, 
32611. 

Ernst,  Carl  H.,  and  Roger  W.  Bar- 
bour. 1972.  Turtles  of  the  United 
States.  The  University  Press  of 
Kentucky.  Lexington,  Kentucky. 
347  p. 

Femer,  John  W.  1979.  A  review  of 
marking  techniques  for  reptiles 
and  amphibians.  Society  for  the 
study  of  amphibians  and  reptiles. 
Herpetological  circular  No.  9.  De- 
partment of  Zoology,  Miami  Uni- 
versity, Oxford,  OH  45056. 

Fisher,  G.C.  1917.  "Gopher  pulling'' 
in  Florida.  American  Museum 
Journal.  17:291-293. 

Franz,  Richard.  1986.  The  Florida  go- 
pher frog  and  the  Florida  pine 
snake  as  burrow  associates  of  the 
gopher  tortoise  in  northern  Ror- 
ida. p.  16-20.  In:  D.R.  Jackson  and 
R.  J.  Bryant  (eds.).  The  gopher  tor- 
toise and  its  community.  Proceed- 
ings of  the  5th  Annual  Meeting  of 
the  Gopher  Tortoise  Council.  93  p. 
Rorida  State  Museum, 
Gainesville,  FL  32611 

Fucigna,  Thomas  F.  Jr.,  and  David  K. 
Nickerson,  Jr.  In  press.  Relocations 
of  two  groups  of  gopher  tortoises 
(Gopherus  polyphemus)  from  Palm 
Beach  County  to  Martin  County, 
Rorida.  In:  J.E.  Diemer,  D.R. 
Jackson,  J.L.  Landers,  J.N.  Layne 
and  D.A.  Wood.  Proceedings  of 
the  gopher  tortoise  relocation 
symposium.  Rorida  Game  and 
Fresh  Water  Fish  Commission, 
Nongame  Wildlife  Program  Tech- 
nical Report.  Tallahassee,  FL. 

Galbraith,  David  A.,  and  Ronald  J. 
Brooks.  1987.  Photographs  and 
dental  casts  as  permanent  records 
for  age  estimates  and  growth 
studies  of  turtles.  Herpetological 
Review  18(4):69-71. 

Germano,  David  J.,  and  Fritts.  in 
press.  Methods  for  age  determina- 
tion of  the  desert  tortoise,  Go- 
pherus agassizii.  Proceedings  of  the 


213 


Desert  Tortoise  Council  1987  Sym- 
posium. 5319  Cerritos  Avenue, 
Long  Beach,  California  90805. 

Graham,  Terry  E.  1979.  Life  History 
Techniques,  p.  73-95.  In:  M.  Har- 
less  and  H.  Morlock  (eds.). 
Turtles:  Research  and  perspec- 
tives. 718  p.  John  Wiley  and  Sons, 
New  York. 

Hallinan,  Thomas.  1923.  Observa- 
tions made  in  Duval  County, 
northern  Florida,  on  the  gopher 
tortoise  (Gopherus  polyphemus). 
Copeia  1923  (115):ll-20. 

Hohman,  Judy  P.,  R.D.  Ohmart,  and 
James  Schwartzmann.  1980.  An 
annotated  bibliography  of  the  des- 
ert tortoise  (Gopherus  agassizi). 
Desert  Tortoise  Council  Sp)ecial 
Publication  No.  1. 121  p. 

Humphrey,  Stephan  R.,  John  F.  Eis- 
enberg,  and  Richard  Franz.  1985. 
Possibilities  for  restoring  wildlife 
of  a  longleaf  pine  savanna  in  an 
abandoned  citrus  grove.  Wildlife 
Society  Bulletin.  13:487-496. 

Jackson,  Dale  R.,  and  Eric  R.  Mil- 
strey.  In  press.  The  fauna  of  go- 
pher tortoise  burrows.  In:  J.E.  Di- 
emer,  D.R.  Jackson,  J.L.  Landers, 
J.N.  Layne  and  D.A.  Wood.  Pro- 
ceedings of  the  gopher  tortoise  re- 
location symposium.  Florida 
Game  and  Fresh  Water  Fish  Com- 
mission, Nongame  Wildlife  Pro- 
gram Technical  Report.  Tallahas- 
see, FL. 

James,  Frances,  and  Henry  H.  Shug- 
art.  1970.  A  quantitative  method 
of  habitat  description.  Audubon 
Field  Notes  24:727-736. 

Laake,  Jeffrey  L.,  Kenneth  P. 
Burnham,  and  David  R.  Ander- 
son. 1979.  User's  manual  for  pro- 
gram TRANSECT.  Utah  State  Uni- 
versity Press.  Logan,  UT. 

Landers,  J.  Larry,  James  A.  Garner, 
and  W.  Alan  McRae.  1980.  Repro- 
duction of  gopher  tortoises  (Go- 
pherus polyphemus)  in  southwest- 
ern Georgia.  Herpetologica 
36(4):353-361. 

Landers,  J.  Larry,  W.  Alan  McRae, 
and  James  A.  Garner.  1982. 
Growth  and  maturity  of  the  go- 


pher tortoise  in  southwestern 
Georgia.  Bulletin  of  the  Florida 
State  Museum,  Biological  Sciences. 
27(2):81-110. 

Legler,  John  M.,  and  Robert  G.  Webb. 
1961.  Remarks  on  a  collection  of 
Bolsbn  tortoises,  Gopherus  fla- 
vomarginatus.  Herpetologica 
17(l):26-37. 

Legler,  Warren  K.  1979.  Telemetry,  p. 
61-72.  In:  M.  Harless  and  H.  Mor- 
lock (eds.).  Turtles:  Research  and 
perspectives.  718  p.  John  Wiley 
and  Sons,  New  York. 

Linley,  Terry  Reed.  1986.  Tortoise 
density,  age/ size  class  distribution 
and  reproductive  parameters  in  a 
central  Horida  population  of  Go- 
pherus polyphemus.  p.  21-32.  In: 
D.R.  Jackson  and  R.  J.  Bryant 
(eds.).  The  gopher  tortoise  and  its 
community.  Proceedings  of  the  5th 
Annual  Meeting  of  the  Gopher 
Tortoise  Council.  93  p.  Florida 
State  Museum,  Gainesville,  FL 
32611. 

Lohoefener,  Ren.  In  press.  Line 
transect  estimation  of  gopher  tor- 
toise burrow  density  using  a 
fourier  series.  Proceedings  of  the 
7th  Annual  Meeting  of  the  Gopher 
Tortoise  Council.  Horida  State 
Museum,  Gainesville,  FL  32611. 

Lohoefener,  Ren.  1982.  Gopher  tor- 
toise ecology  and  land-use  prac- 
tices in  southern  Desoto  National 
Forest,  Harrisn  County,  Missis- 
sippi, p.  50-74.  In:  Franz,  R.  and  R. 
Bryant  (eds.).  The  gopher  tortoise 
and  its  sandhill  habitat.  Proceed- 
ings of  the  3rd  Annual  Meeting  of 
the  Gopher  Tortoise  Council.  78  p. 
Rorida  State  Museum,  Gaines- 
ville, FL.,  32611. 

Lohoefener,  Ren,  and  Lynne  Lo- 
hmeier.  Unpublished  report.  1984. 
The  status  of  Gopherus  polyphemus 
(Testudines,  Testudinidae)  west  of 
the  Tombigbee  and  Mobile  Rivers. 
Report  to  the  U.S.  Fish  and  Wild- 
life Service  to  support  listing  this 
population  under  the  Endangered 
Species  Act  of  1973:1-116. 

Marshall,  John  E.  1987.  The  effects  of 
nest  predation  on  hatchling  go- 


pher tortoises  (Gopherus  pol- 
yphemus Daudin  1802).  M.S.  thesis. 
University  of  South  Alabama.  Bir- 
mingham, AL.  67  p. 

Martin,  Paige  L.,  and  James  A. 
Layne.  1987.  Relationship  of  go- 
pher tortoise  body  size  to  burrow 
size  in  a  south-central  Rorida 
population.  Florida  Scientist 
50(4):264-267. 

McRae,  W.  Alan,  J.  Larry  Landers, 
and  James  A.  Gamer.  1980.  Move- 
ment patterns  and  home  range  of 
the  gopher  tortoise.  American 
Midland  Naturalist  106(1):165-179. 

McRae,  W.  Alan,  J.  Larry  Landers, 
and  James  A.  Gamer.  1981.  Sexual 
dimorphism  in  the  gopher  tortoise 
(Gopherus  polyphemus).  Herpetol- 
ogica 37:46-52. 

Medica,  Philip  A.,  C.L.  Lyons,  and 
Frederick  B.  Tumer.  1986.  Tap- 
ping: A  technique  for  capturing 
tortoises.  Herpetological  Review 
17(1):15-16. 

Mengak,  Michael  T.,  and  David  C. 
Guynn,  Jr.  1987.  Pitfalls  and  snap- 
traps  for  sampling  small  mammals 
and  herptofauna.  The  American 
Midland  Naturalist  118(2):284-288. 

Milstrey,  Eric  G.  1986.  Ticks  and  in- 
vertebrates in  gopher  tortoise  bur- 
rows: Implications  and  impor- 
tance, p.  4-15.  In:  D.R.  Jackson  and 
R.  J.  Bryant  (eds.).  The  gopher  tor- 
toise and  its  community.  Proceed- 
ings of  the  5th  Annual  Meeting  of 
the  Gopher  Tortoise  Council.  93  p. 
Rorida  State  Museum, 
Gainesville,  FL.  32611. 

Nagy,  Kenneth  A.,  and  Philip  A. 
Medica.  1986.  Physiological  ecol- 
ogy of  desert  tortoises  in  southern 
Nevada.  Herpetologica  42(1  ):73- 
92. 

Novotny,  Raymond.  1986.  The  go- 
pher tortoise  at  Momingside  Na- 
ture Center:  A  public  education 
project,  p.  33-37.  In:  D.R.  Jackson 
and  R.  J.  Bryant  (eds.).  The  gopher 
tortoise  and  its  community.  Pro- 
ceedings of  the  5th  Annual  Meet- 
ing of  the  Gopher  Tortoise  Coun- 
cil. 93  p.  Florida  State  Museum, 
Gainesville,  FL.  32611. 


214 


Plummer,  Michael  V.  1979.  Collect- 
ing and  Marking,  p.  45-60.  In:  M. 
Harless  and  H.  Morlock  (eds.). 
Turtles:  Research  and  perspec- 
tives. 718  p.  John  Wiley  and  Sons, 
New  York. 

Recht,  Michael  A.  1981.  A  burrow- 
occluding  trap  for  tortoises.  Jour- 
nal of  Wildlife  Management. 
45(2):557-559. 

Rose,  F.L.  1970.  Tortoise  chin  gland 
fatty  acid  composition:  behavioral 
significance.  Comparative  bio- 
chemistry and  physiology.  32:577- 
580. 

Sauer,  J.R.,  and  N.A.  Slade.  1987. 
Size-based  demography  of  verte- 
brates. Annual  Review  of  Ecology 
and  Systematics.  18:71-90. 

Speake,  Dan  W.  1983.  Unpublished 
report.  Report  of  the  survey  of  in- 
digo snake/ gopher  tortoise  den- 
sity on  the  Margo  Study  Area  of 
Kennedy  Space  Center,  Florida. 
Alabama  Cooperative  Wildlife  Re- 
search Unit,  Auburn,  AL.  24  p. 

Spillers,  D.M.,  and  D.W.  Speake. 
1986.  Status  and  distribution  of  the 
gopher  tortoise  (Gopherus  pol- 
yphemus)  in  southern  Alabama. 
Report  to  U.S.  Fish  and  Wildlife 
Service,  Jackson,  MS.  24  p. 

Stout,  I.  Jack,  Terry  J.  Doonan,  Rich- 
ard E.  Roberts,  and  Donald  R. 
Richardson.  In  press.  A  compari- 
son of  results  of  three  gopher  tor- 
toise relocations  in  central  and 
southeast  Florida.  In:  J.E.  Diemer, 
D.R.  Jackson,  J.L.  Landers,  J.N. 
Layne  and  D.A.  Wood.  Proceed- 
ings of  the  gopher  tortoise  reloca- 
tion symposium.  Florida  Game 
and  Fresh  Water  Fish  Commis- 
sion, Nongame  Wildlife  Program 
Technical  Report.  Tallahassee,  FL. 

Taylor,  Robert  W.  1982.  Human  pre- 
dation  on  the  gopher  tortoise  (Go- 
pherus polyphemus)  in  north-central 
Florida.  Bulletin  of  the  Rorida 
State  Museum,  Biological  Sciences. 
28(4):79-102. 

Turner,  Frederick  B.,  Page  Hayden, 
Betty  L.  Burge,  and  Jan  B. 
Robertson.  1986.  Egg  production 
by  the  desert  tortoise  (Gopherus 


agassizii)  in  California.  Herpetol- 
ogica  42(1):93-104. 

Van  Home,  Bruce.  1983.  Density  as  a 
misleading  indicator  of  habitat 
quality.  Journal  of  Wildlife  Man- 
agement. 47:813-901. 

Vogt,  Richard  C,  and  Ruth  L.  Hine. 
1982.  Evaluation  of  techniques  for 
assessment  of  amphibian  and  repv 
tile  populations  in  Wisconsin,  p. 
201-217.  In:  N.J.  Scott  (ed.).  Herpe- 
tological  Communities.  U.S.  Fish 
and  Wildlife  Service.  Wildlife  Re- 
search Report  13.  Washington, 
D.C 

Wester,  Edward  Earl.  1986.  Demog- 
raphy of  two  insular  populations 
of  the  gopher  tortoise,  Gopherus 
polyphemus  (Daudin),  in  a  reser- 
voir in  southwestern  Georgia. 
M.S.  thesis.  Auburn  University, 
Auburn  AL.  57  p. 

Wilkinson,  Leland.  1985.  Systat:  The 
system  for  statishcs.  SYSTAT,  Inc. 
Evanston,  IL.  417  p. 

Wolcott,  Thomas  G.  1981.  A  radio 
probe  for  mapping  burrows  with- 
out excavation.  Journal  of  Wildlife 
Management  45(4):1069-1072. 

Woodman,  A.  Peter,  and  Kristin  H. 
Berry.  1984.  A  description  of  car- 
cass deterioration  for  the  desert 
tortoise  and  a  preliminary  analysis 
of  disintegration  rates  at  two  sites 
in  the  Mojave  Desert,  California, 
p.  A6/1-A6/31.  In:  K.  H.  Berry, 
(ed.).  The  status  of  the  desert 
tortoise  (Gopherus  agassizii)  in  the 
United  States.  Desert  Tortoise 
Council  report  to  the  U.S.  Fish  and 
Wildlife  Service,  Sacramento,  CA. 
502  p.+  appendixes.  Order  No. 
11310-0083-81. 

Wood,  Don.  1987.  Official  lists  of  en- 
dangered and  potentially  endan- 
gered fauna  and  flora  in  Florida. 
Florida  Game  and  Fresh  Water 
Fish  Conunission,  Tallahassee,  FL. 
19  p. 

Woodruff,  Robert  E.,  and  Bert  C. 
Klein,  in  prep.  A  new  baited  pitfall 
trap  design  for  beetles.  4  p. 


215 


Talus  Use  by  Amphibians  and 
Reptiles  in  the  Pacific 
Northwest^ 

Robert  E.  Herrington^ 


Abstract.— Field  data  and  a  review  of  available 
literature  v/ere  used  to  categorize  the  extent  of  talus 
usage  by  individual  herpetofaunol  species.  Five 
categories  were  recognized  that  ranged  from 
species  essentially  restricted  to  talus  slopes  to  those 
that  were  only  occasionally  observed  there.  More 
than  60%  of  the  amphibian  and  reptile  species  that 
occur  in  the  states  of  Oregon  and  Washington  were 
found  to  utilize  talus  habitats.  In  addition  to  species 
essentially  restricted  to  talus  slopes,  the  most 
frequent  use  patterns  were  to  moderate  the  effects 
of  adverse  seasonal  weather  conditions  and  the  use 
of  talus  slopes  for  reproductive  activities. 


In  recent  years,  biologists  have  em- 
phasized the  importance  of  preserv- 
ing habitats  with  high  species  diver- 
sity (EhrHch  and  Ehrlich  1981).  In 
this  context,  habitats  that  play  a  criti- 
cal role  in  the  life  cycle  of  a  large 
number  of  species  should  also  be 
considered  for  protection.  However, 
there  is  little  information  available 
concerning  habitat  utilization  by 
many  amphibian  and  reptile  species, 
and  even  less  on  the  combined  use  of 
a  single  habitat  by  both  of  these 
groups  (but  see  Scott  and  Campbell 
1982). 

Obtaining  data  on  habitat  use  of 
amphibians  and  reptiles  is  often  hin- 
dered by  the  fact  that  habitat  fidelity 
is  extremely  variable  for  these 
groups.  Most  studies  have  concerned 
eastern  species,  but  some  generaliza- 
tions have  emerged.  Small  species 
may  be  more  or  less  restricted  to  a 
single  habitat  (Ashton  1975,  Barbour 
et  al.  1969,  Fitch  1958,  Gregory  et  al. 
1987,  and  Rose  1982).  Others  rou- 
tinely occupy  two  or  more  distinctly 
different  habitats  over  a  single  sea- 
son. The  latter  group  includes  species 
that  migrate  to  reproduce  and  those 
which  use  a  separate  habitat  for  hi- 
bernation and /or  aestivation  (Brown 
and  Parker  1976,  Duvall  et  al.  1985). 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibiarts.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff, AZ,  July  19-21.  1988.) 

'Robert  E.  Herringfon  is  Assistant  Profes- 
sor of  Biology,  Georgia  Southtwestern  Col- 
lege, Americus,  GA  31709. 


Hov^ever,  the  imjx)rtance  of  a  habitat 
to  the  continued  survival  of  a  popu- 
lation is  not  necessarily  correlated 
with  the  time  that  a  species  spends 
within  it.  Providing  reproductive 
habitat,  refugia  from  adverse 
weather  conditions,  or  protection 
from  predators  can  disproportion- 
ately influence  the  role  that  a  particu- 
lar habitat  plays  in  the  ecology  of  the 
animals  that  use  it. 

Talus  slopes  are  "unique  habitats" 
(Maser  et  al.  1979),  that  represent  the 
gradual  accumulation  of  weathered 
rock  fragments  (mostly  basalt  and 
andesite)  from  a  cliff  face  (Strahler 
1981).  Individual  slopes  are  quite 
variable  in  rock  size,  aspect  and  in 
the  amount  and  type  of  vegetation 
present.  These  factors  interact  in 
complex  ways  to  provide  a  broad 
range  of  thermal  and  moisture  re- 
gimes that  amphibians  and  reptiles 
can  select.  This  study  examines  the 
use  of  talus  slopes  by  amphibians 
and  reptiles  and  compares  these 
findings  with  non-talus  areas. 

Study  Area  and  Mettiods 

Herpetofauna  associated  with  talus 
slopes  and  adjacent  non-talus  areas 
was  determined  by  field  observation 
and  a  review  of  the  literature 
(Campbell  et  al.  1982).  For  the  pur- 
pose of  this  investigation,  talus  habi- 
tats were  those  in  which  the  sub- 
strate was  predominantly  weathered 
rock  fragments  (typically  with  an  as- 


sociated cliff-face)  and  included  a  10 
meter  wide  band  of  transitional  habi- 
tat. Non-talus  habitats  were  those  in 
which  the  substrate  was  not  as  de- 
scribed above  and  were  located  a 
minimum  of  100  meters  from  a  talus 
area.  Aquatic  habitats  were  not  spe- 
cifically sampled;  however,  speci- 
mens observed  under  objects  located 
above  the  high  water  mark  were  in- 
cluded in  the  analysis. 

Field  work  was  conducted  be- 
tween August,  1981  and  August, 
1985.  During  this  period,  more  than 
100  days  were  spent  in  the  Cascade 
Mountains  of  southern  Washington 
and  northern  Oregon.  Additional 
surveys  ranging  from  2-6  days  each, 
were  conducted  in  the  North  Cas- 
cades of  Washington,  the  Coast 
Range  of  southern  Oregon,  and  the 
Wallowa  Mountains  of  northeastern 
Oregon.  A  total  of  183  individual  ta- 
lus slopes  and  adjacent  non-talus  ar- 
eas were  surveyed.  Approximately 
equal  time  was  spent  searching  talus 
and  adjacent  non-talus  habitats.  Ta- 
lus slopes  were  considered  to  have 
been  altered  by  human  activities  if 
there  was  evidence  of  extensive  rock 
or  tree  removal. 

Searches  were  conducted  by  turn- 
ing surface  debris,  raking  through 
leaf  litter,  and  in  the  case  of  talus,  by 
digging  in  the  upper  layers  of  rock 
with  a  potato  rake.  Data  recorded  for 
most  specimens  included  habitat 
type,  the  activity  the  animal  was  en- 
gaged in  when  first  observed  (active 
or  inactive,  surface  or  sub-surface. 


216 


Table  1— Talus  use  by  amphibians  and  reptiies  obsen^ed  during  this  study. 
Use  patterns  by  IndlvlduoJ  species  were:  (1)  species  generally  restricted  to 
Talus  habitats;  (2)  species  which  use  talus  areas  for  reproductive  activities; 

(3)  species  which  use  talus  areas  to  survive  adverse  weather  conditions; 

(4)  species  frequently  associated  with  talus  areas;  (5)  species  occasionally 
observed  In  talus  habitats. 


Species 


Numbers  of  individuals 
obsen/ed  In 
Taius/non-talus  habitats  talus  use  pattern(s) 


AMPHIBIANS 


Frogs 


Hyla  regilla 
Rana  aurora 
Rana  cascade 
Rana  prefiosa 
Bufo  bore  as 
Ascaphus  true} 


Salamanders 


Ambysfom  a  gracile 
Ambysfoma  macrodacfylum 
Dicampfodon  eDsafus 
Rhyacofrifon  olympicus 
Anejdes  ferreus 
Plethodon  eiongafus 
Plefhodon  dunni 
Plethodon  farselli 
Plefhodon  vehiculum 
Plefhodon  vandykei 
Plefhodon  sforrrv 
Ensafina  eschscholf^ 
Bafrachoseps  wrighfi 
Taricha  granulosa 


REPTILES 
Lizards 


Elgaria  coerulea 
Elgaria  mulficarinafa 
Eum eces  skllfonianus 

Snakes 

Confia  tenuis 
Coluber  constrictor 
Charina  boftae 
Diadophis  punctatus 
Hypsigfena  tor  quota 
Pituophis  melanoleucus 
Thamnophis  e/egans 
Thamnophis  ordinoides 
Thamnophis  sirtalis 
Crotalus  viridis 

NUMBER  OF  INDIVIDUALS 
SPECIES  RICHNESS 


3/21 
2/17 

0/6 
0/17 

2/9 
2/18 


3/8 
5/63 
8/5 
26/109 
3/8 
43/3 
123/87 
383/20 
193/146 
31/3 
19/0 
32/26 
13/8 
5/32 


19/9 
3/0 
5/3 


2/0 
10/21 
4/2 
2/2 
2/0 
8/5 
9/11 
16/3 
13/19 
28/5 

1017/686 
31/29 


3 
3 


3 
3 
2,3 
3 
5 
1 

3,4 
1,2,3 
2,3,4 
1 
1 

4,5 
4,5 
3 


3,4 
2,3,4 
4 


3,4 
3 
4 
4 
3,4 
2,3 
3 
2,3 
2,3 
2,3 


foraging  or  involved  in  reproductive 
activities),  and  a  subjective  evalu- 
ation of  the  individual's  approximate 
age  (hatchling,  juvenile,  or  adult). 
The  determination  that  an  individual 
v^as  using  talus  to  avoid  unfavorable 
weather  conditions  was  based  on  the 
season,  prevailing  weather  condi- 
tions, the  behavior  exhibited  by  the 
animal  when  uncovered,  and  the 
depth  at  which  the  specimen  was  lo- 
cated. 

These  observations  were  summa- 
rized in  an  effort  to  categorize  pat- 
terns of  talus  use.  Voucher  specimens 
of  most  species  have  been  deposited 
in  the  vertebrate  collection.  Depart- 
ment of  Zoology,  Washington  State 
University.  However,  the  majority  of 
specimens  were  identified  in  the  field 
and  released  at  the  site  of  capture. 


Results  and  Discussion 
Habitat  Use 

A  total  of  four  species  of  frogs  were 
observed  in  talus  habitats  (table  1), 
with  a  fifth  species  reported  using 
talus  areas  for  feeding  (table  2).  A 
single  Hyla  regilla  and  two  Rana  au- 
rora were  located  under  snow  cov- 
ered talus  and  were  considered  to 
have  been  hibernating  there.  All  frog 
species  were  more  numerous  in  non- 
talus  areas  and  two  species  (Rana  cas- 
cade and  R.  aurora)  observed  in  non- 
talus  areas  were  not  recorded  from 
talus  areas. 

Salamanders  were  numerically 
and  taxonomically  the  most  abun- 
dant amphibians  encountered  during 
the  study.  The  number  of  species  re- 
corded from  talus  and  non-talus 
habitats  were  14  and  13,  respectively 
(table  1).  However,  species  richness 
is  somewhat  misleading,  since  more 
than  90%  of  the  observations  of  Ple- 
thodon elongatus,  P.  larselli,  P.  stormi, 
and  P.  vandykei  were  from  talus  habi- 
tats. I  consider  these  species  to  be 
essentially  restricted  to  forested  talus 
areas.  This  observation  is  supported 
by  the  work  of  Stebbins  and  Rey- 


217 


nolds  (1947)  with  P.  elongatus,  Nuss- 
baum  et  al.  (1983)  with  P.  stormi  and 
P.  vandykei,  and  Harrington  and 
Larsen  (1985)  with  P.  larselli.  Five 
additional  species  (Dicamptodon  en- 
satus,  P.  dunni,  P.  vehiculum,  Ensatim 
eschscholtzi,  and  Batrachoseps  wrighti) 
were  observed  more  frequently  in 
talus  than  in  other  habitats  (table  1). 

All  the  salamanders  mentioned 
above  with  the  exception  of  Dicamp- 
todon ensatus,  are  capable  of  complet- 
ing their  entire  life  cycle  within  talus 
habitats.  I  observed  portions  of  the 
courtship  sequences  of  Plethodon  ve- 
hiculum and  P.  vandykei  only  on 
damp  talus.  Many  of  these  same  spe- 
cies probably  nest  in  deep  recesses 
within  the  talus.  This  is  based  on  two 
observations.  The  first  is  that  given 
the  abundance  of  some  salamander 
species,  very  few  nests  have  ever 
been  located  (Hanlin  et  al.  1979, 
Jones  and  Aubry  1985).  This  suggests 
that  nests  are  located  in  places  gener- 
ally inaccessible  to  investigators.  The 
slope  and  rock  size  associated  with 
talus  fields  generally  precludes  dig- 
ging at  depths  >  50cm  without  the 
talus  caving  in.  Secondly,  I  found 
small  aggregations  (1-3  individuals) 
of  P.  larselli,  P.  vehiculum,  and  P. 
dunni,  that  approached  the  size  re- 
ported for  hatchlings  (Stebbins  1951, 
Peacock  and  Nussbaum  1973,  Her- 
rington  1985)  only  in  loose  talus  ar- 
eas, following  the  first  fall  rains.  This 
is  the  time  that  recent  hatchlings  are 
likely  to  to  emerge  from  their  nests. 

Individuals  uncovered  from  talus 
in  situations  suggesting  that  they 
were  in  winter  dormancy  included 
Ambystoma  gracile,  A.  macrodactylum, 
Dicamptodon  ensatus,  Rhyacotriton 
olympicus,  Plethodon  dunni,  P.  larselli, 
P.  vehiculum,  and  Taricha  granulosa. 
Conversely,  between  June  and  Au- 
gust there  was  reduced  rainfall  and 
elevated  surface  temperatures 
throughout  most  of  the  study  areas. 
Because  of  this,  surface  activity  by 
salamanders  was  greatly  restricted 
and  the  majority  of  observations 
(83%)  were  of  individuals  uncovered 
from  talus  areas. 


A  total  of  5  species  of  lizards  were 
observed  or  reported  from  talus 
habitats  (tables  1  and  2).  Elgaria  coer- 
ula  was  the  most  frequently  observed 
species  and  most  individuals  were 
uncovered  from  the  upper  layers  of 
talus.  T\yo  behavioral  patterns  were 
apparent.  The  first  involved  indi- 
viduals uncovered  before  they  had 
emerged  from  nocturnal  retreats  and 
the  second  was  of  individuals  ther- 
moregulating  under  surface  talus. 
Elgaria  coerula  is  a  live-bearing  spe- 
cies and  this  behavior  may  be  impor- 
tant to  the  developmental  processes 
taking  place.  Talus  habitats  have 
been  identified  as  oviposition  sites 
for  Sceloporus  occidentalis  and  Uta 
stanshuriana  (Maser  et  al.  1979)  and 
Elgaria  multicarinata  (Brodie  et  al. 
1969).  Elgaria  coerula  and  E.  multicari- 
nata were  uncovered  from  talus 
slopes  where  they  appeared  to  be  hi- 
bernating. 

Ten  species  of  snakes  were  ob- 
served (table  1)  and  two  additional 
species  reported  from  talus  habitats 
(table  2).  Taken  as  group,  snakes 
were  most  frequently  observed  bask- 
ing either  on  the  surface  or  between 
exposed  rocks.  Species  that  I  consid- 
ered to  be  entering  or  emerging  from 


hibemacula  located  within  talus  were 
Crotalus  viridis,  Pituophis  melano- 
leucus.  Coluber  constrictor,  Thamnophis 
elegans,  T.  ordinoides,  T.  sirtalis,  Hyp- 
siglena  torquata,  and  Contia  tenuis. 
Both  Hypsiglena  torquata  and  Contia 
tenuis  were  only  observed  in  talus 
habitats  during  the  study,  but  they 
are  known  to  occupy  a  broader  range 
of  habitats  elsewhere  (Cook  1960; 
Diller  and  Wallace  1981). 

Talus  slopes  play  an  important 
role  in  the  reproductive  activities  of 
snakes.  Brodie  et  al.  (1969)  reported 
several  individuals  of  Coluber  con- 
strictor, Diadophis  punctatus,  Contia 
tenuis  and  Pituophis  melanoleucus  ovi- 
positing within  an  exposed  talus 
slope  in  Benton  Co.,  Oregon.  I  ob- 
served gravid  females  of  Thamnophis 
sirtalis,  T.  ordinoides  and  Crotalus 
viridis  basking  on  talus  slopes  during 
late  summer.  Whether  these  snakes 
delivered  their  young  at  the  talus 
slopes  is  not  known.  However, 
gravid  C.  viridis  are  known  to  remain 
in  the  vicinity  of  their  hibernacula  to 
produce  young  (R.  Wallace,  Depart- 
ment of  Biological  Sciences,  Univer- 
sity of  Idaho,  pers.  comm.),  and  I  un- 
covered 7  "yearling"  T,  ordinoides 
from  an  area  of  talus  less  than  2  m^. 


Table  2.— Amphibian  and  reptile  species  not  observed  during  this  study, 
but  which  have  been  reported  to  utilize  talus  habitats.  The  categories  of 
talus  use  are  described  In  table  1. 


Species 


Talus  use  pattern  Reference 


AMPHIBIANS 
Frogs 

Bufo  woodhousei 

REPTILES 

Lizards 

Crofaphyfus  bicincfores 
Sceloporus  occidentalis 
Ufa  sfansburiana 

Snakes 

Lampropelfls  zonafa 
Mastlcophis  taeniafus 


Maser  et  al.  (1979) 


4 

2 
2 


1 

2 


Nussbaum  et  al.  (1983) 
Maser  et  al.  (1979) 
Maser  etal.  (1983) 


Nussbaum  et  al.  (1983) 
Nussbaum  et  al.  (1983) 


218 


where  they  appeared  to  be  in  hiber- 
nation. It  was  not  possible  to  deter- 
mine if  these  snakes  had  independ- 
ently congregated  there,  or  if  they 
represented  a  single  litter  born  at  the 
talus  slope,  but  the  latter  explanation 
seems  more  plausible. 

The  importance  of  talus  slopes  in 
the  feeding  ecology  of  snakes  is  un- 
known. The  relative  abundance  of 
garter  snakes  and  salamanders  on 
talus  slopes  at  certain  times  of  the 


year  could  lead  to  predator-prey  in- 
teractions. This  is  supported  by  evi- 
dence palpated  from  the  stomachs  of 
two  Thamnophis  sirtalis  and  one  T. 
ordinoides  captured  on  talus  slopes. 
Each  of  the  T.  sirtalis  contained  a 
salamander  (1  Plethodon  dunni;  1  En- 
satina  eschscholtzi);  the  single  T.  ordi- 
noides contained  a  large  slug  {Arioli- 
max  sp.).  While  other  interactions 
were  not  observed,  small  mammals 
often  were  observed  in  talus  habitats. 


Alterations  to  Talus  Slopes 

It  became  apparent  after  the  initia- 
tion of  this  study,  that  a  large  num- 
ber of  the  talus  slopes  being  sur- 
veyed had  been  or  were  being  al- 
tered by  human  activities.  Habitat 
modifications  involved  two  not  mu- 
tually exclusive  alterations.  The  first 
was  the  removal  of  rock  from  the 
base  of  talus  slof>es  to  be  used  for 
road  construction  raw  materials  (fig. 
1).  The  second  involved  tree  removal 
(clearcutting)  from  the  talus  slopes. 

I  revisited  talus  slopes  surveyed  in 
the  early  part  of  the  project  to  deter- 
mine the  frequency  and  type  of  al- 
teration. Of  183  talus  slopes  sur- 
veyed, 106  were  altered;  76  had  no- 
ticeable quantities  of  talus  removed, 
13  had  been  deforested,  and  17  had 
been  altered  by  both  events. 

I  was  able  to  document  few  clear 
species  specific  trends  between  al- 
tered and  unaltered  talus  slopes  (see 
Conclusions).  However,  there  were 
differences  in  the  number  of  indi- 
viduals encountered.  Unaltered 
slopes  represented  42%  of  the  habi- 
tats surveyed  but  yielded  73%  of  the 
total  number  of  individuals.  Because 
there  were  differences  in  the  amount 
of  search  effort  (time)  expended  sur- 
veying altered  and  unaltered  talus 
habitats,  I  did  not  statistically  com- 
pare these  results. 

Conclusions 

Talus  slopes  provide  important  habi- 
tat for  a  significant  segment  of  the 
herpetofauna  of  the  Pacific  North- 
west. A  total  37  of  the  58  species  of 
amphibians  and  reptiles  that  occur  in 
the  states  of  Washington  and  Oregon 
are  documented  from  talus  slopes. 
Use  of  this  resource  by  amphibians 
and  reptiles  was  quite  variable,  but 
three  important  patterns  emerged. 
The  first  involves  species  essentially 
restricted  to  talus  habitats.  Four  spe- 
cies of  plethodontid  salamanders  fit 
this  pattern  (Plethodon  larselli,  P.  van- 
dykei,  P.  elongatus,  and  P.  stormi). 


219 


The  second  category  of  talus  use 
consisted  of  species  which  use  talus 
slopes  to  avoid  potentially  lethal 
temperature  extremes.  Nineteen  spe- 
cies (10  reptiles,  9  amphibians)  were 
included  here.  Several  species  of 
snakes  travel  considerable  distances 
to  congregate  at  communal  hibernac- 
ula  (Duvall  et  al.  1985,  Gregory  and 
Stewart  1975,  and  Brown  and  Parker 
1976).  This  behavior  conceivably 
could  put  an  entire  population  at  risk 
if  the  hibemacula  were  irreparably 
altered. 

A  third  use  pattern  of  talus  slopes 
was  for  reproductive  activities.  In 
addition  to  an  egg-laying  aggregation 
of  5  species  of  reptiles  reported  by 
Brodie  et  al.  (1969),  live-bearing  rep- 
tiles were  frequently  observed  in  th- 
ermoregulatory behaviors  on  and 
along  the  edge  of  talus  slopes.  The 
importance  in  this  behavior  to  com- 
pletion of  developmental  processes 
remains  to  be  determined. 

Each  of  these  utilization  patterns 
is  important  to  a  particular  segment 
of  the  herpetofaunal  community. 
Whether  or  not  the  availability  of 
suitable  talus  slopes  is  a  limiting  fac- 
tor for  any  of  these  species  remains 
unknown.  However,  talus  slopes 
typically  make  up  only  a  small  por- 
tion of  the  available  habitat.  In  the 
Gifford  Pinchot  National  Forest 
(where  a  large  part  of  this  work  was 
conducted),  Scharpf  and  Dobler 
(1985)  found  talus  slopes  to  occupy 
less  than  5%  of  the  total  land  area, 
most  other  areas  have  less. 

The  high  frequency  of  altered  ta- 
lus slopes  observed  during  this  study 
may  pose  a  significant  threat  to  the 
long-term  survival  of  many  of  the 
amphibians  and  reptiles  that  use 
them.  Talus  removal  for  road  build- 
ing materials  and  tree  removal  from 
the  slopes  initiate  complex  changes 
in  the  structure  of  the  slope.  Trees, 
through  leaf  fall,  provide  a  major  in- 
put of  nutrients  to  the  slope,  as  well 
as  increasing  the  moisture  retention 
capabilities  of  the  sub-surface  talus. 
Tree  removal  increases  the  solar  ra- 
diation reaching  the  slope  and  this 


results  in  the  rapid  loss  of  moisture 
from  the  upper  layers  of  talus.  In  a 
study  comparing  the  habitat  selection 
of  P.  larselli  and  P.  vehiculum  (Her- 
rington  and  Larsen  1985),  tree  re- 
moval was  implicated  in  rendering  a 
talus  slope  unsuitable  for  habitation 
by  P.  larselli,  but  not  for  P.  vehiculum. 

Talus  removal  results  in  a  major 
shift  of  the  slope  towards  its  base. 
This  results  in  the  extensive  move- 
ment of  both  surface  and  deep  layers 
of  talus.  The  immediate  effect  would 
be  to  kill  or  injure  many  of  the  rep- 
tiles and  amphibians  inhabiting  the 
slope  as  well  as  destroy  any  nests  lo- 
cated there.  A  long  term  consequence 
of  rock  removal  is  that  erosional 
processes  are  increased.  This  results 
in  an  increase  in  the  amount  of  soil 
present  in  the  talus,  and  could  con- 
ceivably close  off  access  and  fill  in 
areas  formerly  used  as  hibemacula. 

Management  Recommendations 

Prior  to  altering  a  particular  talus 
slope,  a  survey  should  be  conducted 
to  detennine  the  presence  of  threat- 
ened, endangered,  or  otherwise  sen- 
sitive species.  Additionally,  it  should 
be  determined  whether  or  not  the 
slope  in  question  serves  as  a  major 
snake  hibernaculum. 

Tree  removal  from  talus  slopes 
should  be  restricted  and  logging 
practices  should  be  modified  to  al- 
low for  leaving  a  sufficient  border  of 
trees  (20-30  m)  along  the  margin  of 
talus  slopes. 

Current  practices  of  removing  ta- 
lus for  road  building  materials  from 
each  slope  encountered  should  be 
discouraged.  Selected  talus  areas 
known  not  to  contain  threatened,  en- 
dangered or  sensitive  species  or  to  be 
major  snake  hibemacula  should  be 
utilized  as  a  source  of  rock  for  con- 
stmction  activities. 

One  area  that  needs  additional 
study  is  the  colonization  and  use  by 
amphibians  and  reptiles  of  artificially 
created  talus  areas.  These  would  in- 
clude areas  such  as  the  banks  of  road 


cuts  with  riprap,  and  rock  piles  asso- 
ciated mining  processes.  Those 
sampled  during  the  study  were 
found  to  have  a  depauperate  fauna 
compared  to  natural  talus  areas  and 
the  fauna  consisted  almost  entirely  of 
species  known  to  have  broad  habitat 
tolerances.  However,  the  possibility 
remains  that  with  adequate  planning, 
suitable  areas  could  be  constructed  in 
such  a  manner  to  benefit  amphibians 
and  reptile  faunas. 

Acknowledgments 

Portions  of  this  study  were  funded 
by  the  Washington  Department  of 
Game,  the  Mazamas,  the  Society  for 
the  Study  of  Amphibians  and  Rep- 
tiles, and  Washington  State  Univer- 
sity. Brian  Miller,  Chris  Davitt,  and 
Linda  Whittlesey  assisted  with  field 
work.  Comments  and  suggestions  by 
Stephen  Corn,  Patrick  Gregory,  and 
Kieth  Severson  substantially  im- 
proved this  manuscript.  Shelia  Hines 
typed  the  numerous  drafts  of  this 
manuscript.  For  all  of  this  help,  I  am 
exceedingly  grateful. 

Literature  Cited 

Ashton,  R.  E.  1975.  A  study  of  the 
movements,  home  range  and  win- 
ter behavior  of  Desmogmthus  fus- 
cus.  Joumal  of  Herpetology  9:85- 
92. 

Barbour,  Roger  W.,  M.  J.  Harvey, 
and  J.  W.  Hardin.  1969.  Home 
range,  movements  and  activity  of 
the  eastern  worm  snake,  Carphohis 
amoenus  amoenus.  Ecology  50:470- 
476. 

Brodie,  Edmund  D.  Jr.,  Ronald  A. 
Nussbaum,  and  Robert  M.  Storm. 
1969.  An  egg-laying  aggregation  of 
Oregon  Reptiles.  Herpetologica 
25:223-227. 

Brown,  William  S.  and  William  S. 
Parker.  1976.  Movement  ecology 
of  Coluber  constrictor  near  commu- 
nal hibemacula.  Copeia  1976:225- 
242. 


220 


Campbell,  R.  Wayne,  Michael  G. 
Shepard,  Brigitta  M.  Van  Der  Raay 
and  Patrick  T.  Gregory.  1982.  A 
Bibliography  of  Pacific  Northwest 
Herpetology.  Heritage  Record  No. 
14.  The  British  Columbia  Provin- 
cial Museum,  Victoria,  British  Co- 
lumbia, 152  p. 

Cook,  Sherburne  F.  1960.  On  the  oc- 
currence and  life  history  of  Contia 
tenuis.  Herpetologica  16:163-173. 

Diller,  Lowell  V.  and  Richard  L.  Wal- 
lace. 1981.  Additional  distribution 
records  and  abundance  of  three 
species  of  snakes  in  southwestern 
Idaho.  Great  Basin  Naturalist 
41:154-157. 

Duvall,  David,  M.  B.  King,  and  M.  J. 
Gutz wilier.  1985.  Behavioral  ecol- 
ogy and  ethology  of  the  prairie 
rattlesnake.  National  Geographic 
Research  1:80-111. 

Ehrlich,  Paul  R.  and  Ann  Ehrlich. 
1981.  Extinction:  The  Cause  and 
Consequence  of  the  Disappear- 
ance of  Species.  Random  House, 
New  York,  305  p. 

Fitch,  Henry  S.  1958.  Home  range, 
territories,  and  seasonal  move- 
ments of  the  vertebrates  of  the 
Natural  History  Reservation. 
Univ.  of  Kansas  Publ.,  Museum  of 
Natural  History  11:63-326. 

Gregory,  Patrick  T.,  J.  Malcolm 
Macartney  and  Karl  W.  Larsen. 
1987.  Spatial  patterns  and  move- 
ments, p.  366-395  In  Snakes:  Ecol- 
ogy and  Evolutionary  Biology. 
Richard  A.  Seigel,  Joseph  T. 
Collins,  and  Susan  S.  Novak,  edi- 
tors. Macmillan  Publishing  Co. 
New  York. 

Gregory,  Patrick  T.  and  K.  W.  Ste- 
wart. 1975.  Long-distance  disper- 
sal and  feeding  strategy  of  the  red- 
sided  garter  snake  (Thamnophis 
sirtalis  parietalis)  in  the  Interlake  of 
Manitoba.  Canadian  Journal  of 
Zoology  53:238-245. 

Hanlin,  Hugh  G.  and  Joseph  J. 
Beatty,  and  Sue  W.  Hanlin.  1979. 
A  nest  site  of  the  western  red- 
backed  salamander  Plethodon  ve- 
hiculum  (Cooper).  Journal  of  Her- 
petology. 13:214-216. 


Herrington,  Robert  E.  1985.  The  ecol- 
ogy, reproductive  biology  and 
management  of  the  Larch  Moun- 
tain Salamander  (Plethodon  larselU 
Burns)  with  comparisons  to  two 
other  sympatric  plethodons.  Ph.D. 
dissertation  Department  of  Zool- 
ogy, Washington  State  University, 
Pullman,  WA.  102  p. 

Herrington,  Robert  E.  and  John  H. 
Larsen,  Jr.  1985.  The  current 
status,  habitat  requirements  and 
management  of  the  Larch  Moun- 
tain salamander.  Biological  Con- 
servation 34:169-179. 

Jones,  Lawrence  L.  C.  and  Keith  B. 
Aubry.  1985.  Ensatina  eschscholtzi 
oregonensis:  Reproduction.  Herpe- 
tological  Review  16:26. 

Maser,  Chris,  Jon  E.  Rodiek,  and  Jack 
W.  Thomas.  1979.  Cliffs,  Talus  and 
Caves.  In  Wildlife  Habitats  in 
managed  Forests  the  Blue  Moun- 
tains of  Oregon  and  Washington. 
Jack  W.  Thomas,  Editor.  U.S.  De- 
partment of  Agriculture,  Agricul- 
ture Handbook  No.  553. 

Nussbaum,  Ronald  A.,  Edmund  D. 
Brodie,  Jr.,  and  Robert  M.  Storm. 
1983.  Amphibians  and  Reptiles  of 
the  Pacific  Northwest.  The  Univer- 
sity of  Idaho  Press,  Moscow, 
Idaho.  332  p. 

Peacock,  Robert  L.,  and  Ronald  A. 
Nussbaum.  1973.  Reproductive 
biology  and  population  structure 
of  the  western  red-backed 
salamander,  Plethodon  vehiculum 
(Cooper).  Journal  of  Herpetology 
7:215-224. 

Rose,  Barbara  1982.  Lizard  home 
ranges:  Methodology  and  Func- 
tions. Journal  of  Herpetology 
16:253-269. 

Scott,  Norman  J.  and  Howard  W. 
Campbell.  1982.  A  chronological 
bibliography,  the  history  and 
status  of  studies  of  herpetological 
communities,  and  suggestions  for 
future  research,  p.  221-239.  In  Her- 
petological Communities.  Norman 
J.  Scott,  editor.  United  States  De- 
partment of  the  Interior  Fish  and 
Wildlife  Service  Wildlife  Research 
Report  13:239  p. 


Scharpf,  Raymond  W.  and  Fred  C. 
Dobler.  1985.  Caves,  Cliffs  and  Ta- 
lus, p.  187-197.  In  Management  of 
Wildlife  and  Fish  Habitats  in  For- 
ests of  Western  Oregon  and  Wash- 
ington. E.  Reade  Brown,  Editor. 
U.S.  Department  of  Agriculture, 
Forest  Service,  Pacific  Northwest 
Region. 

Stebbins,  Robert  C.  1951.  Amphibi- 
ans of  Western  North  America. 
University  of  California  Press, 
Berkeley,  539  p. 

Stebbins,  Robert  C.  and  H.  C.  Rey- 
nolds. 1947.  Southern  extension  of 
the  range  of  the  Del  Norte  sala- 
mander in  California.  Herpetol- 
ogica 4:41-42. 

Strahler,  Arthur  N.  1981.  Physical 
Geology.  Harper  &  Row,  NY. 
612  p. 


221 


Comparison  of 
Herpetofaunas  of  a  Natural 
and  Altered  Riparian 
Ecosystem^ 

K.  Bruce  Jones^ 


Abstract.— Reptile  abundance  and  diversity  were 
greater  on  an  unaltered  riparian  ecosystem  tlian  on 
an  altered  site;  the  former  l^ad  some  species 
typically  found  on  upland  habitats  (e.g.,  chaparral) 
and  the  latter  was  comprised  of  species  from 
adjacent  Sonoron  Desert.  The  distribution  and 
abundance  of  certain  microhobitats  appear  to 
account  for  differences  in  reptile  abundance  and 
diversity  on  the  two  sites. 


Over  the  past  25  years, 
concerns   have   increased  about 
the  impacts  of  population  growth 
and   associated   development  on 
wildlife  habitats  within  the 
southwestern  United  States, 
especially  the  impacts  of 
increased  demand  for  water 
resources  within  arid  regions.  A 
series  of  long-term  studies  on  the 
Colorado  River  have  shown  that 
dam-induced    habitat  alternations 
have  reduced  overall  bird 
abundance  and  diversity  (Ohmart 
et.  al.  1977).  Most  of  the  once  wide- 
spread riparian  woodland  along  the 
Colorado  River  has  been  replaced  by 
non-native  salt  cedar  (Tamarix  spp.) 
and  shrubs  typically  found  in  inter- 
mittent drainages  (Ohmart  et.  al. 
1977).  Many  of  the  birds  requiring 
riparian  woodland  are  no  longer 
found  along  the  Colorado  River. 

Many  studies  demonstrate  how 
water  impoundments  impact  birds 
and  fish  of  riparian  and  aquatic  habi- 
tats, but  little  is  known  about  im- 
pacts on  amphibians  and  reptiles  in- 
habiting these  ecosystems.  Jones  et 
al.  (1985)  and  Jones  and  Glinski 
(1985)  found  that  a  number  of  mesic- 
adapted  or  upland  amphibians  and 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff, Arizona,  July  19-21.  1988.) 

'K.  Bruce  Jones  is  a  Researct)  Ecologist 
with  the  Environmental  Protection  Agency, 
Environmental  Monitoring  Systems  Labora- 
tory, Las  Vegas,  Nevada  89193. 


reptiles  were  restricted  entirely  to 
cotton  wood- willow  riparian  habitats 
within  the  Sonoran  Desert.  Usually 
found  in  habitats  of  the  Upper  Sono- 
ran Life-zone  (e.g..  Chaparral),  these 
species  immigrated  into  lower  eleva- 
tions (<  762  m)  of  the  Sonoran  Desert 
via  riparian  corridors  ([ones  et  al. 
1985).  Upland  species  occur  on  a  few 
riparian  sites  within  the  Sonoran 
Desert  that  have  maintained  mesic 
habitat  conditions  (Jones  et  al.  1985). 
These  conditions  persist  on  these 
sites  due  to  the  moderating  effects  of 
leaf  litter  and  logs  resulting  from  Cot- 
tonwood trees  (Populus  fremonti), 
perennial  waterflow,  shading  of  the 
surface  by  trees,  and  accumulation  of 
large  debris  piles  resulting  from  peri- 
odic flooding  (fones  and  Glinski 
1985).  In  California,  for  example,  ri- 
parian ecosystems  provide  habitat 
for  83  percent  of  the  amphibians  and 
40  percent  of  the  reptiles  known  from 
that  state  (Erode  and  Bury  1984). 

Water  impoundment  structures 
eliminate  periodic  flooding  and  sig- 
nificantly reduce  stands  of  cotton- 
woods  and  willows  (Salix  goodingii) 
along  major  drainages  (Ohmart  et  al. 
1977).  These  structures  may,  there- 
fore, significantly  reduce  mesic  con- 
ditions in  downstream  riparian  eco- 
systems. To  determine  the  possible 
impact  of  impoundment  structures 
on  the  herpetofauna  of  a  desert  ripar- 
ian ecosystem,  I  studied  two  low  ele- 
vation (<  762  m)  sites,  one  with  ma- 
jor water  impoundments  and  one 
without  any  impoundments. 


Figure  1  .—Locations  of  the  study  areas. 

Methods 

To  compare  herpetofaunas  of  an  un- 
altered vs  altered  desert  riparian  eco- 
system, I  chose  study  sites  on  the 
Hassayampa  and  Salt  Rivers.  The 
Hassayampa  River  has  no  major  wa- 
ter impoundments.  It  originates  in 
the  Bradshaw  Mountains  160  km 
north-northwest  of  Phoenix,  Arizona, 
eventually  draining  into  the  Gila 
River  approximately  80  km  south- 
west of  Phoenix.  Lower  reaches  are 
mostly  intermittent,  except  for  a  15 
km  perennial  section  near  Wick- 


222 


enburg,  Arizona.  The  study  site  was 
located  approximately  10  km  south 
of  Wickenburg  near  Palm  Lake,  a  for- 
mer resort  now  owned  by  The  Na- 
ture Conservancy,  in  a  mature  gal- 
lery-type stand  of  cotton  wood  (Popu- 
lus  fremonti)  and  willow  (Salix  good- 
ingii)  (elevation  ca.  585  m,  fig.  1). 

The  Salt  River  originates  in  east- 
central  Arizona,  flowing  southwest 
to  Granite  Reef  Dam  (approximately 
40  km  northeast  of  Phoenix)  where 
water  is  diverted  for  irrigation.  Be- 
low this  point,  the  floodplain  trav- 
erses Phoenix,  eventually  draining 
into  the  Gila  River  approximately  26 
km  southwest  of  Phoenix.  Histori- 
cally, this  river  flowed  perennially 
over  its  entire  course.  However,  sev- 
eral major  water  impoundments,  in- 
cluding dams  forming  Roosevelt, 
Apache,  d^nyon,  and  Saguaro  lakes, 
have  significantly  altered  flows  and 
consequently  physical  characteristics. 
Flows  are  regulated  by  water  re- 
leases at  dams  and  flooding  has 
nearly  been  eliminated;  significant 
flooding  has  occurred  only  when  wa- 
ter releases  from  lakes  have  been 
necessary.  Before  water  impound- 
ment, riparian  vegetation  was  mostly 
Cottonwood  and  willow,  with 
mesquite  (Prosopis  glanulosa)  occur- 


Tabl©  1  —List  of  microhabltats  measured 
at  or  around  each  pit-fall  trap  on  eacti 
river.  Frequency  equals  tt)e  percentage 
of  quarters  around  each  trap  tt^at  had  a 
certain  microhabitat. 


Soil  type  (at  trap) 
Vertical  cover 

(over  trap) 
Distance  to  leaf  litter 
Leaf  litter  depth 
Leaf  litter  frequency 
Distance  to  log 

Log  diameter 
Log  frequency 
Distance  to 

debris  heap 
Debris  heap  width 


Rock  width 
Rock  frequency 

Distance  to  tree 
Tree  height 
Tree  width  (crown) 
Tree  frequency  by 

species 
Distance  to  shrub 
Shrub  height 
Shrub  width 

(crown) 
Shrub  frequency 


Debris  heap  depth     Distance  to  grass 

Patch 

Debris  heap  Grass  height 

frequency 
^Distance  to  rock       Grass  frequency 


ring  primarily  on  vegas  adjacent  to 
the  river(Reference?).  Mesquite  and 
tamarisk  now  dominate  the  riparian 
community,  with  only  a  few  small  (< 
100  m  in  length)  sections  of  cotton- 
wood  and  willow.  The  Salt  River 
sample  site  was  at  Blue  Point,  located 
approximately  6  km  south  of  Sa- 
guaro Lake  (fig.  1).  Cottonwood,  wil- 
low, and  mesquite  trees  were  com- 
mon at  this  site,  although  cotton- 
woods  and  willows  were  not  nearly 
as  common  as  on  the  Hassayampa 
River.  Blue  Point's  tree  gallery  was 
poorly  developed  and  I  found  no  evi- 
dence of  tree  reproduction.  Substrate 
was  dominated  by  sand,  with  gravel 
bars  located  intermittently  through- 
out the  site.  Similar  to  the  Has- 
sayampa River  site,  several  small 
drainages  traversed  this  site. 

The  herpetofauna  on  each  site  was 
sampled  by  using  a  pit-fall  trapping 
grid  consisting  of  110,  double-deep 
1.4  kg  coffee  cans  placed  15  m  apart 
in  a  22  X  5  grid  trapping  configura- 
tion (1.9  ha)  (see  Jones  1987).  Covers 
were  placed  approximately  15  cm 
above  each  trap  to  reduce  loss  of  ani- 
mals due  to  desiccation  and  expo- 
sure. Traps  were  open  continuously 
between  March  and  October,  1984. 
Traps  were  checked  every  three 
days,  and  amphibians  and  reptiles 
captured  in  traps  were  measured 
(snout- vent  length,  SVL),  weighed, 
sexed,  uniquely  marked,  and  re- 
leased into  cover  nearest  to  the  cap- 
ture site. 

While  traveling  between  pit-fall 
traps,  I  recorded  observations  of  all 
frogs,  toads,  lizards,  and  snakes.  I 
also  flipped  rocks  and  logs  to  un- 
cover hidden  herpetofauna. 

In  order  to  determine  amphibian 
and  reptile  composition  in  adjacent 
Sonoran  Desert,  a  modified  array  pit- 
fall trapping  method  was  used  (Jones 
1987).  Five  arrays  were  placed  in 
Sonoran  Desert  habitat  adjacent  to 
each  site,  and  I  checked  these  arrays 
for  animals  whenever  I  checked  the 
main  grids. 

A  point-center  quarter  (plotless) 
sampling  method  (Muller-Dumbois 


and  Ellenberg  1974)  provided  data  to 
characterize  microhabitats  around 
each  trap.  Each  trap  was  a  center 
point  for  quantifying  density  and  fre- 
quency of  microhabitats  within  7m  of 
each  trap.  I  sampled  110  points  or 
440  quarters  on  each  site.  Microhabi- 
tat frequency  was  determined  by  di- 
viding the  number  of  quarters  that  a 
microhabitat  occurred  in  (7  m  or  less 
from  the  trap)  by  the  total  number  of 
quarters  (440).  I  also  estimated  size 
(width,  height,  and  depth)  of  each 
microhabitat  and  frequency  of  can- 
opy cover  as  the  percentage  of  pit- 
fall traps  that  were  covered  by  vege- 
tation (table  1). 

Relative  abundance  equaled  the 
number  of  an  individual  species 
trapped  during  a  24-hour  period.  I 
estimated  the  diversity  of  herpe- 
tofaunas  and  microhabitats  on  each 
site  using  a  modified  Shannon- 
Weaver  diversity  index  (H')  (Hair 
1980):  H'  =  fiPjdogjoP,),  where  s  = 
number  of  species  and  pj  =  the  pro- 
portion of  the  total  number  of  indi- 
viduals consisting  of  the  i^j^  species.  I 
used  a  Student's  t-test  to  determine 
differences  between  herpetofaunas 
and  microhabitats  on  the  two  sites. 
Finally,  I  compared  herpetofaunas  of 
the  two  riparian  sites  and  adjacent 
Sonoran  Desert  by  calculating  Jac- 
card  Similarity  Coefficients  and  then 
clustered  them  using  an  unweighted 
pair  group  average  (Pimental  1979). 


Results 

Microhabitats 

The  Hassayampa  River  had  greater 
amounts  and  diversity  of  microhabi- 
tats than  the  Salt  River  (table  2).  Of 
these  differences,  the  frequency  of 
downed  litter  on  the  two  sites  was 
the  greatest  (table  2).  Leaf  litter  was  3 
times  more  common,  debris  heaps  10 
times  more  common,  and  logs  and 
limbs  twice  as  common  on  the  Has- 
sayampa River  than  on  the  Salt  River 
(table  2).  Rock  substrate  and  grasses 
were  more  common  on  the  Has- 


223 


sayampa  River  and  shrubs  on  the 
Salt  River  (table  2).  Trees  were  com- 
mon on  the  Hassayampa  River  and 
sand  substrate  on  the  Salt  River,  al- 
though neither  of  these  differences 
were  significant  (table  2).  In  addition, 
average  leaf  litter  depth  was  signifi- 
cantly greater  on  the  Hassayampa 
River  than  on  the  Salt  River  (table  2). 

Of  the  specific  types  of  canopy 
covering  pit-fall  traps,  trees  were  by 
far  the  most  common  on  both  rivers, 
although  the  Salt  River  had  more  pit- 
fall traps  with  no  canopy  cover  (fig. 
2). 

Tree  composition  varied  consid- 
erably between  sites.  The  Has- 
sayampa River  had  more 
cottonwoods  (Populus  fremonti)  and 
willows  (Salix  goodingi)  and  the  Salt 
River  more  salt  cedars  {Tamarix  spp.) 
(fig.  3).  Mesquite  (Prosopis  glandulosa) 
was  the  most  common  tree  on  both 
sites  (fig.  3). 

The  Hassayampa  River  had  more 
trees  in  the  0-1.9, 5.0-9.9,  and  10.0- 
14.9  m  height  ranges,  but  most  at  the 
Salt  River  were  in  the  2.0-4.9  m  range 
(fig.  4).  Cottonwood  height  distribu- 
tion was  relatively  even  on  the  Has- 
sayampa River,  but  most  Salt  River 
cottonwoods  were  greater  than  10  m, 
with  none  less  than  5  m,  hence  no 
reproduction  (fig.  5). 

Herpetofaunas 

The  abundance  and  diversity  of  her- 
petofauna  was  greater  on  the  Has- 
sayampa River  than  on  the  Salt 
River.  The  Hassayampa  River  had 
nearly  twice  as  many  species,  more 
than  twice  the  number  of  individu- 
als, and  a  greater  species  diversity 
(1.05  vs.  0.86)  than  the  Salt  River  (fig. 
6). 

All  but  three  species  (Bufo  mi- 
croscaphus  x  woodhousei,  B.  punctatus, 
and  Cnemidophorus  tigris)  were  more 
abundant  on  the  Hassayampa  River, 
and  this  site  had  five  "upland" 
species  (Cophosaurus  texanum,  Diado- 
phis  punctatus,  Eumeces  gilberti,  Masti-- 
cophis  bilineatus,  and  Tantilla 


hobartsmithii)  usually  found  in  habi- 
tats of  the  Upper  Sonoran  Life-zone 
(e.g.,  chaparral).  These  upland  spe- 
cies were  absent  from  the  Salt  River 
and  adjacent  Sonoran  Desert  (table 
3).  C.  tigris  had  the  same  abundance 
on  both  rivers,  C.  tigris  was  the  most 
abundant  sp)ecies  on  the  Salt  River, 
and  E.  gilberti  was  the  most  abundant 


species  on  the  Hassayampa  River 
(table  3).  The  Hassayampa  River  also 
had  4  species  with  abundances 
greater  than  1.0,  whereas  the  Salt 
River  only  had  one  (table  3). 

A  cluster  analysis  of  Jaccard  Simi- 
larity Coefficients  using  data  in  table 
3  revealed  that  the  Salt  River  riparian 
site  had  a  herpetofauna  more  similar 


labia  2*— Co  ^btiot  uDundu  -^en  the  Salt  und 

l^assayampa  .v  ^..^^  Is  ihe  m^an  rw.HK.^*  n^^arters  !n  which  a 

tmitcfohablfcsf  Was  found  csrcmd  each  frcp  <v'?thin  7  m)  +  SD 


Mlcrohab8at 


$o\i  River 


Hassoyampa  River  $lgnlficar>t 

Difference 
'p<m 


Scald  substfot 

(4ie) 

<209) 
<326) 

lectf  litter 

Logs/downec  tree  (fnrtbs 

(^65) 

<407) 

Debris  heaps 

<123 

Trees 

1.9±DJ 

(213) 

<315 

Shajbs 

(297) 

<275, 

Grass 

0.4±0,3 

(48) 

<12S) 

IVticrohdDitot  diversity  (H') 


J7 


CO 

Q. 

O 


a> 
o 

a> 

Q_ 


Shrub  Canopy 
Tree  Canopy 
Shrub/Tree  Canopy 
Open  Conopy 


Salt  River  Hcssayompo  River 

Canopy  Type 

Figure  2.— Comparison  of  canopy  types  on  the  Salt  and  Hassayampa  Rivers. 
224 


100 


80  - 


-2  70 


60 
50  - 
W  h 
30 
20 
10 
0 


E2L 


Cottonwoods 

MIon 

M«s(|uite 

ToTHinsk 

Other 


Salt  FSver  Hossoyompa 
Tree  Types 


1^ 


Figure  3.— Comparison  of  tree  composition  on  the  Salt  and  Has- 
soyampa  Rivers. 


o 


100 
90 
80 
70 
60 
50 
40 
X 
20 
10 
0 





■  0.0  -  1.9  m 

2.0  -  4.9  m 
^  5.0  -  9.9  m 
10.0  -  14.9  m 
>  15.0  m 


Solt  River  Hossoyompo 
Size  Classes 


River 


Figure  4.— Comparison  of  tree  hteight  distribution  on  ttie  Salt  and 


100 


90  - 
80  - 
70  - 
60 

50  I- 
40 

30  h 
20 
10 

0 


H  0.0  -  1.9  m 

2.0 -4.9  m 
\/'^  5.0  -  9.9  m 

I  1  10.0  -  14.9  I 

^8881  >  15.0  m 


Sott  Rivw  Hossoyompo 

Tree  Height  Glosses 


River 


8 

1 

20 

bcrof ! 

IS 

E 

10 

5 

0 

3c, 

1 

Spec 

a75 

0.5 

025 

0 

SaHRw 


Figure  6,— Comporison  of  ttie  total  number  of  amphibians  and  rep- 
tiles, total  relative  abundance,  and  species  diversity  on  the  Salt 
and  Hassayampa  Rivers. 


3,^<wtjpori«)ft  of  »2ard  cft>undance  and  diversity  be- 
tween the  Salt  stir.  Hdssoyqmpo  R|v^.  Abundance  is  the  number 
of  8tardsccttight/Qrld/24  Ytotttt.  Aitiphlblans  and  reptiles  occu- 
pying adjacent  Sonofon  Desert  habllals  also  are  Indicated, 


Figure  5.— Comparison  of  size  classes  of  cottonwoods  on  the  Salt 
and  Hassayampa  Rivers. 


Spe<>ies 


Salt 


Nassayampo 

Rivef 


Sonoian 
Pe$ert 


Cafkaurus  draconotd^s 
Cmmidophoras  ifgfis 
■X^ofeonyx  voftegaius 
ieptofypfitcm  hm)i{($ 

ijrpsautu^  orn^tus 

Bufo  mkifoicuphui  x 

Bufd  puncfatw 
ScapNopu$  couchi 
Cophosourus  t^xonum 
Dkfdophf^  punckrfus 
fl^^Ss  gUberff 

Smofn  $^mharnJafa 
TantSia  hobaffsmma 
Bufo  ofvorius 


0,27 
147 

0.07 
D.47 

()q7 

DM' 

&7 


XX 


0,60^ 
1,47 

0. 03 
OW 

1.  W 

0,23 

0,40 

0<63 

0  20^ 

0<TO* 

2,17* 

0,07* 

0,07* 

o,m* 

0,37* 


X 
X 
X 


X 
X 


X 

X 


X 

X 


*S}gnfifcanf}ygf&:3fef  a}:?Undance  (ifp<  iJ5, 

X  V^m^  In  <3C^Qc:^nf  ^ondrati  DesGrf  habifats  viCj  ptf- 
fdU  trapping. 

XX  V^ifiBtd  on  th&  Satt  River  $ft&  vio  field  s&atch. 


225 


to  adjacent  Sonoran  Desert  than  to 
the  herpetofauna  of  the  Hassayampa 
River  riparian  site,  although  the  two 
riparian  herpetofaunas  were  rela- 
tively similar  (fig.  7). 

Discussion 

The  distribution,  abundance,  and  di- 
versity of  herpetofauna  on  the  Salt 
River  correlate  with  impoundment- 
induced  changes  in  microhabitats. 
On  the  unaltered  riparian  ecosystem 
on  the  Hassayampa  River,  many  mi- 
crohabitats were  more  abundant  and 
diverse  than  on  the  Salt  River,  espe- 
cially surface  litter  and  trees.  These 
differences  in  microhabitats  correlate 
with  differences  in  species  diversity 
and  abundance  on  the  two  rivers. 
Species  that  were  most  abundant  on 
the  Hassayampa  River  (Eumeces 
gilberti,  Sceloporus  magister,  and  Uro- 
saurus  omatus)  prefer  sites  with 
downed  vegetative  litter  and  vertical 
structure  (e.g.,  trees)  (Jones  and 
Glinski  1985,  Jones  1986).  These  rep- 
tiles were  not  nearly  as  common  on 
the  Salt  River  and  this  may  result 
from  lower  surface  litter  and  vegeta- 
tion structure  (higher  percentage  of 
salt  cedar,  Tamarix  spp.,  and  a  lower 
percentage  of  cotton  woods,  Populus 
fremonti,  and  willows,  Salix  goodingii) 
on  this  site. 

The  greatest  difference  between 
herpetofaunas  on  the  two  rivers  was 
presence  of  five  upland  species  on 
the  Hassayampa  River  and  the  ab- 
sence of  these  species  on  the  Salt 
River.  Jones  and  Glinski  (1985)  sug- 
gested these  species  occur  in  riparian 
habitats  within  low  elevation  Sono- 
ran Desert  because  of  the  moderating 
effects  of  certain  microhabitats,  espe- 
cially surface  litter  and  debris  heaps. 
Surface  litter  and  debris  heaps  are 
considerably  less  common  on  the  Salt 
River,  and  this  probably  accounts  for 
the  lack  of  any  upland  species  in  this 
river's  herpetofauna.  Szaro  et  al. 
(1985)  suggest  that  debris  heaps  are 
the  principal  source  of  food  and 
cover  for  Thamnophis  elegans,  and 


that  grazing-caused  reduction  in  this 
n\icrohabitat  caused  decline  of  this 
snake  in  a  high  elevation  riparian 
community. 

The  relatively  low  amounts  of  sur- 
face litter  and  lack  of  smaller  size 
classes  of  trees  (especially  cotton- 
woods  and  willows)  on  the  Salt  River 
appear  to  result  from  dam-induced 
changes  in  water  flow  and  flooding. 
Periodic  flooding  is  essential  in  the 
long-term  maintenance  of  southwest- 
ern U.S.  riparian  ecosystems  (Brady 
et  al.  1985).  Flooding  also  provides 
the  physical  mechanism  by  which 
large  debris  piles  are  built  (Jones  and 
Glinski  1985).  Water  impoundment 
structures  on  the  Salt  River  appear  to 
prevent  flooding  regimes  necessary 
to  maintain  cottonwood  reproduc- 
tion and  debris  piles. 

Over  the  past  10  years,  the  major 
emphasis  in  riparian  management 
has  been  to  manage  trees,  particu- 
larly cottonwoods.  Several  tech- 
niques, such  as  planting  live  trees 
and  tree  poles,  have  been  used  on 
drainages  with  major  water  im- 


Similarity 


poundment  structures  to  improve 
reproduction  and  survival  of  cotton- 
woods  (Swenson  and  Mullins  1985). 
Although  these  techniques  generally 
increase  nesting  habitat  for  birds, 
they  do  not  provide  enough  surface 
litter  to  support  litter-dwelling  spe- 
cies, such  as  upland  herpetofauna. 
Szaro  and  Belfit  (1986)  studied  a  arti- 
ficially created  stand  of  riparian 
vegetation  on  Queen  Creek  in  south- 
central  Arizona.  This  stand  of  mostly 
willows  resulted  from  accumulation 
of  water  behind  a  dike.  Although  the 
stand  emulated  vegetation  structure 
of  natural  riparian  sites,  it  had  a 
depauperate  herpetofauna,  even  after 
20  years. 

This  study  suggests  surface  litter 
is  important  in  determining  abun- 
dance and  diversity  of  herpetofaunas 
in  riparian  communities.  If  we  are  to 
conserve  riparian  ecosystems,  we 
must  increase  our  emphasis  on  pro- 
tecting all  habitat  components,  in- 
cluding microhabitats  such  as  surface 
litter.  Like  the  Salt  River  site,  riparian 
areas  will  loose  litter-dwelling  and 


0     .2     .4     .6     .8  1.0 


Sonoran  Desert 


Salt  River 


Hassayampa  River 


Figure  7.— Dendrogram  comparing  herpetofaunas  of  the  Sonoran  Desert  and  Salt  and  Has- 
sayampa Rivers. 


226 


mesic-adapted  species  unless  we 
consider  these  other  components. 

Acknowledgments 

I  thank  Pattie  Glinski,  Scott  Belfit, 
Richard  GUnski,  Chuck  Hunter,  John 
McConnaughey,  Dan  Abbas,  and  my 
son  Justin  Jones  for  helping  with  data 
collection.  Special  thanks  to  Dan 
James,  Mike  Bender,  James  P. 
Collins,  and  David  J.  Germano  for 
review  of  this  manuscript. 

Literature  Cited 

Brady,  Ward,  David  R.  Patton,  and 
Jay  Paxson.  1985.  The  develop- 
ment of  Southwestern  riparian 
gallery  forests.  U.S.  Forest  Service 
Gen.  Tech.  Report  No.  RM-120.  p. 
39-43. 

Brode,  John  M.  and  R.  Bruce  Bury. 
1984.  The  importance  of  riparian 
systems  to  amphibians  and  rep- 
tiles, p.  30-36.  In:  R.E.  Warner  and 
K.M.  Hendrix  (eds.),  California 
riparian  systems:  ecology,  conser- 
vation, and  productive  manage- 
ment. Univ.  California  Press, 
Berkeley.  1035  p. 
Germano,  David  J.  and  C.  Roger 
Hungerford.  1981.  Reptile  population 
changes  with  manipulation  of  Sono- 
ran  Desert  shrub.  Great  Basin  Nat. 
41(1):129-138. 
Hair,  Jay  D.  1980.  Measurements  of 
ecological  diversity,  p.  269-275.  In: 
S.D.  Schemnitz  (ed.).  Wildlife 
Management  Techniques  Manual. 
The  Wildlife  Society,  Washington, 
D.C. 

Jones,  K.  Bruce.  1986.  Amphibians 
and  reptiles,  p.  267-290.  In:  A.Y. 
Cooperrider,  R.J.  Boyd,  and  H.R. 
Stuart  (eds.).  Inventory  and  moni- 
toring of  wildlife  habitat.  U.S.  Bu- 
reau of  Land  Man.,  Denver,  Colo- 
rado xviii.  858  p. 

Jones,  K.  Bruce  and  Patricia  C. 

Glinski.  1985.  Microhabitats  of  liz- 
ards in  a  southwestern  riparian 
community,  p.  355-358.  In:  R.  Roy 


Johnson  et.  al..  Riparian  ecosys- 
tems and  their  management:  rec- 
onciling conflicting  uses.  First 
North  American  riparian  confer- 
ence. Rocky  Mountain  Forest  and 
Range  Experimental  Station,  Gen- 
eral Technical  Report  Number 
RM-120.,  Fort  Collins,  Colorado. 

Jones,  K.  Bruce,  Lauren  P.  Kepner, 
and  Thomas  E.  Martin.  1985.  Spe- 
cies of  reptiles  occupying  habitat 
islands  in  western  Arizona:  a  de- 
terministic assemblage.  Oecologia 
66:595-601. 

Mueller-Dombois,  Dieter  and  Heinz 
Ellenberg.  1974.  Aims  and  meth- 
ods of  vegetation  ecology,  p.  110- 
118.  John  Wiley  and  Sons,  New 
York. 

Ohmart,  Robert  D.,  Wayne  O.  Dea- 
son,  and  C.  Burke.  1977.  A  ripar- 
ian case  history:  the  Colorado 
River,  p.  35-46.  In:  Importance, 
preservation  and  management  of 
riparian  habitat:  a  symposium. 
U.S.  Forest  Service  Gen.  Tech.  Re- 
port RM-43,  Fort  Collins,  CO. 

Pimental,  Roger  A.  1979.  Morphom- 
etries: the  multivariate  analysis  of 
biological  data.  Kendall/Hunt 
Publ.  Co.,  Dubuque,  Iowa. 

Swenson,  E.A.  and  Charles  L. 

Mullins.  1985.  Revegetating  ripar- 
ian trees  in  Southwestern 
floodplains.  p.  135-138.  In:  R.  Roy 
Johnson  et.  al..  Riparian  ecosys- 
tems and  their  management:  rec- 
onciling conflicting  uses.  First 
North  American  riparian  confer- 
ence. Rocky  Mountain  Forest  and 
Range  Experimental  Station,  Gen- 
eral Technical  Report  Number 
RM-120.,  Fort  Collins,  Colorado. 

Szaro,  Robert  C.  and  Scott  C.  Belfit. 
1986.  Herpetofaunal  use  of  a  des- 
ert riparian  island  and  its  adjacent 
scrub  habitat.  J.  Wildl.  Man. 
50(4):752-761. 

Szaro,  Robert  C,  Scott  C.  Belfit,  and 
J.  Kevin  Aitkin.  1985.  Impact  of 
grazing  on  a  riparian  garter  snake, 
p.  359-363.  In:  R.  Roy  Johnson  et. 
al..  Riparian  ecosystems  and  their 
management:  reconciling  conflict- 
ing uses.  First  North  American 


riparian  conference.  Rocky  Moun 
tain  Forest  and  Range  Exf)erimen 
tal  Station,  General  Technical  Re- 
port Number  RM-120.,  Fort 
Collins,  Colorado. 


227 


Critical  Habitat,  Predator 
Pressures,  and  the 
IVlanagement  of  Epicrates 
monensis  (Serpentes: 
Boidae)  on  the  Puerto  Rico 
Banl<:  A  IVIultivariate  Analysis^ 


Abstract.— Ep/crofes  monensis  is  a  endangered 
boa  endemic  to  the  Puerto  Rico  Bank.  Principal 
components  anaiysis,  based  on  data  collected 
during  five  years  of  study  and  200  captures  of  this 
species,  was  used  to  identify  predator,  prey,  and 
habitat  variables  critical  to  survival  of  the  snake. 
Management  recommendations  ore  discussed. 


Peter  J.  Tolson^ 


Epicrates  monensis  is  a  small  (ca.  <  1 
m  snout-vent  length)  senni-arboreal 
boid  snake  (fig.  1)  that  exhibits  an  ex- 
tremely disjunct  distribution  on  the 
Puerto  Rico  Bank.  The  Mona  boa  (E. 
m.  monensis)  is  endemic  to  Isla  Mona, 
a  large  island  in  the  Mona  Passage 
between  Hispaniola  and  Puerto  Rico 
(Schmidt  1926).  The  other  subspecies, 
the  Virgin  Islands  boa  (E.  m.  granti), 
is  found  on  scattered  islands  and 
cays  from  La  Cordillera  eastward 
through  the  Virgin  Islands,  including 
St.  Thomas,  Tortola,  and  Virgin 
Gorda  (Shill  1933;  Nellis  et  al.  1984; 
Mayer  and  Lazell  1988).  The  boa  is 
apparently  absent  from  Puerto  Rico 
and  the  other  large  islands  on  the 
bank.  Judging  from  the  present  dis- 
tributions, the  historical  range  of  Ep- 
icrates monensis  encompassed  virtu- 
ally the  whole  length  of  the  Puerto 
Rico  Bank.  Today,  unfortunately,  the 
snake  is  endangered  (USFWS  1980) 
and  absent  from  far  more  islands  on 
the  bank  than  it  is  resident — doubt- 
less the  result  of  a  long  history  of  ex- 
tirpation. It  is  improbable  that  the 
decline  of  the  boa  can  be  traced  to  a 
single  causative  factor;  more  likely 
the  survival  of  the  snake  at  certain 
localities  is  due  to  a  complex  series  of 
biotic,  environmental,  and  stochastic 

'Paper  presented  at  symposium.  I^an- 
agement  of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff, AZ  July  19-21,  1988.) 

'Peter  J.  Tolson  is  Curator  of  Amphibians 
and  Reptiles,  Toledo  Zoological  Society. 
2700  Broadway,  Toledo,  OH  43609. 


interactions.  The  rarity  of  the  snake 
has  made  habitat  analysis  difficult; 
one  cannot  define  critical  habitat  if 
the  snake  cannot  be  observed.  Prior 
to  my  work,  fewer  than  13  specimens 
of  the  boa  had  been  encountered,  and 
habitat  descriptions  were  largely  an- 
ecdotal with  no  attempts  to  quantify 
those  factors  important  in  determin- 
ing population  levels  (Div.  of  Fish 
and  Wildlife,  USVI  1983;  USFWS 
1984, 1986). 

The  parameters  dictating  the  dis- 
tribution and  abundance  of  animal 
species  within  a  habitat  are  often  di- 
verse. They  include  not  only  the 
physical  structure  of  habitat,  such  as 
vegetational  composition  and  spatial 
heterogeniety  (Rotenberry  and  Wiens 
1980),  but  also  species  composition 
(Matthews  1985;  Moulton,  1985)  and 
other  aspects  of  community  structure 
which  are  less  easily  defined,  such  as 
competition  (Cody  1974)  or  preda- 
tion  pressure.  In  the  West  Indies, 
particularly  on  the  Puerto  Rico  Bank, 
utilization  of  a  particular  habitat  by 
the  endemic  herpetofauna  is  not  only 
dependent  on  the  structural  attrib- 
utes of  vegetative  cover  and  the  com- 
position of  the  endemic  animal  com- 
munities, but  also  on  the  number  and 
severity  of  feral  and  exotic  animal 
introductions  that  have  occurred. 
Colonizations  (accidental  or  other- 
wise) of  the  roof  rat,  Rattus  rattus,  the 
house  cat,  Felis  catus,  and  the 
mongoose,  Herpestes  auropunctatus, 
have  profoundly  influenced  the  sur- 
vival and  distribution  of  endemics  on 


Figure  1  .—Epicrates  monensis  granti. 
Above— adult  female,  Cayo  Diablo,  Puerto 
Rico.  Below—juveniles  bom  at  the  Toledo 
Zoological  Gardens  14  July  87. 

the  Puerto  Rico  Bank  (Barbour  1917, 
1930;  USFWS  1986;  Div.  of  Fish  and 
Wildlife,  USVI  1983). 

Principal  components  analysis 
(PCA)  is  a  multivariate  statistical 
technique  that  has  been  used  by  com- 
munity ecologists  to  model  distribu- 
tions of  animal  populations  in  a 
multidimensional  habitat  space  de- 
fined by  a  correlation  matrix  of  habi- 
tat variables  (See  Wiens  and  Roten- 
berry 1981  and  Matthews  1985).  My 
current  work  with  Epicrates  monensis 
utilizes  PCA  to  correlate  the  abun- 


228 


A 


Cayo  fcacos 


\ 


Cayo  Lobos 


La  Cordillera 


t8'19' 


Cayo  Diablo 
1 


N  i  1 

1  km 


Figure  2.— Location  of  sampiing  plofs  in  Puerto  Rico.  Above— plots  on  Isia  Mono.  Beiow— plots 
on  La  Cordillera. 


dance  of  the  boa  with  certain  critical 
elements  of  habitat  structure  and  in- 
dices of  p>opulation  densities  of  pre- 
ferred prey  species  and  predators. 
Compilation  of  such  data  is  ex- 
tremely important  in  establishing  the 
critical  dimensions  of  the  boa  niche, 
the  identification  of  suitable  release 
sites  for  snakes  bom  in  captivity,  and 
the  selection  of  likely  search  localities 
for  surveys  of  previously  unde- 
scribed  populations  of  the  snake.  By 
using  PCA,  we  also  hoped  to  extract 
indep)endent  patterns  of  covariation, 
such  as  the  degree  of  niche  overlap 
with  Alsophis,  which  might  explain 
certain  distributional  anomalies  of 
the  boa  populations. 

Methods 

Study  Areas 

This  study  is  based  on  habitat  analy- 
sis of  24  different  localities  on  the  fol- 
lowing islands  and  cays  of  the  Puerto 
Rico  Bank:  Buck  Is.,  Cas  Cay,  Cayo 
Diablo,  Cayo  Icacos,  Cayo  Lobos, 
Congo  Cay,  Great  St.  James  Cay,  Isla 
Mona,  Outer  Brass  Cay,  Salt  Cay, 
Saba  Cay,  and  Steven  Cay  from  Feb- 
ruary 1986  through  April  1988.  Some 
islands  had  several  plots.  Sites  were 
chosen  at  random  without  regards  to 
presence  or  absence  of  boas,  but  an 
attempt  was  made  to  select  sites  so 
that  sampling  included  the  full  spec- 
trum of  habitat  available  to  the  boa. 
Figures  2  and  3  illustrate  the  location 
of  sampling  plots  included  in  the 
study. 

Vegetational  Profiles  of  Study  Sites 

Subtropical  dry  forest  is  the  habitat 
where  E.  monensis  is  most  commonly 
observed,  particularly  on  Isla  Mona 
and  St.  Thomas.  It  is  characterized  by 
small  (<  5  m)  deciduous  trees  with 
small,  coriaceous  or  succulent  leaves 
and  thorns,  spines,  and  secondary 
defensive  compounds  (Ewel  and 
Whitmore  1973).  Examination  of  the 


229 


present  range  of  the  boa  indicates 
that  it  matches  the  occurrence  of  dry 
subtropical  forest  on  the  Puerto  Rico 
Bank  (Ewel  and  Whitmore  1973). 
This  is  most  apparent  on  St.  Thomas, 
where  E.  monensis  is  restricted  to  the 
dry  eastern  end  of  the  island  despite 
presumably  suitable  habitat  else- 
where (Nellis  et  al.  1984).  Common 
tree  species  include  Burseria 
simaruba,  Cephalocereus  royenii, 
Pidetia  aculeata,  Bucida  buceras, 
Guaiacum  officinale,  Leucaena  glauca, 
Tamarindis  indica,  Melicoccus  bijuga- 
tus,  Acacia  ssp.,  and  Capparis  cynoph- 
allophora  (Little  and  Wadsworth 
1964).  In  addition,  on  our  dry  forest 
plots  (Cas,  Icacos  1,  Congo  1,  Outer 
Brass  1,  and  Gt.  St.  James  1),  we  en- 
countered many  Byrsonima  lucida, 
Euphorbia  petiolaris,  and  Metopium 
toxiferum.  On  Buck  1,  Diablo  1,  Gt.  St 
James  3,  and  Mona  2  the  vegetation 
consisted  of  tree  species  with  com- 
pound trunks,  primarily  Coccoloba 
uvifera,  Hippomane  mancinella,  and 
Thespesia  populnea.  Sabal  palm  groves 
were  present  on  Outer  Brass  2  and 
Salt  2.  Salt-tolerant  shrublands  pri- 
marily composed  oiSuriam  and 
Tournefortia  just  above  the  high  tide 
line  was  the  dominant  vegetation  on 
Diablo  2,  while  Diablo  3  primarily 


consisted  of  Cassythia/Opuntia 
tangles,  f  icws-dominated  forest  was 
present  on  Mona  1  and  Congo  2. 
Guinea  grass,  Panicum  maximum, 
dominated  the  transect  on  Buck  2 
and  Acacia  macracantha  on  Buck  3.  A 
basic  summary  of  the  vegetation  of 
the  smaller  cays  is  given  in  Heatwole 
et  al.  (1981).  Figures  4  through  7  il- 
lustrate four  typical  vegetational 
types  at  transect  sites:  Coccoloba 
grove  (Buck  1),  mixed  palm/ shru- 
bland  (Diablo  2)  Opuntia/Cassythia 
tangles  (Diablo  3)  and  grassland 
(Buck  2). 


Geomorphology  and  Topography 
of  Study  Sites 

Geomophology  of  the  various  islands 
and  cays  studied  varied  considera- 
bly, from  the  steep-sided  metamor- 
phic  topography  of  St.  Thomas  and 
associated  cays  (Heatwole  et  al.  1981) 
to  the  cemented  dune  structure  of  La 
Cordillera  (Kaye  1959a).  Isla  Mona  is 
composed  primarily  of  a  Pleistocene 
limestone  plateau  surrounded  by 
sheer  cliff  (Kaye  1959b).  In  fact,  most 
islands  of  the  bank  have  significant 
limestone  deposits,  with  varying 
amounts  of  metamorphic  rock,  in- 


Outer  Brass  Is 


St.  Thomai"^  Steven  Cay 


t8»20 


Congo  Cay 


Saba  Cay 


U.  S.  Virgin  Islands 


,65* 


Figure  A.— Coccoloba  uvifera  habitat  on 
Buckl. 


Figure  5.— Mixed  Cocos  and  scrubland 
hiabitat  on  Cayo  Diabio.  Ttie  vegetation  at 
thie  center  of  \he  island  is  primarily  Cas- 
sythla  vine  growing  over  Opuntia  cactus. 


Figure  6.—  Aromatic  beactifrontshirubiand, 
primarily  Surlana  and  Tournefortia,  near 
Diablo  2. 


Figure  3.— Location  of  sampling  plots,  U.S.  Virgin  Islands. 


Figure  7.— Guinea  grass,  Panicum  maxi- 
mum, tiabitat  on  Buck  2. 


230 


eluding  gneiss  and  basalt,  present  as 
well.  The  cays  of  La  Cordillera  are 
exceedingly  low,  with  maximum  ele- 
vations under  15  m.  In  the  Virgin  Is- 
lands the  cays  are  of  moderate  eleva- 
tion with  eroded  limestone  hills  ap- 
proaching 50-300  m  in  height.  An 
overview  of  the  geology  of  the  Virgin 
Islands  is  given  in  Schuchert  (1935). 

Climate 

The  climate  of  the  Bank  is  essentially 
subtropical  to  tropical.  Temperatures 
of  the  coastal  areas  range  from  over- 
night lows  of  ca.  15°  C  to  daytime 
highs  approaching  35°  C.  Rainfall, 
especially  on  Puerto  Rico,  is  geo- 
graphically variable  (Briscoe  1966). 
Areas  within  the  range  of  E.  monensis 
typically  receive  <  750  mm  of  rainfall 
per  year. 

Sampling  Techniques 

The  presence  or  probable  absence  of 
the  boa  on  a  particular  cay  was  de- 
termined by  active  searching  of  all 
habitat  types  during  surveys  (carried 
out  independently  of  habitat  analy- 
sis) from  April  1983  to  September 
1987.  Typically  2  weeks  or  more  were 


spent  searching  larger  islands  and 
three  to  five  days  for  smaller  cays. 
Only  1  night  was  spent  on  Cayo  Lo- 
bos,  as  the  native  vegetation  was  all 
but  completely  destroyed  by  human 
activity  and  all  densely  vegetated  ar- 
eas could  be  searched  repeatedly  in  a 
single  night.  Our  experience  with 
multiple  recaptures  of  the  same  indi- 
vidual indicates  that  the  snakes  for- 
age every  night  under  most  circum- 
stances. Within  each  24-hour  period 
4  hours  per  night  were  spent  search- 
ing likely  foraging  sites  such  as  vine 
tangles,  terminal  branches  of  trees, 
palm  crowns,  and  beachfront  vegeta- 
tion. Ehiring  the  daylight  hours,  refu- 
gia  sites  such  as  debris  piles,  termite 
nests,  and  palm  axils  were  exanrdned. 
After  capture,  the  time,  capture 
height,  habitat  description,  ambient 
temperature,  refugium  temperature, 
and  cloacal  temperature  of  each 
snake  were  recorded.  Later,  sex, 
body  mass,  snout-vent  length  (SVL), 
and  caudal  length  (CL)  were  re- 
corded. The  snakes  were  examined 
for  reproductive  condition,  presence 
of  injuries,  and  parasite  infestation. 
Snakes  were  marked  using  the  tech- 
nique of  Brown  and  Parker  (1976) 
and  released  at  the  point  of  capture. 

Habitat  variables  recorded  in- 
cluded both  physical  and  biological 


parameters  (table  1).  Predator  den- 
sity estimates  include  indices  of 
abundance  for  likely  predators  of  E. 
monensis:  the  roof  rat,  Rattus  rattus, 
the  pearly-eyed  thrasher,  Margarops 
fuscatus,  and  the  Puerto  Rican  racer, 
Alsophis  portoricensis.  Rattus  densities 
were  estimated  using  removal  trap- 
ping over  a  3-day  span  on  100-m 
transects  with  Victor  snap  traps 
spaced  every  5  m.  Presence  of  Felis 
catus  was  determined  by  direct  ob- 
servation. Because  of  the  extreme 
wariness  and  trap-shy  nature  of  the 
Felis  on  study  plots,  only  their  pres- 
ence or  absence  was  recorded. 

Prey  density  data  includes  of 
population  densities  for  Anolis  cris- 
tatellus  and  Ameiva  exsul.  Anolis,  Also- 
phis, Ameiva,  and  Margarops  were 
counted  by  having  two  observers 
slowly  walk  the  transects  and  count- 
ing the  individuals  of  each  species 
observed  within  a  5  m  distance  on 
each  side  of  the  transect  line.  On 
Cayo  Diablo,  independent  estimates 
of  Ameiva  and  Anolis  cristatellus 
populations  were  gathered  by  sur- 
veys of  5  m^  quatrats.  Anolis  cristatel- 
lus perch  heights  were  measured 
with  a  metric  tape  except  on  Cayo 
Lobos  and  Salt  Cay.  Canopy  height 
was  estimated  for  each  habitat  with 
the  help  of  a  metric  tape.  Vegetative 
composition  was  determined  by  sub- 
jective stratified  sampling  using  10 
m^  quadrat  plots  (Clarke  1986);  plant 
samples  were  taken  for  species  iden- 
tification from  each  island.  Vegeta- 
tion coverage  data  indicates  the  per- 
centage composition  of  five  different 
classes  of  vegetation:  trees  (trunk  cir- 
cumference at  shoulder  height  >  25 
cm),  palms,  Opuntia  cactus,  shrubs 
and  small  trees  (trunk  circumfer- 
ences <  25  cm),  and  grasses.  Vegeta- 
tion structural  data  includes  the 
number  of  dominant  plant  species, 
the  height  of  the  canopy,  and  the 
continuity  of  the  vegetation  (a  meas- 
ure of  the  difficulty  for  the  boa  to 
crawl  from  one  plant  to  another 
without  going  to  the  ground).  Plants 
were  identified  by  David  W.  Nellis 
and  the  author. 


r 


Table  L— Factor  patterns  of  the  original  variables  on  each  of  the  first  six 
principal  components. 


Principal  component 


Variable 

1 

11 

III 

IV 

V 

VI 

Rat  density 

-0.6795 

-0.3582 

0.0162 

-0.2453 

0.4298 

-0.0379 

Cot  presence 

0.4976 

0.1703 

0,0895 

-0.4592 

-0.3190 

-0.2217 

Racer  density 

-0.4702 

0,0838 

-0,1051 

0.5491 

-0.1627 

-0.0437 

Thrasher  density 

-0.5268 

0,2052 

0,0410 

0,0091 

-0.6211 

0.4070 

Anolis  density 

-0.1400 

0.3391 

0.4005 

0.4582 

0.1560 

-0.4850 

Ameiva  density 

-0.0657 

0.6091 

-0,5302 

0.3176 

-0.0890 

-0.0628 

/Anofe  perch  height 

0.7972 

0.1830 

-0.1991 

-0.0386 

0.2946 

-0.0233 

Compound  tree  density 

-0.0810 

0,7175 

-0,4546 

-0.2118 

0.2106 

0.0407 

Singletree  density 

-0.5156 

-0.0546 

0,3885 

0.1877 

0.5079 

0.2290 

Palm  der^ity 

0.4478 

0.1154 

0.3690 

-0.5638 

-0.2063 

0.0771 

Shrub  der-isily 

0.6215 

-0.3240 

0,4724 

0.4028 

-0.1720 

-0.0718 

Grass  density 

-0,1656 

-0.5586 

-0.5118 

-0.1659 

-0,0554 

-0.4613 

Cactus  density 

0,3458 

-0.3437 

-0.1861 

0.3130 

0.1391 

0.6288 

Vegetational  continuity 

0.6882 

0.3989 

-0.0270 

-0,1124 

0.3701 

0.0988 

Canopy  height 

-0.3978 

0.7730 

0,4023 

-0.1372 

0.0139 

0.0257 

231 


I  attempted  to  use  continuously 
distributed  standarized  environ- 
mental variables  whenever  possible. 
Absence  of  a  particular  predator  or 
prey  species  on  a  given  a  sample  plot 
did  not  always  indicate  its  absence 
from  the  island  on  which  the  plot 
was  situated.  Only  male  Anolis  perch 
heights  were  used  for  the  statistical 
analysis,  as  female  and  juvenile  A. 
cristatellus  tend  to  frequent  the 
ground  under  all  circumstances  (Ki- 
ester  et  al.  1975).  Mean  male  Anolis 
perch  height  data  were  pooled  for 
each  island  for  character  16  of  the 
PCA  data  matrix,  as  some  plots  were 
completely  devoid  of  Anolis. 

Statistical  Analysis 

Principal  components  analysis  was 
performed  using  the  Statistical 
Analysis  System  '"SAS"  release  5.16 
(SAS  Institute  1985).  Significant  habi- 
tat components,  which  included  both 
biotic  and  structural  variables  of  the 
collecting  localities  (e.g.  those  which 
accounted  for  >  10%  of  the  total  vari- 
ance in  the  data),  were  clustered  on 
the  basis  of  their  association  within 
the  PCA  data  matrix.  The  second 
step  of  the  analysis  compared  the 
relative  abundance  of  E.  monensis  at 
each  collecting  locality  with  habitats 
described  by  the  significant  axes  of 
the  principal  components.  Regression 
analysis,  ANOVA,  and  descriptive 
statistics  (mean,  standard  deviation, 
etc.)  were  performed  using  Statview 
512^  on  an  Apple  Macintosh  Plus. 

Results 

Multivariate  Analysis  of  Habitats 

The  PCA  indicates  that  biotic  factors, 
plant  composition,  and  structural  at- 

^7he  use  of  trade  and  company  names 
is  for  the  benefit  of  the  reader:  such  use 
does  not  constitute  an  official  endorsement 
or  approval  of  any  service  or  product  by 
the  U.  S.  Department  of  Agriculture  to  the 
exclusion  of  others  that  may  be  suitable. 


tributes  of  vegetation  are  all  impor- 
tant contributors  to  variance  in  the 
PCA  patterns.  Factor  patterns  for  the 
first  six  principal  components  are 
given  in  table  1.  Principal  component 
I  accounts  for  23.4%  of  the  variance. 
This  component  clusters  habitats 
with  high  shrub  and  palm  densities, 
low  numbers  of  single  trees,  vegeta- 
tional  continuity,  and  low  canopy 
heights.  Important  biotic  characteris- 
tics of  this  space  include  Felis  pres- 
ence and  low  Rattus,  Margarops,  and 
Alsophis  densities  with  high  Anolis 
perch  heights.  Principal  component  II 
accounts  for  an  additional  17.0%  of 
the  variance  observed.  This  axis  de- 
scribes sites  having  low  grass  den- 
sity, high  compound  tree  densities, 
canopy  height  >  3  m ,  and  high 
Ameiva  densities.  Principal  compo- 
nent III  accounted  for  11.2%  of  the 
variance  and  suggested  an  associa- 
tion between  low  Ameiva  density, 
low  compound  tree  density,  low 
grass  density,  high  shrub  density, 
and  canopy  height  >  3  m.  Factor  IV 
accounted  for  another  10.7%  of  the 
variance  and  clustered  high  Alsophis 
and  Anolis  densities  with  Felis  ab- 
sence and  low  palm  density.  Compo- 
nents V-VI  were  less  significant  in 
the  PCA  (e.g.  each  accounted  for  <  10 
%  of  the  variance)  but  added  some 
interesting  ecological  information  to 
the  habitat  analysis.  Principal  compo- 
nent V  clustered  high  Rattus  density 
with  low  Margarops  density;  princi- 
pal component  VI  grouped  high  Mar- 
garops density  with  low  Anolis  den- 
sity. 

Habitat  Utilization  by  Epicrafes 
monensis 

The  vegetational  profiles  of  climax 
plant  communities  (and  E.  monensis 
collection  localities)  in  the  dry  forest 
may  differ  considerably  depending 
on  island  size,  geology,  geomorphol- 
ogy,  rainfall,  and  history  of  human  or 
feral  mammal  disturbance.  However, 
most  dry  forest  habitats  on  the  Bank 
are  structurally  simple,  with  usually 


only  two  to  five  dominant  plant  spe- 
cies (table  2).  Captures  and  sightings 
of  the  Mona  boa  have  been  limited  to 
three  distinct  localities:  dry  plateau 
forest  adjacent  to  Uvero  and  Pajaros 
(Campbell  and  Thompson  1978;  Riv- 
ero  et  al.  1982)  Coccoloba  uvifera 
groves  of  Pajaros  (M.  Frontera,  Pers. 
Comm.),  and  Cocos  groves  and 
nearby  vegetation  adjacent  to  Playa 
Sardinera  (G.  Rodriguez  pers. 
comm.).  The  Virgin  Islands  boa  has 
been  encountered  repeatedly  on  only 
two  islands:  St.  Thomas  and  Cayo 
Diablo.  All  specimens  from  St.  Tho- 
mas were  captured  on  the  east  end  of 
the  island  near  Red  Hook.  Two  speci- 
mens were  found  beneath  a  lime- 
stone slab  during  construction  of  the 
Vessup  Bay  Estates  housing  subdivi- 
sion, another  was  taken  from  a  stone 
wall,  and  a  third  was  found  as  a 
roadkill  near  Smith  Bay.  R.  Thomas 
captured  a  specimen  crawling  in  a 
viney  tangle  ca.  2.4  m  high  (Sheplan 
and  Schwartz  1974). 

The  Red  Hook  area  is  dominated 
by  xeric  forest  composed  primarily 
of  Burseria,  Croton,  and  Acacia.  No 
habitat  data  is  available  for  E.  m. 
granti  on  Tortola.  I  have  received  re- 
ports that  the  boa  was  present  in  the 
palm  forest  of  Outer  Brass  Island  (J. 
LaPlace  pers.  comm.)  but  I  was  un- 
able to  find  it  there  even  after  five 
trips  to  the  island.  Virgin  Islands 
residents  also  report  the  boa  as  in- 
habiting Great  St.  James  Is.  (D.  Nellis 
pers.  comm.).  Great  Camanoe, 
Necker  Is.,  and  Virgin  Gorda,  (Mayer 
and  Lazell  1988),  but  these  sightings 
have  not  been  confirmed  by  biolo- 
gists. Grant  (1932)  mentioned  anec- 
dotally  (he  did  not  capture  the 
holotype  himself)  that  "the  boa  is 
found  on  rocky  cliffs  on  Tortola  and 
Guana  Islands." 

On  Cayo  Diablo,  Coccoloba  uvifera 
is  the  habitat  most  commonly  associ- 
ated with  foraging  E.  monensis.  Of  the 
79  active  snakes  we  captured,  51 
were  found  in  Coccoloba,  ten  in  Cae- 
salpinea,  nine  on  Cassythia,  seven  in 
Suriana,  and  two  in  Opuntia.  Twenty- 
three  percent  of  the  snakes  were  ac- 


232 


tive  at  heights  >  2  m.  Of  these,  67% 
had  SVLs  >  400  mm.  Seventy-five 
percent  of  juvenile  snakes  (under  300 
mm  SVL)  foraged  at  heights  <  1.5  m, 
but  regression  analysis  indicated  that 
these  differences  were  not  statisti- 
cally significant.  Of  the  149  inactive 
snakes  taken  from  refugia,  43%  were 
in  Cocos  or  Sahal  axils,  36%  were  in 
termite  nests,  and  21%  were  under 
rocks  or  debris.  Fifty-one  percent  of 
snakes  taken  from  termite  nests  were 
females;  over  half  of  these  were 
gravid.  Gravid  females  use  termite 
nests  or  sun-baked  debris  to  ther- 
moregulate  and  may  elevate  their 
body  temperatures  to  over  33°  C. 

Prey  Density  and  Epicrate$ 
monensis  Distributions 

The  greatest  concentrations  of  Ep- 
icrates  monensis  are  in  areas 
(particularly  Coccoloha  groves)  with 
AnoUs  densities  >  60  Anolis/ 100  m^. 


This  Anolis/Epicrates  association  is 
reinforced  by  PCA  (see  below).  My 
field  logs  indicate  that  the  greatest 
success  in  finding  foraging  Epicrates 
occurs  when  observations  of  sleeping 
Anolis  are  >  12  lizards/ h.  Numerical 
counts  of  sleeping  Anolis  and  the 
times  between  sightings  are  regularly 
noted  in  my  field  book  as  a  rough 
guide  to  potential  hunting  success  in 
a  study  locality. 

AnoUs  cristatellus  is  the  primary 
prey  species  of  E.  monensis,  and  the 
mean  foraging  height  of  the  snake  (x 
=  1.356,  SD  =  1.079  N  =  54)  is  close  to 
the  mean  perch  height  of  sleeping 
Anolis  (x  males  =  1.816  m,  SD  =  0.993, 
N  =  17;  X  females  =  1.323  m,  SD  = 
.681,  N  =  14;  X  juveniles  =  1.417  m, 
SD  =  0.169,  N  =  5). 

High  Ameiva  densities  are  also  a 
common  component  of  localities 
with  high  boa  densities,  although  I 
observed  only  one  instance  of  a  boa 
feeding  on  Ameiva,  which  are 
strongly  diurnal. 


Feral  Mammal  Abundance  and 
Epicrates  monensis  Distributions 

Of  the  10  islands  surveyed  for  this 
study,  only  three  were  completely 
devoid  of  rats:  Cayo  Diablo,  Cayo 
Icacos,  and  Steven  Cay.  These  islands 
have  high  Ameiva  and  Anolis  densi- 
ties, but  only  Diablo  Cay  harbors  a 
population  of  the  boa.  It  also  has  the 
highest  densities  of  Epicrates  monensis 
found  anywhere  on  the  bank,  >  100 
snakes/hc  at  some  localities.  Those 
islands  with  heavy  rat  densities  (ca. 
20  rats/ hectare) — Buck  Is.,  Cas  Cay, 
and  Salt  Cay — have  lower  Ameiva  and 
Anolis  densities  and  apparently  no 
boa  populations,  despite  suitable 
habitat.  Rat  densities  are  not  always 
correlated  with  low  Anolis  densities, 
however.  Some  islands,  such  as 
Outer  Brass  and  Congo,  have  Anolis 
densities  apparently  high  enough  to 
support  populations  of  the  boa,  but 
their  perch  heights  (table  2)  are  sig- 
nificantly different  from  those  Anolis 


Table  2.— PCA  habitat  matrix  for  the  Puerto  Rico  Bank. 


Rat  Cat  Racer  Thrasher  Anolis 
dens,  presence  dens,  abund,  dens. 


Ameiva  Comp, 
dens,  tree 
dens. 


Single 
tree 
dens. 


Palm 
dens. 


Shrub/ 
small 
tree 
dens. 


Grass  Opuntla  Contlg.  Plant 
dens,    dens/    veg.  diver, 
cover 


Canopy  Anolis 
height  perch 
height 


;  Cayo  Diablo  1 

0.00 

0 

0.00 

0,00 

1,48 

1,50 

0.99 

0.00 

0.01 

0.00 

0,00 

0,00 

1.00 

1 

1,00 

1.70 

Cayo  Diablo  2 

0.00 

0 

0.00 

0.00 

0.25 

0.05 

0,00 

0.00 

0,03 

0,95 

0,00 

0,00 

1.00 

2 

0,50 

1.70 

Cayo  Diablo  3 

0.00 

0 

0.00 

0.00 

0,00 

0.10 

0,00 

0.00 

0.00 

1,00 

0.00 

0,60 

1.00 

3 

0.26 

1.70 

Cayo  icacos  1 

o.oc 

1 

0.00 

0.05 

0,36 

0.76 

0,46 

0.06 

0.00 

0,50 

0.00 

0,00 

0.75 

4 

1.00 

1,57 

Cayo  icacos  2 

0.00 

1 

0.00 

0.00 

0.91 

0.00 

0,14 

0.10 

0.13 

0.63 

0.00 

0,00 

0.60 

6 

1.00 

1.57 

i  Cayo  Icacos  3 

0,00 

1 

0,00 

0.00 

0,28 

0.00 

0,02 

0.00 

0.00 

0,98 

0.00 

0,00 

0.75 

4 

0,60 

0.50 

;  Cayo  Lobos 

0.08 

0 

0.00 

0,20 

0.03 

0.02 

0,75 

0.10 

0.10 

0.16 

0,00 

0,00 

0.26 

3 

1.00 

0.18 

Congo  Cay  1 

0,08 

0 

0.50 

0.10 

2.10 

0,00 

0,23 

0,01 

0.00 

0.76 

0.00 

0,00 

0.10 

1 

1.00 

0.18 

Congo  Cay  2 

0.09 

0 

0.60 

0.08 

2.00 

0.00 

0.04 

0,42 

0.13 

0.46 

0.00 

0,00 

0.25 

4 

1.00 

0,32 

Outer  Brass  1 

0.04 

0 

2.60 

0.30 

1.16 

1,00 

0.69 

0,07 

0.00 

0.35 

0.00 

0.00 

0.50 

2 

1,00 

0.32 

Outer  Brass  2 

0.02 

0 

0.50 

0.73 

0.36 

0,60 

0.00 

0,10 

0.86 

0.00 

0.00 

0,00 

0.00 

2 

1.00 

1.00 

Salt  Cay  i 

o.n 

0 

0.33 

0.10 

0.42 

0,16 

0.90 

0.10 

0.00 

0.00 

0.00 

0.00 

1,00 

5 

1.00 

1.00 

Salt  Cay  2 

0.18 

0 

0.67 

0.05 

0,84 

0,08 

0.00 

1.00 

1.00 

0.00 

0.00 

0.00 

0.60 

1 

1.00 

0.67 

Isia  Mono  1 

0,02 

0 

0.00 

0.21 

0,58 

0,00 

0.14 

0.15 

0.00 

0.71 

0.00 

0,00 

1.00 

5 

1.00 

0.67 

Isia  Mono  2 

0.02 

1 

0.00 

O.n 

0.25 

0,00 

0.58 

0.07 

0.10 

0.26 

0.00 

0.00 

0,76 

3 

1,00 

0.42 

Gt,  St.  James  1 

0.06 

0 

0.50 

0,10 

0,68 

0,00 

0,68 

0.29 

0.00 

0.00 

0.00 

0,13 

0,26 

1 

1.00 

0.42 

Gt,  St.  James  2 

0.06 

0 

1.00 

0.16 

0.53 

0,00 

0,00 

0.27 

0.00 

0.68 

0.00 

0.05 

0.25 

3 

1,00 

0.42 

Gt,  St.  James  3 

0.05 

0 

0.35 

0.15 

0.00 

0,00 

0,27 

0.45 

0,00 

0,24 

0.00 

0.04 

0,50 

4 

1.00 

0.60 

Buck  Is.  1 

0.04 

0 

0.00 

0.00 

0.60 

0,00 

0,28 

0.06 

0.00 

0,66 

0.00 

0,00 

1,00 

4 

1.00 

060 

Bucl<  Is,  2 

0.13 

0 

0.33 

0,00 

0.00 

0,00 

0.00 

0.00 

0,00 

1.00 

0.00 

0,00 

0,25 

2 

0.50 

0.50 

Bucl<  Is.  3 

0.10 

0 

0.33 

0,00 

0.05 

0,00 

000 

0.00 

0.00 

0.00 

1.00 

0.00 

0,00 

2 

0.26 

0.50 

Steven  Cay 

0.00 

0 

0.00 

0,01 

2.79 

0,00 

0.16 

0.08 

0.00 

0.76 

0.00 

0.00 

1,00 

3 

1.00 

1.26 

Saba  Cay 

0.00 

0 

2.00 

0.00 

1.06 

0,00 

0.08 

0.08 

0.00 

0.84 

0,00 

0.00 

1,00 

2 

1.00 

1.02 

Cos  Cay 

0.04 

0.00 

0.00 

0.0476 

0.12 

0.00 

0.78 

0.09 

0.00 

0.13 

0,00 

0.00 

1,00 

3 

1.00 

1.40 

'See  appendix  A  for  variable  descriptioiis. 

233 


inhabiting  rat  free  islands.  ANOVA 
performed  on  the  regression  line  (y  = 
-5.548x  +  1.127)  which  plots  Anolis 
perch  height  vs.  rat  density  on  my 
study  islands  (Tolson  and  Campbell 
in  prep)  shows  a  negative  correlation 
(p  =  .0137)  between  rat  density  and 
Anolis  perch  height.  This  is  not 
surprising.  Anolis  cristatellus  resident 
on  rat-infested  islands  exhibit  a  typi- 
cal escape  behavior.  Male  Puerto 
Rican  A.  cristatellus  escape  to  the  can- 
opy when  threatened  (Heatwole 
1968),  but  those  on  Congo  Cay,  Outer 
Brass,  and  Salt  Cay  all  run  to  the 
ground  when  disturbed,  even  when 
suitable  cover  on  the  ground  is  lack- 
ing. At  night,  the  Anolis  are  not  usu- 
ally found  sleeping  exposed  on  vege- 
tation, but  rather  under  rocks.  This  is 
extremely  unusual  behavior  for  A. 
cristatellus  (E.  Williams  pers.  comm.). 

Although  one  does  not  often 
discover  E.  monensis  on  islands 
which  are  infested  with  rats,  some 
sympatry  does  occur.  Isla  Mona  and 
St.  Thomas  are  islands  with  moder- 
ate rat  densities  and  extant  (although 
apparently  dwindling)  populations 
of  E.  monensis.  Interestingly,  at  locali- 
ties where  Epicrates  coexists  with  Rat- 
tus,  there  are  also  significant  num- 
bers of  introduced  mammalian 
predators  such  as  Felis  and  Herpestes 
(table  2). 


Discussion 

PCA  and  E.  monensis  Habitat 
Utilization 

The  Puerto  Rico  Bank  encompasses  a 
total  land  area  in  excess  of  9,300  km^, 
of  which  1700+  km^  (or  17.6%)  is  cov- 
ered with  subtropical  dry  forest 
(Ewel  and  Whitmore,  1973).  This 
xeric  forest  is  widely  distributed 
throughout  the  range  of  Epicrates 
monensis,  yet  the  boa,  as  far  as  we 
know,  occupies  only  seven  islands  of 
the  243  that  make  up  the  banks — ef- 
fectively exploiting  only  0.04%  of  the 
land  area  available  to  it.  PCA  helped 
to  identify  those  factors  which  seem 


to  define  critical  boa  habitat.  Several 
vegetative  parameters  which  cluster 
together  in  the  PCA  are  descriptive 
of  habitat  where  I  or  others  have 
encountered  E,  monensis  repeatedly. 
These  include  areas  with  high  shrub 
and  palm  densities  coupled  with  a 
low  canopy  and  vegetational  conti- 
nuity. These  values  describe  plot 
habitat  on  Diablo  2,  Icacos  2,  and  cer- 
tain sites  within  the  Red  Hook  area 
of  St.  Thomas.  Either  high  shrub  or 
high  palm  densities  coupled  with 
vegetational  continuity  and  lower 
canopy  are  found  on  Diablo  3,  Icacos 
3,  and  Mona  1.  Of  these  two  subsets 
of  PC  I,  boas  occur  on  Diablo  2  and  3, 
Mona  1,  St.  Thomas,  and  almost  cer- 
tainly inhabited  Icacos  1  and  3  at  one 
time. 

In  PC  II,  habitat  correlates  include 
high  compound  tree  density,  high 
canopy  height,  vegetational  continu- 
ity, and  low  grass  density.  This  is  a 
perfect  structural  and  compositional 
description  of  Diablo  1,  which  has 
the  highest  population  of  E.  monensis 
I  have  ever  encountered,  and  Mona 
2 — another  locality  where  E.  monensis 
has  been  observed  (Campbell  and 
Thompson  1978).  It  seems  clear  from 
these  data  that  the  unifying  variable 
which  causes  an  intersection  of  these 
two  differing  habitat  types  is  vegeta- 
tional continuity — an  interlocking  of 
the  branches  of  shrubs  or  the  tree 
canopy.  I  believe  this  vegetational 
characteristic  is  essential  to  E.  monen- 
sis foraging  success  and  survival.  It 
probably  not  only  decreases  the 
search  time  between  encounters  with 
sleeping  Anolis  while  foraging,  but  it 
also  potentially  limits  the  encounters 
between  the  boa  and  Felis  and  Her- 
pestes. Fortunately,  at  least  some 
tracts  of  subtropical  dry  forest  and 
Coccoloba  have  remained  relatively 
undisturbed  on  the  Virgin  Islands, 
Isla  Mona,  and  Puerto  Rico  and  its 
offshore  satellites.  Much  suitable 
habitat  does  exist — even  near  popu- 
lated areas. 

While  habitats  throughout  the 
Bank  are  presumably  underutilized 
by  E.  monensis,  and  suitable  areas  for 


reintroduction  apparently  exist  in  a 
number  of  localities,  the  extant  boa 
populations  are  so  fragmented  and 
reduced  in  numbers  that  it  is  crucial 
to  protect  those  areas  now  support- 
ing the  boa.  This  may  be  difficult. 
Historically,  vegetation  on  Puerto 
Rico  and  the  Virgin  Islands  has  been 
severely  disrupted,  and  17th-18th 
century  land  use  patterns  on  the  U.S. 
Virgin  Islands  may  partially  explain 
the  limited  distribution  of  the  boa  on 
the  east  end  of  St.  Thomas  and  its 
absence  from  St.  John.  Even  now 
enormous  pressures  exist  for  contin- 
ued development  on  the  east  end  of 
St.  Thomas.  Construction  around 
Red  Hook  seems  to  have  accelerated 
in  recent  months,  perhaps  in  re- 
sponse to  the  decline  of  interest  rates 
in  the  United  States,  and  three  rela- 
tively undeveloped  areas  on  the  east 
end — Red  Hook  Mountain,  Cabrita 
Point,  and  Water  Point — all  have  proj- 
ects in  progress  that  do  not  involve 
federal  funding.  The  management 
authority  on  St.  Thomas,  U.S.  Virgin 
Islands — the  Division  of  Fish  and 
Wildlife — has  no  control  over  such 
development. 

In  contrast,  Puerto  Rican  islands 
with  populations  of  Epicrates  monen- 
sis are  in  no  imminent  danger  of  de- 
velopment. Cayo  Diablo  is  part  of  the 
Reserva  Forestal  de  La  Cordillera, 
and  Isla  Mona  is  likewise  a  Forest 
Preserve  (although  it  was  once  pro- 
posed to  develop  the  island  as  a 
deep-water  oil  fX)rt).  A  problem  does 
exist,  however,  with  habitat  destruc- 
tion on  isolated  cays  caused  by 
campers  and  fishermen  (Heatwole 
and  Mackenzie  1967).  Coccoloba  trees 
in  the  larger  groves — areas  where  the 
greatest  densities  of  E,  monensis  are 
found — are  often  used  as  firewood  by 
visitors.  A  survey  done  in  1987  of 
damage  to  Coccoloba  stands  on  Cayo 
Diablo  showed  that  many  trees  sus- 
tained some  sort  of  damage  caused 
by  human  activity,  primarily  ma- 
chete cuts  and  burns  from  fires 
started  at  the  bases  of  the  trees. 


234 


Effects  of  Feral  Mammals 

My  analysis  shows  that  Rattus  and 
Felis  are  a  primary  influence  on  com- 
munity composition  on  the  Puerto 
Rico  Bank.  Felis  presence  is  associ- 
ated with  low  Alsophis,  Margarops, 
and  Rattus  density  (table  1:  PC  I);  Fe- 
lis absence  is  associated  high  Anolis 
and  Alsophis  densities  in  PC  IV  (table 
1).  Clearly  the  presence  of  Felis  in  E. 
monensis  habitat  is  a  mixed  blessing. 
Cats  present  a  great  danger  to  Ep- 
icrates  because  they  hunt  at  night. 
Several  instances  of  cat  predation  of 
Epicrates  have  been  reported  on  St. 
Thomas  (D.  Nellis  pers.  comm.)  In 
fact,  in  April  and  May  of  1988  two  E. 
monensis  were  rescued  from  cats  on 
St.  Thomas  and  were  incorf>orated 
into  the  captive  breeding  program  at 
the  Toledo  Zoological  Gardens.  In 
contrast,  however,  on  islands  where 
boas  and  rats  coexist — Isla  Mona  and 
St.  Thomas — there  are  also  significant 
populations  of  Felis.  Cats  feed  on 
Rattus  and  may  keep  rat  populations 
at  levels  low  enough  to  permit  sur- 
vival of  the  boa.  Their  apparent  ad- 
verse affect  on  Alsophis  and  Marga- 
rops density — two  potential  predators 
of  E.  monensis — may  also  be  of  some 
small  benefit  in  certain  circum- 
stances. Weiwandf  s  (1977)  observa- 
tion of  cat  predation  of  Alsophis  on 
Isla  Mona  corroborate  the  PC  I  link- 
age of  cat  presence  with  low  Alsophis 
density. 

I  cannot  be  certain  whether  Rattus 
affect  boa  populations  by  acting  pri- 
marily as  a  constraint  on  their  re- 
source levels  or  by  direct  predation. 
Although  I  have  been  unable  to  dem- 
onstrate that  rats  forage  on  boas,  I 
have  every  reason  to  suspect  that 
they  do.  Rattus  is  known  to  prey  on 
lizards  (Whitaker  1978).  While  sur- 
veying for  boa  populations  on  the 
Bank  I  found  habitat  (Congo  Cay, 
Outer  Brass  Cay)  which  provides  op- 
timal foraging  opportunities  for  the 
boa  (e.g.  vegetation  associated  with 
population  densities  of  >  60  Anolis/ 
100  m^  on  rat-free  islands)  but  had  no 
or  few  boas  and  were  virtually  over- 


run with  rats  at  night.  Rats  may  also 
affect  boa  populations  by  preying  on 
Anolis  directly  or  by  influencing  their 
perching  behavior,  (indicated  by  the 
negative  correlation  between  rat  den- 
sity and  Anolis  perch  height  (table  1: 
PC  I)  or  selection  of  sleeping  sites.  If 
lizards  rarely  rest  in  the  canopy  at 
night  but  rather  seek  refuge  sites  on 
the  ground,  there  would  be  poten- 
tially disastrous  consequences  for 
boa  foraging  success.  Rattus  also  ap- 
parently affect  Margarops  density 
(table  1:  PC  V). 

There  can  be  little  doubt  that  the 
Indian  mongoose,  Herpestes  auropunc- 
tatus,  threatens  Epicrates  monensis  di- 
rectly as  well,  but  I  believe  the  risk  to 
Epicrates  is  sometimes  exaggerated. 
Herpestes  predation  on  endemic  West 
Indian  snakes  is  well  documented 
(Maclean  1982),  but  the  mongoose  is 
a  strictly  diurnal,  terrestrial  predator; 
Epicrates  monensis  is  nocturnal  and 
arboreal.  Herpestes  poses  the  greatest 
danger  to  the  diurnal  West  Indian 
racers,  genus  Alsophis,  and  are  di- 
rectly responsible  for  the  extinction 
of  Alsophis  sancticrucis  on  St  Croix 
and  the  extirpation  of  A.  portoricensis 
from  St.  Thomas  and  St.  John.  In  con- 
trast, I  have  found  Epicrates  monensis 
abroad  during  the  daylight  hours  on 
only  two  occasions  over  a  period  of 
several  years.  It  seems  that  Herpestes 
would  have  the  greatest  chance  of 
capturing  Epicrates  when  the  latter  is 
resting  in  some  moderately  acces- 
sible location  during  the  day — in 
loose  sections  of  termite  nests,  for 
example.  Feral  pigs  (Sus  scrofa)  may 
also  threaten  the  Mona  boa  to  some 
degree,  either  by  eating  them  or  by 
destroying  vegetation,  such  as  terres- 
trial bromeliads,  that  may  act  as 
snake  refugia.  I  have  no  data  on  the 
magnitude  of  this  threat. 

Natural  Predators 

The  Puerto  Rico  Bank  has  no  extant 
species  of  native  mammalian  preda- 
tors, but  two  nocturnal  avian  preda- 
tory species  may  pose  a  limited 


threat  to  Epicrates  monensis.  The  yel- 
low-crowned night  heron,  Nyctanassa 
violacea,  and  the  Puerto  Rican  screech 
owl,  Otus  nudipes,  are  two  potential 
predators  of  the  boa.  While  popula- 
tions of  Otus  are  declining  on  the 
bank  (lUCN  1981)  those  of  the  heron 
seem  quite  stable.  I  have  repeatedly 
observed  herons  foraging  at  night  in 
boa  habitat  on  both  Isla  Mona  and 
Cayo  Diablo.  Examination  of  the  de- 
bris beneath  heron  rookeries  on  Cayo 
Diablo  has  revealed  numerous  frag- 
ments of  Anolis  and  Ameiva  skin  and 
skeletal  materials,  usually  ribs,  verte- 
brae, and  jaw  elements.  No  snake 
remains  have  been  found,  but  my  co- 
workers and  I  are  continuing  to  in- 
vestigate this  potential  problem.  I 
also  found  that  Anolis  densities  and 
perch  heights  are  reduced  (table  2) 
on  plots  with  high  pearly-eyed 
thrasher  densities.  In  PC  I  (table  1) 
high  Anolis  perch  heights  are  associ- 
ated with  low  thrasher  density. 
These  birds  also  prey  on  Anolis,  and 
are  so  common  in  some  areas  they 
could  easily  depress  Anolis  popula- 
tion numbers.  Principal  component 
VI  (table  1)  couples  high  thrasher 
density  with  low  Anolis  density. 

Two  arthropods  are  potential 
predators  of  E.  monensis:  the  land 
crab  Gecarcinus  and  the  hermit  crab 
Caenobita  clypeatus.  Searches  of  ter- 
restrial refugia  for  Epicrates  have  re- 
vealed that  these  snakes  are  nearly 
always  absent  from  areas  occupied 
by  Gecarcinus  and  Caenobita.  This  is 
especially  true  in  termite  nests. 
Snakes  only  occupy  areas  of  the  nest 
that  are  inaccessible  to  crabs.  If 
weathering  or  disturbance  causes  a 
section  of  termite  nest  to  become 
habitable  for  crabs  it  is  abandoned  by 
Epicrates,  despite  their  prior  use  of 
the  refugium  for  several  past  field 
seasons.  In  hundreds  of  examinations 
of  refugia  over  the  past  five  field  sea- 
sons, I  found  Epicrates  in  association 
with  Caenobita  on  only  one  occasion:  I 
found  a  gravid  female  thermoregu- 
lating  under  a  discarded  tarpaulin  in 
the  midst  of  several  Caenobita  on  7 
September  1987.  Evidence  for  preda- 


235 


tion  by  the  aforementioned  species  is 
strictly  circumstantial,  but  the  fact 
remains  that  over  17%  of  the  Ep- 
icrates  captured  have  obvious 
wounds,  scars,  or  partially  ampu- 
tated tails.  This  is  strong  evidence 
that  some  form  of  natural  predation 
is  occurring. 

Climatic/Stochastic  Events 

The  apparent  extirpation  of  the  snake 
from  the  majority  of  the  islands  on 
the  Bank  relate  not  only  to  the  arrival 
of  European  man  on  the  Bank  and 
the  habitat  destruction  which  fol- 
lowed, but  also  to  climatic,  eustatic, 
and  stochastic  events,  many  of  which 
had  profound  influences  on  habitat. 
During  the  late  Pleistocene  several 
climatic  and  eustatic  events  occurred 
that  apparently  set  the  stage  for  the 
decline  of  E.  monensis  on  the  Bank. 
Foremost  among  these  was  a  dra- 
matic change  in  the  climate  of  Puerto 
Rico.  From  a  relatively  xeric  climate, 
Puerto  Rico  became  progressively 
more  mesic  during  the  late  Pleisto- 
cene. Today,  over  81%  of  Puerto 
Rico's  vegetation  is  classified  as 
moist  or  wet  forest  (Ewel  and  Whit- 
more  1973).  Pregill  (1981)  and  Pregill 
and  Olson  (1982)  describe  the  effect 
this  climatic  change  had  on  the  xeric- 
adapted  Puerto  Rican  herpetofauna. 
This  extreme  climatic  shift  may  have 
resulted  in  the  extirpation  of  E. 
monensis  on  Puerto  Rico.*  In  addi- 
tion, sea  levels  rose  nearly  100  m 
about  8,000-10,000  years  ago  and 
separated  the  Virgin  Islands  from 
one  another  and  from  Puerto  Rico, 
transforming  what  was  a  contiguous 
land  mass  into  a  scattered  series  of 
islets  and  cays  spread  over  nearly 
400  km.  Many  of  these  cays  now 
have  extremely  low  elevations  (Heat- 
wole  and  Mackenzie  1967). 

''it  is  unclear  why  E.  monensis  is  absent 
from  the  dry  forest  in  southwestern  Puerto 
Rico.  Hab  'ftat  in  the  Guanica  forest  seems 
quite  suitable  for  the  boa:  perhaps  further 
survey  work  will  result  in  its  discovery  there. 


The  fragmentation  of  E.  monensis 
into  several  small  demes  may  have 
left  several  populations  without  the 
genetic  resources  to  survive  changing 
environments,  and  doubtless  allowed 
stochastic  processes  such  as  disease, 
prey  fluctuations,  or  storms  to  extir- 
pate many  isolated  populations.  I  as- 
sume that  the  influences  of  random 
events  on  the  present  distribution  of 
the  native  herpetofauna  complicates 
the  multivariate  analysis  by  introduc- 
ing more  variance  into  the  correla- 
tion matrix.  These  factors  may  ex- 
plain the  absence  of  snakes  from  is- 
lets with  suitable  habitat,  as  some  of 
these  islands  may  have  inadequate 
food  resources  or  lower  probabilities 
of  recolonization. 


Management  Recommendations 

The  forces  threatening  Epicrates 
monensis  are  complex.  Solutions  for 
the  recovery  of  the  boa  will  not  be 
simple,  but  I  am  optimistic  about  the 
chances  of  success.  My  management 
recommendations  are  summarized 
below. 


Saving  Boa  Habitat 

This  may  be  impossible  on  St.  Tho- 
mas, but  with  luck  the  boa  may  coex- 
ist with  man  (as  it  now  does)  at  some 
relatively  developed  localities.  Con- 
tinued protection  of  Isla  Mona  and 
La  Cordillera  are  absolutely  neces- 
sary. 

Continued  protection  and  man- 
agement should  be  extended  to  those 
cays  now  protected  by  the  Division 
of  Fish  and  Wildlife,  U.S.  Virgin  Is- 
lands— particularly  Congo  Cay,  Outer 
Brass  Cay,  Salt  Cay,  Savana  Island 
and  Steven  Cay — as  these  sites  might 
eventually  be  utilized  for  reintroduc- 
tion  programs.  The  smaller  islands 
should  be  off  limits  to  casual  visitors 
to  prevent  habitat  damage  and  hu- 
man persecution  of  the  snakes. 


Predator  Eradication  on  Suitable 
Offsliore  Islets 

Rat  control  programs  should  be  initi- 
ated immediately  on  those  islands 
with  habitat  suitable  for  E.  monensis. 
Preliminary  studies  of  rat  eradication 
using  anticoagulant  poisons  on  some 
small  cays  near  St.  Thomas  have  pro- 
duced promising  results  (Division  of 
Fish  and  Wildlife,  USVI 1983).  It  is 
critical,  however,  that  time  and  fund- 
ing be  committed  for  follow  up  stud- 
ies on  any  islands  made  the  subject 
for  a  rat  control  program.  This  must 
be  done  to  ensure  that  immunity  to 
poisons  has  not  evolved  or  that 
populations  are  being  replenished  by 
recolonization  from  St.  Thomas. 

It  is  unlikely  that  Felis  or  Herpestes 
will  ever  be  eradicated  from  larger 
islands  such  as  Isla  Mona  or  St.  Tho- 
mas, but  Felis  control  programs  now 
in  force  on  Mona  should  be  contin- 
ued to  further  reduce  populations 
and  should  be  expanded  to  include 
Cayo  Icacos.  It  is  important  to  con- 
vince management  authorities  that 
feral  mammal  control  measures  on 
the  Bank  must  be  increased,  and 
quickly. 

It  is  a  credit  to  the  evolutionary  re- 
silience of  this  little  snake  that  it  has 
survived  at  all.  Few  endangered  spe- 
cies have  been  exposed  to  such  a 
wide  range  of  adverse  effects  and 
have  still  survived.  It  is  my  fervent 
hope  that  this,  and  other  endemic 
species  of  the  Caribbean,  will  not  be 
exterminated  in  the  wake  of  the  liv- 
ing human  debris,  such  as  Rattus  rat- 
tus,  that  we  have  allowed  to  pollute 
the  islands  of  the  West  Indies. 


Captive  Breeding  for 
Reintroduction  Purposes 

Captive  propagation  can  figure  sig- 
nificantly in  the  recovery  of  this 
snake  (USFWS  1986)  The  current  co- 
operative breeding  plan  for  E.  monen- 
sis should  be  expanded  to  more 
American  Association  of  Zoological 
Parks  and  Aquarium  member  institu- 


236 


tions,  and  Species  Survival  Plan  des- 
ignation should  be  sought  for  the 
snake  immediately  to  facilitate  ge- 
netic management  of  the  captive 
population. 

For  the  present,  until  genetic 
analysis  has  been  completed,  the 
strategy  of  deme  integrity  mainte- 
nance should  be  continued,  with  St. 
Thomas  founders  and  La  Cordillera 
founders  managed  as  separate  popu- 
lations. Continuous  outcrossing 
within  demes  facilitated  by  a  random 
pair  mating  scheme  should  be  en- 
couraged. Fortunately,  the  first  cap- 
tive breeding  has  already  taken 
place,  the  proximate  factors  critical 
to  reproduction  have  been  identified 
(Tolson  and  Tuebner  1987),  and  there 
is  no  reason  why  the  captive  popula- 
tion cannot  be  expanded  quickly  for 
reintroduction  attempts  within  five 
years. 

I  firmly  believe  that  we  are  finally 
at  the  point  where  we  can  look  for- 
ward to  augmenting  boa  popula- 
tions, rather  than  helplessly  watch 
them  decline. 


Acknowledgments 

This  research  was  conducted  as  part 
of  USFWS  recovery  activities  under 
the  support  of  the  Institute  of  Mu- 
seum Services  conservation  program 
(Grant  IC-70095-87)  and  the  Toledo 
Zoological  Society.  I  am  extremely 
grateful  to  Dr.  David  W.  Nellis,  Divi- 
sion of  Fish  and  Wildlife,  U.S  Virgin 
Islands,  Drs.  Eduardo  R.  Cardona 
and  Jose  A.  Vivaldi,  Departmento  de 
Recursos  Naturales,  Commonwealth 
of  Puerto  Rico,  and  to  Hilda  Diaz- 
Soltero  and  Robert  Pace  of  the 
USFWS  Caribbean  Field  Office  for 
their  counsel  and  logistical  support 
during  the  execution  of  this  project. 

I  thank  Earl  W.  Campbell  III,  Jorge 
L.  Pinero,  and  Carlos  Diez  for  their 
assistance  in  the  field,  which  was  of- 
ten given  under  difficult  conditions. 
C.  Ray  Chandler  and  Earl  W. 
Campbell  III  aided  in  the  statistical 
analysis. 


Literature  Cited 

Barbour,  Thomas  1917.  Notes  on  the 
herpetology  of  the  Virgin  Islands. 
Proceedings  of  the  Biological  Soci- 
ety of  Washington  30:97-104. 

Barbour,  Thomas  1930.  Some  faunis- 
tic  changes  in  the  Lesser  Antilles. 
Proceedings  of  the  New  England 
Zoological  Club  11:73-85. 

Briscoe,  C.B.  1966.  Weather  in  the 
Luquillo  Mountains  of  Puerto 
Rico.  U.S.D.A.  Forest  Service  Re- 
search Paper  ITF-3.  250  p.  Institute 
of  Tropical  Forestry,  Rio  Piedras, 
Puerto  Rico. 

Brown,  William  S.  and  William  S. 
Parker  1976.  A  ventral  scale  clip- 
ping system  for  permanently 
marking  snakes  (Reptilia,  Serpen- 
tes).  Journal  of  Herpetology 
10(3)247-249. 

Campbell,  Howard  W.  and  Frederick 
G.  Thompson  1978.  Observations 
on  a  captive  Mona  Island  boa,  Ep- 
icrates  monensis  monensis  Zenneck. 
Bulletin  of  the  Maryland  Herpeto- 
logical  Society  14(2)98-99. 

Clarke,  Robert  1986.  The  handbook 
of  ecological  monitoring.  298  p. 
Clarendon  Press,  Oxford. 

Cody,  Martin  L.  1974.  Competition 
and  the  structure  of  bird  commu- 
nities. Monographs  in  Population 
Biology  7.  Princeton  University 
Press,  Princeton,  New  Jersey. 

Division  of  Fish  and  Wildlife,  U.S. 
Virgin  Islands  1983.  Rat  control  on 
small  tropical  cays.  26  p.  Unpub- 
lished report. 

Ewel,  Jacob  J.  and  John  .L.  Whitmore 
1973.  The  ecological  life  zones  of 
Puerto  Rico  and  the  U.S.  Virgin 
Islands.  U.S.D.A.  Forest  Service 
Res.  Paper  ITF-18.  71  p.  Institute 
of  Tropical  Forestry,  Rio  Piedras, 
Puerto  Rico. 

Grant,  Chapman  1932.  Herpetology 
of  Tortola;  notes  on  Anegada  and 
Virgin  Gorda,  British  Virgin  Is- 
lands. Journal  of  the  Department 
Agriculture  of  Puerto  Rico  16:327- 
329. 

Heatwole,  Harold,  Levins,  Robert, 
and  Michael  D.  Byer  1981.  Biogeo- 


graphy  of  the  Puerto  Rican  Bank. 
Atoll  Research  Bulletin  No.  251:1- 
55. 

Heatwole,  Harold,  and  Frank  Mack- 
enzie 1967.  Herpetogeography  of 
Puerto  Rico.  IV.  Paleogeography, 
faunal  similarity,  and  endemism. 
Evolution  21:429-438. 

Kaye,  Clifford  A.  1959a.  Shoreline 
features  and  quaternary  shoreline 
changes  in  Puerto  Rico.  U.S.  Geo- 
logical Survey  Professional  Paper 
317-B:49-140. 

Kaye,  Clifford  A.  1959b.  Geology  of 
Isla  Mona,  Puerto  Rico,  and  notes 
on  the  age  of  the  Mona  Passage. 
U.S.  Geological  Survey  Profes- 
sional Paper  317-E:141-178. 

Kiester,  A.  Ross,  George  C.  Gorman, 
and  David  Colon  Arroyo  1975. 
Habitat  selection  behavior  of  three 
species  of  Anolis  lizards.  Ecology 
56:220-225. 

Little,  Elbert  L.  Jr.  and  Frank  H. 
Wads  worth  1964.  Common  trees 
of  Puerto  Rico  and  the  Virgin  is- 
lands. U.S.  Department  of  Agricul- 
ture, Agriculture  Handbook  249. 
548  p.  Washington,  D.C. 

MacLean,  William  P.  1982.  Reptiles 
and  amphibians  of  the  Virgin  Is- 
lands. 54  p.  Macmillan  Caribbean, 
London. 

Mayer,  Gregory  C.  and  James  D, 
Lazell  1988.  Distributional  records 
for  reptiles  and  amphibians  from 
the  Puerto  Rico  Bank.  Herpeto- 
logical  Review  19(l):23-24. 

Matthews,  William  J.  1985.  Distribu- 
tion of  midwestem  fishes  on 
multivariate  environmental  gradi- 
ents, with  emphasis  on  Notropis 
lutrensis.  American  Midland 
Naturalist  113(2):225-237. 

Moulton,  Michael  P.  1985.  Morpho- 
logical similarity  and  coexistence 
of  congeners:  an  experimental  test 
with  introduced  Hawaiian  birds. 
Oikos  44:301-305. 

Nellis,  David  W.,  Norton,  Robert  L., 
and  William  P.  MacLean  1984.  On 
the  biogeography  of  the  Virgin  Is- 
lands boa,  Epicrates  monensis 
granti.  Journal  of  Herpetology, 
17(4):413-417. 


237 


Pregill,  Gregory  K.  1981.  Pleistocene 
herpetofaunas  from  Puerto  Rico. 
72  p.  Miscellaneous  Publications 
Museum  of  Zoology,  University  of 
Michigan,  No.  71. 

Pregill,  Gregory  K.  and  Storrs  L. 
Olson  1981.  Zoogeography  of 
West  Indian  land  vertebrates  in 
relation  to  Pleistocene  cli- 
maticcycles.  Annual  Review  of 
Ecology  and  Systematics  12:75-98. 

Rivero,  Juan  A.,  Rafael  Joglar,  and 
Idabella  Vazquez.  1982.  Cinco 
nuevos  ejemplares  del  culebron  de 
La  Mona  Epicrates  m.  monensis 
(OphidiaiBoidae).  Caribbean  Jour- 
nal of  Science  17(l-4):7-13. 

Rotenberry,  John  T.  and  John  A. 
Wiens  1980.  Habitat  structure, 
patchiness,  and  avian  communi- 
ties in  North  American  shrub- 
steppe  vegetation:  a  multivariate 
analysis.  Ecology  61:1228-1250. 

SAS  Institute  1985.  SAS  users  guide: 
statistics  Version  5.  957  p.  SAS  In- 
stitute Inc.  Gary,  North  Carolina. 

Schmidt,  Karl  P.  1926.  The  amphibi- 
ans and  reptiles  of  Mona  Island, 
West  Indies.  Field  Museum  of 
Natural  History  Publications 
12(12):149-163. 

Sheplan,  Bruce  R.  and  Albert 

Schwartz.  1974.  Hispaniolan  boas 
of  the  genus  Epicrates  and  their 
Antillean  relationships.  Annals  of 
the  Carnegie  Museum  45(5):57- 
143. 

Stull,  Olive  G.  1933.  Two  new  sub- 
species of  the  family  boidae.  4  p. 
(Occasional  Papers  of  the  Museum 
of  Zoology,  University  of  Michi- 
gan, Number  267. 

Tolson,  Peter  J.  1987.  Phylogenetics 
of  the  boid  snake  genus  Epicrates 
and  Caribbean  vicariance  theory. 
68  p.  Occasional  Papers  of  the  Mu- 
seum of  Zoology,  University  of 
Michigan,  Number  115. 

Tolson,  Peter  J.  and  Victoria  A. 
Tuebner  1987.  The  role  of  social 
manipulation  and  environmental 
cycling  in  propagation  of  the  boid 
genus  Epicrates:  Lessons  from  the 
field  and  laboratory.  American 
Association  of  Zoological  Parks 


and  Aquariums  Regional  Confer- 
ence Proceedings  1987:606-613. 

U.S.  Fish  and  Wildlife  Service  1980. 
Status  of  Virgin  Islands  boa  clari- 
fied. Endangered  Species  Techni- 
cal Bulletin,  Department  of  Inte- 
rior, U.S.  Fish  and  Wildlife  Serv- 
ice, Endangered  Species  Program, 
Washington,  D.C.  5:12. 

U.S.  Fish  and  Wildlife  Service  1984. 
Mona  boa  recovery  plan.  14  p. 
U.S.  Fish  and  Wildlife  Service,  At- 
lanta, Ga. 

U.S.  Fish  and  Wildlife  Service  1986. 
Virgin  Island  tree  boa  recovery 
plan.  23  p.  U.S.  Fish  and  Wildlife 
Service,  Atlanta,  Ga. 


Weiwandt,  Thomas  A.  1977.  Behav- 
ior, ecology,  and  management  of 
the  Mona  ground  iguana,  Cyclura 
stegnegeri.  Ph.  D.  dissertation,  Cor- 
nell University,  Ithaca,  New  York. 

Whitaker,  A.  H.  1973.  Lizard  popula- 
tions on  islands  with  and  without 
Polynesian  rats,  Rattus  exulans 
(Peale).  Proceedings  of  the  New 
Zealand  Ecological  Society  20:121- 
130. 

Wiens,  John  A.  and  John  T.  Roten- 
berry 1981.  Habitat  associations 
and  community  structure  of  birds 
in  shrubsteppe  environments.  Eco- 
logical Monographs  5(1)21-41. 


Appendix  A. 
PCA  Variables  Measured  on  island  Study  Plots. 


Variable 


Predator 

Rattus  density 
Felis  presence 
Alsophis  density 
Margarops  density 

Prey 

Anolis  density 
Ameiva  density 
Anolis  perch  height 

Coverage 

Percent  cover  C  trees 
Percent  cover  S  trees 
Percent  cover  palms 
Percent  cover  Opuntia 
Percent  cover  grasses 

Structural 

Vegetational  continuity 

Canopy  height 
Plant  diversity 


Description 


Rats  captured/ trap  hour 

Present  =  1,  absent  =  0 

Mean  no.  Alsophis  observed /day  on  transect 

Mean  no.  Margarops  observed /day  on  transect 

Mean  no.  Anolis/ 5  m  of  transect 
Mean  no.  Ameiva /S  m  of  transect 
Mean  perch  height  in  m  of  male  Anolis 

No.compound  trees/ no.  woody  plants 
No.  single  trees/no.  woody  plants 
No.  palms/no.  woody  plants 
No.  Opuntia/no.  woody  plants 
Grassland  area /total  area 

Contiguous  =  1,  high  =  .75,  Moderate  =  .5 

low  =  .25,  absent  =  0 

>3  m  =  1, 1-2  m  =  .5,  <1  m  =  0 

No.  of  dominant  plant  species  on  plot 


238 


The  Use  of  Timed  Fixed-Area 
Plots  and  a  Mark-Recapture 
Technique  In  Assessing 
Riparian  Garter  Snake 
Populations^ 

Robert  C.  Szaro,^  Scott  C.  Belfit,^  J.  Kevin 
Aitkin/  and  Randall  D.  Babb^ 


Abstract.— Wandering  garter  snake  (Thamnophis 
elegans  vagrans)  populations  along  a  thin-leaf  alder 
(AInus  tenuifolia)  riparian  community  in  northern 
New  Mexico  were  sampled  using  timed  fixed-area 
plots  and  a  mark- recapture  method.  Both  methods 
served  to  determine  yearly  differences  and  relative 
magnitude  of  snake  density  between  years.  But 
population  estimates  determined  by  timed  fixed- 
area  plots  were  inconsistent  between  study  plots  in 
the  same  year. 


Research  studies  often  attempt  to  de- 
termine the  effects  of  disturbance  or 
management  regimes  on  the  abun- 
dance of  wildlife  species  (Cooper- 
rider  et  al.  1986,  Fitch  1987,  Parker 
and  Plummer  1987,  Ralph  and  Scott 
1980).  How  well  the  method  of  data 
collection  and  analyses  reflect  actual 
populations  is  critically  important  for 
assessing  the  validity  of  these  stud- 
ies. Snakes  are  difficult  subjects  for 
field  studies  because  of  their  secre- 
tive and  cryptic  habits  (Fitch  1987). 

Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptile,  and 
Small  Mammais  in  North  America.  (Flag- 
staff, AZ,  July  19-21,  1988.) 

^Robert  C.  Szaro  is  Research  Wildlife  Bi- 
ologist. USD  A  Forest  Service.  Roclcy  Moun- 
tain Forest  and  Range  Experiment  Station. 
Arizona  State  University  Campus.  Tempe,  AZ 
85287-1304. 

^ Scott  C.  Belfitis  Wildlife  Biologist,  De- 
partment of  the  Army,  Wildlife  Manage- 
ment Section,  Fort  Huachuca,  AZ  856 13- 
6000.  Beint's  current  address  is  P.O.  Box  336, 
Fort  Belvoir,  VA  22060-0336. 

^  J.  Kevin  Aitkin  is  Wildlife  Technician, 
USDA  Forest  Service.  Rocky  Mountain  Forest 
and  Range  Experiment  Station.  Arizona 
State  University  Campus,  Tempe.  AZ  85287- 
1304. 

^Randall  D.  Babb.  formerly  Wildlife  Tech- 
nician. USDA  Forest  Service.  Rocky  Moun- 
tain Forest  and  Range  Experiment  Station. 
Arizona  State  University  Campus.  Tempe.  AZ 
85287- 1304  is  currently  Sportfishing  Program 
Coordinator,  Arizona  Game  and  Fish  De- 
partment, 2222  West  Greenway  Road, 
Phoenix,  AZ  85023. 


Many  attempts  to  census  snakes  have 
been  inaccurate  (Turner  1977,  Fitch 
1987).  Population  estimates  can  be 
influenced  by  sex,  reproductive  con- 
dition, and  stage  of  maturity,  all  of 
which  are  critical  determinants  of 
activity  within  species  (Gibbons  and 
Semlitsch  1987).  Differences  among 
juveniles  and  breeding  and  non- 
breeding  females,  and  males  often 
lead  to  much  different  risks  of  cap- 
ture at  various  stages  of  the  season 
and  time  of  day.  Overall  population 
estimates  can  be  distorted  as  a  result, 
requiring  separate  estimates  by  sex 
and  age  class  (Fitch  1987). 

Two  methods  often  used  to  esti- 
mate snake  density  are  direct  counts 
and  mark-recapture  analyses.  Sys- 
tematic searches  of  defined  areas  (di- 
rect counts)  yield  species  occurrence 
data,  and  usually  require  less  time 
and  effort  than  mark-recapture  meth- 
ods (Jones  1986).  Using  direct  counts. 
Bury  and  Luckenbach  (1977)  success- 
fully censused  desert  tortoise  (Go- 
pherus  agassizii)  populations  with  a 
quartet  and  grid  location  system. 
Bury  (1982)  used  a  removal  method 
to  assess  reptile  community  structure 
in  the  Mohave  Desert  (Zippin  1956, 
1958).  Bury  and  Raphael  (1983)  refer 
to  searches  conducted  per  unit  effort 
of  time  as  time-constraint  proce- 
dures. Usually  it  is  impossible  to  find 
every  snake  in  an  area,  making  it 
necessary  to  estimate  population  size 


from  capture-recapture  ratios  (Fitch 
1987).  Yet,  when  several  density  esti- 
mates become  available  from  the 
same  area  at  different  times,  they  of- 
ten show  such  drastic  discrepancies 
that  the  basic  methods  have  been 
thought  invalid  (Turner  1977). 
Turner  (1977)  had  no  confidence  in 
the  density  estimates  for  snakes  de- 
rived from  mark-recapture  tech- 
niques. However,  since  his  critical 
review,  estimation  techniques  have 
greatly  improved  with  the  develop- 
ment of  models  and  computer  pro- 
grams that  test  model  assumptions 
and  estimate  standard  errors  (Ar- 
nason  and  Baniuk  1980,  White  et  al. 
1978, 1982,  Otis  et  al.  1978,  Brownie 
et  al.  1985). 

Although  time  consuming,  deter- 
mining accurate  pxDpulation  estimates 
is  necessary  to  develop  management 
polices  not  only  for  abundant  spe- 
cies, such  as  the  wandering  garter 
snake  (Thamnophis  elegans  vagrans), 
but  also  for  aquatic  or  semi-aquatic 
endangered  snake  species  such  as  the 
Concho  water  snake  (Nerodia  harteri 
paucimaculata)  (Scott  and  Fitzgerald 
1985)  and  the  narrow-headed  garter 
snake  (Thamnophis  rufipunctatus) 
(Lowe  1985).  Flowever,  because  the 
wandering  garter  snake,  is  less  secre- 
tive than  most  kinds  of  snakes,  and  is 
concentrated  in  riparian  habitats,  it  is 
probably  one  of  the  best  adapted  to 
this  sort  of  investigation  (Fitch,  per- 


239 


sonal  communication).  The  results  of 
this  work  should  be  directly  appli- 
cable to  other  snake  species  normally 
concentrated  in  riparian  ecosystems 
and  may  be  especially  useful  for  cen- 
susing  endangered  species  where 
large  samples  to  determine  the  accu- 
racy of  sampling  techniques  are  not 
available.  Our  previous  work 
showed  the  inadequacy  of  simple 
transects  and  depletion  sampling  in 
determining  garter  snake  popula- 
tions along  the  Rio  de  las  Vacas,  New 
Mexico  (Szaro  et  al.  1985).  The  objec- 
tive of  this  study  was  to  compare 
timed  fixed-area  plots  and  a  mark- 
recapture  technique  in  assessing  the 
impacts  of  management  regimes  on 
riparian  ecosystems  in  the  arid 
Southwest  by  sampling  wandering 
garter  snake  populations  along  the 
Rio  de  las  Vacas. 


Methods  and  Study  Areas 

The  Rio  de  las  Vacas,  is  a  montane 
stream  draining  the  San  Pedro  Parks 
Wilderness  Area,  Santa  Fe  National 
Forest,  New  Mexico.  Under  low  flow 
conditions,  stream  width  ranges 
from  2.8  to  10.5  m  and  averages  7.6 
m.  The  study  area  is  17  km  southeast 
of  Cuba,  in  Sandoval  County,  at  2600 
m.  Two  cattle  exclosures  enclosing 
stream  reaches  (each  about  1  km  long 
by  50  m  wide)  were  installed  in  the 
early  1970's  (Szaro  et  al.  1985).  Con- 
tiguous, downstream  areas,  privately 
owned  and  grazed  by  livestock,  were 
used  for  comparison.  The  most  ap- 
parent difference  between  the  grazed 
and  exclosed  stream  segments  was 
the  band  of  small  riparian  trees  and 
shrubs  in  the  exclosures  (figs.  1  and 
2).  Thin-leafed  alder  (Alnus  tenuifolia) 
and  a  mixture  of  willow  species 
{Salix  spp.)  edged  the  exclosure 
streambanks  but  were  widely  scat- 
tered where  the  streambanks  were 
grazed  (9.5  ±  1.16,  7.5  ±  1.23,  and  0.3 
+  0.14  trees/250  m^  in  exclosures  1,  2, 
and  grazed  areas,  respectively). 

Snake  populations  were  estimated 
by  timed  fixed-area  plot  sampling. 


Figure  1  .—Grazed  section  of  the  Rb  de  las  Vacas,  New  Mexico.  Notice  the  lack  of  shrub 
growth  and  the  unstable  stream  banks. 


and  mark-recapture  sampling  in  both 
grazed  and  ungrazed  areas.  For  the 
former,  16  plots  (10  x  25  m),  with  the 
long  edge  being  defined  by  the 
stream  bank,  were  intensively 
sampled  for  20  minutes  in  each  of  the 


two  ungrazed  exclosures  and  one 
grazed  stream  segment  along  the  Rio 
de  las  Vacas,  for  a  total  of  48  plots 
(fig.  3).  During  sample  periods  we 
turned  rocks,  logs,  debris  piles,  and 
generally  searched  the  area.  All  plots 


Figure  2.— Shrubby  growth  in  Exclosure  2  along  the  Rio  de  las  Vacas,  New  Mexico. 


240 


were  sampled  once  between  0900 
and  1300  hours  (MST)  within  a  3-day 
period  each  month.  Sampling  times 
were  determined  from  preliminary 
activity  period  sampling  that  showed 
two  distinct  periods  of  activity 
(morning  and  late  afternoon).  All 
snakes  captured  were  placed  in  a 
cloth  sack  at  their  point  of  capture, 
until  the  end  of  the  sampling  p>eriod. 
Plot  sampling  began  in  June  1984  and 
was  replicated  in  July,  August,  and 
September  of  that  year  and  in  the 
same  months  in  1985.  Total  time 
spent  sampling  was  approximately 
64  hours  per  year,  excluding  time  be- 
tween samples  to  process  snakes. 

For  mark-recapture  estimates,  we 
searched  the  entire  extent  of  both  ex- 
closures  and  a  similarly  sized  down- 
stream grazed  stream  area.  The  plots 
used  for  the  timed-fixed  plot  sam- 
pling were  a  subset  of  the  area  used 
for  the  mark-recapture  sampling.  All 
captured  snakes  were  marked  by 
clipping  three  subcaudal  scales  (Blan- 
chard  and  Finster  1933,  Woodbury 
1956).  Mark-recapture  sampling  peri- 
ods occurred  in  the  same  months  as 
the  plot  sampling;  but  snakes  were 
captured,  marked,  and  released  dur- 


ing intensive  searches  for  6  consecu- 
tive days  by  3  to  4  collectors.  All 
snakes  were  released  where  cap- 
tured. Approximately  equal  time  and 
effort  was  spent  searching  for  snakes 
in  each  of  the  three  areas.  Time  of 
day  bias  was  minimized  by  alternat- 
ing starting  areas  daily.  Sampling 
began  at  0900  hours  (MST)  and  con- 
tinued until  dusk.  Only  captures 
within  10  m  of  the  stream  were  used 
in  the  mark-recapture  analyses  to  al- 
low a  direct  comparison  to  plot  sam- 
pling estimates.  Thus,  the  plot  sam- 
pling represents  a  sample  within  the 
exclosures  and  the  grazed  stream 
area,  whereas  the  mark-recapture 
sampling  represents  an  ''open" 
population  estimate  of  each  study 
area.  Total  time  sp>ent  sampling  and 
marking  snakes  was  approximately 
450  hours  per  year  including  time  to 
process  snakes. 

The  approach  to  mark-recapture 
analysis  was  to  analyze  each  year 
separately  using  closed  population 
models  calculated  by  program  CAP- 
TURE, which  allows  unequal  catch- 
ability  (Otis  et  al.  1978,  White  et  al. 
1978, 1982)  as  recommended  by  Pol- 
lock (1981, 1982).  Because  we  were 


unable  to  estimate  survival  using  the 
timed  fixed-area  plots,  we  do  not 
present  these  estimates  here  for  the 
mark-recapture  analysis.  However, 
all  sampling  periods  were  pooled 
and  survival  estimators  between 
years  estimated  using  the  Jolly-Seber 
Model  (Seber  1986,  Szaro  et  al.,  in 
preparation). 

Inferences  about  differences  be- 
tween years  and  exclosures  were 
based  on  Bonferroni's  method  for 
multiple  comparisons  by  fixing  the 
experimentwise  error  rate  at  0.05 
(Milliken  and  Johnson  1984).  Thus, 
the  overall  experimentwise  error  rate 
is  less  than  P  (in  this  case  0.05);  but 
for  each  comparison,  the  compari- 
son wise  error  rate  is  equal  to  P/n, 
where  n  is  the  number  of  compari- 
sons. For  example,  with  3  compari- 
sons the  actual  P  value  per  compari- 
son would  be  0.05/3  or  0.017. 


Results 

We  are  confident  the  mark-recapture 
estimates  accurately  reflect  popula- 
tion densities  on  the  three  study  ar- 
eas and  use  these  as  the  basis  for 


Rio  de  los  Voces 

% 

Iidosure  1 

Eidosvt  2               1  n 

1    ol  i 
1     ^10  I 
1    bIo  I 

1      K  1 

Grazed 

1  KU 

1  1 

/  Dev.  2600  m 

Figure  3.— Study  areas  and  sample  plot  lay- 
out along  thie  Rio  de  las  Vacas,  New  Mex- 
ico. 


Table  1  .—Population  estimates  of  the  wandering  garter  snake  (Thamnophis 
elegrans  vagrans)  In  1984  and  1985  within  10  m  of  the  streambank  at  Rio 
de  las  Vacas,  New  Mexico. 


Mark-recapture' 


Times  fixed-area  plot^ 


Study  area 

Year 

Mean 

S.E.  Sig.3 

Mean 

S.E. 

SIg. 

Exclosure  1 

1984 

282 

+ 

23.53  (3.86)^  a 

1.28 

+ 

0.18 

a 

1985 

166 

15.51  (2.28)  b 

0.88 

0.11 

a 

Exclosure  2 

1984 

296 

24.42(4.53)  a 

1.30 

0.17 

a 

1985 

146 

+ 

13.92  (2.23)  b 

0.45 

± 

0.09 

c 

Grazed 

1984 

67 

+ 

10.49(1.00)  c 

0.28 

+ 

0.07 

cd 

1985 

26 

+ 

5.22  (0.39)  d 

0.11 

± 

0.04 

d 

'Mark-recapture  estimates  for  each  stvdy  area  are  for  ttie  total  population  using 
tt)e  best  model  in  CAPTURE  for  whicti  solutions  exist.  The  total  area  sampled  in  each 
area  was  16,240  m^  in  Exclosure  L  16,340     in  Exclosure  2,  and  16,760  m^  in  the 
grazed  area. 

'Plot  samples  are  mean  number  of  snakes  caught  per  250  m', 
^Population  estimates  by  each  method  that  do  not  have  a  letter  in  common  are 
significantly  different  (Bonferroni's  method.  P<0.05). 

^Number  in  parenthesis  is  estimated  number  of  snakes  per  250  mP  using  the  mark- 
recapture  population  estimate. 


241 


comparison  for  the  timed  fixed-area 
plot  results.  Mark-recapture  esti- 
mates were  based  on  118  individuals 
and  35  recaptures  (118/35)  in  exclo- 
sure  1  in  1984,  72/28  in  1985, 127/30 
in  exclosure  2  in  1984,  74/26  in  1985, 
12/2  in  the  grazed  area  in  1984,  and 
10/1  in  1985. 

We  asked  two  questions  of  the 
sampling  methods.  First,  were  there 
any  differences  in  population  esti- 
mates between  years?  Both  methods 
indicated  decreases  in  populahon 
size  on  all  three  areas  between  1984 
and  1985.  However,  yearly  differ- 
ences were  significant  only  for  mark- 
recapture  estimates  and  for  the  timed 
fixed-area  plot  estimates  in  exclosure 
2  (P  <  0.05)  (table  1).  Mark-recapture 
estimates  revealed  that  snake  popu- 
lations decreased  by  41%  to  54% 
from  1984  to  1985  in  all  study  areas. 
Decreases  in  mean  number  of  snakes 
per  fixed-area  plot  were  not  as  uni- 
form, varying  from  31%  on  exclosure 
1  to  65%  on  exclosure  2  and  the 
grazed  stream  segment. 

Second,  were  there  differences  be- 
tween the  study  areas?  Population 
estimates  between  exclosures  and  the 
grazed  stream  segment  within  a 
given  year  were  significantly  differ- 
ent by  both  census  methods  and  for 
both  years  (P  <  0.05)  (table  1).  Popu- 
lation estimates  by  both  methods 
were  not  significantly  different  be- 
tween exclosures,  except  in  1985 
when  the  estimate  determined  by 
timed  fixed-area  plots  for  exclosure  2 
was  50%  of  that  on  exclosure  1  (P  < 
0.05)  (table  1). 

Estimating  population  size  by  re- 
stricting the  mark-recapture  esti- 
mates to  a  10  m  band  on  either  side 
of  the  stream  served  a  twofold  pur- 
pose. First,  it  allowed  us  to  estimate 
the  number  of  snakes  per  unit  area. 
Second,  it  made  estimates  by  both 
techniques  more  readily  comparable, 
because  all  plot  sampling  was  con- 
fined to  the  10-m  band  next  to  the 
stream  where  most  of  the  available 
down  litter,  grass  clumps,  and 
shrubby  vegetation  was  concen- 
trated. In  exclosure  1,  there  were  3.86 


and  2.28  snakes  per  250  m^  in  1984 
and  1985,  respectively.  In  exclosure 
2,  there  were  4.53  and  2.23  snakes 
per  250  m^  in  1984  and  1985,  respec- 
tively. Along  the  grazed  stream  reach 
there  were  1.00  and  0.38  snakes  per 
250  m^  in  1984  and  1985,  respectively. 
Based  on  these  estimates,  we  caught 
between  20.2%  (exclosure  2, 1985) 
and  38.6%  (exclosure  1,  1985)  of  the 
snakes  present  in  the  exclosures.  On 
the  grazed  area  we  caught  28%  of  the 
snakes  in  both  1984  and  1985. 


Discussion 

Apparent  short-term  downward 
population  fluctuations  averaging 
about  50%  have  been  found  in  sev- 
eral mark-recapture  studies  (Fukada 
1969,  Piatt  1969,  Fitch  1975,  Feaver 
1977,  Gregory  1977).  Many  studies  of 
snakes  have  related  population 
changes  over  several  years  to  succes- 
sional  changes  (Clark  1970,  Fitch 
1982)  or  to  environmental  factors, 
such  as  decreases  in  annual  precipi- 
tation (Clark  1974,  Clark  and  Fleet 
1976).  Another  possibility,  is  that  a 
study  like  this  actually  destroys  hid- 
ing places  (turning  rocks,  logs,  etc.); 
and  even  if  each  piece  is  put  back 
carefully,  the  site  has  opened  up  and 
changed  (Clark,  personal  communi- 
cation). 

We  undoubtedly  had  some  impact 
on  the  quality  of  the  available  habitat 
by  our  intensive  searching  tactics; 
but  we  did  try  to  be  as  careful  as  pos- 
sible to  return  moved  objects  back 
into  their  original  positions.  Parker 
and  Plummer  (1987)  suggest  that 
these  apparent  fluctuations  in  den- 
sity result  from  changes  in  activity 
level  (which  affect  recapture  proba- 
bilities) rather  than  from  actual 
changes  in  density  (Lillywhite  1982, 
Pough  1983).  There  are  three  possible 
explanations  for  these  results:  (1) 
snakes  simply  moved  out  of  the  plot 
and  exclosure  areas;  (2)  snakes  be- 
came inactive  in  burrows  or  cover 
sites  because  of  environmental  condi- 
tions; or  (3)  snakes  died. 


Activity  periods  of  wandering  gar- 
ter snakes  varied  between  individu- 
als from  our  preliminary  sample  of 
wandering  garter  snake  populations 
along  the  Rio  de  las  Vacas  in  July 
1983.  We  failed  to  decrease  signifi- 
cantly the  total  numbers  of  animals 
caught  per  plot  even  after  3  days  of 
removal  sampling  (Szaro  et  al.  1985); 
but  at  other  times  snakes  were  diffi- 
cult to  find.  However,  we  feel  the  in- 
tensive sampling  effort  of  at  least  1 
week  each  month  minimized  the  ef- 
fect of  changes  in  snake  behavior  on 
population  estimates. 

The  almost  50%  difference  in  1985 
between  exclosures  in  mean  number 
of  snakes  caught  while  plot  sampling 
was  probably  a  result  of  a  shift  in  ar- 
eas used  by  the  snakes  and  not  dif- 
ferences in  mortality  between  the 
two  exclosures.  Monthly  trends  in 
total  number  of  snakes  caught  also 
showed  a  dramatic  difference  in  the 
number  of  snakes  caught  per  month 
while  plot  sampling  in  both  exclo- 
sures. However,  this  difference  was 
not  reflected  in  the  overall  number  of 
snakes  caught  during  mark-recapture 
sampling  (fig.  4).  In  fact,  overall  we 
caught  more  snakes  in  exclosure  2 
than  in  exclosure  1  in  all  months  in 
1985. 

The  difference  in  plot  sampling  es- 
timates between  exclosures  in  1985 
was  not  a  result  of  changes  in  daily 
activity  patterns,  because  equal  pro- 
portions of  snake  captures  in  both 
exclosures  were  before  1300  (63%  in 
exclosure  1  and  59%  in  exclosure  2, 
chi-square,  P  >  0.05).  Furthermore, 
differences  in  captures  between  years 
and  methods  were  not  sex-based,  be- 
cause there  were  no  significant  dif- 
ferences in  sex  ratios  between  years 
or  method  in  a  given  study  section 
(chi-square,  P  >  0.05)  (fig.  5).  How- 
ever, there  were  distributional  differ- 
ences in  snake  captures  between 
years  and  exclosures. 

In  1984,  34.6%  and  34.7%  of  all 
captures  on  exclosures  1  and  2,  re- 
spectively were  made  on  the  plot  ar- 
eas. In  contrast,  42.1%  and  20.6%  of 
all  captures  on  exclosures  1  and  2,  re- 


242 


Plot  Sampling 


June      July       Aug.      Sept.      June      July       Aug.  Sept 
1984  1985 


Sampling  Period 


Figure  4.— Total  numbers  of  wandering  garter  snakes  caught  in  June,  July,  August,  and  Sep- 
tember 1984  and  1985  along  \he  Rio  de  las  Vacas,  New  Mexico  during  timed  fixed-area  plot 
and  mark-recapture  sampling. 


specrively,  were  made  on  the  plot 
areas  in  1985. 

We  cannot  explain  this  distribu- 
tional shift  in  exclosure  2.  Although 
we  did  not  plot  sample  in  1986  and 
1987,  mark-recapture  efforts  in  those 
years  showed  a  similar  distributional 
pattern  (Szaro  et  al.,  unpublished).  In 
exclosure  1,  33.0  %  and  37.3%  of  all 
captures  in  1986  and  1987,  respec- 
tively were  on  the  old  plot  areas, 
whereas  in  exclosure  2,  these  values 
were  10.07o  and  9.8%. 

We  feel  that  the  distributional 
changes  in  exclosure  2  were  not  an 
artifact  of  plot  sampling,  because 
snakes  in  exclosure  2  did  not  return 
to  plot  areas  after  plot  sampling  had 
stopped.  In  any  case,  our  sampling 
potentially  would  have  been  more 
destructive  in  exclosure  1  than  in  ex- 
closure  2  because  of  the  higher  inci- 
dence of  turnable  rocks  in  that  exclo- 
sure. 

Whatever  the  cause,  these  changes 
in  distribution  indicate  that  initial 
randomized  selection  of  plots  did  in- 
fluence density  estimates  for  exclo- 
sure 2.  Although  it  would  increase 
substantially  the  amount  of  time  nec- 
essary to  adequately  sample  vegeta- 
tion, a  better  approach  would  be  to 
randomly  select  plots  within  exclo- 
sures  each  sampling  period  rather 
than  repeatedly  sampling  the  same 
plots. 

In  conclusion,  the  use  of  timed 
fixed-area  plots  enabled  us  to  quan- 
tify dramatic  differences  in  snake 
abundance  between  exclosures  and 
the  grazed  area.  However,  this  sam- 
pling method  is  of  questionable  merit 
because  of  the  significant  difference 
in  exclosure  population  estimates  for 
1985.  Further  study  incorporating 
newly  randomized  plots  for  each 
sampling  period  may  solve  this  prob- 
lem. Care  should  be  taken  to  deter- 
mine if  snakes  are  distributing  them- 
selves in  a  nonrandom  pattern.  At 
this  time,  we  recommend  the  more 
labor-intensive  mark-recapture  esti- 
mators for  assessing  the  impacts  of 
riparian  management  regimes  on 
snake  populations. 


243 


Acknowledgments 

We  thank  D.  R.  Clark,  H.  S.  Fitch,  K. 
B.  Jones,  and  N.  J.  Scott,  Jr.  for  their 
constructive  reviews  of  this  paper.  H. 
Berna,  M.  Cady,  C.  Engel- Wilson,  X. 
Hernandez,  D.  Johnson,  M.  Lane,  W. 
Legarde,  L.  Simon,  and  D.  Smith 
aided  in  the  collection  of  the  field 
data.  Special  thanks  to  Jim  and  Mary 
Bedeaux  for  their  gracious  hospital- 
ity and  allowing  us  to  sample  on 
their  property. 

Liferature  Cited 

Arnason,  A.  Neil  and  Leonard 
Baniuk.  1980.  A  computer  system 
for  mark-recapture  analysis  of 
open  populations.  Journal  of  Wild- 
life Management  44:325-332. 

Blanchard,  Frank  N.  and  Ethel  B.  Fin- 
ster.  1933.  A  method  of  marking 
living  snakes  for  future  recogni- 
tion, with  a  discussion  of  some 
problems  and  results.  Ecology 
14:334-347. 

Brownie,  Cavell,  David  R.  Anderson, 
Kenneth  P.  Bumham,  and  Douglas 
S.  Robson.  1985.  Statistical  Infer- 
ence from  band  recovery  data — a 
handbook.  USDI  Fish  and  Wildlife 
Service  Resource  Publication 
156:1-305. 

Bury,  R.  Bruce.  1982.  Structure  and 
composition  of  Mojave  Desert  rep- 
tile Communities  determined  with 
a  removal  method.  Pages  135-142. 
In  Norman  J.  Scott,  Jr.,  Editor, 
Herpetological  Communities. 
USDI  Fish  and  Wildlife  Service 
Wildlife  Research  Report  13. 

Bury,  R.  Bruce  and  R.  A.  Lucken- 
bach.  1977.  Censusing  desert  tor- 
toise populations  using  a  quartet 
and  grid  location  system.  Desert 
Tortoise  Council  Symposium  Pro- 
ceedings 1977:169-178. 

Bury,  R.  Bruce  and  Martin  G.  Ra- 
phael. 1983.  Inventory  methods 
for  amphibians  and  reptiles.  Pages 
416-419.  In  J.  F.  Bell  and  T.  Atter- 
bury.  Editors,  Renewable  resource 
inventories  for  monitoring 


changes  and  trends:  Proceedings 
of  an  international  conference. 
[Corvallis,  Oregon,  August  15-19, 
1983]  Society  of  American  Forest- 
ers 83-14.. 
Clark,  Donald  R.,  Jr.  1970.  Ecological 
study  of  the  worm  snake,  Car- 
phophis  vermis  (Kennicott).  Univer- 
sity of  Kansas  Publications  of  the 


Museum  of  Natural  History  19:85- 
194. 

Clark,  Donald  R.,  Jr.  1974.  The  west- 
em  ribbon  snake  (Thamnophis 
proximus):  ecology  of  a  Texas 
population.  Herpetologica  30:372- 
379. 

Clark,  Donald  R.,  Jr.  and  R.  R.  Fleet. 
1976.  The  rough  earth  snake  (Vir- 

Exclosure  1 


1984 


1985 


1984 


o 


E 
o 


o 
>- 


1985 


1984 


1985 


40  60 
Exclosure  2 


Grazed  Stream  Reach 


0  20  40  60 

Percent  of  Total  Observations 


•..  1  Femole 


100 


Figure  5.— Proportion  of  totai  captures  by  sex  during  timed  fixed-area  plot  and  mark-recap- 
ture sampling  in  two  exclosures  and  a  grazed  stream  reach)  along  \he  Rio  de  las  Vacas, 
New  Mexico,  1984  and  1985. 


244 


ginia  striatula):  ecology  of  a  Texas 
population.  Southwestern  Natu- 
ralist 20:467-478. 

Cooperrider,  Allen  Y.,  Raymond  J. 
Boyd,  and  Hanson  R.  Stuart,  Edi- 
tors. 1986.  Inventory  and  monitor- 
ing of  wildlife  habitat.  U.S.  De- 
partment of  Interior,  Bureau  of 
Land  Management  Service  Center, 
Denver,  Colorado.  858  p. 

Feaver,  P.  E.  1977.  The  demography 
of  a  Michigan  population  of 
(Natrix  sipedon)  with  discussions  of 
ophidian  growth  and  reproduc- 
tion. Ph.D.  Dissertation.  Univer- 
sity of  Michigan,  Ann  Arbor. 

Fitch,  Henry  S.  1975.  A  demographic 
study  of  the  ringneck  snake  (Di- 
adophis  punctatus)  in  Kansas.  Uni- 
versity of  Kansas  Museum  of 
Natural  History  Miscellaneous 
Publications  62:1-53. 

Fitch,  Henry  S.  1982.  Resources  of  a 
snake  community  in  prairie-wood- 
land habitat  of  northeastern 
Kansas.  In  Norman  J.  Scott,  Jr., 
Editor,  Herpetological  Communi- 
ties. USDI  Fish  and  Wildlife  Serv- 
ice Wildlife  Research  Report  13:83- 
97. 

Fitch,  Henry  S.  1987.  Collecting  and 
life-history  techniques.  Pages  143- 
164.  In  Richard  A.  Seigel,  Joseph  T. 
Collins,  and  Susan  S.  Novak,  Edi- 
tors, Snakes:  Ecology  and  evolu- 
tionary biology.  MacMillan  Pub- 
lishing Company,  New  York. 
529  p. 

Fukada,  H.  1969.  Biological  studies 
on  the  snakes.  XIII.  Preliminary 
estimate  of  population  size  in 
Tambabashi  study  area.  Bulletin  of 
Kyoto  University  Educational  Se- 
ries B  Mathematics  and  Natural 
Sciences  36:3-9. 

Gibbons,  J.  Whitfield  and  Raymond 
D.  Semlitsch.  1987.  AcUvity  pat- 
terns. Pages  396-421.  In  Richard  A. 
Seigel,  Joseph  T.  Collins,  and 
Susan  S.  Novak,  Editors,  Snakes: 
Ecology  and  evolutionary  biology. 
MacMillan  Publishing  Company, 
New  York.  529  p. 

Gregory,  P.  T.  1977.  Life  history 
parameters  of  the  red-sided  garter 


snake  (Thamnophis  sirtalis  parietalis) 
in  the  Interlake  of  Manitoba.  Na- 
tional Museum  of  Canada  Publica- 
tions in  Zoology  13:1-44. 

Jones,  K.  Bruce.  1986.  Amphibians 
and  reptiles.  Pages  267-290.  In  Al- 
len Y.  Cooperrider,  Raymond  J. 
Boyd,  and  Hanson  R.  Stuart,  Edi- 
tors, Inventory  and  monitoring  of 
wildlife  habitat.  U.S.  Department 
of  Interior,  Bureau  of  Land  Man- 
agement Service  Center,  Denver, 
Colorado.  858  p. 

Lillywhite,  Harvey  B.  1982.  Tracking 
as  an  aid  in  ecological  studies  of 
snakes.  In  Norman  J.  Scott,  Jr., 
Editor,  Herpetological  Communi- 
ties. USDI  Fish  and  Wildlife  Serv- 
ice Wildlife  Research  Report 
13:181-191. 

Lowe,  Charles  H.  1985.  Amphibians 
and  reptiles  in  southwest  riparian 
ecosystems.  Pages  339-341.  In  R. 
Roy  Johnson,  Charles  D.  Ziebel, 
David  R.  Patton,  Peter  F.  Ffolliott, 
and  Robert  H.  Hamre,  Technical 
Coordinators,  Riparian  ecosys- 
tems and  their  management:  rec- 
onciling conflicting  uses.  First 
North  American  Riparian  Confer- 
ence. USDA  Forest  Service  Gen- 
eral Technical  Report  RM-120. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Ft.  Collins, 
Colo.  523  p. 

Milliken,  George  A.  and  Dallas  E. 
Johnson.  1984.  Analysis  of  messy 
data.  Volume  1:  designed  experi- 
ments. Van  Nostrand  Reinhold 
Company,  New  York.  473  p. 

Otis,  David  L.,  Kenneth  P.  Burnham, 
Gary  C.  White,  and  David  R.  An- 
derson. 1978.  Statistical  inference 
from  capture  data  on  closed  ani- 
mal populations.  Wildlife  Mono- 
graphs 62:1-135. 

Parker,  William  S.  and  Michael  V. 
Plummer.  1987.  Population  ecol- 
ogy. Pages  253-301.  In  Richard  A. 
Seigel,  Joseph  T.  Collins,  and 
Susan  S.  Novak,  Editors,  Snakes: 
Ecology  and  evolutionary  biology. 
MacMillan  Publishing  Company, 
New  York.  529  p. 

Piatt,  D.  R.  1969.  Natural  history  of 


the  hognose  snakes,  Heterodon  plat- 
yrhinos  and  Heterodon  nasicus.  Uni- 
versity of  Kansas  Publications  of 
the  Museum  of  Natural  History 
18:253-420. 

Pollock,  K.  H.  1981.  Caphire-recap- 
ture  models:  a  review  of  current 
methods,  assumptions,  and  ex- 
perimental design.  Studies  in 
Avian  Biology  6:426-435. 

Pollock,  Kenneth  H.  1982.  A  capture- 
recapture  design  robust  to  un- 
equal probability  of  capture.  Jour- 
nal of  Wildlife  Management 
46:752-757. 

Pough,  F.  H.  1983.  Amphibians  and 
reptiles  as  low-energy  systems. 
Pages  141-188.  In  W.  P.  Aspey  and 
S.  I.  Lustick,  Editors,  Behavioral 
energetics:  the  cost  of  survival  in 
vertebrates.  Ohio  State  University 
Press,  Columbus. 

Ralph,  C.  John  and  J.  Michael  Scott, 
Editors.  1980.  Estimating  numbers 
of  terrestrial  birds.  Studies  in 
Avian  Biology  6:1-630. 

Scott,  Norman  J.,  Jr.  and  Lee  H.  Fitz- 
gerald. 1985.  Status  survey  of 
Nerodia  harteri,  Brazos  and  Chon- 
cho — Colorado  Rivers,  Texas. 
Status  Survey  Report.  U.S.  Fish 
and  Wildlife  Service,  National 
Ecology  Center,  Museum  of 
Southwestern  Biology,  University 
of  New  Mexico,  Albuquerque, 
New  Mexico.  44  p. 

Seber,  G.  A.  F.  1986.  A  review  of  esti- 
mating animal  abundance.  Bio- 
metrics 42:267-292. 

Szaro,  Robert  C,  Scott  C.  Belfit,  J. 
Kevin  Aitkin,  and  John  N.  Rinne. 
1985.  Impact  of  grazing  on  a  ripar- 
ian garter  snake.  Pages  359-363.  In 
R.  Roy  Johnson,  Charles  D.  Ziebel, 
David  R.  Patton,  Peter  F.  Ffolliott, 
and  Robert  H.  Hamre,  Technical 
Coordinators,  Riparian  ecosys- 
tems and  their  management:  rec- 
onciling conflicting  uses.  First 
North  A.merican  Riparian  Confer- 
ence. USDA  Forest  Service  Gen- 
eral Technical  Report  RM-120. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Ft.  Collins, 
Colo.  523  p. 


245 


Turner,  F.  B.  1977.  The  djmamics  of 
populations  of  squamates,  croco- 
dilians  and  rhynchocephalians. 
Pages  157-264.  In  C.  Cans  and  D. 
W.  Tinkle,  Editors,  Biology  of  Re|> 
tilia,  volume  7.  Academic  Press. 


I 

Wh 
1 
1 

c 

\ 

J 
8 
2 

Wh 
I 

d 


3 


F 
L 

4 

Wo( 
n 

sU 

reptiles.  Ecology  37:670-674. 

Zippin,  C.  1956.  An  evaluation  of  the 
removal  method  of  estimating  ani- 
mal populations.  Biometrics 
12:163-189. 

Zippin,  C.  1958.  The  removal  method 
of  population  estimation.  Journal 
of  Wildlife  Management  22:82-90. 


246 


Design  Considerations  for  the 
Study  of  Amphibians,  Rep- 
tiles, and  Small  Mammals  in 
California's  Oak  Woodlands: 
Temporal  and  Spatial 
Patterns^ 

William  M.  Block,  Michael  L  Morrison,  John 
C.  Slaymaker,  and  Gwen  Jongejan^ 


Abstract.— We  monitored  pitfall  traps  for  >50,000 
trap  nights  among  three  study  areas  in  California's 
oak  woodlands.  Numbers  of  captures  and  trap 
success  varied  spatially  in  comparisons  of  grids 
within  and  among  stand  types,  as  well  as  among 
study  areas.  Capture  numbers  also  varied 
temporally,  both  within  and  between  the  years  of 
study.  Differences  in  capture  rates  varied  among 
taxa  (amphibians,  reptiles,  and  small  mammals)  and 
also  varied  among  species  within  a  taxon. 
Researchers  should  design  studies  to  sample 
temporal  and  spatial  variations  in  activity  patterns  to 
provide  a  more  complete  understanding  of  the 
habitat  associations  of  the  species  studied. 

position  and  structure  of  the  vegeta- 
tion among  the  study  areas. 

Sierra  Foothill  Range  Field  Station 
(SFRFS),  Yuba  County,  was  located 
in  the  foothills  of  the  Sierra  Nevada 
about  25  km  NE  of  Marysville.  Eleva- 
tion ranged  from  2(X)  to  700  m  on  a 
general  west-northwest  facing  slope. 
Blue  oak  (Quercus  douglasii),  interior 
live  oak  (Q.  wislizenii),  and  digger 
pine  (Pinus  sabiniam)  were  the  major 
species  of  trees  with  lesser  amounts 
of  California  black  oak  (Q.  kelloggii), 
California  buckeye  (Aesculus  californi- 
cus),  and  ponderosa  pine  (Pinus  pon- 
derosa).  Major  components  of  the 
shrub  layer  included  buckbrush 
(Ceanothus  cuneatus),  coffeebcrry 
(Rhamnus  californica),  and  poison  oak 
(Toxicodendron  diversiloba).  Annual 
and  perennial  grasses  and  forbs 
dominated  cover  within  a  meter  of 
the  ground,  although  there  were  spa- 
tial and  temporal  variations  in  spe- 
cies compositions  and  also  in  amount 
of  ground  cover.  Further,  the  compo- 
sition and  structure  of  the  canopy, 
shrub,  and  ground  layers  have  all 
been  modified  by  historic  land-use 
practices  at  the  Station.  Except  for  60 
ha  of  fenced  areas,  the  remaining 
1800  ha  are  used  for  varied  research 
projects  usually  entailing  cattle  graz- 
ing and  often  entailing  tree  removal. 

San  Joaquin  Experimental  Range 
(SJER),  Madera  County,  was  located 
in  the  foothills  of  the  Sierra  Nevada 
about  40  km  N  of  Fresno.  Elevation 
ranged  from  200  to  500  m;  the  aspect 
was  in  a  general  southwest  direction. 


The  hardwood  rangelands  of  Califor- 
nia are  coming  under  increasing 
land-use  pressures.  Cattle  grazing, 
fuelwood  removal,  hydro-electric 
projects,  urban  sprawl,  and  countless 
other  factors  are  impacting  these 
woodlands  at  local,  regional,  and 
geographical  levels  (see  papers 
within  Plumb  and  Pillsbury  1987). 
Unfortunately,  little  is  known  of  the 
distributions  and  ecologies  of  many 
of  the  vertebrates  occurring  in  these 
areas  (Vemer  1987).  As  a  conse- 
quence, resource  managers  fre- 
quently have  too  little  information 
upon  which  to  base  land-use  deci- 
sions. Thus,  a  research  agenda  is  re- 
quired first  to  obtain  baseline  infor- 
mation on  distributions  and  habitat 
associations  of  these  animals,  and 
then  to  use  these  data  to  predict  the 
presence  or  absence  of  these  species, 
and  ultimately  to  predict  the  effects 
of  habitat  change  on  their  popula- 
tions. Research  should  encompass  a 
hierarchy  of  spatial  scales  to  account 
for  variations  in  patterns  of  habitat 
use,  and  also  to  determine  if  a  spe- 
cies' habitat  exhibits  consistent  and 
measurable  features  (Allen  and  Starr 
1982,  Block,  in  press).  Study  must 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  t^ommals  in  Northi  America.  (Flag- 
staff. AZ.  July  19-21.  1988.) 

'Project  Leader.  Associate  Professor. 
Researct)  Associate,  and  Research)  Associ- 
ate, respectively.  Department  of  Forestry 
and  Resource  Management.  University  of 
California.  Berkeley.  CA  94720. 


also  be  done  year-round  to  sample 
habitat-use  by  species  during  differ- 
ent stages  of  their  life  histories,  and  it 
also  should  be  done  over  a  number 
of  years  to  include  annual  variations 
in  environmental  conditions  (Halvor- 
son  1984,  Morrison,  this  volume). 

As  part  of  an  ongoing  study  to  de- 
termine habitat  relationships  of  ver- 
tebrates in  California's  oak  wood- 
lands, we  have  been  using  pitfall 
traps  to  sample  populations  of  small 
mammals,  reptiles,  and  amphibians 
at  three  distinct  areas.  To  date  we 
have  collected  data  from  greater  than 
50,000  trap  nights  distributed  among 
20  trapping  grids.  This  general  de- 
sign has  allowed  us  to  examine  spa- 
tial patterns  of  habitat-use  both 
within  and  among  areas.  Further, 
more  intensive  study  has  been  done 
at  one  area  to  examine  temporal  pat- 
terns in  habitat  use  both  within  and 
between  years.  In  this  paper  we  pres- 
ent these  data  to  examine  spatial  and 
temporal  patterns  of  habitat  use  and 
discuss  our  results  in  relation  to  the 
general  design  of  studies  of  small 
mammal,  reptile,  and  amphibian 
populations. 

STUDY  AREAS 

The  study  was  done  at  three  areas, 
all  oak  or  pine-oak  woodlands.  Study 
areas  were  distributed  along  a  latitu- 
dinal gradient  of  about  600  km,  and 
consequently  there  were  notable  dif- 
ferences in  topography  and  in  com- 


247 


Blue  oak,  interior  live  oak,  and  dig- 
ger pine  were  the  major  tree  species. 
These  species  occurred  in  mixed-spe- 
cies stands,  stands  of  blue  oak  wood- 
land, or  as  blue  oak  savannas.  An- 
nual and  perennial  forbs  and  annual 
grasses  dominated  the  ground  layer. 
About  20  ha  of  SJER  have  been 
fenced  to  exclude  cattle  grazing. 
Cattle  grazing  on  the  remaining  1500 
ha  has  resulted  in  a  sparser  shrub 
understory  at  SJER  than  of  that 
found  at  SFRFS  (Duncan  et  al.  1987). 
Major  shrubs  include  buckbrush, 
whitethorn  ceanothus  (Ceanothus  leu- 
codermis),  redberry  (Rhamnus  crocea), 
coffeberry,  poison  oak,  and  white 
lupine  (Lupinus  alba).  The  shrub  un- 
derstory is  restricted  mostly  to 
widely  scattered  stands  of  mature 
shrubs  which  have  grown  above  the 
deer-cattle  browse  line. 

Tejon  Ranch  (TR),  Kern  County, 
was  located  about  50  km  south  of 
Bakersfield  in  the  Tehachapi  Moun- 
tains. Elevation  ranged  from  1100  to 
1700  m;  aspects  included  all  cardinal 
directions.  Major  trees  found  on  TR 
included  blue  oak,  valley  oak  (Quer- 
cus  lobata),  California  black  oak,  inte- 
rior live  oak,  canyon  live  oak  ( Q. 
chrysolepis),  Brewer's  oak  (Q.  garryam 
var.  breweri),  and  California  buckeye. 
At  lower  elevations,  these  trees  gen- 
erally occurred  in  pure  stands  of 
single  species,  with  mixed-stands  of 
California  black,  canyon  live,  interior 
live,  and  Brewer's  oaks  occurring  at 
higher  elevations.  Buckbrush, 
redberry,  and  mountain  mahogany 
(Cercocarpus  betuloides)  were  the  ma- 
jor shrubs  with  annual  and  perennial 
grasses  and  forbs  comprising  the 
ground  canopy.  Cattle  grazing  and 
fuelwood  harvest  have  modified  the 
composition  and  structure  of  the 
tree,  shrub,  and  herbaceous  layers. 

METHODS 

Field  Methods 

At  TR  we  placed  three  grids  in  each 
of  three  different  stand  types — blue 


oak,  valley  oak,  and  canyon  live  oak 
woodlands — and  we  placed  four 
grids  in  four  different  stands  of 
mixed-oak  woodlands  (California 
black,  interior  live,  canyon  live,  and 
Brewer's  oaks).  At  SJER  we  placed 
four  grids,  one  each  in  a  blue  oak  and 
an  interior  live  oak  stand,  and  two  in 
mixed  blue  oak-interior  live  oak-dig- 
ger pine  stands.  The  three  grids  at 
SFRFS  sampled  three  stands  of 
mixed  blue  oak-interior  live  oak-dig- 
ger pine  woodlands.  Selection  of 
stands  was  not  entirely  random  be- 
cause we  needed  to  consider  accessi- 
bility during  inclement  weather,  and 
possible  conflicts  with  other  research 
projects  or  with  certain  management 
practices  (e.g.,  excessive  cattle  graz- 
ing, fuelwood  harvest,  road  con- 
struction) when  selecting  stands.  The 
actual  selection  of  the  grid  location 
within  a  stand  was  by  a  series  of  ran- 
dom procedures  to  determine  dis- 
tance of  the  grid  from  the  stand  edge 
(>100  m  from  the  stand  edge  to  mini- 
mize edge  effects)  and  the  direction 
of  the  grid  array. 

Each  grid  consisted  of  36  2-gal, 
plastic  buckets  arrayed  in  a  6  x  6 
square  with  20-m  interstation  spac- 
ings.  Buckets  were  placed  within  2  m 
of  each  grid  point  at  a  suitable  trap- 
ping location.  Buckets  were  sunk  to 
ground  level  and  left  closed  (a  piece 
of  plywood  secured  with  a  rock)  for 
at  least  one  month  prior  to  being 
opened.  This  period  enabled  germi- 
nation of  grasses  and  forbs  to  occur 
thus  making  the  area  near  the  trap 
appear  less  disturbed  and  also  al- 
lowed small  mammals  and  herpe- 
tofauna  to  become  accustomed  to  the 
presence  of  the  traps.  Traps  were 
opened  by  propping  a  plywood  lid  5- 
10  cm  above  the  lip  of  the  bucket  us- 
ing small  branches  or  small  rocks 
and  then  placing  3-6  cm  of  water  in 
the  bottom  of  the  bucket.  Traps  were 
checked  once  a  week  and  were  left 
open  for  1-2  months  at  a  time.  We 
noted  the  species,  date,  and  trap  lo- 
cation of  all  captures.  Dead  animals 
were  removed  from  traps;  live  ani- 
mals were  removed  and  relocated  to 


a  similar  habitat  at  least  one  km  from 
the  nearest  trapping  grid. 

We  monitored  pitfall  traps  at  TR 
from  4  January  to  20  May  1987  and 
from  10  December  1987  to  20  June 
1988.  We  regarded  the  first  year  of 
monitoring  as  a  pilot  study  to  evalu- 
ate and  refine  our  methods.  Traps 
were  opened  and  monitored  for  30 
days  using  the  methods  described 
above.  However,  in  light  of  a  recent 
article  by  Bury  and  Corn  (1987),  we 
increased  our  trapping  period  from 
30  to  60-65  days  per  grid.  Thus,  our 
design  at  TR  for  the  second  year  con- 
sisted of  opening  one  grid  of  each 
stand  type  for  60-65  days,  closing 
those,  and  then  opening  another  set 
of  four  grids.  We  repeated  this  de- 
sign three  times.  We  opened  the  four 
grids  at  SJER  and  the  three  grids  at 
SFRFS  each  for  60  days  from  mid- 
January  through  mid-March  1988. 

Data  Analyses 

We  compared  standardized  capture 
numbers  among  stand  types  at  TR 
and  among  the  three  study  areas  (TR, 
SJER,  and  SFRFS)  to  determine  gen- 
eral distributional  patterns  of  the  ani- 
mals caught.  Capture  numbers  were 
standardized  by  pooling  all  captures 
of  a  species  within  a  stand  type  or 
within  a  study  area  and  dividing  this 
number  by  the  total  number  of  trap 
nights  for  each  grid  within  that  stand 
type  or  study  area.  We  calculated 
Spearman  rank-order  coefficients 
(Marascuilo  and  McSweeny  1977)  to 
test  for  differences  in  rankings  of 
captures  of  species  among  stand 
types  at  TR  and  then  of  captures 
among  the  three  areas.  We  tested  for 
species-specific  differences  in  capture 
rates  among  stand  types  and  among 
study  areas  using  Kruskal-Wallis 
analyses  (Marascuilo  and  McSweeny 
1977). 

We  used  log-linear  analyses  (Fien- 
berg  1980)  to  determine  the  sources 
of  variation  in  trap  success  within 
and  among  years,  stand  types,  and 
study  areas.  We  used  data  only  for 


248 


the  presence  or  absence  of  a  species 
at  each  trapping  station,  regardless 
of  the  number  of  individuals  of  the 
species  that  were  captured  at  the  sta- 
tion. Because  the  number  of  trap 
nights  varied  between  grids,  we  used 
this  variable  as  a  covariate  in  all 
analyses  to  factor  out  the  bias  this 
might  have  entered  in  our  analyses. 

To  test  for  within-year,  spatial- 
temporal  patterns,  we  restricted  our 
analyses  to  data  collected  in  1988. 
Analyses  were  done  for  common 
species  (i.e.,  those  for  which  we  had 
adequate  numbers  of  samples)  and 
taxon  variables  of  mammals,  am- 
phibians, and  reptiles.  We  used  data 
from  TR  to  examine  seasonal  and 
stand  associations  of  common  spe- 
cies of  each  taxon. 

To  examine  geographic  patterns  of 
captures,  we  compared  trap  success 
among  the  three  study  areas.  Be- 
tween-year  analyses  were  done  by 


comparing  trap  success  at  TR  from 
1987  and  1988. 


RESULTS 
General  Patterns 
Tejon  Rancti 

The  ranking  of  species  captured  in 
canyon  live  oak  woodlands  was  not 
significantly  correlated  with  the 
rankings  of  species  found  in  the 
other  woodland  types  (all  values 
were  nonsignificant,  n  =  21,  P  >  0.05). 
These  differences  were  attributable 
to  a  stronger  association  of  amphibi- 
ans, particularly  Ensatim  and  Batra- 
choseps  salamanders,  with  canyon 
live  oak  stands  than  with  the  other 
types  of  woodlands  (Kruskal-Wallis 
Analyses,  df  =  2,  P  <  0.10)  (table  1). 
Differences  among  stands  were  also 


Table  1.— Capture  numbers  of  amphibians,  reptiles,  and  small  mammals 
within  four  different  oak  woodlarxl  types  at  Tejon  Ranch,  Kem  County, 
California  from  1  January  1987  through  20  June  1988. 


Species 


Bafrachoseps  nigrivenfris^ 
Ensafina  eschscholfzii^ 
Rana  boylii 

Sceloporus  occidentalism 
Eumeces  gilberfF 
Gerrhonofus  mulficarinafus 
Anniella  pulchra 
Diadophis  pulchellus 
Peromyscus  maniculafus^ 
P.  boylii 
P.  frueP 

Perognafhus  californicus 
Microfus  californicus 
Thomomys  botfae 
Reifhrodonfomys  megalofis 
Scapanus  lafimanus 
Sorex  omafus 

Total  captures 

Species  richness 


Valley 
oak 
<n=7848)' 

Blue 
oak 
(n=8828) 

Canyon 
live  oak 
(n=7848) 

f^lxed 
oak 
(n=8828) 

19 

1 

39 
34 

38 
53 

3 
13 

20 
28 

1 

31 
4 
3 

1 

42 
33 
14 

2 
8 


1 

168 
10 


10 
20 

1 
2 
4 
1 


112 
9 


22 

3 
4 
2 

1 

13 
138 
10 


3 
24 
6 

1 

6 
1 
1 

6 
102 
13 


'Number  of  trap  nights. 

^Significant  difference  (P  <0. 10)  of  captures  among  stand  types. 


noted  for  captures  of  Peromyscus 
maniculatus,  P.  truei,  Sceloporus  oc- 
cidentalis,  and  Eumeces  gilberti,  which 
were  captured  more  frequently  in 
blue  and  valley  oak  stands  than  in 
canyon  live  or  mixed-species  oak 
stands  (table  1).  In  comparisons  of 
rankings  of  taxonomic  groups  among 
stand  types,  we  found  a  significant 
positive  correlation  between  mixed- 
species  and  valley  oak  stands,  but  a 
significant  negative  correlation  be- 
tween blue  and  canyon  live  oak 
stands  (r^  significant,  n  =  3,  P  <  0.01) 
(fig.  1).  All  other  pair-wise  compari- 
sons between  stand  types  were  non- 
significant. 


All  Study  Areas 

Rankings  of  captures  of  species  were 
weakly  correlated  only  between  TR 
and  SFRFS  (r^  =  0.37,  n  =  21,  P  = 
0.052);  Spearman  rank-order  correla- 
tions were  nonsignificant  in  all  other 
comparisons.  Significant  differences 
were  found  among  areas  in  the  cap- 
ture rates  of  Sceloporus  occidentalis, 
Eumeces  gilberti,  E.  sJdltonianus,  Batra- 
choseps  attenuatus,  Batrachoseps  ni- 
griventris,  and  Ensatina  eschscholtzii 
(table  2).  In  contrast,  rankings  of  taxa 
were  significantly  correlated  between 
SJER  and  SFRFS  (r^  =  1.00,  n  =  3,  P  = 
1.00),  but  nonsignificant  (P  >  0.05)  in 
all  other  between-area  comparisons. 
The  differences  were  primarily  be- 
cause of  differences  in  capture  rates 
of  reptiles  and  amphibians  (fig.  2). 


Log-linear  Analyses 

Trap  success  at  TR  for  small  mam- 
mals, reptiles,  and  amphibians  dif- 
fered with  stand  type  and  trapping 
period  (likelihood  ratio  chi-squares, 
P  <  0.01).  Similar  results  were  found 
for  the  selected  common  species.  In 
contrast,  fewer  differences  were 
found  between  years  for  captures  of 
amphibians,  reptiles,  and  small 
mammals.  Only  captures  of  reptiles 
in  blue  oak  stands  and  captures  of 


249 


r  .•nil 


small  mammals  within  valley  oak 
stands  were  significantly  different 
between  years  (likelihood  ratio  chi- 
squares,  P  <  0.01).  We  noted  signifi- 
cant differences  (P  <  0.01)  in  capture 
frequencies  of  reptiles  and  amphibi- 
ans among  study  areas,  but  differ- 
ences were  nonsignificant  (P  >  0.05) 
for  captures  of  small  mammals. 

DISCUSSION 

Intra-year  differences  in  trap  success 
at  TR  were  observed  for  all  common 
species  and  taxonomic  groups  tested. 
Much  of  the  intra-year  variation  in 
trap  success  was  probably  because  of 
differences  in  activity  patterns  dur- 
ing different  times  of  the  year  (Welsh 
1987).  Our  results  further  suggested 
that  activity  patterns  varied  within 
and  among  taxa.  For  example,  few 
reptiles  were  captured  from  Decem- 
ber through  March;  capture  rates 
then  increased  dramatically  after 
March.  In  contrast,  fewer  salaman- 
ders were  caught  in  December,  Janu- 
ary, May,  and  June  than  were  caught 
during  March  and  April.  Similar  re- 
sults emerge  when  comparing  activ- 
ity patterns  of  species  within  a  taxon. 
Thus,  activity  patterns  of  a  species  or 
of  a  taxon  tend  to  be  somewhat  spe- 
cific to  the  animal  or  group  studied. 

Differences  in  trap  success  were 
not  as  apparent  for  interyear  com- 
parisons, however.  In  fact,  the  only 
differences  that  we  noted  were  in- 
creases from  1987  to  1988  in  trap  suc- 
cess for  reptiles  in  blue  oak  and  for 
mammals  in  valley  oak  stands.  These 
results  might  be  interpreted  in  two 
ways.  First,  species  compositions  are 
fairly  consistent  from  year  to  year,  or 
the  2  years  of  data  that  we  compared 
were  possibly  insufficient  to  detect 
population  or  habitat  shifts  (Halvor- 
son  1984,  Morrison,  this  volume). 
Undoubtedly,  a  long-term  study  is 
required  to  determine  if  these  results 
remain  valid  with  time  or  if  they  are 
an  artifact  of  the  sampling  period. 

Species  distributions  also  varied 
spatially  among  the  different  stand 


types  at  TR  and  among  the  three 
study  areas.  For  example,  canyon 
live  oak  stands  contained  more  am- 
phibians and  fewer  reptiles  than 
other  types  of  stands,  whereas  few 
amphibians  and  more  reptiles  were 
captured  in  blue  oak  stands.  Valley 
oak  and  mixed-species  oak  stands 
contained  intermediate  numbers  of 
amphibians  and  reptiles.  We  also 
noted  differences  of  captures  among 
grids  of  the  same  woodland  type. 
However,  given  the  short  duration  of 
this  study  (2  years  to  date),  these  dif- 
ferences may  reflect  temporal  differ- 
ences between  sampling  periods 
more  than  variation  within  stand 
types.  Variation  was  also  noted  on  a 
broader  geographical  scale  of  be- 
tween study  areas. 

Pitfall  traps  are  one  of  many  tech- 
niques used  to  sample  vertebrate 
populations  (Day  et  al.  1980).  As 
with  each  technique,  however,  pitfall 
traps  are  not  without  limitations 
(Bury  and  Com  1987).  Inter-  and  in- 
traspecific  differences  in  motility, 
mode  of  travel,  and  activity  range  all 


influence  the  probability  of  an  animal 
being  captured.  Because  of  probable 
species-specific  biases  in  catchability, 
a  study  design  should  consider  alter- 
native methods  (e.g.,  live  traps  for 
small  mammals,  and  active  searches 
for  reptiles  and  amphibians)  to 
sample  the  population(s)  of  the  spe- 
cies of  interest  (Halvorson  1984,  Ra- 
phael and  Rosenberg  1983,  Welsh 
1987). 

For  example,  results  from  our  pit- 
fall data  do  not  completely  agree 
with  preliminary  results  from  >6,000 
trap  nights  using  live  traps  or  from 
20  time-constraint  searches,  both 
done  at  TR  (Block,  unpubl.  data).  In 
particular,  we  captured  more  Per- 
ognathus  californicus  and  Reithrodonto- 
mys  megalotis  using  live  traps  than 
we  did  using  pitfall  traps,  but  have 
captured  no  Microtus,  Sorex,  Tho- 
momys,  or  Scapanus  in  live  traps 
whereas  we  have  caught  them  in  the 
pitfalls. 

Thus,  researchers  should  compare 
and  evaluate  results  from  alternative 
methods  to  determine  the  most  effec- 


8- 


1 


Valley  ook 


Bkie  ook 


Conyon  ook 


Mixed  ook 


Amphibians 


Reptiles 


Mafranols 


Figure  1  .—Relative  numbers  of  captures  using  pitfall  traps  wittiin  four  oak  woodland  types  at 
Tejon  Rancti,  Kern  County,  California  from  5  Decemt^er  1987  to  20  June  1988. 


250 


tive  method  or  combination  of  meth- 
ods to  use  for  the  species  under 
study. 

We  evaluated  our  data  in  two  dif- 
ferent ways:  comparisons  of  capture 
numbers  and  comparisons  of  trap 
success.  Results  from  both  analyses 
were  generally  consistent,  although 
in  some  cases  we  found  differences 
in  comparisons  of  trap  success,  but 
failed  to  do  so  in  comparisons  of  cap- 
ture numbers.  The  discrepancies  be- 
tween these  results  may  be  attribut- 
able to  both  statistical  and  biological 
factors. 

Statistical  factors  stem  from  the 
fact  that  continuous  data  were  re- 
corded for  capture  numbers  whereas 
categorical  data  were  recorded  for 
trap  success.  Consequently,  different 
statistical  tests  were  required  to  ana- 
lyze the  different  types  of  data.  The 


lack  of  concordance  between  results 
may  be  the  result  of  different  as- 
sumptions of  the  different  tests  and 
of  different  powers  of  the  associated 
statistics. 

For  example,  in  comparisons  of 
capture  numbers,  our  use  of  all  cap- 
tures from  a  trap  for  a  given  species 
may  have  violated  assumptions  of  in- 
dependence of  samples;  assumptions 
underlying  most  parametric  and 
nonparametric  statistical  tests  (e.g., 
see  Sokal  and  Rohlf  1969,  Marascuilo 
and  McSweeny  1977).  Conversely, 
using  presence-absence  data  as  we 
did  in  analyses  of  trap  success  avoids 
the  problem  of  dependency.  A  short- 
coming of  using  only  presence-ab- 
sence data,  however,  is  that  informa- 
tion of  the  numbers  and  hence  rela- 
tive abundance  of  animals  captured 
might  be  lost. 


r 


Table  2.— Capture  numbers  of  amphibians,  reptiles,  and  small  mammals  at 
three  California  oak  woodlands:  Tejon  Ranch,  Kern  County;  San  Joaquin 
Experimental  Range,  Madera  County;  and  Sierra  Foothill  Range  Reld  Sta- 
tion, Yuba  County,  from  mid-January  through  mld-f^arch  1988. 


Species 


Tejon 
Ranch 
<n=8828)' 


San  Joaquin 
Exp.  Range 
{n=8828) 


Sierra  Foothill 
Range  Field  Stn. 
(n=6912) 


Bafrachoseps  attenuafus^ 
Bafrachoseps  nigrivenfris^ 
Ensafina  eschscholfzP 
Taricha  forosa 
Ran  a  boy  Hi 

Scaphiopus  hammondii 
Sceloporus  occidentalism 
Bum  eces  gilb  erfF 
Eumeces  si<ilfonianu$^ 
Gerrhonotus  mulficarinatus 
Peromyscus  maniculafus 
P.  boy  Hi 
P.  truei 

Perognafhus  californicus 
P.  inomatus 
Microfus  californicus 
Thomomys  botfae 
Scapanus  latimanus 
Sorex  omafus 

Total  captures 

Species  richness 


19 
3 

1 

20 
9 

1 

6 
13 
5 
1 

1 
1 
1 

82 
13 


8 


1 

3 
31 
46 


7 
6 
9 

1 

4 


113 
10 


1 


96 

8 
1 
3 
5 
4 


6 
1 

3 
128 
9 


'Number  of  trap  nigtits. 

^Significant  difference  (P  <0. 10)  of  captures  among  study  areas. 


CONCLUSIONS 

Using  pitfall  traps  to  sample  amphib- 
ian, reptile,  and  small  mammal 
populations,  we  found  pronounced 
variation  within  and  among  study 
areas,  and  within  and  between  years 
in  capture  rates  of  all  taxa  and  of 
many  of  the  species  studies.  Implica- 
tions of  these  results  apply  both  to 
the  design  of  studies  for  these  ani- 
mals as  well  as  for  their  manage- 
ment. First,  we  recognize  biases  by 
using  only  pitfall  traps  to  sample 
populations  of  free-ranging  verte- 
brates, and  we  suggest  that  research- 
ers evaluate  all  possible  methods  to 
determine  the  best  one  or  combina- 
tion of  methods  for  the  study  of  a 
particular  organism(s).  Second, 
within-year  variation  in  capture  rates 
suggests  that  researchers  should  de- 
sign a  study  to  sample  seasonal  vari- 
ations in  activities  and  in  habitat  use. 
Similarly,  spatial  variation,  both 
within  and  among  stand  types  and 
among  distinct  geographic  locations, 
should  be  studied  to  better  identify 
distributional  limits  of  the  species 
studied  and  to  determine  how  spe- 
cific habitats  contribute  to  the  sur- 
vival and  reproduction  of  the  spe- 
cies. From  a  management  perspec- 
tive, understanding  temporal  and 
spatial  variability  in  habitat  use  is 
critical  when  trying  to  provide  suit- 
able conditions  for  the  animal  to  sur- 
vive and  reproduce.  All  oak  wood- 
lands cannot  be  managed  in  the  same 
way  for  all  species.  Each  oak-wood- 
land type  contains  a  unique  set  of 
factors  that  predispose  species  to  use 
the  area  for  some  aspect  of  their  life 
histories.  Management  for  a  species 
should  be  based  on  information  that 
considers  the  spatial  and  temporal 
variability  in  habitat  use  to  provide 
for  all  life  requisites. 


ACKNOWLEDGMENTS 

We  thank  J.  Bartolome,  L.  Brennan,  J. 
Dunne,  and  T.  Pruden  for  construc- 
tive comments  on  earlier  versions  of 


251 


this  paper.  M.  Dixon,  S.  Kee,  W. 
Maynard,  and  T.  Tennant  assisted  in 
the  field.  We  thank  R.  Barrett  for  use 
of  field  equipment,  and  S.  Lee,  L. 
Merkle,  and  I.  Timossi  for  technical 
assistance.  D.  Geivet,  D.  Duncan,  and 
J.  M.  Conner  provided  logistical  as- 
sistance at  Tejon  Ranch,  San  Joaquin 
Experimental  Range,  and  Sierra  Foot- 
hill Range  Field  Station,  respectively. 
University  of  California,  Division  of 
Agriculture,  Integrated  Hardwood 
Program;  U.S.  Forest  Service,  Pacific 
Southwest  Forest  and  Range  Experi- 
ment Station;  and  California  Depart- 
ment of  Forestry,  Forest  and  Range 
Resource  Assessment  Program  pro- 
vided funding  for  this  research. 

LITERATURE  CITED 

Allen,  T.  F.  H.,  and  Thomas.  B.  Starr. 
1982.  Hierarchy:  perspectives  for 
ecological  complexity.  310  p.  Uni- 
versity of  Chicago  Press,  Chicago, 
111. 

Block,  William  M.  In  press.  Geo- 
graphical variation  in  foraging 
ecologies:  breeding  and  nonbreed- 
ing  birds  in  oak  woodlands.  Pro- 
ceedings of  the  symposium  on 
food  exploitation  by  terrestrial 
birds  [Asilomar,  CA,  March  19-20, 
1988]. 

Bury,  R.  Bruce,  and  Paul  Stephen 
Corn.  1987.  Evaluation  of  pitfall 
trapping  in  northwestern  forests: 
trap  arrays  with  drift  fences.  Jour- 
nal of  Wildlife  Management 
51:112-119. 

Day,  Gerald  I.,  Sanford  D.  Schem- 
nitz,  and  Richard  D.  Tabor.  1980. 
Capturing  and  marking  wild  ani- 
mals, p.  61-88.  In  Wildhfe  Manage- 
ment Techniques  Manual.  686  p. 
Sanford  D.  Schemnitz,  editor. 
Wildlife  Society,  Washington,  D.C. 

Duncan,  Don  A.,  Neil  K.  McDougald, 
and  Stanley  E.  Westfall.  1987. 
Long-term  changes  from  different 
uses  of  foothill  hardwood  range- 
lands,  p.  367-372.  In  Proceedings 
of  the  symposium  on  multiple-use 
management  of  California's  hard- 


wood resources  [San  Luis  Obispo, 
CA,  November  12-14, 1986]. 
USDA  Forest  Service  General 
Technical  Report  PSW-100, 462  p. 
Pacific  Southwest  Forest  and 
Range  Experiment  Station, 
Berkeley,  CA. 

Fienberg,  Stephen,  E.  1980.  The 
analysis  of  cross-classified  cate- 
gorical data.  151  p.  Second  edition. 
MIT  press,  Cambridge. 

Halvorson,  Curtis,  H.  1984.  Long- 
term  monitoring  of  small  verte- 
brates: a  review  with  suggestions, 
p.  11-25.  In  Research  natural  areas: 
baseline  monitoring  and  manage- 
ment. Proceedings  of  a  sympo- 
sium [Missoula,  Mont.,  March  21, 
1984).  USDA  Forest  Service  Gen- 
eral Technical  Report  INT-173, 84 
p.  Intermountain  Forest  and 
Range  Experiment  Station,  Ogden, 
Utah. 

Marascuilo,  Leonard  A.  and  Ma- 
ryellen  McSweeny.  1977.  Non-par- 
ametric and  distribution-free 


methods  of  the  social  sciences.  580 
p.  Brooks-Cole,  Monterey,  CA. 

Plumb,  Timothy  R.,  and  Norman  H. 
Pillsbury,  technical  coordinators. 
1987.  Proceedings  of  the  sympo- 
sium on  multiple-use  management 
of  California's  hardwood  re- 
sources [San  Luis  Obispo,  CA, 
November  12-14,  1986].  USDA 
Forest  Service  General  Technical 
Report  PSW-100, 462  p.  Pacific 
Southwest  Forest  and  Range  Ex- 
periment Station,  Berkeley,  CA. 

Raphael,  Martin  G.,  and  Kenneth  V. 
Rosenberg.  1983.  An  integrated 
approach  to  wildlife  inventories  in 
forested  habitats,  p.  219-222.  In 
Renewable  resources  inventories 
for  monitoring  changes  and 
trends.  731  p.  Oregon  State  Uni- 
versity, Corvallis,  Oregon. 

Sokal,  Robert  R.,  and  F.  James  Rohlf. 
1969.  Biometry:  the  principles  and 
practice  of  statistics  in  biological 
research,  776  p.  W.  H.  Freeman, 
San  Francisco,  Calif. 


Tqon  Rondi  Son  Jooquin  Sam  FooM 


Figure  2.— Relative  numbers  of  captures  using  pitfall  traps  wittiin  thiree  oalc  woodland  study 
areas  in  California:  Tejon  Rancti,  Kern  County,  San  Joaquin  Experimental  Range,  Madera 
County;  and  Sierra  Foottiill  Range  Experimental  Field  Station,  Yuba  County.  Trapping  oc- 
curred from  5  December  1987  to  20  Marcti  1988. 


252 


Vemer,  Jared.  1987.  The  importance 
of  hardwood  habitats  for  wildlife 
in  California,  p.  162.  In  Proceed- 
ings of  the  symposium  on  mul- 
tiple-use management  of  Califor- 
nia's hardwood  resources  [San 
Luis  Obispo,  CA,  November  12- 
14, 1986].  USDA  Forest  Service 
General  Technical  Report  PSW- 
100, 462  p.  Pacific  Southwest  For- 
est and  Range  Experiment  Station, 
Berkeley,  CA. 

Welsh,  Hartwell,  H.  Jr..  1987.  Moni- 
toring herpetofauna  in  woodland 
habitats  of  northwestern  Califor- 


nia and  southwestern  Oregon:  a 
comprehensive  strategy,  p.  203- 
213.  /n  Proceedings  of  the  sympo- 
sium on  multiple-use  management 
of  California's  hardwood  re- 
sources [San  Luis  Obispo,  CA, 
November  12-14, 1986].  USDA 
Forest  Service  General  Technical 
Report  PSW-100, 462  p.  Pacific 
Southwest  Forest  and  Range  Ex- 
f>eriment  Station,  Berkeley,  CA. 


253 


The  Importance  of  Biological 
Surveys  in  Managing  Public 
Lands  in  the  Western  United 
States' 

Michael  A.  Bogan,^  Robert  B.  Finley,  Jr.,^  and 
Stephen  J.  Petersburg^ 


Abstract.— Despite  previous  studies,  incomplete 
knowledge  of  the  mammalian  fauna  of  many 
national  parks  hinders  our  ability  to  understand  the 
consequences  of  either  management  actions  or 
natural  disasters  to  such  preserves.  Fauna!  losses 
have  occurred  and  con  be  expected  to  continue 
(Newmark  1986a,  1986b).  Our  studies  in  and  near 
Dinosaur  National  Monument,  one  of  the  parks 
studied  by  Nev^mark  (1986a,  1986b),  have  added  1 1 
species  to  the  known  fauna.  Some  species  have 
increased  with  human  impact;  other  species  hove 
either  disappeared  or  are  declining.  Finally,  many 
species,  which  are  uncommon  and  poorly  known, 
may  hove  rather  specific  habitat  needs. 

era!  information  is  available  in  only  a 
few  sources  (Cary  1911,  Warren  1942, 
Lechleitner  1969,  Armstrong  1972), 
each  of  which  treats  all  Coloradoan 
mammals.  Detailed  studies  of  this 
area  are  not  common  and  may  be  dif- 
ficult to  obtain  (Durrant  1963,  Bogan 
et  al.  1983).  This  paucity  of  knowl- 
edge is  frustrating  not  only  to  mam- 
malogists,  but  also  to  land  managers 
seeking  to  protect  the  resources  un- 
der their  care.  In  the  absence  of  reli- 
able information,  land  stewards  may 
end  up  managing  for  a  relatively 
small  portion  of  the  total  fauna,  pri- 
marily those  that  are  rare  or  endan- 
gered, highly  visible  or  popular, 
pests,  or  those  of  importance  to  hunt- 
ers and  trappers. 

Our  studies  in  DNM  and  adjacent 
Browns  Park  National  Wildlife  Ref- 
uge, conducted  since  1980,  have  pro- 
vided new  information  on  the  mam- 
mals of  northwestern  Colorado.  In 
addition,  our  data  can  provide  a  per- 
spective on  1)  the  severity  of  the 
problem  of  faunal  loss  as  shown  by 
Newmark  (1986a,  1986b)  for  one  area 
(DNM);  and  2)  the  continuing  need 
for  a  better  data  base  from  which  to 
manage  parks  and  their  fauna  and 
flora.  We  summarize  the  gradual  ac- 
quisition of  knowledge  about  mam- 
mals in  DNM,  the  contribution  of  re- 
cent detailed  studies  to  the  faunal 
data  base,  and  how  some  species 
seem  to  be  responding  to  human  ac- 
tivity. Finally,  we  comment  on  some 
of  Newmark's  (1986a,  1986b)  data 
and  conclusions  for  DNM. 


The  equilibrium  model  of  island  bio- 
geography  (MacArthur  and  Wilson 
1963, 1967)  spawned  a  plethora  of 
studies  that  examined  ways  in  which 
various  kinds  of  insular  faunas  be- 
have (for  mammals  see  Heaney  and 
Patterson  1986).  Some  of  the  most 
interesting  applications  of  the  model 
have  been  to  animals  in  islands  of 
habitat,  such  as  mountains  in  the 
Great  Basin  (Brown  1971,  1978). 
These  studies  revealed  that  such  fau- 
nas often  behave  in  contrast  to  the 
model,  which  predicts  that  the  num- 
ber of  species  on  an  island  reflects  an 
equilibrium  between  processes  of  ori- 
gin, i.e.,  species  emigrating  to  the  is- 
land as  a  function  of  island  size  and 
distance  from  the  mainland,  and 
processes  of  extinction  on  the  island. 
Such  studies  lend  support  to  the  con- 
tention that  montane  mammalian 
faunas  in  the  Southwest  are  not  in 
equilibrium  (Brown  1986);  rather, 
they  are  relicts  derived  by  extinction 

'Paper  presented  of  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Norfti  America.  (Flag- 
staff, AZ,  July  19-21,  1988.) 

^Michael  A.  Bogan  is  Wildlife  Researcti 
Biologist,  U.S.  Fisti  &  Wildlife  Sen/ice.  Na- 
tional Ecology  Research)  Center,  1300  Blue 
Spruce  Drive,  Fort  Collins,  CO  80524-2098. 

^Robert  B.  Finley,  Jr.,  retired  from  the  U.S. 
Fish  &  Wildlife  Service,  is  Research  Associ- 
ate, The  Museum,  University  of  Colorado, 
Boulder,  CO  80309. 

^Stephen  J.  Petersburg  is  Resource  Man- 
agement Specialist,  National  Park  Service, 
Dinosaur  National  Monument,  P.O.  Box  210, 
Dinosaur,  CO  81610. 


from  a  set  of  colonizing  species  that 
reached  the  mountains  when  life 
zones  were  lowered  during  the  Pleis- 
tocene. 

Newmark  (1986a,  1986b,  1987)  re- 
cently examined  ways  in  which  west- 
ern North  American  national  parks 
also  behave,  biologically,  as  islands. 
Newmark's  (1986a,  1986b)  analysis  of 
data  for  29  parks  (data  from  only  24 
were  used  in  most  analyses)  in  the 
United  States  and  Canada  showed 
that  the  number  of  mammalian  spe- 
cies in  these  parks  is  declining. 

Newmark  (1986a,  1986b)  pre- 
dicted that  western  national  parks, 
under  a  program  of  minimal  man- 
agement, could  lose  up  to  100%  of 
the  extant  species  of  lagomorphs, 
carnivores,  and  artiodactyls  in  the 
next  100  to  200  years.  This  loss  of 
species  would  be  dependent  upon 
the  original  size  of  the  park  (larger 
areas  have  more  species  and  larger 
populations  that  persist  better 
through  time),  the  degree  of  insulari- 
zation  of  the  parks  (although  most 
parks  presently  are  not  completely 
isolated,  the  more  isolated  they  are, 
the  less  likely  they  will  be  colonized 
from  outside),  and  intensity  of  man- 
agement both  within  and  outside 
park  boundaries. 

One  of  the  mammalian  faunas  in- 
cluded by  Newmark  (1986a,  1986b) 
in  his  analysis  was  that  of  Dinosaur 
National  Monument  (DNM),  located 
in  northwestern  Colorado  and  adja- 
cent Utah,  where  few  studies  of 
mammals  have  been  conducted.  Gen- 


254 


Methods 

Data  were  obtained  from  our  studies 
conducted  in  northwestern  Colorado 
since  1980.  These  studies,  conducted 
in  riparian  and  upland  habitats  in 
and  near  DNM,  involved  biological 
surveys  for  mammals  and  their  sign. 
Mammals  were  observed,  trapped 
and  released,  and  collected.  Speci- 
mens form  a  major  part  of  our  data 
base,  confirming  the  actual  presence 
of  a  species  at  a  point  in  time. 

Most  habitats  were  sampled  from 
one  to  three  nights  with  250  to  300 
live  or  snap  traps  each  night.  Traps 
were  set  both  in  linear  transects  and 
opportunistically;  mist  nets  and  other 
methods  were  used  for  some  species. 
Our  study  sites  included  camp- 
grounds, subjectively  categorized 
according  to  use  by  humans,  as  well 
as  isolated  areas  rarely  visited  by 
humans.  Although  data  from  some 


sites  are  directly  comparable  and  sta- 
tistically testable  due  to  standardiz- 
ing numbers  of  traps  and  techniques, 
our  purpose  here  is  to  present  an 
overview  of  the  mammals  at  DNM 
using  all  available  information. 

Data  on  distribution  and  abun- 
dance of  mammals  in  this  part  of 
Colorado  came  from  four  primary 
sources;  these  are  Cary  (1911),  War- 
ren (1942;  a  slightly  revised  version 
of  Warren  1910),  Lechleitner  (1969), 
and  Armstrong  (1972).  Studies  of 
nearby  areas  were  consulted 
(Kirkland  1981,  Finley  et  al.  1984,  Fin- 
ley  et  al.  1976).  Original  surveys  of 
DNM  by  Durrant  (1963)  and  Bogan 
et  al.  (1983)  were  of  value,  as  were 
observations  and  reports  by  knowl- 
edgeable park  visitors  and  specimens 
in  collections.  Historic  accounts  (e.g., 
Wishart  1979)  of  fur  trappers  and  ex- 
plorers of  the  nineteenth  century 
were  reviewed  for  additional  infor- 


Table  1.— Numbers  of  species  of  mammals  at  Dinosaur  Nationai  IVlonument 
per  order  as  given  in  various  reports  on  Colorado  mammals  (see  text).  Per- 
centages in  parentheses  are  the  proportion  of  the  total  mammal  fauna 
that  a  given  order  represents. 


REFERENCE 


ORDER 

Gary 

Warren 

Lechleitner  Armstrong  NewmarkThis  pap 

1911 

1942 

1969 

1972 

1986a 

1986 

INSECTIVORA 

0 

0 

? 

0 

0 

1 

(1.5%) 

CHIROPTERA 

4 

3 

7 

8 

13 

14 

(21.5%) 

LAGOMORPHA 

3 

3 

3 

3 

4 

4 

(6.1%) 

RODENTIA 

18 

18 

21 

20 

19 

25 

(38.5%) 

(SCIURIDAE 

6 

7 

7 

8 

6 

9) 

(GEOMYIDAE 

1 

1 

1 

1 

1 

1) 

(HETEROMYIDAE 

1 

2 

2 

2 

2 

3) 

(CRICETIDAE 

8 

6 

9 

8 

8 

10) 

(OTHER 

2 

2 

2 

1 

2 

2) 

CARNIVORA 

13 

n 

15 

12 

19 

16 

(24.6%) 

ARTIODACP/LA 

4 

4 

4 

4 

6 

5 

(7.6%) 

TOTALS 

42 

39 

50 

47 

62* 

65 

(%) 

65 

60 

77 

72 

95 

100 

'Includes  nine  species  that  are  not  l<nown  from  DNM. 


mation  on  the  occurrence  and  disap- 
pearance of  some  game  species  un- 
documented by  specimens. 

Specimens  of  mammals  from 
DNM  are  contained  in  the  University 
of  Utah  Museum  of  Natural  History 
(UU),  the  University  of  Colorado 
Museum  (UCM),  the  Denver  Mu- 
seum of  Natural  History  (DMNH), 
and  the  Biological  Surveys  Collection 
of  the  U.  S.  Fish  and  Wildlife  Service 
in  Washington,  DC  (USNM),  and 
Fort  Collins,  CO  (BS/FC).  Original 
field  notes,  photographs,  and  cata- 
logs form  an  important  part  of  this 
data  base  and  are  available  for  in- 
spection. Names  of  mammals  follow 
Banks  etal.  (1987). 


Results  and  Discussion 

IHistoric  Data  Acquisition 

The  growth  in  knowledge  of  the 
mammals  of  DNM  is  shown  in  table 
1.  Data  in  Cary  (1911),  who  worked 
just  east  of  the  present  Monument 
and  used  both  specimen  data  and  his 
own  and  others'  reports,  suggest  that 
about  42  species  (65%  of  the  species 
listed  in  appendix  1)  occurred  in  or 
near  DNM.  Warren  (1942),  who  did 
limited  work  in  northwestern  Colo- 
rado, provides  information  suggest- 
ing that  perhaps  39  species  occurred 
there.  Lechleitner's  (1969)  general 
treatise  on  Coloradoan  mammals, 
although  not  intended  to  provide  de- 
tailed information  on  distribution, 
supports  an  expected  fauna  of  about 
50  species.  Armstrong  (1972),  in  the 
first  comprehensive  study  of  Colora- 
doan mammals,  and  building  upon  a 
sixty-year  data  base,  relied  on  speci- 
men data  to  confirm  the  presence  or 
absence  of  mammals  in  a  given  area 
and  recorded  47  species  (72%  of 
those  currently  known)  for  DNM  or 
nearby  areas.  Although  some  of  these 
references  perhaps  should  not  be 
used  to  infer  the  specific  occurrence 
of  species  in  a  given  area,  we  think 
they  are  so  used  by  land  managers 
and  others. 


255 


During  the  period  covered  by 
these  references  little  actual  work  on 
the  mammals  of  DNM  was  con- 
ducted. Exceptions  were  the  work  of 
Hayward  et  al.  (1958),  Durrant  and 
Dean  (1959, 1960),  and  Durrant 
(1963)  who  chronicled  the  only  extant 
baseline  data  for  many  riparian  areas 
along  the  Colorado  River  and  its  ma- 
jor tributaries  (Green,  Yampa)  prior 
to  the  impoundments  at  Flaming 
Gorge  and  Glen  Canyon. 

Durrant  (1963)  surveyed  for  mam- 
mals in  DNM  and  reported  24  spe- 
cies collected  or  observed,  about  37% 
of  the  known  fauna.  Two  later  sur- 
veys for  mammals  and  other  verte- 
brates in  the  Monument  produced  29 
(Bogan  et  al.  1983)  and  27  (Bogan 
unpubl.  data)  species,  45%  and  42% 
of  the  presently  known  fauna.  Many 
of  the  same  species  were  obtained  on 
both  trips. 

Contributions  of  Recent  Surveys 

The  known  fauna  of  DNM  includes 
65  species  (appendix  1)  based  on 
specimens  and  reliable  sight  records. 
Three  species  (Canis  lupus,  Ursus 
arctos,  and  Bison  bison)  are  now  extir- 
pated; we  have  omitted  one  species 
of  dubious  occurrence  (Mustela  ni- 
gripes).  The  percentage  of  mammal- 
ian species  at  DNM  by  order  is  Insec- 
tivora,  1.5%;  Chiroptera,  21.5%;  La- 
gomorpha,  6.1%;  Rodentia,  38.5%; 
Carnivora,  24.6%;  and  Artiodactyla, 
7.7%.  Horses  (Equus  caballus)  and 
house  mice  (Mus  musculus)  occur  at 
DNM;  we  have  excluded  these  intro- 
duced species  from  our  list. 

What  result  have  enhanced  levels 
of  faunal  surveys  had  on  the  known 
fauna  of  DNM?  Our  work  has  added 
11  species  to  the  known  fauna.  These 
include  two  state  records  [Per- 
ognathus  parvus  and  Euderma  macula- 
turn  (Finley  and  Creasey  1982)  from 
Browns  Park  National  Wildlife  Ref- 
uge, about  8  mi  from  DNM];  one 
county  record  (Lepus  californicus) 
from  DNM;  seven  Monument  rec- 
ords in  1982  (Myotis  californicus,  M, 


thysanodes,  Lasionycteris  noctivagans, 
Pipistrellus  hesperus,  Perognathus  par- 
vus, Microtus  longicaudus,  and  M. 
montanus);  and  three  records  for  the 
Monument  in  1987  (Sorex  monticolus, 
Euderma  maculatum,  and  Lemmiscus 
curtatus). 

These  1 1  species  represent  an  in- 
crease of  20.3%  over  the  number  pre- 
viously known  from  DNM.  Much  of 
this  increase  (five  species)  has  come 
by  acquiring  a  better  understanding 
of  the  bats.  This  has  been  possible 
because  of  better  techniques  of  sur- 
veying for  bats,  an  improved  under- 
standing of  continental  and  regional 
distributions  of  bats,  and  an  en- 
hanced effort  in  surveying  for  bats  at 
DNM.  Additional  knowledge  of 
some  other  groups  has  come  more 
slowly,  primarily  because  we  are  ap- 
proaching the  asymptote  with  re- 
spect to  species  occurring  in  DNM. 
The  number  of  cricetid  rodents 
known  or  suspected  to  occur  has  in- 
creased from  eight  to  ten  in  75  years; 
that  for  sciurids  has  increased  from 
six  to  nine.  Armstrong  (1972)  re- 
ported 20  rodents  known  from 
DNM;  our  records  reveal  a  rodent 
fauna  of  25  species.  For  bats  the  fig- 
ures are  8  in  1972  and  14  in  1987,  an 
increase  of  75 7o. 

The  extent  to  which  surveys  reveal 
previously  unknown  faunal  compo- 
nents is  both  fortuitous  and  regu- 
lated by  biological  phenomena.  The 
capture  of  the  first  records  of  shrews 
and  spotted  bats  from  DNM  is  partly 
luck,  by  being  in  the  right  place  at  the 
right  time.  Yet  this  ability  to  "test" 
distributions  of  mammals  by  examin- 
ing (trapping)  suitable  habitats  re- 
quires training,  skill,  and  knowledge. 
In  addition,  the  ability  to  find  rare 
animals  often  requires  removing  the 
more  abundant  and  common  species. 

For  example,  of  the  1,469  speci- 
mens of  small  mammals  that  we 
have  captured  at  DNM,  52.6%  have 
been  Peromyscus  maniculatus.  We 
have  taken  1,049  Peromyscus  {7\A% 
of  the  total  trapped)  as  follows:  P. 
maniculatus,  772;  P.  truei,  175;  P.  crini- 
tus,  102.  There  may  be  many  reasons 


why  so  many  Peromyscus  are  taken; 
our  techniques  may  be  biased  in  fa- 
vor of  them,  they  are  easily  trapped, 
etc.  Still,  they  are  abundant  relative 
to  other  species  of  mammals  on  the 
Monument. 

We  have  no  exact  density  figures 
for  P.  maniculatus  in  DNM  but  ex- 
trapolations are  possible.  The  area  of 
DNM  is  827  km^  or  82,700  ha;  an  av- 
erage density  for  P.  maniculatus 
nnight  be  20/ha  (French  et  al.  1975), 
or  1,654,000  deer  mice.  We  suspect 
that  the  densities  at  DNM  are  higher, 
at  least  seasonally.  A  higher  density 
of  50/ha  (French  et  al.  1975)  would 
yield  4,135,000  deer  mice.  If  the  aver- 
age deer  mouse  weighs  20  g  (a  low 
estimate),  then  the  deer  mouse  bio- 
mass  at  DNM  is  33,080  kg  to  82,700 
kg;  the  equivalent  of  144  to  360  adult 
elk  (Cervus  elaphus)  weighing  230  kg 
each.  The  current  resident  elk  popu- 
lation of  DNM  is  150  to  200;  up  to 
600  may  be  resident  seasonally. 

This  abundance  has  several  impli- 
cations. One  is  that  the  common  spe- 
cies can  fill  the  traps,  reducing  the 
possibility  of  captures  of  other  spe- 
cies, and  thus  biasing  the  catch.  More 
interestingly,  an  accurate  under- 
standing that  there  are  a  few  abun- 
dant species  and  many  uncommon 
ones  can  provide  information  of 
value  in  assessing  impacts  of  human 
activities  and  management  of  the 
park,  e.g.,  what  species  appear  to  be 
increasing,  those  that  are  decreasing 
or  extirpated,  those  that  are  adjusting 
their  ranges,  and  those  for  which  we 
have  insufficient  information.  Ex- 
amples for  these  categories  are  dis- 
cussed below. 

Management  Implications 

Species  Increasing  in  Abundance. — 

Peromyscus  maniculatus  has  been  sug- 
gested (Armstrong  1977,  1979)  as  one 
species  that  increases  in  areas  dis- 
turbed by  humans.  It  is  a  widespread 
and  adaptable  species;  whether  it  has 
actually  increased  in  some  situations, 
such  as  in  campgrounds,  may  be  de- 


256 


batable.  Armstrong  (in  litt)  has 
noted  that  deer  mice  are  weed  spe- 
cies and  that  rather  than  representing 
a  moral  failure,  they  represent  a  suc- 
cessful evolutionary  strategy.  P.  man- 
iculatus  apparently  always  has  been 
common  in  this  part  of  Colorado; 
Gary  (1911:103)  stated  that  this  spe- 
cies was  "exceedingly  numerous  de- 
spite coyotes,  hawks,  and  owls... in 
western  Routt  [now  Moffat]  and  Rio 
Blanco  Counties  in  1906..."  He  re- 
ports (1911:103)  that  in  one  case  their 
"excessive  numbers  all  but  pre- 
vented my  securing  topotypes"  of 
another  species,  and  that  near  Lo- 
dore  they  were  everywhere  a  "great 
nuisance." 

Our  data  from  DNM  reveal  that 
the  canyon  mouse  (P.  crinitus)  is  a 
specialist  of  rocky  canyon  areas.  It 
does  penetrate  to  the  upper  reaches 
of  some  canyons  but  rarely  does  it 
spread  much  further.  The  pinon 
mouse  (P.  truei)  is  a  specialist  of  pi- 
non-juniper  forests  and  occasionally 
becomes  moderately  abundant. 
Conversely,  P.  maniculatus  is  com- 
mon in  sagebrush  (Artemisia  sp.) 
flats,  a  common  upland  habitat  at 
DNM.  A  comparison  of  relative 
abundance  of  this  species  in  subjec- 
tively categorized  "natural"  and 
"campground"  situations  reveals  an 
average  of  22.4  animals/locality  (n  = 
16)  in  areas  where  camping  is  of  low 
intensity  or  absent,  versus  an  average 
of  29.6  deer  mice/locality  in  14  heav- 
ily-used areas.  Although  these  num- 
bers cannot  be  tested  for  significance, 
due  to  non-uniform  trapping  proce- 
dures, there  is  a  difference  in  relative 
abundance  of  P.  maniculatus. 

Another  species  that  appears  to 
show  a  "campground"  effect  is  the 
golden-mantled  ground  squirrel 
(Spermophilus  lateralis).  We  have 
taken  this  species  in  many  areas  and 
it  is  widespread.  Cary  (1911:84)  re- 
ported that  this  species  was  "said  to 
be  abundant"  near  Lily  (just  outside 
the  present  Monument),  and  7  mi  N 
of  Lily  they  were  reported  to  be  "tol- 
erably common,"  but  Cary  saw  none 
there  the  previous  year.  They  are  so 


common  in  campgrounds  of  the 
Monument  now  that  they  are  a  nui- 
sance, albeit  an  attractive  one.  They 
are  fed  by  visitors  and  thus  are  en- 
couraged to  remain  near  the  camp- 
grounds. Our  data  from  areas  subjec- 
tively categorized  in  terms  of  human 
use  reveals  an  average  of  7.1  ground 
squirrels  from  eight  areas  heavily 
used  by  humans  versus  1.2  animals/ 
locality  in  six  little-used  areas.  In  ar- 
eas where  golden-mantled  ground 
squirrels  are  very  common  we  rou- 
tinely close  our  traps  during  the  day 
to  prevent  being  overrun  with  these 
animals. 

Species  Declining  or 
Disappearing. — Those  elements  of  a 
fauna  that  disappear  over  time  are 
clearly  of  concern,  and  may  provide 
clues  to  habitat  changes  or  other  fac- 
tors leading  to  faunistic  changes.  At 
least  three  mammalian  species  are 
now  extirpated  from  DNM,  and 
likely  from  Colorado.  These  are  the 
gray  wolf  (Canis  lupus),  the  grizzly 
bear  (Ursus  arctos),  and  the  bison  (Bz- 
son  bison).  Armstrong  (1972)  cites  a 
specimen  of  C.  lupus  from  Douglas 
Spring,  near  the  present-day  Monu- 
ment. That  gray  wolves  were  com- 
mon is  shown  by  the  fact  that  about 
50  were  killed  by  hired  trappers  in 
Brown's  Park  in  the  winter  of  1906-07 
(Cary  1911).  C.  lupus  was  not  in- 
cluded in  the  DNM  fauna  by  New- 
mark  (1986a). 

No  specimen  of  17.  arctos  from  or 
near  the  Monument  is  known  to  us, 
but  there  are  reports  of  sightings  in 
the  1800s.  About  60  fur  trappers  and 
800  Indians  wintered  in  Brown's 
Park  in  1839-40,  during  which  time 
they  killed  six  grizzlies  and  100  bison 
for  meat  (Dunham  and  Dunham 
1977).  Fresh  tracks  of  grizzlies  were 
seen  in  1871  by  members  of  the  sec- 
ond Powell  expedition  in  Lodore 
Canyon,  a  few  miles  above  Echo  Park 
(Dellenbaugh  1926);  and  in  1891  Ann 
Willis  was  rescued  from  a  female 
grizzly  with  two  cubs  in  Zenobia  Ba- 
sin (Murie  and  Penfold  1983). 

Remains  of  B.  bison  were  exca- 
vated from  Hell's  Midden,  an  occu- 


pation site  of  the  Fremont  Culture  in 
Castle  Park  (Lister  1983).  In  addirion. 
Walker  (1983)  reports  the  recovery  of 
remains  of  bison,  as  well  as  black 
bear  (U.  americanus),  pronghorn 
(Antilocapra  americana),  mule  deer 
(Odocoileus  hemionus),  wapiti  (Cervus 
elaphus),  and  bighorn  sheep  (Oms  ca- 
nadensis), from  Fort  Davy  Crockett  in 
Brown's  Park.  These  remains  date 
from  between  1836  and  1842.  Ashley 
saw  several  bison  in  Island  Park  in 
1825  (Murie  and  Penfold  1983). 

The  dates  of  disappearance  of 
these  species  are  speculative.  B.  bison, 
which  wintered  in  Brown's  Park,  was 
already  in  decline  west  of  the  Conti- 
nental Divide  in  the  late  1830s,  as  ob- 
served by  concerned  fur  trappers 
(Wishart  1979).  According  to  Wishart 
(1979),  the  Rocky  Mountain  trapping 
system  in  Wyoming  and  Colorado 
decayed  not  only  because  its  main 
fur-bearer,  the  beaver,  was  depleted 
but  also  because  the  main  source  of 
provisionment,  the  mountain  bison, 
was  destroyed.  Termination  of  the 
fur  trade  in  1840  allowed  mountain 
bison  to  persist  for  several  decades. 
The  last  bison  killed  in  northwestern 
Colorado  was  at  Cedar  Springs  west 
of  Craig  in  1884  (Armstrong  1972). 

C.  lupus  seems  to  have  disap- 
peared by  1935-40  (Young  1944, 
Lechleitner  1969).  The  last  report  of 
U.  arctos  in  northern  Colorado  was 
in  1920  in  the  Medicine  Bow  Range 
(Armstrong  1972).  Both  species  were 
victims  of  increasing  human  en- 
croachment and  active  predator  con- 
trol campaigns. 

We  have  chosen  to  exclude  the 
black-footed  ferret,  Mustek  nigripes, 
from  the  known  fauna  of  the  Monu- 
ment, for  lack  of  specimens  and 
sightings,  although  it  was  included 
by  Newmark.  Generally,  the  ferret 
appears  to  have  been  a  victim  of  the 
active  poisoning  of  its  principal  prey, 
prairie  dogs  (Cynomys  spp.)  in  addi- 
tion to  other  factors  (Clark  1986, 
Rath  and  Clark  1986). 

Newmark  (1986a)  stated  that 
wapiti  (Cervus  elaphus)  should  be 
added  to  the  list  of  mammals  extir- 


257 


pated  from  DNM.  Wapiti  did  occur 
in  the  Monument  in  the  early  nine- 
teenth century  and  are  there  today, 
but  their  origin  is  questionable. 

The  present  animals  may  be  de- 
scended from  remnant  populations 
from  elsewhere  in  parts  of  northern 
Colorado  or  Utah,  or  from  later  in- 
troduced wapiti  from  Wyoming.  We 
suspect  they  may  be  of  mixed  de- 
scent. 

Ovis  canadensis  occurring  on  the 
Monument  today  may  likewise  be  of 
mixed  descent.  As  noted  by  Fillmore 
(unpubl.  ms.)  bighorn  were  common 
and  highly  desired  for  food  by  trap- 
pers and  explorers  in  northwestern 
Colorado  in  the  first  half  of  the  1800s, 
but  were  greatly  reduced  by  the 
1880s,  when  they  were  protected  by 
the  first  game  laws.  Thereafter  the 
herds  slowly  increased  until  heavy 
die-offs  were  caused  by  diseases 
from  domestic  sheep.  Such  losses  oc- 
curred in  Lodore  Canyon  between 
1936  and  1945.  By  1947  the  superin- 
tendent at  DNM  was  ready  to  "write 
them  off."  In  1954  the  Colorado 
Game  and  Fish  Department  made 
two  transplants  in  Lily  Fark  and 
Zenobia  Feak,  and  numbers  since 
have  increased  in  the  Monument 
(Murie  and  Fenfold  1983). 

At  least  two  species  may  be  ad- 
justing their  ranges  relative  to  each 
other  in  reciprocal  fashion.  We  are 
aware  of  no  reports  of  Lepus  californi- 
cus  in  Moffat  County  prior  to  about 
1980,  although  both  specimens  and 
sightings  of  L.  townsendii  exist.  In 
1972  in  western  Colorado,  the  north- 
ernmost locality  for  L.  calif ornicus 
was  Mesa  County  (Armstrong  1972). 
In  the  summer  of  1987,  we  captured 
both  species,  in  close  proximity,  in 
DNM.  Based  on  the  pattern  of  re- 
placement seen  elsewhere,  including 
the  eastern  plains  of  Colorado  (Arm- 
strong 1972),  it  is  possible  that  the 
range  of  L.  townsendii  is  contracting 
to  the  north  and  that  of  L.  californicus 
is  expanding  to  the  north.  This  re- 
placement is  commonly  tied  to  land 
use  practices,  especially  breaking  the 
ground  for  cultivation,  or  over- 


grazing, which  may  lead  to  increased 
amounts  of  Opuntia  (Armstrong 
1972).  Whether  L.  californicus  is  actu- 
ally replacing  L,  townsendii  at  DNM 
is  debatable;  what  is  not  arguable  is 
that  L.  californicus  is  extending  its 
range  northward  in  western  Colo- 
rado. 

Species  for  Which  Information  is 
Inadequate. — There  are  many  species  , 
for  which  scant  information  exists. 
These  species  include  most  of  the  in- 
sectivores,  bats,  and  rodents,  to- 
gether composing  61.5%  of  the  mam- 
malian fauna  of  the  Monument.  Of 
the  40  species  in  this  category,  almost 
one-third  were  unknown  at  DNM 
just  15  years  ago.  Much  of  this  in- 
crease comes  from  a  better  under- 
standing of  the  bats,  but  knowledge 
of  their  presence  does  not  tell  us  if 
there  are  important  hibernacula  for 
bats  on  DNM,  what  proportion  of  the 
bats  may  be  migratory,  or  how  best 
to  manage  for  this  significant  compo- 
nent (22%)  of  the  fauna.  Similar  com- 
ments can  be  made  for  most  of  the 
other  small  mammals,  although  few 
are  as  vulnerable  to  mismanagement 
and  destruction  as  are  bats  (Hill  and 
Smith  1984). 

Cottontails  {Sylvilagus  spp.)  are 
commonly  seen,  even  abundant  at 
times,  but  it  is  difficult  to  identify 
animals  with  certainty  as  the  two 
species  (S.  audubonii  and  S.  nuttallii) 
occurring  at  DNM  are  externally 
similar.  The  two  species  overlap  in 
northwestern  Colorado  between  ap- 
proximately 6500  ft  and  7000  ft  and 
specimens  of  both  were  collected  by 
Warren  at  Douglas  Spring.  The  na- 
ture of  interactions  between  the  two 
species  of  cottontail  at  DNM  is  un- 
known and  studies  based  on  speci- 
mens are  needed. 

The  raccoon  was  likely  absent 
from  the  park  and  probably  the  en- 
tire upper  Colorado  River  basin  prior 
to  the  1950s  (Durrant  1952,  Long 
1965).  Specimens  (BS/FC)  indicate 
that  they  moved  into  the  upper 
Green  River  and  Brown's  Fark  in  the 
1960s  and  1970s,  probably  from  east- 
ern Wyoming. 


Newmark's  Analysis  Applied  to 
Dinosaur  National  Monunnent 

Newmark's  (1986a)  analysis  is  im- 
portant because  it  stimulates  us  to 
consider  a  problem  and  assess  its 
magnitude,  and  also  because  he  sug- 
gests some  solutions.  He  predicts  a 
depressing  picture  for  some  species 
in  national  parks  and  there  is  clear 
cause  for  concern.  Still,  it  is  useful  to 
put  his  analysis  in  perspective.  New- 
mark  (1986a)  lists  62  species  of  mam- 
mals as  occurring  in  DNM,  including 
E.  caballus  but  not  M.  musculus.  He 
(1986a:21)  confined  his  analysis  to 
only  three  orders,  lagomorphs,  carni- 
vores, and  artiodactyls  'l^ecause 
these  orders  had  the  most  complete 
park  sighting  records.  Species  of 
these  orders  tend  to  be  more  fre- 
quently reported  because  of  their 
relatively  large  body  size,  non-fosso- 
rial  nature,  and  popularity."  He  also 
used  park  sighting  records  as  well  as 
continental  (Hall  1981),  statewide 
(Armstrong  1972),  and  local  (Ander- 
son 1961)  reports. 

Those  orders  used  by  Newmark 
(1986a)  in  his  analysis  include  39%  of 
the  known  mammalian  species  at 
DNM.  The  most  diverse  order 
(Rodentia)  and  the  third  most  di- 
verse order  (Chiroptera)  at  DNM  are 
excluded.  Furthermore,  the  22  spe- 
cies he  does  consider  include  the 
only  faunal  losses  (5)  he  believes  oc- 
curred in  DNM.  We  believe  that  only 
three  species  are  extirpated  from 
DNM,  and  further  suspect  that  most 
of  the  extinctions  occurred  prior  to 
major  expansion  of  the  Monument' s 
boundaries  (1938). 

However,  the  best  management 
decisions  will  be  derived  from  the 
most  accurate  data,  and  we  should 
try  to  obtain  such  data.  We  also  be- 
lieve that  a  holistic  approach  to  ani- 
mal management  on  public  lands  is 
needed.  This  means  including  small 
and  secretive  species  in  our  plans,  as 
well  as  the  large  "glamorous"  ones. 
Newmark  recognizes  this  in  his  rec- 
ommendations; he  notes  the  need  to 
develop  a  more  extensive  monitoring 


258 


program  for  vertebrate  populations, 
including  key  species  of  every  order. 

An  examination  of  Newmark's 
(1986a)  data  reveals  that  nine  of  the 
62  species  he  lists  for  DNM  do  not 
occur  there:  Plecotus  rafinesquii  (an 
eastern  bat  perhaps  listed  due  to  a 
misunderstanding  of  its  taxonomy), 
Tadarida  brasiliensis  (accidental  at 
best,  no  records  for  northern 
Colorado),  Lepus  americanus  (perhaps 
confused  by  an  observer  with  L, 
townsendii  in  all-white  winter  pelage), 
Glaucomys  sabrinus  (may  possibly  oc- 
cur in  higher  areas  of  Douglas  Moun- 
tain at  DNM  but  presently 
unknown),  Peromyscus  boylii  (perhaps 
mistaken  by  an  observer  for  the 
large-eared  P.  truei),  Vulpes  velox  (no 
specimens  north  of  Mesa  County), 
Gulo  gulo  (there  is  a  specimen  from 
near  the  Utah-Colorado  stateline, 
outside  the  Monument),  Mustek  er- 
minea,  and  Alces  alces  (accidental 
stragglers  only). 

Why  some  of  these  species  were 
included  by  Newmark  is  unknown, 
but  in  some  cases  it  may  have  been 
because  they  were  listed  in  park  rec- 
ords, compiled  from  observations  by 
visitors  and  staff.  We  reexamined  the 
records  at  DNM  and  also  found  rec- 
ords (mostly  sightings)  of  Sorex  cin- 
ereus,  Tamias  umbrinus,  Perognathus 
flavescens,  Ammospermophilus  leucurus, 
Neotoma  lepida  (perhaps  juveniles  of 
N.  cinerea),  and  Zapus  hudsonius.  We 
know  of  no  specimens  to  substantiate 
these  records  and  do  not  include 
them  in  the  fauna  of  the  Monument. 

These  errors  are  not  necessarily 
Newmark's,  although  he  may  have 
been  uncritical  in  some  instances,  but 
likely  stem  from  several  sources. 
Among  these  are  inadequate  or  lack- 
ing baseline  surveys,  inaccurate  rec- 
ord-keeping by  park  staff,  misunder- 
standings of  current  nomenclature  by 
observers  or  recorders,  unreliable 
observations,  and  human  error. 
Nonetheless,  these  errors  cloud  our 
understanding  of  mammals  at  DNM 
and  the  management  problems  they 
present.  Additionally,  although  all 
data  and  results  age  with  time.  New- 


mark  did  not  have  the  most  current 
information  in  many  cases  and  thus 
was  unaware  of  recent  records  of 
mammals  from  DNM. 


Conclusion 

Lists  of  species  from  a  given  area  are 
subject  to  interpretation.  We  have 
taken  a  conservative  approach  rely- 
ing on  specimens  (and  giving  reasons 
for  inclusions  and  exclusions  where 
appropriate)  and  have  added  signifi- 
cantly to  the  known  mammalian 
fauna  of  DNM.  Such  lists  are  not 
trivial  exercises  because  they  are  the 
raw  materials  for  making  land  man- 
agement decisions.  Incorrect  or  miss- 
ing data  will  diminish  our  ability  to 
manage  these  lands  and  their  faunas. 
We  believe  that  biological  surveys, 
resulting  in  verified  records  (prefera- 
bly specimens,  but  sometimes  other 
data),  are  the  only  reliable  means  to 
determine  the  presence  of  a  species 
and  to  monitor  population  trends 
over  time.  We  agree  with  Newmark 
(1986a)  that  such  surveys  need  to  be 
undertaken  immediately,  because  the 
information  is  needed  now;  and 
where  surveys  have  been  initiated 
they  should  be  continued  on  a  regu- 
lar basis.  Monitoring  of  animal  popu- 
lations and  the  incorporation  of  accu- 
rate data  into  rational  management 
plans  is  the  only  way  to  ensure  that 
our  public  lands  continue  to  support 
a  diverse  fauna  that  is  as  complete  as 
possible. 

Acknowledgments 

Many  people  have  contributed  their 
time  and  efforts  to  learn  more  about 
the  mammals  of  Dinosaur  National 
Monument  and  vicinity.  Chief 
among  them  are:  R.  B.  Bury,  G.  H. 
Clemmer,  R.  D.  Fisher,  D.  Finnic,  K. 
Hammond,  D.  Hogan,  J.  Hogan,  M. 
L.  Killpack,  C.  A.  Langtimm,  D.  Lan- 
ning,  B.  Lapin,  D.  Leibman,  S.  J.  Mar- 
tin, C.  A.  Ramotnik,  B.  R.  Riddle,  A. 
L.  Riedel,  D.  E.  Wilson,  and  D. 


Worthington.  Our  boatpersons,  who 
also  provided  research  support  on 
several  trips,  included  R.  Buram,  H. 
DeWitt,  C.  Frye,  J.  Rucks,  and  S. 
Walker.  Reports  from  Cloudridge 
Naturalists,  supplied  by  D.  M.  Arm- 
strong, added  materially  to  our 
understanding  of  mammals  in  DNM. 
J.  Creasey,  then-Refuge  Manager  of 
Browns  Park  NWR,  provided  much 
information  and  cooperative  support 
for  field  work.  Early  inspirational 
support  for  this  work  came  from  A. 
R.  Weisbrod  of  the  National  Park 
Service.  We  appreciate  the  comments 
of  D.  M.  Armstrong,  R.  B.  Bury,  C. 
Jones,  and  F.  L.  Knopf  on  earlier  ver- 
sions of  this  paper. 

Literature  Cited 

Anderson,  Sydney.  1961.  Mammals 
of  Mesa  Verde  National  Park, 
Colorado.  University  of  Kansas 
Publications,  Museum  of  Natural 
History,  14:29-67. 

Armstrong,  David  M.  1972.  Distribu- 
tion of  mammals  in  Colorado. 
Monograph,  University  of  Kansas 
Museum  of  Natural  History  3:1- 
415. 

Armstrong,  David  M.  1977.  Ecologi- 
cal distribution  of  small  mammals 
in  the  Upp)er  Williams  Fork  Basin, 
Grand  County,  Colorado.  South- 
western Naturalist  22:289-304. 

Armstrong,  David  M.  1979.  Distribu- 
tional checklist  of  rodents  in  Can- 
yonlands  National  Park,  Utah.  Oc- 
casional Papers,  The  Museum, 
Texas  Tech  University  59:1-44. 

Banks,  Richard  D.,  Roy  W.  McDi- 
armid,  and  Alfred  L.  Gardner 
(eds.).  1987.  Checklist  of  verte- 
brates of  the  United  States,  the 
U.S.  Territories,  and  Canada.  U.S. 
Fish  and  Wildlife  Service,  Re- 
source Publication  166:1-79. 

Bogan,  Michael  A.,  Stephen  J.  Martin, 
and  R.  Bruce  Bury.  1983.  Baseline 
studies  of  riparian  vertebrates  of 
the  Yampa  and  Green  river  corri- 
dors within  Dinosaur  National 
Monument,  June  and  July  1982. 


259 


Final  (unpublished)  report  submit- 
ted to  Regional  Office,  Rocky 
Mountain  Region,  National  Park 
Service,  16  p. 

Brown,  James  H.  1971.  Mammals  on 
mountaintops:  nonequilibrium  in- 
sular biogeography.  American 
Naturalist  105:467-478. 

Brown,  James  H.  1978.  The  theory  of 
insular  biogeography  and  the  dis- 
tribution of  boreal  birds  and  mam- 
mals. Great  Basin  Naturalist  Mem- 
oirs 2:209-227. 

Brown,  James  H.  1986.  Two  decades 
of  interaction  between  the 
MacArthur  Wilson  model  and  the 
complexities  of  mammalian  distri- 
butions. Biological  Journal  of  the 
Linnean  Society  28:231-251. 

Gary,  Merritt.  1911.  A  biological  sur- 
vey of  Golorado.  North  American 
Fauna  33:1-256. 

Glark,  Tim  W.  1986.  Technical  intro- 
duction. The  black-footed  ferret. 
Great  Basin  Naturalist  Memoirs 
8:8-10. 

Dellenbaugh,  Frederick  S.  1926.  A 
canyon  voyage.  Yale  University 
Press,  New  Haven,  277  p. 

Dunham,  Dick,  and  Vivian  Dunham. 
1977.  Flaming  Gorge  country. 
Eastwood  Printing  and  Publishing 
Company,  Denver,  384  p. 

Durrant,  Stephen  D.  1952.  Mammals 
of  Utah.  University  of  Kansas  Pub- 
lications, Museum  of  Natural  His- 
tory 6:1-549. 

Durrant,  Stephen  D.  1963.  Mammals 
of  Dinosaur  National  Monument, 
p.  69-77.  In  Woodbury,  Angus  M. 
(ed.).  Studies  of  biota  in  Dinosaur 
National  Monument  Utah  and 
Colorado.  University  of  Utah,  Di- 
vision of  Biological  Sciences  Mis- 
cellaneous Papers  1:1-77. 

Durrant,  Stephen  D.,  and  Nowlan  K. 
Dean.  1959.  Mammals  of  Glen 
Canyon,  p.  73-103.  In  Woodbury, 
Angus  M.  (ed.).  Ecological  studies 
of  the  flora  and  fauna  in  Glen  Can- 
yon. University  of  Utah,  Anthro- 
pological Papers  (Glen  Canyon 
Series  Number  7)  40:1-228. 

Durrant,  Stephen  D.,  and  Nowlan  K. 
Dean.  1960.  Mammals  of  Flaming 


Gorge  Reservoir  Basin.  University 
of  Utah,  Anthropological  Papers 
48:209-243. 

Finley,  Robert  B.,  Jr.,  William  Caire, 
and  Dallas  E.  Wilhelm.  1983 
[1984].  Bats  of  the  Colorado  oil 
shale  region.  Great  Basin  Natural- 
ist 43:554-560. 

Finley,  Robert  B.,  Jr.,  and  James 
Creasey.  1982.  First  specimen  of 
the  spotted  bat  (Euderma  macula- 
turn)  from  Colorado.  Great  Basin 
Naturalist  42:360. 

Finley,  Robert  B.,  Jr.,  Dallas  E.  Wil- 
helm, and  William  Caire.  1976. 
Mammals  of  the  Colorado  oil 
shale  region.  Annual  Progress  Re- 
port (Unpubl.),  National  Fish  and 
Wildlife  Laboratory,  U.S.  Fish  and 
Wildlife  Service,  Fort  CoUins,  xx  p. 

Rath,  Dennis  L.  and  Tim  W.  Clark. 
1986.  Historic  status  of  black- 
footed  ferret  habitat  in  Montana. 
Great  Basin  Naturalist  Memoirs 
8:63-71. 

French,  Norm  R.,  Dennis  M.  Stod- 
dart,  and  B.  Bobek.  1975.  Patterns 
of  demography  in  small  mammal 
populations,  p.  73-102.  In  Golley, 
Frank  B.,  K.  Petrusewicz,  and  L. 
Ryszkowski  (eds.).  Small  mam- 
mals: their  productivity  and  popu- 
lation dynamics.  Cambridge  Uni- 
versity Press,  Cambridge.  451  p. 

Hall,  E.  Raymond.  1981.  The  mam- 
mals of  North  America.  Volumes 
I,  II.  Second  Edition.  John  Wiley 
and  Sons,  New  York.  1181+90  p. 

Hayward,  C.  Lynn,  D.  Elden  Beck, 
and  Wilmer  W.  Tanner.  1958.  Zo- 
ology of  the  upper  Colorado  River 
basin.  Brigham  Young  University 
Science  Bulletin,  Biological  Series 
1:1-74. 

Heaney,  Larry  R.,  and  Bruce  D.  Pat- 
terson. Eds.  1986.  Island  biogeo- 
graphy of  mammals.  Academic 
Press,  New  York.  271  p. 

Hill,  John  E.,  and  James  D.  Smith. 
1984.  Bats,  a  natural  history.  Univ. 
of  Texas  Press,  Austin.  243  p. 

Kirkland,  Gordon  L.,  Jr.  1981.  The 
zoogeography  of  the  mammals  of 
the  Uinta  Mountains  region.  The 
Southwest.  Naturalist  26:325-339. 


Lechleitner,  Robert  R.  1969.  Wild 
mammals  of  Colorado:  their  ap- 
pearance, habits,  distribution,  and 
abundance.  Pruett  Publishing 
Company,  Boulder.  254  p. 

Lister,  Robert.  1983.  The  ancients  of 
the  canyons,  p.  48-57.  In  Stegner, 
Wallace  (ed.).  This  is  Dinosaur. 
Roberts  Rinehart  Inc.  Boulder, 
Colorado,  93  p. 

Long,  Charles  A.  1965.  The  mammals 
of  Wyoming.  University  of  Kansas 
Publications,  Museum  of  Natural 
History  14:493-758. 

MacArthur,  Robert  H.,  and  Edward 
O.  Wilson.  1963.  An  equilibrium 
theory  of  insular  zoogeography. 
Evolution  17:373-387. 

MacArthur,  Robert  H.,  and  Edward 
O.  Wilson.  1967.  The  theory  of  is- 
land biogeography.  Monographs 
in  Population  Biology,  Princeton 
University  Press  1:1-203. 

Murie,  Olaus,  and  Joseph  W.  Pen- 
fold.  1983.  The  natural  world  of 
Dinosaur,  p.  31-47.  In  Stegner, 
Wallace  (ed.).  This  is  Dinosaur. 
Roberts  Rinehart,  Inc.,  Boulder, 
Colorado,  93  p. 

Newmark,  William  D.  1986a.  Mam- 
malian richness,  colonization,  and 
extinction  in  western  North 
American  national  parks.  Unpub- 
lished Ph.D.  dissertation.  Univer- 
sity of  Michigan,  Ann  Arbor.  172 

P- 

Newmark,  William  D.  1986b.  Spe- 
cies-area relationship  and  its  de- 
terminants for  mammals  in  west- 
ern North  American  national 
parks.  Biological  Journal  of  the 
Linnean  Society  28:83-98. 

Newmark,  William  D.  1987.  A  land- 
bridge  island  perspective  on  mam- 
malian extinctions  in  western 
North  American  parks.  Nature 
325:430-432. 

Walker,  Danny  N.  1983.  Faunal  re- 
mains and  subsistence  practices  at 
5MF605,  a  fur  trading  post  in 
northwestern  Colorado.  South- 
western Lore  49:6-29. 

Warren,  Edward  R.  1910.  The  mam- 
mals of  Colorado.  G.  P.  Putnam's 
Sons,  New  York,  300  p. 


260 


Warren,  Edward  R.  1942.  The  mam- 
mals of  Colorado.  University  of 
Oklahoma  Press,  Norman.  Second 
revised  edition.  330  p. 

Wishart,  David  J.  1979.  The  fur  trade 
of  the  American  west,  1807-1840. 
University  of  Nebraska  Press,  Lin- 
coln, 237  p. 

Young,  Stanley  P.  1944.  Their  history, 
habits,  economic  status  and  con- 
trol, p.  1-173.  In  Young,  Stanley  P., 
and  Edward  A.  Goldman.  The 
wolves  of  North  America.  Ameri- 
can Wildlife  Institute,  Washington. 
636  p. 

Appendix  1 

List  of  mammalian  species  from 
Dinosaur  National  Monument. 
Species  are  represented  by 
specimens  in  collections  unless 
otherwise  noted  in  parentheses. 
Those  specimens  not  in  the  U.  S. 
Fish  and  Wildlife  Service's 
Biological  Surveys  Collections  in 
Fort  Collins  (BS/FC),  or  known  only 
from  near  the  Monument,  are  so 
noted  in  parentheses.  See  text  for 
species  excluded  from  this  list. 
Additional  information  on 
specimens  or  sight  records  is 
available  from  the  authors. 

Sorex  monticolus  (Montane  shrew) 
Myotis  californicus  (California  myotis) 
Myotis  ciliolabrum  (Western  small- 
footed  myotis) 

Myotis  evotis  (Long-eared  myotis) 
Myotis  lucifugus  (Little  brown  bat;  5 
mi  SE  Elk  Springs,  UCM) 
Myotis  thysanodes  (Fringed  myotis) 
Myotis  volans  (Long-legged  myotis) 
Myotis  yumanensis  (Yuma  myotis) 
Lasiurus  cinereus  (Hoary  bat) 
Lasionycteris  noctivagans  (Silver- 
haired  bat) 

Pipistrellus  hesperus  (Western  pipis- 
trelle) 

Eptesicus  fuscus  (Big  brown  bat) 
Euderma  maculatum  (Spotted  bat) 
Plecotus  townsendii  (Townsend's  big- 
eared  bat) 

Antrozous  pallidus  (Pallid  bat) 


Sylvilagus  audubonii  (Desert  cotton- 
tail) 

Sylvilagus  nuttallii  (Nuttall's  cotton- 
tail) 

Lqjus  californicus  (Black-tailed  jack- 
rabbit) 

Lepus  townsendii  (White-tailed  jack- 
rabbit) 

Tamias  dorsalis  (Cliff  chipmunk) 
Tamias  minimus  (Least  chipmunk) 
Tamias  quadrivittatus  (Colorado  chip- 
munk) 

Marmota  flaviventris  (Yellow-bellied 
marmot;  Castle  Park,  UCM) 
Spermophilus  lateralis  (Golden- 
mantled  ground  squirrel) 
Spermophilus  elegans  (Wyoming 
ground  squirrel;  Two  Bar  Spring, 
DMNH) 

Spermophilus  tridecemlineatus  (Thir- 
teen-lined  ground  squirrel) 
Spermophilus  variegatus  (Rock  squir- 
rel) 

Cynomys  leucurus  (White-tailed  prai- 
rie dog) 

Thomomys  talpoides  (Northern  pocket 
gopher;  Pot  Creek,  DMNH) 
Perognathus  fasciatus  (Olive-backed 
pocket  mouse) 

Perognathus  parvus  (Great  Basin 
pocket  mouse) 

Dipodomys  ordii  (Ord's  kangaroo  rat) 
Castor  canadensis  (Beaver) 
Reithrodontomys  megalotis  (Western 
harvest  mouse) 

Peromyscus  crinitus  (Canyon  mouse) 
Peromyscus  maniculatus  (Deer  mouse) 
Peromyscus  truei  (Pinon  mouse) 
Onychomys  leucogaster  (Northern 
grasshopper  mouse) 
Neotoma  cinerea  (Bushy- tailed 
woodrat) 

Microtus  longicaudus  (Long- tailed 
vole) 

Microtus  montanus  (Montane  vole) 
Lemmiscus  curtatus  (Sagebrush  vole) 
Ondatra  zibethicus  (Muskrat;  Castle 
Park,  UCM) 

Erethizon  dorsatum  (Porcupine;  Pot 
Creek  near  Pat's  Hole,  DMNH) 
Canis  latrans  (Coyote) 
Canis  lupus  (Gray  wolf,  +;  Douglas 
Spring,  UCM) 

Vulpes  vulpes  (Red  fox;  ca.  Zenobia 
Peak,  Gary  1911) 


Urocyon  cinereoargenteus  (Gray  fox; 

Castle  Park,  UCM) 

Ursus  americanus  (Black  bear) 

Ursus  arctos  (Grizzly  bear,  +) 

Bassariscus  astutus  (Ringtail;  Castle 

Park,  UCM) 

Procyon  lotor  (Raccoon) 

Mustela  frenata  (Long-tailed  weasel; 

Castle  Park,  UCM) 

Mustela  vison  (Mink;  sightings  in  Lo- 

dore  Canyon) 

Spilogale  gracilis  (Western  spotted 
skunk;  Irish  Canyon,  ca.  Lodore) 
Mephitis  mephitis  (Striped  skunk) 
Taxidea  taxus  (Badger;  Two  Bar 
Spring,  DMNH) 

Lutra  canadensis  (River  otter;  Yampa 
Canyon,  Warren  1942) 
Felis  concolor  (Mountain  lion;  Grey- 
stone,  UCM) 
Felis  rufus  (Bobcat) 
Cervus  elaphus  (Wapiti) 
Odocoileus  hemionus  (Mule  deer;  Pot 
Creek,  USNM) 

Antilocapra  americana  (Pronghorn) 

Bison  bison  (Bison,  +) 

Ovis  canadensis  (Bighorn  sheep) 

C+  =  species  is  extirpated  from  the  Monu- 
ment) 


261 


Sampling  Problems  in 
Estimating  Small  Mammal 
Population  Size^ 

George  E.  Menkens,  Jr.^  and  Stanley  H. 
Anderson^ 


Abstract.— Estimates  of  population  size  are 
influenced  by  four  sources  of  error:  measurement, 
sampling,  missing  data,  and  gross  errors. 
Measurement  error  can  be  reduced  by  using  the 
correct  estimator,  reducing  variation  in  capture 
probabilities,  and  by  increasing  sample  size  and  trap 
period  length.  Sampling  error  can  be  decreased  by 
increasing  the  number  of  grids  trapped. 


Species  conservation  and  manage- 
ment or  analysis  of  environmental 
impacts  require  accurate  estimates  of 
population  size.  Because  censusing 
entire  populations  is  difficult,  if  not 
impossible,  a  sampling  program  is 
generally  employed  to  estimate  ani- 
mal abundance.  In  small  mammal 
studies,  sampling  is  frequently  per- 
formed using  live  traps  placed  in 
grids.  Numerous  approaches  have 
been  used  to  estimate  animal  abun- 
dance on  trapping  grids  (e.g.,  catch- 
per-unit  effort,  removal  methods)  but 
capture-mark-recapture  techniques 
are  the  most  commonly  used  (Seber 
1986). 

Four  sources  of  error  may  influ- 
ence an  estimator's  bias  and  preci- 
sion (Cochran  1977,  McDonald  1981). 
Two,  missing  data  and  gross  errors 
(e.g.,  misreading  tag  numbers)  are 
"human"  errors  and  can  be  avoided 
by  using  careful  field  and  laboratory 
techniques.  The  remaining  sources, 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  North  America  (Flagstaff. 
AZ.July  19-21,  1988). 

^George  E.  Menkens,  Jr.,  is  a  Research 
Associate  with  the  Wyoming  Cooperative 
Fish  and  Wildlife  Research  Unit,''  Laramie, 
WY  82071. 

'Stanley  H.  Anderson  is  Leader,  Wyo- 
ming Cooperative  Fish  and  Wildlife  Re- 
search Unit,"  Laramie,  WY  82071. 

"Cooperators  in  the  Wyoming  Coopera- 
tive Fish  and  Wildlife  Research  Unit  include: 
the  Department  of  Zoology  and  Physiology, 
University  of  Wyoming;  Wyoming  Game 
and  Fish  Department:  and  the  U.S.  Fish  and 
Wildlife  Service. 


measurement  and  sampling  error, 
may,  in  many  cases  greatly  affect  an 
estimate  (McDonald  1981).  Measure- 
ment error  is  the  error  resulting  from 
the  use  of  imprecise  or  biased  (or  a 
combination  of  these)  data  collection 
methods  (McDonald  1981).  In  mark- 
recapture  studies,  measurement  er- 
ror influences  the  bias  and  precision 
of  an  estimate  for  any  single  grid. 
Sampling  variance  is  considered  to 
be  a  measurement  error  in  mark-re- 
capture studies  (White  et  al.  1982). 
Sampling  error  is  error  introduced 
by  natural  variation  between  sam- 
pling units,  i.e.,  trap  grids. 

Potentially  large  sources  of  meas- 
urement error  in  mark-recapture 
studies  may  result  from  capture 
probability  variation  and  model  se- 
lection. All  mark-recapture  estima- 
tors make  specific  assumptions  about 
capture  probability  variation  within 
and  among  animals  and  trapping 
days.  Three  factors  influencing  indi- 
vidual capture  probability  variation 
have  received  attention  (Burnham 
and  Overton  1969,  Otis  et  al.  1978, 
Pollock  1981,  Seber  1982)  and  are 
time,  behavior,  and  individual 
heterogeneity.  Models  assuming  time 
variation  allow  all  animals  to  have 
the  same  capture  probability  on  a 
given  day,  but  this  probability  may 
change  between  days.  Models  allow- 
ing behavioral  responses  to  trapping 
assume  all  animals  initially  possess 
identical  capture  probabilities,  but 
these  probabilities  may  change  upon 
first  capture.  Capture  probabilities 


may  increase  (animals  become  trap 
happy)  or  decrease  (animals  become 
trap  shy)  after  initial  capture.  Models 
assuming  that  individual  heterogene- 
ity is  present  allow  each  animal  to 
have  a  unique  capture  probability 
that  does  not  change  over  time.  (Com- 
binations of  these  factors  may  also 
occur.  For  example,  an  animal's  cap- 
ture probability  may  be  influenced 
by  both  time  and  behavioral  effects. 

Model  selection  is  another  source 
of  measurement  error.  Selection  of  an 
inappropriate  or  incorrect  model  for 
data  analysis  results  in  estimates 
with  unknown  degrees  of  bias  and 
unacceptably  large  or  unrealistically 
small  standard  errors  (Otis  et  al. 
1978,  White  et  al.  1982).  CAPTURE 
(Otis  et  al.  1978)  is  a  widely  used 
computer  program  for  estimating 
population  size  using  mark-recapture 
data  that  also  provides  an  objective 
method  for  selecting  the  correct 
model  when  any  of  the  above 
sources  of  capture  probability  vari- 
ation are  present. 

In  this  paper,  we  investigate  the 
effects  that  variation  in  capture 
probabilities  due  to  time,  behavior, 
and  individual  heterogeneity  have  on 
estimates  of  animal  abundance  and 
model  selection.  We  also  discuss  im- 
provement of  an  estimate  using  data 
pooling.  We  use  these  results  to 
show  how  reducing  trap  period 
length  influences  estimator  bias, 
standard  error,  and  confidence  inter- 
val coverage  rate,  and  discuss  how 
this  may  help  reduce  the  number  of 


262 


grids  required  to  detect  a  given  dif- 
ference between  yearly  estimates  of 
population  sizes. 


Material  and  Methods 

To  investigate  effects  of  both  capture 
probability  variation  and  trap  period 
reduction,  we  used  program  CAP- 
TURE (Otis  et  al.  1978)  to  randomly 
generate  and  analyze  data  sets  with 
known  population  characteristics 
(see  Menkens  1987  for  details).  CAP- 
TURE contains  eight  models,  five 
with  estimators,  for  estimating  popu- 
lation size  for  closed  populations 
when  capture  probabilities  do  not 
vary  (model  M(o)),  or  when  they 
vary  with  time  (model  M(t)),  behav- 


ioral response  (model  M(b)),  individ- 
ual heterogeneity  (model  M(h))  or  a 
combination  of  the  behavioral  and 
individual  heterogeneity  models 
(model  M(bh)).  Using  CAPTURE,  we 
specified  the  number  of  trapping  pe- 
riods, population  size,  and  capture 
probabilities,  and  patterns  of  vari- 
ation. CAPTURE  was  then  used  to 
analyze  each  data  set. 

We  analyzed  the  same  data  sets 
using  Chapman's  unbiased  version 
of  the  Lincoln-Petersen  estimator  and 
its  variance  estimator  (Seber  1982). 
Because  the  Lincoln-Petersen  estima- 
tor uses  data  from  only  two  periods, 
each  data  set  was  split  prior  to  esti- 
mation. Thus  in  a  5  day  trapping 
study,  the  first  3  days  constituted  the 
marking  period,  and  the  second  2 


Table  1  .—Capture  probability  patterns  used  in  simulations  (from  Menkens 
1987  and  Menkens  and  Anderson  in  press).  Good  capture  probabilities  are 
defined  as  being  large  (generally  >  0.30)  with  little  difference  (about  0.15) 
between  the  highest  and  lowest  capture  probability.  Poor  capture  proba- 
bilities are  defined  as  being  low  with  large  differences  between  the  highest 
and  lowest  capture  probability.  "Model"  refers  to  the  CAPTURE  model  un- 
der which  the  data  were  generated.  See  the  text  for  description  of  the 
model  abbreviations,  p  =  capture  probability,  c  =  recapture  probability 
(trap  shyness  =  p(0.50),  trap  happiness  =  p(1.50)),  all  simulations  were  run 
for  5  and  1 0  day  capture  periods  (t). 


Model 


Poor 


Good 


M(o) 
M(h) 
M(b) 
M(bh) 


M(t) 


P=.l 

p  =  0,05,  0.10,  0.25' 
p  =  0.10,  c  =  0.50 
p  =  0.05,  0.20,  0.40 
c  =  0.50 

p  =  0.05.  0.20.  0.40 
c  =  0.50  or  1.502 
p  =  0.10. 0.15,  0.05, 
0.15,0.10 
t  =  5 

p  =  0.10,  0.10,  0.15, 

0.15,0.05,0.05, 

0.15.0.15,0.10, 

0.10 

t=  10 


p  =  ,5. 

p  =  0.40,  0.50,  0.60' 
p  =  0.50,  c  =  1.50 
p  =  0.20,  0,30,  0.40 
c=:  1.50 

p  =  0.20.  0,15,  0.25 
c  =  0.50  or  1.502 
p  =  0.50,  0,55,  0,40, 
0.55,  0.50 
t  =  5 

p  =  0.50,  0.50,  0.55, 
0.55,0.40,  0.40, 
0,55,0,55,0.50, 
0.50 
t=  10 


'Three  groups  of  animals  were  assumed  to  be  present  in  the  population,  the  first  group 
was  associated  with  the  first  capture  probability,  the  second  group  with  the  second 
capture  probability,  the  third  group  with  the  third  capture  probability.  For  N=  50,  ani- 
mals 1-20  were  in  group  1,  21-40  in  group  2,  41-50  were  in  group  3.  For  N=  100,  animals 
1-40  were  in  group  1,  4 1-80  were  in  group  2,  81-100  were  in  group  3. 

'When  a  heterogeneous  recapture  probability  was  assumed,  half  of  the  animals 
became  trap  shy,  half  became  trap  happy. 


days  was  the  recapture  period.  In 
studies  10  days  long,  the  first  5  days 
were  the  marking  period,  the  second 
5  days  the  recapture  period. 

Data  were  generated  for  a  wide 
range  of  conditions.  We  used  trap 
periods  of  5  and  10  days,  population 
sizes  of  50  and  100  and  a  wide  vari- 
ety of  capture  probability  patterns 
(table  1).  One  thousand  data  sets 
were  generated  for  each  combination 
of  these  conditions.  In  this  paper,  we 
only  generated  data  meeting  the  as- 
sumptions of  one  of  the  five  models 
with  estimators  in  CAPTURE.  For 
each  data  set,  CAPTURE  was  forced 
to  perform  the  analysis  using  the  cor- 
rect model.  For  example,  if  data  were 
generated  under  the  assumption  of 
time  variation,  CAPTURE  was  forced 
to  use  model  M(t)  for  the  analysis. 
Simulations  were  also  performed  us- 
ing the  same,  and  additional,  capture 
probabilities  (table  1),  with  CAP- 
TURE being  allowed  to  select  an  esti- 
mator using  its  model  selection  pro- 
cedure. 


Results 

Performance  of  both  the  Lincoln-Pe- 
tersen estimator  and  CAPTURE  is 
dependent  upon  the  size  and  magni- 
tude of  the  variation  in  capture 
probabilities  (table  2).  Estimators 
have  lower  degrees  of  bias,  smaller 
standard  errors,  and  higher  confi- 
dence interval  coverage  rates  when 
capture  probabilities  are  high  and 
their  variation  is  low  (tables  1  and  2) 
over  all  population  sizes.  When  cap- 
ture probability  variation  is  constant, 
the  estimator's  bias  tends  to  decrease 
and  confidence  interval  coverage 
rates  increase  with  increasing  popu- 
lation size  (table  2).  Although  this 
pattern  is  evident  for  standard  er- 
rors, patterns  of  change  with  increas- 
ing sample  size  are  not  as  clear  (table 
2). 

In  general,  the  estimator's  bias  de- 
creases and  confidence  interval  cov- 
erage rates  increase  as  trapping  pe- 
riod length  increases  (table  2).  This 


263 


pattern  is  not  as  obvious  for  standard 
errors,  although  they  do  tend  to  im- 
prove with  increasing  trap  period 
length  (table  2).  In  most  cases  the 
magnitude  of  change  in  bias  is 
smaller  for  good  capture  probabili- 
ties than  for  poor  capture  probabili- 
ties when  the  trapping  period  in- 
creases (table  2).  Although  estimated 
standard  errors  tend  to  decrease 
with  lengthening  trap  periods  (more 
so  with  good  capture  probabilities), 
the  magnitude  of  this  change  is  gen- 
erally smaller  than  is  change  in  bias 
(table  2).  As  with  bias  and  standard 
error,  confidence  interval  coverage 
rates  improve  as  trapping  period  in- 
creases; the  magnitude  of  change 
tends  to  be  larger  when  capture 
probabilities  are  poor  (table  2). 

Except  when  data  were  generated 
under  model  M(o),  CAPTURE  se- 
lected the  correct  model  less  than 
11%  of  the  time  (table  3).  The  Lin- 
coln-Petersen  estimator  failed  to  pro- 
vide an  estimate  at  most  7%  of  the 
time  (table  3). 


Discussion 

In  small  mammal  studies,  measure- 
ment errors  can  significantly  influ- 
ence an  estimator's  bias  and  preci- 
sion. This  study  shows  the  impor- 
tance of  both  reducing  capture 
probability  variation  and  increasing 
the  size  of  those  probabilities  on 
measurement  error.  Decreasing  cap- 
ture probability  variation  reduces  the 
estimate's  bias  and  coefficient  of 
variation,  and  increases  its  confi- 
dence interval  coverage  rate.  This 
result  has  also  been  stressed  by 
Burnham  and  Overton  (1969),  Menk- 
ens (1987),  Menkens  and  Anderson 
(in  press),  Otis  et  al.  (1978),  and 
White  et  al.  (1982).  Of  particular  sig- 
nificance is  the  need  to  reduce  vari- 
ation due  to  behavioral  responses 
(i.e.,  trap-happiness  and  shyness) 
and  individual  heterogeneity,  espe- 
cially when  these  factors  act  in  con- 
cert (Menkens  and  Anderson  in 
press,  Otis  et  al.  1978,  White  et  al. 


Table  2. -r Simulation  results  for  N  =  50  and  100  when  CAPTURE  was  forced  to  use  tiie 
correct  estimator.  CAPTURE  refers  to  the  appropriate  CAPTURE  nrtodel  for  analysis,  L-P  = 
lincoln-Petersen  estimate,  Model  is  the  model  under  which  the  data  were  generated 
by  CAPTURE,  P  =  poor  capture  probabilities,  G  =  good  capture  probabilities  (see  table 
1  for  definitions),  t  =  length  of  trapping  period  <in  days),  PRB  =  percent  relative  bias,  SE 
=  empirical  standard  error,  CIC  =  confidence  interval  coverage  rate. 


IModel 


L-P 


CAPTURE 


50 


100 


SO 


100 


M(o) 
t  =  5 


M(h) 

t  =  5 
PRB 
SE 
CIC 


-55.4 
1.1 
16,2 


G 


PRB 

-52.3 

0.02 

-34.9 

-0,5 

-9.0 

1,4 

17.3 

-0.2 

SE 

0.5 

0.2 

0.9 

0,3 

1,0 

0,2 

2.3 

0.3 

CIC 

27.6 

87,0 

57.1 

91,2 

80.7 

90.7 

87.5 

92.1 

t=  10 

PRB 

-11.5 

0.0 

-2,5 

0,0 

18.3 

-0.10 

7.1 

-0,50 

SE 

0.6 

0.1  ■ 

1,1 

0,1 

1,1 

0.1 

1.2 

0.1 

CIC 

'  73.8  ■ 

92,3 

83,8 

93,4 

89,1 

92.1 

91.6 

93,2 

-1.2 
0.2 
86.4 


-40.9 
1.6 
40.1 


1.6 
0.3 
89.0 


-49.4 
0.3 
22.8 


12.8 
0.3 
81.1 


-43.0 
0.4 
6.0 


16.5 
0.5 
67.1 


PRB 

-28,2 

-1,0 

-24.9 

-0.7 

-16,1 

7.2 

-10,0 

9.5 

SE 

0.6 

0,1  , 

1.0 

0.1 

0.5 

0,2 

0,8 

0.2 

CIC 

48.0 

87.9 

52.4 

91.4 

52.8 

92.3 

59,9 

63.2 

M(b) 

t  =  5 

PRB 

-52.0 

-12.2 

-30.5 

-12.1 

-70.0 

2.2 

-59.1 

3.2 

SE 

1.6 

0.2 

1.9 

0.4 

0.3 

0,6 

0.7 

0.9 

CIC 

9.4 

52,4 

34.1 

40.0 

13.3 

81.4 

36.4 

87.3 

t=  10 

PRB 

6.2 

-4,0 

43,7 

-3.7 

-33.3 

.  -1.4 

-12.3 

-0.6 

SE 

0.7 

0,1 

2.3 

0.1 

0.6 

0.1 

1.7 

0.1 

CIC 

66,6 

54.3 

91.3 

43.6 

497 

82.4 

64.8 

89.9 

M(bh)  (set  1) 

t  =  5 

PRB 

-23.5 

26.0 

-23.1 

45.8 

-17.8 

-17.8 

-1.5 

-1.5 

SE 

0,5 

0.9 

0.8 

2.2 

0.6 

0.6 

1.7 

1.7 

CIC 

49.5 

88,0 

41.3 

99.1 

60.7 

60.7 

70,9 

70.9 

t=  10 

PRB 

-0.4 

-13.8 

-13.7 

27.9 

-28.8 

-4.1 

-24.6 

-3.5 

SE 

0.5 

0.3 

0.5 

1.1 

0.4 

0,4 

0.8 

0.6 

CIC 

83.4 

43,1 

27.5 

92.5 

40.5 

69,6 

43.2 

75.2 

(set  2) 

t  =  5 

PRB 

-32.5 

-17,8 

-31.3 

-3.6 

-31.3 

-40,6 

-18.8 

-22,1 

SE 

0.6 

0.6 

1.1 

1.1 

0.4 

0,4 

1.2 

1.2 

CIC 

33.6 

64,8 

24.4 

82.4 

46.2 

47,6 

67.2 

64,9 

t=  10 

PRB 

-20.5 

0,2 

-20.2 

-0.2 

-28.8 

-4,9 

-24.6 

1,7 

SE 

0,4 

0,4 

0.7 

0.5 

0.4 

0,6 

0.8 

1,4 

CiC 

32.1 

86,7 

17.2 

91.2 

40.5 

67,7 

43.2 

73,5 

M(t) 

t  =  5 

PRB 

-0.2 

0,0 

■0.4 

0,0 

3,8 

-1,2 

3.5 

-0,6 

SE 

0.6 

0,1 

0.6 

0,1 

0.9 

0,1 

1.0 

0.1 

CiC 

85.6 

91,1 

90.6 

93,7 

79.8 

78,1 

89.0 

87.4 

t=  10 

PRB 

-0.2 

0,0 

-0.1 

0,0 

-6.4 

0,0 

-0.8 

-0.1 

SE 

0.3 

0,1 

0.2 

0,1 

0.3 

0,1 

0.3 

0.1 

CIC 

87.4 

95.3 

92.4 

92,7 

74.6 

95,3 

90.7 

90.6 

264 


1982).  Reduction  of  rime  variarion, 
parricularly  if  the  Lincoln-Petersen 
estimator  is  used,  is  important,  but 
not  as  crirical  (Menkens  1987,  Menk- 
ens and  Anderson  in  press).  Again, 
reducing  variarion  in  capture  proba- 
bilities leads  to  estimates  that  have 
lower  bias  and  increased  precision. 

Methods  for  reducing  variation  in 
capture  probabilities  are  numerous 
(see  Oris  et  al.  1978,  Seber  1986, 
White  et  al.  1982).  Behavioral  re- 
sponses may  be  reduced  by  the  use 
of  different  capture  and  recapture 
techniques.  For  example,  animals 
could  be  captured  using  live  traps 
and  marked,  and  then  "recaptured" 
visually  using  spotting  scopes. 
(Fagerstone  and  Biggins  1986).  In 
addirion,  use  of  traps  not  avoided  by 
animals,  and  use  of  non-intrusive 
marking  techniques  (e.g.,  ear  tags 
instead  of  toe  clipping)  may  also  help 
reduce  behavioral  responses.  Use  of 
traps  not  avoided  by  animals  may 
help  increase  capture  probabilities.  If 
sample  sizes  are  large,  heterogeneity 
may  be  reduced  by  stratifying  the 
data  into  sex  and  age  groups  with 
separate  analyses  performed  on  each 
group  (Oris  et  al.  1978,  White  et  al. 
1982).  If  data  are  strarified  however, 
the  effects  of  small  sample  size  on  the 
esrima tor's  properries  must  be  con- 
sidered. 

Capture  probabiliries  may  be  in- 
creased and  their  variarion  reduced 
after  study  completion  by  fxx)ling 
individual  trap  periods  into  single 
marking  and  recapture  periods  as 
was  done  in  our  simulations  (Menk- 
ens 1987,  Menkens  and  Anderson  in 
press).  When  data  are  pooled  in  this 


way  and  the  Lincoln-Petersen  estima- 
tor used,  capture  probabilities  are  20 
to  25%  higher  than  those  for  individ- 
ual days  (Menkens  1987).  In  most 
cases,  data  pooling  results  in  esti- 
mates with  improved  properries. 

Use  of  the  wrong  model  for  analy- 
sis leads  to  estimates  with  unknown 
degrees  of  bias  and  unacceptably 
large  or  unreasonably  small  standard 
errors  (Oris  et  al.  1978,  White  et  al. 
1982),  thus  contributing  significantly 
to  measurement  error.  In  this  study, 
we  forced  CAPTURE  to  use  the  cor- 
rect model  for  analysis.  This,  is  unre- 
alisric  however,  in  that  biologist 
never  know  which  model  is  appro- 
priate. CAPTURE  provides  a  objec- 
tive model  selection  procedure,  how- 
ever this  procedure  works  poorly 
with  the  small  sample  sizes  typically 
encountered  in  many  field  studies 
(Menkens  1987,  Menkens  and  Ander- 
son in  press.  Oris  et  al.  1978,  White  et 
al.  1982).  In  most  cases,  the  Lincoln- 
Petersen  estimator  is  a  valid  alterna- 
tive to  CAPTURE  when  sample  sizes 
are  small,  except  when  capture 
probabilities  are  influenced  by  severe 
behavioral  responses  or  large  de- 
grees of  individual  heterogeneity 
(Menkens  1987,  Menkens  and  Ander- 
son in  press).  Because  use  of  the  most 
appropriate  model  is  critical,  CAP- 
TURE should  be  used  to  determine 
the  type  and  magnitude  of  capture 
probability  variation  in  a  data  set, 
and  if  variation  is  low,  the  Lincoln- 
Petersen  estimator  should  be  used  in 
analysis  (Menkens  and  Anderson  in 
press). 

Many  additional  factors  contribute 
to  measurement  error.  Eliminating 


these  requires  detailed  knowledge  of 
species  behavior  and  ecology,  and 
use  of  compatible  techniques.  For  ex- 
ample, baits  identical  to,  or  that 
closely  approximate  natural  food 
items,  should  be  used  (Dobson  and 
Kjelgaard  1985).  Tags  that  are  easily 
lost  will  lead  to  severe  overestimates 
of  population  size  and  should  not  be 
used.  Other  factors  that  could  con- 
tribute to  measurement  error  include 
use  of  traps  or  other  activities  that 
decrease  survival  or  increase  emigra- 
tion or  immigration,  and  use  of  im- 
proper traps  for  the  species. 

Sampling  error  is  the  error  that 
results  from  natural  variation  be- 
tween sampling  units;  the  larger  this 
variation,  the  larger  the  number  of 
units  that  must  be  sampled  to  detect 
a  difference  in  population  size.  For 
example,  when  environmental  im- 
pacts are  being  assessed,  sampling 
error  would  be  decreased  by  increas- 
ing the  number  of  grids  in  the  control 
and  experimental  groups.  Reducing 
variation  in  capture  probabilities  al- 
lows decreasing  the  number  of  days 
each  grid  is  trapped  without  large 
increases  in  bias  or  standard  errors 
or  decreases  in  confidence  interval 
coverage  rates.  By  reducing  the  trap 
period,  more  grids  can  be  sampled  in 
a  shorter  period  of  time,  thereby  re- 
ducing sampling  error  and  improv- 
ing the  estimate  of  overall  population 
size.  Trapping  in  as  short  a  time 
interval  as  possible  will  also  decrease 
variation  caused  by  temporal  popu- 
lation effects. 

One  approach  to  reducing  sam- 
pling error  is  to  reduce  intergrid 
variation  by  using  a  stratified  sam- 
pling approach.  In  this  case,  investi- 
gators could  stratify  the  habitat 
based  on  some  characteristic  that  is 
correlated  with  animal  density  and 
trap  within  these  strata.  Sample  sizes 
would  be  estimated  for  each  strata. 


Conclusions 

Reduction  of  capture  probability 
variation  and  maximizing  their  mag- 


Table  3.— The  percentage  of  times  the  Lincoln-Petersen  estlnnator  (LP)  or  the 
appropriate  CAPTURE  model  (CAPTURE)  was  selected  by  CAPTURE'S  model 
selection  routine  (from  Menkens  1987).  Model  refers  to  the  CAPTURE  model 
under  which  the  data  were  generated. 

Model 


Estimator 


M(o) 


M(h) 


M(b) 


M(bh) 


M(t) 


LP 

CAPTURE 


96 
100 


98 
9 


93 
7 


99 
11 


97 
6 


265 


nitude  are  critical  to  obtaining  unbi- 
ased and  precise  estimates  of  popula- 
tion size,  and  also  allow  the  selection 
of  the  proper  model  for  use  in  analy- 
sis. Although  we  have  concentrated 
on  small  mammals,  our  points  con- 
cerning reduction  of  both  variation  in 
capture  probabilities  and  of  measure- 
ment and  sampling  error,  pertain  to 
other  studies  using  mark-recapture 
techniques  (e.g.,  papers  in  Ralph  et 
al.).  Our  conclusions  will  hopefully 
force  investigators  to  realize  that 
their  techniques,  particularly  in 
poorly  designed  and  carelessly  per- 
formed studies,  may  not  provide  as 
detailed  and  profound  conclusions  as 
they  might  expect.  We  reiterate  that 
care  in  designing  a  study  can  mini- 
mize many  (but  not  all)  of  the 
sources  of  measurement  and  sam- 
pling errors  we  have  discussed. 

Acknowledgments 

We  thank  Mark  Boyce,  Marc  Evans, 
Richard  Greer,  Lyman  McDonald, 
and  Brian  Miller  for  comments  on 
earlier  versions  of  this  manuscript. 
The  Department  of  Zoology  and 
Physiology,  the  University  of  Wyo- 
ming provided  the  computer  time 
used  for  the  simulations  reported 
here. 


Literature  Cited 

Burnham,  Kenneth  P.  and  William  S. 
Overton.  1969.  A  simulation  study 
of  live-trapping  and  estimation  of 
population  size.  Oregon  State  Uni- 
versity, Department  of  Statistics, 
Technical  Report  14.  60  p.  and  ap- 
pendix. 

Cochran,  William  G.  1977.  Sampling 
Techniques.  John  Wiley  and  Sons, 
New  York. 

Dobson,  F.  Stephen  and  Julia  D. 
Kjelgaard.  1985.  The  influence  of 
food  resources  on  population  dy- 
namics in  Columbian  ground 
squirrels.  Canadian  Journal  of  Zo- 
ology 63:2095-2104. 


Fagerstone,  Katheleen  A.  and  Dean 
E.  Biggins.  1986.  Comparison  of 
capture-recapture  and  visual 
count  indices  of  prairie  dog  densi- 
ties in  black-footed  ferret  habitat. 
Great  Basin  Naturalist  Memoirs 
8:94-98. 

McDonald,  Lyman  L.  1981.  "Summa- 
rizing remarks:  Observer 
variability.  In  Estimating  numbers 
of  birds.  Studies  in  Avian  Biology 
Number  6,  p.  426-435.  J.M.  Scott 
and  C.J.  Ralph,  editors.  The  Coo- 
per  Ornithological  Society. 

Menkens,  George  E.,  Jr.  1987.  Tempo- 
ral and  spatial  variation  in  white- 
tailed  prairie  dog  (Cynomys 
leucurus)  populations  and  life  his- 
tories in  Wyoming.  Unpublished 
Ph.D.  Dissertation,  University  of 
Wyoming,  Laramie. 

Menkens,  George  E.,  Jr.  and  Stanley 
H.  Anderson,  in  press.  Estimating 
small  mammal  population  sizes. 
Ecology. 

Otis,  David  L.,  Kenneth  P.  Burnham, 
Gary  C.  White,  and  David  R.  An- 
derson. 1978.  Statistical  inference 
from  capture  data  on  closed  ani- 
mal populations.  Wildlife  Mono- 
graphs 62:1-135. 

Pollock,  Kenneth  H.  1981.  Capture- 
recapture  models:  a  review  of  cur- 
rent methods,  assumptions  and 
experimental  design.  In  Estimating 
numbers  of  birds.  Studies  in  Avian 
Biology  Number  6,  p.  426-435.  J.M. 
Scott  and  C.J.  Ralph,  editors.  The 
Cooper  Ornithological  Society. 

Ralph,  C.  John  and  J.  Michael  Scott 
(eds).  1981.  Estimating  numbers  of 
birds.  Studies  in  Avian  Biology 
Number  6,  p.  426-435.  The  Cooper 
Ornithological  Society. 

Seber,  G.A.F.  1982.  The  estimation  of 
animal  abundance.  MacMillan 
Publishing  Company,  New  York. 

Seber,  G.A.F.  1986.  A  review  of  esti- 
mating animal  abundance.  Bio- 
metrics 42:267-292. 


White,  Gary  C,  David  R.  Anderson, 
Kenneth  P.  Burnham,  and  David 
L.  Otis.  1982.  Capture-recapture 
and  removal  methods  for  sam- 
pling closed  p)opulations.  Los 
Alamos  National  Laboratory,  LA- 
8787-NERP. 


266 


The  Design  and  Importance 
of  Long-Ternn  Ecological 
Studies:  Analysis  of 
Vertebrates  In  the  Inyo- White 
Mountains,  California^ 

Michael  L.  Morrison^ 


Abstract.— This  paper  reviews  the  importance  of 
duration  in  the  design  of  studies  of  wildlife-habitat 
relationships.  Long-term  studies  are  especially  suited 
to  examining  slow  processes,  rare  events,  subtle 
processes,  and  complex  phenomena.  Four  major 
alternatives  to  long-term  studies— retroactive  studies, 
substitution  of  space  for  time,  use  of  systems  with  fast 
dynamics  as  analogues  for  systems  wrth  slow 
dynamics,  and  modeling— are  discussed.  All  studies 
should  justify  their  results  and  (especially) 
conclusions-recommendations  with  regard  to  study 
duration.  A  suggested  design  for  a  long-term  study 
of  small  vertebrates  is  presented,  including 
preliminary  data  (as  an  example)  from  the  Inyo- 
White  Mountains  of  eastern  California. 


A  fundamental  quesrion  that  should 
arise  early  in  the  design  process  of 
any  investigation  is  the  duration  of 
study.  Along  with  questions  of  sam- 
pling methods,  sample  size,  seasons 
of  study,  and  the  like,  is  the  central 
question  of  how  long  to  collect  data: 
is  1  week  or  1  month  ample?  Or 
should  the  study  extend  for  1  or 
more  years?  Naturally,  this  is  a 
study-specific  question  based  largely 
on  the  objectives  of  the  research.  As  I 
show  in  this  paper,  however,  a  study 
of  insufficient  length  may  fail  to  at- 
tain its  objectives  regardless  of  the 
strength  of  the  design  components 
(e.g.,  sample  size).  Unfortunately,  the 
researcher  and  manager  may  not 
even  realize  that  the  study  gave  only 
a  partial  picture  of  the  system  under 
study;  this,  then,  raises  the  issue  of 
study  length. 

As  outlined  elsewhere  (e.g..  Likens 
1983,  Wiens  1984,  Strayer  et  al.  1986), 
a  tradition  has  developed  over  the 
past  several  decades — especially 
among  North  American  scientists — of 
the  pursuit  of  short-term  studies. 
This  situation  arose  from  constraints 
imposed  by  funding  duration,  the 
need  to  finish  graduate  programs 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibiarxs,  Reptiles,  and 
Small  Mammals  in  North  America  (Flagstaff. 
AZ.July  19-21.  1988). 

^Associate  Professor  of  Wildlife  Biology. 
Department  of  Forestry  and  Resource  Man- 
agement. University  of  California.  Berkeley, 
CA  94720. 


within  short  periods  of  time,  the 
pressure  placed  on  researchers  to 
publish,  and,  of  course,  human  na- 
ture. A  quote  from  John  A.  Wiens 
(1984)  in  his  review  of  long-term 
studies  in  ornithology  is  appropriate 
here:  "...an  excessive  preoccupation 
with  short-term  studies  can  lead  to 
short-term  insights.  By  restricting  the 
duration  of  investigation,  we  adopt  a 
snapshot  approach  to  studying  na- 
ture. We  can  only  hope  that  the 
glimpses  of  patterns  and  processes 
that  we  obtain  depict  reality  accu- 
rately and  that  something  critical  has 
not  been  missed  because  we  looked 
at  the  system  too  briefly."  These  final 
thoughts — that  the  pattern  we  saw 
may  not  depict  reality,  and  that  a 
critical  factor  may  have  been 
missed — have  direct  implications  for 
the  design  of  future  wildlife-habitat 
relationships  studies.  Such  studies 
are  usually  of  only  1-3  years  in  dura- 
tion. At  best,  they  give  only  a  partial 
view  of  most  ecological  systen\s;  and, 
at  worst,  lead  to  false  interpretations. 

My  objectives  in  this  paper  are  (1) 
to  compare  and  contrast  short-  and 
long-term  studies,  including  discus- 
sion of  when  each  type  of  study  can 
be  most  useful;  I  will  draw  heavily 
from  the  comprehensive  review  of 
long-term  ecological  studies  by 
Strayer  et  al.  (1986).  (2)  Using  a  shidy 
recently  implemented  in  the  Inyo- 
White  mountains  of  eastern  Califor- 
nia, I  will  suggest  a  design  for  long- 
term  studies  that  seeks  to  deternrune 


trends  in  abundance  and  habitat  rela- 
tionships of  small  vertebrates. 

LONG-TERM  STUDIES 

Conceptual  Framework 

As  sunrunarized  by  Strayer  et  al. 
(1986),  long-term  studies  are  espe- 
cially suited  to  exploring  four  major 
classes  of  ecological  phenomena: 
slow  processes,  rare  events,  subtle 
processes,  and  complex  phenomena. 

Slow  Processes 

Long-term  studies  obviously  can 
contribute  to  the  understanding  of 
ecological  processes  that  exceeds  that 
gained  from  studies  of  only  1-3  years 
in  duration.  The  importance  of  this 
contribution  depends  on  the  magni- 
tude of  the  process:  results  obtained 
from  any  several-year  period  of  the 
hypothetical  25-year  curve  (fig.  lA) 
could  differ  substantially  from  other 
periods  (e.g.,  showing  an  increasing 
or  decreasing  trend).  Data  obtained 
during  any  short  period  could  be  ac- 
curate, but  only  for  that  period.  Al- 
though continuous  sampling  may  not 
be  necessary  to  identify  such  a  rela- 
tionship, certainly  regularly-repeated 
sampling  is.  Prominent  examples  of 
such  slow  processes  given  by  Strayer 
et  al.  (1986)  are  forest  succession,  in- 
vasion of  exotic  species,  and  verte- 


267 


brate  population  cycles.  Several  spe- 
cific examples  of  obvious  long-term 
relationships  or  cycles  are  given  in 
Halvorson  (1984):  the  23-fold  differ- 
ence between  peaks  and  low  num- 
bers of  snowshoe  hares  (Lepus  ameri- 
canus)  during  a  15-year  study  by 
Keith  (1983);  and  it  took  12  years  for 
a  relationship  between  conifer  seed- 
crop  and  red  squirrel  (Tamiasdurus 
hudsonicus)  abundance  to  be  repeated 
(Halvorson,  unpubl.  data). 

Rare  Events 

Ecological  phenomena  can  occur  at 
regular  intervals  (fig.  IB);  such 
events  include  catastrophes  (e.g., 
fires,  floods),  population  eruptions, 
and  various  environmental  'lx)ttle- 
necks"  or  "crunches."  Shorter-term 
studies  are  often  used  to  study  such 
events  after  their  occurrence,  focus- 
ing on  the  response  or  recovery  of 
the  system.  Studies  of  post-fire  suc- 
cession (e.g..  Bock  and  Lynch  1970, 
Raphael  et  al.  1987),  and  changes  in 
bird  populations  following  oceanic  El 
Nino  conditions  (e.g..  Barber  and 
Chavez  1983,  Schreiber  and  Schreiber 
1984),  are  a  few  examples.  Short-term 
studies,  cannot,  however,  be  used  to 
study  the  frequency  and  reason  (con- 
text) for  the  event. 


Subtle  Processes 

Here  Strayer  et  al.  (1986)  identified 
processes  that  change  over  time  in  a 
regular  fashion  (e.g.,  monotonic 
change,  a  step-function),  but  where 
the  year-to-year  variance  is  large 
relative  to  the  magnitude  of  the 
longer-term  trend  (as  depicted  in  fig. 
IC).  According  to  Strayer  et  al.,  "A 
short-term  study  will  be  unable  to 
discern  the  long-term  trend,  or,  even 
worse,  will  suggest  a  completely  in- 
correct conclusion  about  the  magni- 
tude and  direction  of  the  change... A 
short-term  record  simply  lacks  the 
statistical  power  to  detect  subtle 
long-term  trends..." 


Complex  Phenon^no 

Evaluation  of  biological  phenomena 
are  often  complicated  by  the  intercor- 
related  nature  of  associated  environ- 
mental factors.  Further,  relationships 
between  dependent  and  independent 
variables  may  be  characterized  by 
both  linear  and  nonlinear  responses 
(e.g.,  Meents  et  al.  1983).  Long-term 
data  are  often  necessary  to  sort  out 
such  relationships  for  several  rea- 
sons. First,  it  may  simply  take  many 
years  for  the  phenomenon  to  reveal 
enough  of  its  characteristics  to  allow 
meaningful  analysis  (e.g.,  to  model 
the  system).  Further,  it  may  be  neces- 
sary to  accumulate  data  for  many 
years  to  provide  the  necessary  statis- 
tical degrees  of  freedom  to  conduct 
complex  analyses  (e.g.,  multivariate 
statistics;  Strayer  et  al.  1986). 

Other  Considerations. — A  myriad 
of  other,  often  related,  factors  indi- 
cate the  need  for  long-term  studies. 
Many  of  these  factors  are  related  to 
the  basic — albeit  complex — biology  of 
the  organism.  Vertebrates  have  long 
generation  time  and  long  life  spans, 
which  tends  to  mask  a  population 
response  to  environmental  change. 
Site  fidelity,  another  common  charac- 
teristic of  adult  vertebrates,  may 
cause  a  time-delay  in  the  response  of 
an  animal  to  perturbation. 

How  Long  is  Long-Term? 

Strayer  et  al.  (1986)  gave  two,  rather 
different,  definitions  to  the  concept 
of  "long-term."  The  first  definition 
considers  the  length  of  study  in 
terms  of  natural  processes.  Quoting 
them,  a  study  is  long-term  "...if  it 
continues  for  as  long  as  the  genera- 
tion time  of  the  dominant  organism 
or  long  enough  to  include  examples 
of  the  important  processes  that  struc- 
ture the  ecosystem  under  study.. .the 
length  of  study  is  measured  against 
the  dynamic  speed  of  the  system 
being  studied." 

A  different  approach  is  to  view 
the  length  of  studies  relatively,  with 


long-term  studies  being  those  that 
have  continued  for  a  longer  time 
than  most  other  such  studies.  By  fol- 
lowing this  definition,  we  are  accept- 
ing human  institutions  and  con- 
straints (e.g.,  human  life  span,  length 
of  graduate  education,  pressure  to 
publish),  and  not  the  rate  of  natural 
processes  (Strayer  et  al.  1986). 

A 


16 

YEARS 


Figure  1  .—Situations  where  long-term  stud- 
ies rTKiy  k>e  useful.  (A)  slow  processes,  (B) 
rare  events,  and  (C)  subtle  changes.  The 
record  in  (C)  is  a  long-term  trend  beginning 
at  Y  =  4  and  increasing  at  5%  per  year  (dot- 
ted line)  to  which  a  random  error  with  a 
variance  equal  to  the  trend  line  has  been 
added.  Redrawn  with  permission  following 
strayer  et  al.  (1986:  fig.  3). 


268 


To  illustrate  the  difficulty  in  defin- 
ing the  length  of  time  necessary  for  a 
study  to  be  considered  long-term, 
Strayer  et  al.  (1986)  contrasted  the 
classic  experiment  on  competitive 
exclusion  in  Paramecium  with  the  for- 
est ecosystem  studies  at  the  Hubbard 
Brook  Experimental  Forest:  Cause 
took  about  20  days  to  elucidate  the 
dynamics  of  the  Paramecium  system; 
the  recovery  of  a  forest  ecosystem 
from  clearcutting  has  been  underway 
for  20  years,  which  is  perhaps  only 
1  / 20  of  the  time  necessary  for  the 
forest  to  reach  steady-state.  By  the 
first  definition.  Cause's  work  is  long 
term,  while  the  20-year  Hubbard 
Brook  work  is  not;  the  latter  becomes 
"long-term"  under  the  second  defini- 
tion. 

Thus,  one  cannot  establish  a  for- 
mal definition  for  "long  term."  Re- 
searchers should  recognize,  however, 
that  conclusions  drawn  from  any 
study  should  consider  the  dynamic 
speed  of  the  system  being  studied. 
As  reviewed  by  Likens  (1983),  there 
are  numerous  examples  which  illus- 
trate that  5  to  20  years  of  baseline 
data  are  required  to  characterize  the 
complexity  of  ecological  interactions 
and  systems. 

Length  of  Study:  Advantages  and 
Disadvantages 

Not  all  studies  must  be  "long  term" 
to  provide  reliable  results.  Descrip- 
tive studies  of  essentially  static  pat- 
terns (e.g.,  morphology,  genetic  char- 
acteristics of  species),  of  processes  at 
the  individual  level  (e.g.,  growth,  be- 
havior), or  evolutionary  patterns  or 
systematic  relationships  do  not  nec- 
essarily require  long-term  study. 
These  phenomena  occur  on  time 
frames  that  are  either  very  short  or 
very  long  relative  to  the  normal  du- 
ration of  a  short-term  study  (Wiens 
1984).  The  principal  disadvantages  of 
long-term  studies  are  not  ecological, 
but  practical.  The  need  for  continued 
support  of  money,  time,  staff,  and 
facilities;  the  problems  associated 


with  the  study  falling  into  unproduc- 
tive complacency;  and  environmental 
concerns  that  often  require  immedi- 
ate, even  if  incomplete,  answers. 

As  pointed  out  by  Wiens  (1984), 
long-term  studies,  because  of  the  in- 
tense and  continued  commitment  of 
time  and  money,  must  focus  on  just  a 
few  specific  situations.  Long-term 
work,  therefore,  must  sacrifice  the 
breadth  possible  with  a  series  of 
short-term  studies,  in  exchange  for 
this  increased  detail  and  intensity. 
This,  of  course,  reduces  the  potential 
for  generalizing  from  such  (long- 
term)  studies.  A  degree  of  compro- 
mise between  these  extremes  (short- 
term  vs.  long-term  studies)  is  dis- 
cussed below. 


Alternatives  to  Long-Term  Studies 

There  are  four  classes  of  short-term 
studies  that  can  potentially  provide 
insight  into  long-term  phenomena: 
(1)  retrospective  studies,  (2)  substitu- 
tion of  space  for  time,  (3)  use  of  sys- 
tems with  fast  dynamics  as  ana- 
logues for  systems  with  slow  dynam- 
ics, and  (4)  modeling  (Strayer  et  al. 
1986).  They  raise  the  important  point 
that  such  short-term  approaches  can 
be  integrated  into  an  overall,  longer- 
term,  study,  thus  "...extending  the 
temporal  and  spatial  scales  of  the  in- 
vestigation and  allowing  the  ecolo- 
gist  to  explore  a  wider  range  of  eco- 
logical phenomena  than  nnight  be 
practical  in  a  direct  long-term 
study." 

Retrospective  Studies 

The  record  of  past  conditions  can  be 
used  to  help  reconstruct  a  long-term 
trend.  Obvious  examples  of  such  ap- 
proaches are  tree-rings  and  pollen 
deposition.  Unfortunately,  conclu- 
sions regarding  past  conditions  re- 
lated to  or  even  causing  the  pattern 
remaining  can  only  be  inferred;  fur- 
ther, only  persistent  structures  re- 
main to  be  analyzed. 


Substitution  of  Space  for  Time 

This  is  an  often-used  substitute  for  a 
long-term  study.  Here  sites  with  dif- 
fering characteristics  are  used  in- 
stead of  following  the  course  of  a 
single  or  a  few  sites  for  an  extended 
period.  For  example,  evaluating  suc- 
cession by  simultaneously  using  sites 
of  different  age  (e.g.,  1, 5, 15,  30  years 
post-harvest).  This  approach,  how- 
ever, requires  the  assumption  that  all 
important  environmental  processes 
are  independent  of  space  and  time 
(i.e.,  all  sites  must  have  the  same  en- 
vironmental characteristics  and  his- 
tory). To  provide  valid  results 
through  this  approach  requires,  then, 
that  many  sites  with  very  similar  his- 
tories and  characteristics  be  used.  An 
obvious  problem,  of  course,  is  deter- 
mination of  how  "similar"  sites  must 
be.  Although  results  of  such  studies 
may  theoretically  approach  those  of  a 
long-term  study,  they  can  only  do  so 
with  a  large  number  of  replicates. 
Further,  such  substitutions  cannot 
capture  the  historical  events  that 
shaped  each  site,  but  can  only  mask 
or  "swamp"  the  effect  through  a 
large  sample  size  (which  may  yield 
adequate  results  for  many  applica- 
tions). 

These  problems  can  best  be  dealt 
with  in  studies  combining  direct 
long-term  studies  with  space-for- 
time  substitutions.  Long-term  studies 
done  in  parallel  with  carefully 
matched,  short-term  "substitutes," 
can  factor  out  the  year-to-year  vari- 
ation that  may  mask  general  trends. 

Ottier  Mettiods 

Applying  the  results  of  a  simple  sys- 
tem with  rapid  generation  time  can 
give  insight  into  how  a  system  with  a 
slower  generation  might  behave:  for 
example,  applying  the  results  of  labo- 
ratory studies  on  rodents  to  evalu- 
ations of  population  dynamics  of 
larger  mammals.  Such  extensions  of 
results  have  obvious  drawbacks,  but 
can  be  useful  in  the  development  of 
general  theories  used  to  guide 


269 


longer-term  studies. 

Mathematical  modeling  can,  of 
course,  be  used  to  predict  the  longer- 
term  behavior  of  a  system.  Such 
models  are  often  based  on  guides 
provided  by  various  short-term  stud- 
ies. Obviously,  the  predictive  ability 
of  models  can  only  be  determined 
through  long-term  studies,  and/ or  a 
series  of  short-term  f)erturbations 
that  experimentally  test  them;  the 
latter  will  fail  unless  all  likely  catas- 
trophes and  conditions  can  be  ade- 
quately simulated.  Here  again,  such 
modeling  can  provide  valuable  in- 
sight into  the  design  and  conduct  of 
parallel  long-term  studies. 

Ecological  Monitoring 

Monitoring  of  environmental  condi- 
tions is  a  closely  aligned  aspect  of 
long-term  studies.  When  a  manage- 
ment agency  such  as  the  USDA  For- 
est Service  discusses  the  need  for 
monitoring  of  wildlife  population 
numbers,  they  are  essentially  de- 
scribing a  long-term  study,  the  goal 
of  which  is  to  identify  trends.  Unfor- 
tunately, "monitoring"  has  a  low 
status  in  ecology,  being  widely  re- 
garded as  possessing  little  originality 
and  as  unproductive  of  new  scientific 
knowledge  (Strayer  et  al.  1986). 
Monitoring  data  can  provide,  how- 
ever, essential  support  for  many  re- 
search projects  and  publications  aris- 
ing from  long-term  studies.  In  addi- 
tion, monitoring  programs  can  lead 
to  important  and  unexpected  discov- 
eries (e.g.,  first  report  of  acid  rain  in 
North  America;  Strayer  et  al.  1986). 

Sutcliffe  and  Shachak  (in  Strayer  et 
al.  1986)  outlined  several  elements 
that  are  essential  in  the  conduct  of 
monitoring  programs:  (1)  the  initial 
sampling  design,  variables  to  be 
measured,  and  methodology  must  be 
carefully  chosen;  and  (2)  a  scientist 
capable  of  interpreting  the  data 
should  be  closely  involved  with  all 
aspects  of  the  study,  allowing  modi- 
fication of  design  to  take  advantage 
of  the  ever-increasing  knowledge 


about  the  system  under  study.  A 
critical  aspect  of  any  monitoring  pro- 
gram is  to  eliminate  the  unproduc- 
tive parts  of  the  program  to  allow  for 
more  fruitful  analyses  without  de- 
stroying some  part  of  the  long-term 
core  data  (Strayer  et  al.  1986). 

STUDY  DESIGN 

Introduction 

The  design  of  a  long-term  study 
must  be  sufficiently  simple  to  persist 
over  a  long  period  of  time.  Thus,  es- 
sential measurements  must  be  simple 
enough  to  be  repeatable  by  workers 
with  varying  degrees  of  experience 
(Strayer  et  al.  1986).  There  are  also 
numerous  specific  aspects  of  site  pro- 
tection and  management,  manage- 
ment of  data,  quality  control,  chang- 
ing methodologies,  and  the  like  that 
are  all  critical  to  a  successful  study; 
these  concerns  are  discussed  by 
Strayer  et  al.  (1986)  and  will  not  be 
repeated  here. 

The  design  of  a  long-term  study 
must  also  be  sufficiently  flexible  to 
accommodate  short-term  investiga- 
tions. Long-term  data  often  suggest 
questions  that  can  be  investigated 
through  short-term  exp>erimentation 
or  observations.  A  benefit  of  such  an 
approach  is  that  overall  productivity 
can  be  increased;  the  longer-term  ob- 
jectives of  a  study  can  also  be  more 
easily  funded  as  a  result  of  such 
shorter-term  efforts.  In  summary, 
studies  of  varying  lengths  can  usu- 
ally complement  one  another. 

I  have  designed  and  implemented 
a  study  to  evaluate  both  short-  and 
long-term  responses  of  vertebrates  to 
abiotic  and  biotic  conditions  in  the 
Inyo-White  mountains  (Inyo  and 
Mono  counties)  of  eastern  California. 
The  design  represents  a  compromise 
among  the  many  different  methods 
necessary  to  sample  different  groups 
of  small  vertebrates  on  the  same  site. 
Below  I  briefly  describe  the  sampling 
design,  and  provide  data  on  initial 
surveys.  I  present  this  design  as  a 


possible  template  for  other  studies 
that  seek  to  determine  wildlife-habi- 
tat relationships  and  responses  to 
environmental  changes  (i.e.,  monitor- 
ing) over  the  short-  and  long-term. 

Rationale 

The  overall  objective  of  this  study  is 
to  determine  long-term  behavioral 
and  ecological  attributes  and  interre- 
lationships of  vertebrates  in  the  Inyo- 
White  mountains  of  eastern  Califor- 
nia. Amphibians,  reptiles,  small 
mammals,  and  birds  will  be  censused 
on  a  series  of  sites  in  the  pinyon-juni- 
per  (Pinus  monophylla-Juniperus  os- 
teosperma)  plant  community  on  a 
year-round  basis.  Abiotic  factors  and 
food  resources  will  also  be  sampled. 
Reproductive  physiology  of  small 
mammals  will  be  addressed. 

Numerous  hypotheses  can  be 
evaluated  depending  upon  the  taxo- 
nomic  group(s)  (e.g.,  species  level, 
class  level,  guild  level)  chosen  for 
analysis;  for  example: 

1.  H^l:  The  p>opulation  num- 
bers of  the  group  are  not  re- 
lated to  (a)  food  resources, 

(b)  abiotic  conditions,  and /or 

(c)  population  numbers  of 
other  groups. 

2.  H^2:  The  behavior  (e.g.,  for- 
aging behavior)  of  the  group 
does  not  vary  with  fluctua- 
tions in  (a)  food,  (b)  abiotic 
conditions,  and /or  (c)  popu- 
lation numbers  of  other 
groups. 

3.  H^3:  Population  numbers  of 
the  group  during  spring  are 
not  related  to  (a)  food,  (b) 
abiotic  conditions,  and/ or  (c) 
number  of  other  animals 
during  a  previous  season. 

4.  H  4:  Guild  structure  cannot 

o 

be  identified  on  any  tempo- 
ral basis. 


270 


4a.  H^4:  The  guild  structure 
identified  does  not  vary  with 
variation  in  (a)  food,  (b)  abi- 
otic conditions,  (c)  popula- 
tion numbers  of  other 
groups,  and/ or  (d)  tempo- 
rally. 

This  study  is  designed  to  address 
these  and  numerous  other  null  hy- 
p>otheses.  The  data  set  necessary  to 
answer  any  one  hypothesis  is  very 
similar  to  that  required  to  address 
another  hypothesis.  Thus,  the  num- 
ber of  hypotheses  generated  is,  in  a 
sense,  independent  of  the  effort  ex- 
pended to  collect  the  data. 

This  study  will  contribute  to  our 
understanding  of  the  ecology  of  this 
system  in  several  major  ways.  First, 
it  will  provide  data  on  fluctuations  in 
population  numbers  of  vertebrates, 
thus  serving  a  monitoring  role  (espe- 
cially important  to  the  USDA  Forest 
Service).  Second,  it  will  provide  data 
which  will  allow  development  of 
multi-species  population  models  (by 
myself  and  other  workers),  allow  de- 
velopment of  habitat-relationships 
models  that  incorporate  both  short- 
and  long-term  responses  to  biotic 
and  abiotic  factors,  and  allow  devel- 
opment and  subsequent  testing  of 
models  of  multi-species  interrelation- 
ships at  various  taxonomic  levels. 
Third,  and  possibly  the  most  impor- 
tant aspect  of  the  study,  it  will  result 
in  the  accumulation  of  vast  amounts 
of  ecological  information  on  the  ver* 
tebrate  (and  invertebrate)  commu- 
nity. To  date,  only  brief  and  sporadic 
surveys  have  been  conducted  in  the 
Inyo- White  mountains.  Finally,  it  is 
my  goal  to  use  preliminary  results  to 
generate  specific  hypotheses  that  can 
be  tested  by  my,  or  others',  students. 
For  example,  if  initial  data  indicate 
rejection  of  the  null  hypothesis  of  no 
relationship  between  a  certain  small 
mammal  and  their  prey  base,  then  a 
student  could  select  additional  sites 
where  food  supplementation  and /or 
removal  experiments  could  be  con- 
ducted. Additional,  study-specific 
funding  will  be  sought  for  such  stud- 
ies. This  study  will  thus  generate 


short-term  results  under  the  general 
framework  of  its  long-term  design 
and  goals. 

The  Inyo- White  mountains  were 
chosen  as  the  study  location  for  sev- 
eral reasons.  First,  my  intent  was  to 
select  a  type  of  habitat  that  offered 
structural,  especially  vertical,  diver- 
sity intermediate  between  that  of  a 
grass-  or  shrubland  and  that  of  a  ma- 
ture, hardwood  or  coniferous  forest. 
With  a  canopy  rarely  exceeding  10  m, 
I  will  be  able  to  sample  arthropod 
populations  from  the  upper  canopy. 
This  is  not  conveniently  possible  in 
mature  conifer  forest,  where  the  can- 
opy extends  to  20-30  m  or  more  in 
height.  Second,  I  desired  an  area  that 
offered  only  several  dominant  tree 
and  shrub  species:  this  allows  inten- 
sive sampling  of  all  major  species, 
while  allowing  some  diversity  of 
plant  species  beyond  that  evident  in 
more  monotypic  habitats. 

The  pinyon-juniper  woodland  was 
chosen  because  of  its  extensive  cov- 
erage throughout  the  intermountain 
west.  Further,  the  pinyon-juniper 
woodland  undergoes  few  significant 
changes  in  plant  species  frequency 
and  density  relative  to  earlier  succes- 
sional  communities.  Austin  (1987), 
for  example,  showed  virtually  no 
change  in  a  pinyon-juniper  commu- 
nity in  Utah  between  1974-84.  In  con- 
trast, seeds  and  berries  undergo  of- 
ten marked,  interyear  changes  in  pro- 
duction. Thus,  barring  some  cata- 
strophic change,  the  gross  composi- 
tion of  the  plant  community  used  in 
this  study  should  remain  relatively 
stable,  helping  to  control  for  at  least 
some  of  the  variance  likely  to  be  en- 
countered in  animal  communities. 

A  definition  of  what  I  mean  by 
"long-term"  in  this  study  is  not  yet 
possible,  but  I  have  committed  my- 
self to  this  study  for  an  indefinite  pe- 
riod; 15-20  years  seems  a  minimum. 
The  study  is  designed  to  be  con- 
ducted, at  a  minimum,  by  myself  and 
one  assistant.  Additional  personnel, 
primarily  undergraduate  volunteers 
during  summer,  will  also  be  avail- 
able. Thus,  the  ability  to  adequately 


conduct  the  study  over  the  long-term 
is  considered  in  the  design,  and  will 
be  possible  given  my  focus  (concen- 
tration) on  this  study.  The  initial  de- 
sign can  accommodate  expansion  in 
size  both  through  the  enlargement  of 
each  site  (using  the  original  area  as  a 
standard  core),  and /or  the  addition 
of  additional  sites  (e.g.,  to  sample 
from  a  wider  range  of  ecological  con- 
ditions). Various  ancillary  studies 
will  add  to  my  understanding  of  this 
system,  although  my  primary  goal  is 
to  examine  the  interrelationships 
among  vertebrates  and  their  environ- 
ment. 

Sampling  intensity  will  not  be  in- 
creased beyond  that  discussed  herein 
(see  Methods)  to  avoid  substantial 
impact  (e.g.,  trampling)  by  observers 
on  the  study  sites.  Thus,  an  increase 
in  effort  (given  adequate  time  and 
funding)  will  be  directed  towards  an 
increase  in  site  size,  number  of  sites, 
and/or  towards  ancillary  studies,  the 
decision  based  on  preliminary  data. 

I  will  be  intensively  involved  with 
all  aspects  of  this  study,  including 
establishment  of  the  sites  (already 
accomplished)  and  collection  of  data 
throughout  the  duration  of  the  study. 
It  is  essential,  in  any  long-term  effort, 
that  methodology  be  standardized, 
and  a  high  level  of  quality  control  be 
maintained.  My  involvement  will 
serve  as  the  standard  upon  which 
new  assistants  will  be  trained.  Any 
changes  in  methods,  whether  this  in- 
volves modification  of  sampling  in- 
tensity or  a  change  in  trap  type,  will 
be  fully  documented.  If  any  proce- 
dures must  be  changed,  the  old  and 
new  methods  will  be  run  simultane- 
ously to  allow  for  intercalibration  of 
methods  (as  described  by  Strayer  et 
al.  1986).  All  field  notes  and  data  will 
be  duplicated  or  triplicated  and 
stored  in  several  locations  for  safety. 

Design  Considerations 

The  study  sites — their  number,  size, 
and  location — chosen  for  this  study 
were  selected  to  restrict  samples  to 


271 


convenient  and  modest-sized  popu- 
lations. They  will  be  low  in  cost  to 
sample  and  are  located  in  practical 
locations  for  year-round  access.  With 
the  few  (3)  sites  chosen,  it  would  be 
foolhardy  to  attempt  representation 
with  probability  sampling  of  entire 
populations.  The  study  is  designed  to 
spread  effort  across  important  vari- 
ables to  obtain  some  measure  of  con- 
formation of  results.  I  follow  the 
philosophical  view  of  Popper  (1959), 
as  summarized  by  Kish  (1987):  "The 
choice  of  the  sites  should  strain  to 
increase  the  possibilities  for  falsifica- 
tion." Similar  and  consistent  results 
from  the  replications  yield  stronger 
confirmation  than  a  single  site 
would.  But  if  the  results  are  discor- 
dant, the  replications  are  too  few  to 
yield  dependable  inference;  then  fur- 
ther research  is  indicated.  Discordant 
results  yield  a  healthy  skepticism 
that  naive  "success"  from  a  single 
site  would  obscure  (taken  from  Kish 
1987).  As  discussed  earlier,  the  study 
is  designed  to  allow  an  increase  in 
the  number  of  sites  (or  their  size, 
etc.)  should  early  results  so  indicate. 

The  general  locations  of  the  sites 
were  not  chosen  at  random.  A  gen- 
eral area  was  identified  based  on  (1) 
ease  of  access  during  winter  (e.g., 
within  0.5-1.0  km  of  a  maintained, 
although  usually  dirt,  road),  but  also 
(2)  isolated  from  access  by  off-road 
vehicles.  Using  these  general  guides, 
specific  sites  were  chosen  to  repre- 
sent a  sampling  of  slope,  aspect,  and 
longitudinal  location  in  the  Inyo- 
White  mountains.  My  intent  was  to 
increase  the  likelihood  of  "falsifica- 
tion," which  is  better  served  with 
tests  obtained  in  contrasting  condi- 
tions, as  opposed  to  selecting  more 
or  less  "average"  sites  (see  Kish  1987 
for  a  development  of  this  strategy). 
The  extremes  of  a  relationship  are 
more  informative  than  either  random 
or  modal  or  centralized  selection. 
Three  sites  were  considered  a  mini- 
mum, because  two  sites  might  indi- 
cate a  false,  linear  relationship  in  cer- 
tain factors.  Survey  data  will  be  col- 
lected on  other  areas  throughout  the 


Inyo-White  mountains.  Such  data 
will  provide  useful  information  re- 
garding the  overall  distribution  and 
habitat  associations  of  vertebrates 
throughout  the  ranges.  Further,  these 
sites  will  serve  as  'l3ack-ups"  should 
a  catastrophic  event  occur  on  one  of 
the  three  main  sites.  Because  this 
study  is  largely  exploratory,  such  a 
strategy  was  warranted  (with  the  op- 
tion of  later  expansion). 

METHODS 

Terminology 

I  have  attempted  to  standardize  the 
terms  used  to  describe  the  study  ar- 
eas described  beyond;  they  are: 

"Permanent  site":  a  400  x  400  m 
(16  ha)  area  that  forms  the  long-term 
"study  sites"  used. 

"Point":  a  trap  location  within  a 
site. 

"Ancillary  site":  additional  areas 
(of  various  shapes  and  sizes) 
sampled  at  an  intensity  less  than  on 
the  permanent  sites;  established  to 
increase  scope  of  sampling  effort. 

"Sampling  f>eriod":  a  5-7-day  pe- 
riod in  which  a  permanent  site  is 
sampled. 

"Transect":  the  parallel,  400-m 
long  lines  ("transect")  forming  the 
study  sites. 

"Core  trapping  area":  the  central, 
200  X  200  m  area,  location  within  a 
study  site  where  small-mammal  and 
pitfall  traps  are  placed. 

Sampling  Schedule 

Study  sites  (described  below)  were 
established  during  fall  1987.  All 
methods  outlined  herein  were  evalu- 
ated during  fall  1987,  winter  1987-88, 
and  spring  1988.  Each  permanent  site 
will  be  visited  for  a  5-7-day  period 
on  two  occasions  per  season.  Seasons 
are  defined  as:  spring  (1  Mar.-31 
May);  summer  (1  June-31  Aug.);  fall 
(1  Sept.-30  Nov.);  and  winter  (1  Dec- 
28  Feb.).  The  exact  length  of  visit  will 


be  based  on  trapping  results.  Initial 
order  of  visit  to  sites  will  be  random- 
ized; this  order  then  followed  on  the 
subsequent  visit  in  that  season. 

Remaining  time  available  during  a 
season  will  be  spent  sampling  the  an- 
cillary sites.  These  ancillary  sites  will 
increase  my  knowledge  about  the 
distribution  and  relative  abundance 
of  vertebrates  and  invertebrates  in 
the  Inyo- White  mountains.  Not  all 
data  from  ancillary  sites  will  be  di- 
rectly comparable — to  permanent  or 
other  ancillary  sites — because  of  the 
lower  sampling  intensity.  Neverthe- 
less, they  will  supply  information 
important  to  the  long-term  success  of 
the  study. 

Permanent  (n  =  3)  Study  Sites 

Each  will  be  established  as  a  400  x 
400  m  (16  ha)  site.  A  16-ha  site  was 
chosen  because:  (1)  an  observer  can 
travel  this  distance,  even  over  rough 
terrain,  in  a  short  period  of  time;  (2) 
the  utilized  area  of  most  small  verte- 
brates can  be  sampled  within  a  16-ha 
area;  and  (3)  this  area  allowed  estab- 
lishment of  sites  protected  from 
roadways,  trails,  and  other  human 
activity  (e.g.,  fit  between  cliffs  and 
gullies  that  form  barriers  to  illegal 
vehicle  access).  Each  site  will  be 
sampled  repeatedly  within  a  season 
to  provide  measures  on  within-site 
variability.  The  effective  n  per  per- 
manent site  is,  of  course,  one  (for 
comparison  among  permanent  sites, 
n  =  3).  Each  site  will  have  permanent 
grid  points  marked  at  25-  or  50-m 
intervals  (using  rubber  cattle  ear 
tags). 

SAMPLING 

Amptiibian,  Reptile,  and  Snnall- 
Mammal  Trapping 

One-hundred-one  Sherman  live  traps 
and  41  pit-falls  (two  3.2 1  (3  lb)  rin 
cans  taped  together)  were  estab- 
lished on  each  site.  Live  traps  were 


272 


12.5  m  apart  in  the  center  100  x  100 
m  section  of  a  site,  and  at  25-m  spac- 
ings  in  the  remaining  trapping  area. 
The  closer  trap  spacing  helps  deter- 
mine actual  population  density, 
whereas  the  wider  spacing  in  the  sur- 
rounding area  provides  information 
on  animal  movements. 

Live  traps  are  baited  with  seed 
mixtures  and  checked  each  morning 
and  late  afternoon.  Certain  rodents 
(e.g.,  Peromyscus)  are  active  through- 
out the  year.  During  winter,  there- 
fore, traps  are  provided  with  insulat- 
ing material  (e.g.,  wool).  All  captures 
are  toe-clipped.  Trapping  continues 
until  new  captures  are  minimal  (usu- 
ally 5  days).  Pitfalls  are  not  baited, 
but  captures  are  marked  and  re- 
leased. Traps  are  run  "dry":  holes 
were  drilled  in  each  trap,  rocks 
placed  in  the  bottom  of  each  hole  to 
provide  drainage,  and  a  wooden  lid 
placed  over  the  trap  to  reduce  expo- 
sure. 


Bird  Activity 

The  spot-map  method  (e.g.,  see 
Ralph  and  Scott  1981)  is  used  to  de- 
termine bird  abundance  and  territory 
(during  breeding)  size.  Following  a 
census,  the  observer  slowly  walks 
though  the  entire  site  and  records 
foraging  birds  as  encountered.  Data 
are  recorded  on  activity  and  sub- 
strate used.  (The  specific  methods 
used  for  birds  will  not  be  detailed  in 
this  paper.) 

Vegetation  Sampling 
General  site 

Trees  and  shrubs  will  be  sampled 
once  per  year,  and  grass  and  herba- 
ceous cover  will  be  sampled  once  per 
season.  Changes  in  plant  phenology 
will  be  recorded  as  they  occur.  Vege- 
tation will  be  sampled  using  circular 
plots  and  line  intercepts  centered  at 
each  of  the  81, 50-m-transect  inter- 
cepts. Pinyon  and  juniper  will  be 


counted  and  measured  (e.g.,  dbh, 
height,  vigor,  canopy  cover)  within 
20-m-radius  plots.  Shrubs,  grass,  and 
herbaceous  cover  will  be  measured 
along  40-m-long  line  intercepts  bi- 
secting each  circular  plot  (and  run- 
ning parallel  to  the  main  transect 
line). 


Trap  Locations 

Vegetation  and  soil  characteristics 
will  be  measured  at  each  trap  site. 
The  nearest  tree  in  each  quarter  from 
the  trap  will  be  measured.  Two  5-m- 
p)erpendicular  transects  will  be 
placed  over  each  trap;  shrub  and  her- 
baceous cover  will  be  measured 
along  each  transect.  Soil  moisture, 
compactability,  texture,  and  pH  will 
be  measured  on  each  arm  (2.5  m)  of 
the  trap-site  transects:  at  0.5, 1.0, 1.5, 
and  2.0  m  (one  of  these  distances  per 
arm,  randomized,  for  four  measure- 
ments per  trap).  These  samples  will 
be  gathered  once  during  each  season. 

Abiotic  Factors 

A  weather  station  will  be  established 
near  the  center  of  each  study  site. 
Temperature,  humidity,  and  rain  fall 
will  be  automatically  recorded 
throughout  the  year.  Snowfall  will  be 
measured  by  visiting  each  site  fol- 
lowing snowstorms. 

Ottier  Sampling 

Data  on  arthropod  abundance 
(branch  sampling  and  pan  traps)  and 
cone-seed  production  (of  pinyon,  ju- 
niper and  major  shrub  species)  will 
also  be  collected  (but  not  detailed  in 
this  paper). 

RESULTS 

Only  one  site  has  been  sampled  with 
adequate  intensity  for  presentation  of 
data  at  this  time.  Sampling  occurred 


during  fall  (8  days  during  two  trap- 
ping sessions),  winter  (7  days  during 
two  trapping  sessions),  and  spring 
(12  days  during  three  trapping  ses- 
sions). Pitfalls  were  used  only  during 
the  first  trapping  session  in  the  fall 
and  the  last  session  in  the  spring  (be- 
cause of  snow  and  little  or  no  lizard 
activity).  Chipmunks  and  Great  Ba- 
sin pocket  mice  were  not  active  from 
October-November  until  early 
March. 

The  sagebrush  and  western  fence 
lizards  were  the  most  frequently  cap- 
tured animals  in  pitfalls  (table  1).  A 
single  deer  mouse  (immature)  was 
also  captured  in  a  pitfall  trap.  Inten- 
sity of  pitfall  trapping  has  been  in- 
adequate to  date  to  make  conclusions 
on  their  effectiveness. 

Seven  small  mammal  species  and 
a  skink  were  captured  in  the  live 
traps  (table  1).  The  pinyon  mouse 
was  the  most  abundant  species  cap- 
tured during  fall.  Relatively  few 
pinyon  mice  were  captured  during 
winter,  however  (a  77.9%  decline  be- 
tween fall  and  winter).  The  decline  of 
pinyon  mice  continued  into  spring, 
with  abundance  dropping  64%  be- 
tween winter  and  spring.  (Dnly  a  few 
deer  and  pinyon  mice  were  captured 
during  winter. 

The  highest  overall  abundance  of 
small  mammals  was  found  during 
spring  (table  1).  The  two  species  of 
chipmunks  were  the  most  abundant 
animals  captured.  The  Great  Basin 
pocket  mouse  and  the  deer  mouse 
were  also  captured  frequently  during 
spring. 

DISCUSSION 

Live-trapping  data  indicate  the  im- 
portance of  repeated  sampling  over 
time:  pinyon  mice  apparently  suf- 
fered substantial  winter  mortality. 
Thus,  trapping  in  only  fall  or  spring 
would  have  falsely  indicated  a  rela- 
tively high  or  low  population  size, 
respectively.  Although  this  study  can 
hardly  be  considered  "long-term," 
initial  results  do  highlight  the  need 


273 


for  repeated  sampling  even  over  the 
short  term.  Only  continued  sampling 
will  elicit  the  frequency  and  reasons 
for  such  a  decline.  My  initial  trap- 
ping configuration  contained  a  dense 
trap  placement  (12.5  m  trap  inter- 
vals) in  the  middle  of  the  grid  rela- 
tive to  the  outer  traps  (25-m  spacing). 
My  intent  was  to  use  the  outer  traps 
to  determine  movement  of  animals  in 
and  out  of  the  smaller  100  x  100  m 
area.  Cursory  examination  of  trap- 
ping results  (unpubl.  data)  indicate, 
however,  that  even  the  total  200  x 
200  m  grid  is  not  sufficiently  large  to 
quantify  movements  (i.e.,  animals 
moving  >200  m).  Therefore,  I  suggest 
the  following  modifications  in  trap 
placement:  10  x  10  trapping  grid  with 
15-m  spacing.  This  placement  should 
adequately  sample  the  animals  pres- 
ent. To  detect  movement  (e.g.,  dis- 
persal), trap  lines  can  be  established 
periodically  that  run  perpendicular 
from  the  edge  of  the  trapping  grid. 


ACKNOWLEDGMENTS 

I  thank  David  Strayer  and  Robert  C. 
Szaro  for  reviewing  earlier  drafts, 
and  Lorraine  M.  Merkle  for  prepar- 
ing the  text.  C.  John  Ralph,  Redwood 
Sciences  Laboratory,  Areata,  Califor- 
nia, Jared  Verner,  Forestry  Sciences 
Laboratory,  Fresno,  California  (both 
Pacific  Southwest  Forest  and  Range 
Experiment  Station,  USDA  Forest 
Service),  and  Reginald  H.  Barrett, 
Dept.  Forestry  and  Resource  Man- 
agement, University  of  California, 
Berkeley,  provided  equipment.  The 
director  ((Tlarence  Hall),  superinten- 
dent (David  Trydahl),  and  staff  of 
the  White  Mountain  Research  Sta- 
tion, University  of  California,  Los 
Angeles,  are  thanked  for  supplying 
logistical  support.  John  A.  Keane, 
Martin  L.  Morton,  and  Kimberly  A. 
With  assisted  with  field  work.  Kathy 
Noland,  White  Mountain  Ranger  Sta- 
tion, USDA  Forest  Service,  is 
thanked  for  arranging  access  to 
study  sites.  Harold  Klieforth,  Atmos- 
pheric Sciences  Center,  Desert  Re- 


search Institute,  University  of  Ne- 
vada, Reno,  helped  establish  weather 
stations.  This  study  was  funded,  in 
part,  by  the  Committee  on  Research 
and  the  Department  of  Forestry  and 
Resource  Management,  University  of 
California,  Berkeley. 


LPTERATURE  CITED 

Austin,  Dennis  D.  1987.  Plant  com- 
munity changes  within  a  mature 
pinyon-juniper  woodland.  Great 
Basin  Naturalist  47:96-99. 

Barber,  Richard  T.,  and  Francisco  P. 
Chavez.  1983.  Biological  conse- 
quences of  El  Nino.  Science 
222:1203-1210. 


Bock,  Carl  E.,  and  James  F.  Lynch. 
1970.  Breeding  bird  populations  of 
burned  and  unburned  conifer  for- 
est in  the  Sierra  Nevada.  Condor 
72:182-189. 

Halvorson,  Curtis  H.  1984.  Long- 
term  modeling  of  small  verte- 
brates: a  review  with  suggestions, 
p.  11-25  In  Janet  L.  Johnson,  Jerry 
F.  Franklin,  and  Richard  G.  Krebill 
(coord.).  Research  natural  areas: 
baseline  monitoring  and  manage- 
ment. USDA  Forest  Service  Gen- 
eral Technical  Rep.  INT-173, 84  p. 
Intermountain  Forest  and  Range 
Experiment  Station,  Ogden,  Utah. 

Keith,  Lloyd  B.  1983.  Role  of  food  in 
hare  populations  cycles.  Oikos 
40:385-395. 


Table  1  .—Abundance  (no./lOO  trap-nights)  of  reptiles  and  small  mammals, 
Inyo-White  mountains,  California,  during  fall  (Sept.-Nov.)  1987,  winter  (Dec. 
Feb.)  1987-88,  and  spring  (Mar.-Moy)  1988. 


Fall= 


Winter'' 


Spring' 


Species 


Captures  Abund.    Captures  Abund.    C^tures  AlHjnd. 


Western  fence  lizard 

(Sceloporous  ocddentalis) 
Sagebrush  lizard  1^ 

(S.  gradosus) 
Gilbert  skink 

(Eumeces  gilberfi) 
Golden-mantled  ground  squirrel 

(Spermophifus  lateralis) 
Least  chipmunk  3 

(Eufamias  minimus) 
Panamint  chipmunk 

(E.  panaminfinus) 
Great  Basin  pocket  mouse 

(Ferognafhus  parvus) 
Deer  mouse  9 

(Peromyscus  maniculafus) 
Pinyon  mouse  56 

(P.  true!) 
Desert  woodrat  1 

(Neofoma  lepida) 

Total  69 


1.3 


0.4 


1.2 
7.7 
0.1 
9.5 


12 
12 

24 


1.7 
1.7 

3.4 


2^ 
3^ 
1* 

I 

39 
41 
29 
26 
7 


1.4 
2.0 
0.1 
0.1 
3.2 
3,4 
2.4 
2.2 
0.6 


143  11.8 


"Trap-nlghfs:  75  for  pitfalls,  728  for  small-mammal  traps. 
^Trap-nigt)ts:  0  for  pitfalls,  707  for  small-mammal  traps. 
'^Trap-nightts:  147  for  pitfalls.  1211  for  small-mammal  traps. 
'^Captured  in  pitfall. 
^Captured  in  small-mammal  trap. 


274 


Kish,  Leslie.  1987.  Statistical  design 
for  research.  Wiley  and  Sons,  N.Y. 

Likens,  Gene  E.  1983.  A  priority  for 
ecological  research.  Bulletin  Eco- 
logical Society  of  America  64:234- 
243. 

Meents,  Julie  K.,  Jake  Rice,  Bertrin  W. 
Anderson,  and  Robert  D.  Ohmart. 
1983.  Nonlinear  relationships  be- 
tween birds  and  vegetation.  Ecol- 
ogy 64:1022-1027. 

Ralph,  C.  John,  and  J.  Michael  Scott 
(editors).  1981.  Estimating  num- 
bers of  terrestrial  birds.  Studies  in 
Avian  Biol.  No.  6. 

Raphael,  Martin  G.,  Michael  L.  Mor- 
rison, and  Michael  P.  Yoder-Wil- 
liams.  1987.  Breeding  bird  popula- 
tions during  twenty-five  years  of 
postfire  succession  in  the  Sierra 
Nevada.  Condor  89:614-626. 

Schreiber,  Ralph  W.,  and  Elizabeth 
Anne  Schreiber.  1984.  Central  Pa- 
cific seabirds  and  the  El  Nino 
southern  oscillation:  1982-1983 
perspectives.  Science  225-713-716. 

Strayer,  David,  Jeff  S.  Glitzenstein, 
Clive  G.  Jones,  Jerzy  Kolasa,  Gene 
E.  Likens,  Mark  J.  McDonnell, 
Geoffrey  G.  Parker,  and  Steward 
T.  A.  Pickett.  1986.  Long-term  eco- 
logical studies:  an  illustrated  ac- 
count of  their  design,  operation, 
and  importance  to  ecology.  Insti- 
tute of  Ecosystem  Studies,  Occa- 
sional Paper  No.  2.  38  p. 

Wiens,  John  A.  1984.  The  place  of 
long-term  studies  in  ornithology. 
Auk  101:202-203. 


275 


An  Ecological  Problem- 
Solving  Process  for  Managing 
Special-Interest  Species^ 

Henry  L.  Short^  and  Samuel  C.  Williamson^ 


Abstract.— We  present  a  structured  problem- 
solving  process  that  con  help  resolve  v/ildlife 
management  issues.  Management  goals  for  wildlife 
species  are  expressed  in  terms  of  populations  to  be 
attained  and  maintained.  Habitat  quantity  and 
quality  necessary  to  achieve  those  population  goals 
can  then  be  determined.  Proposed  land-use 
changes  are  evaluated  in  terms  of  how  they  will 
contribute  toward  recovery  or  extinction  of  the 
species  of  interest. 


Land-use  problems  associated  with 
the  need  to  protect  wildlife  habitat 
and  the  desire  to  develop  resources 
can  sometimes  be  resolved  using  an 
ecological  problem-solving  process. 
The  process  requires  development  of 
a  management  goal  for  individual 
wildlife  species,  determination  of  the 
quantity  of  habitat  required  to 
achieve  that  management  goal,  and 
an  appraisal  of  how  development 
scenarios  will  affect  the  management 
goal. 

We  describe  how  the  process 
might  work  using  available  data 
about  the  endangered  Mount  Gra- 
ham red  squirrel  (Tamiasciurus 
hudsonicus  grahamensis).  The  exercise 
is  relevant  because  the  squirrel  exists 
entirely  as  a  disjunct  population  in 
the  high  elevation  coniferous  forest 
community  of  the  Pinaleno  Moun- 
tains of  southeastern  Arizona,  and  a 
new  astrophysics  observatory  has 
been  proposed  within  important 
squirrel  habitat.  Our  process  was  not 
applied  in  the  development  of  the 
Environmental  Impact  Statement 
(EIS)  prepared  for  the  red  squirrel 
and  its  habitat  nor  in  negotiations  for 
the  future  management  of  the  squir- 
rel. An  extensive  and  current  infor- 

' Paper  presented  of  symposium,  Mon- 
agement  of  Amphibians.  Reptiles,  and 
small  Mammals  in  Northi  America.  Ragstaff. 
Ariz.  July  19-21.  1988. 

'Ecologist.  U.S.  Fisti  and  Wildlife  Service. 
National  Ecology  Researct^  Center.  2627 
Redwing  Road.  Fort  Collins.  Colorado. 
60526. 


mation  base  (Spicer  et  al.  1985;  U.S. 
Forest  Service  1987, 1988)  recently 
has  been  developed  for  the  Mount 
Graham  red  squirrel  in  order  to  de- 
velop the  EIS  for  the  proposed  astro- 
physics observatory.  We  applied 
these  data  to  a  cumulative  impacts 
assessment  process  being  developed 
by  the  U.S.  Fish  and  Wildlife  Service. 
We  assume  that  species-habitat  man- 
agement goals  can  be  developed  and 
that  these  goal  statements  can  drive 
habitat  management  plans  and  ac- 
tivities. We  have  not  analyzed  the 
merits  of  any  development  scenarios 
proposed  for  the  astrophysics  obser- 
vatory. 

The  Pinaleno  Mountains  are  an 
isolated  range  that  supp)orts  one  of 
the  southernmost  spruce-fir  forests  in 
North  America  (Spicer  et  al.  1985). 
The  Mount  Graham  red  squirrel  is 
endemic  to  the  small  patches  of  co- 
niferous forests  that  occur  at  the 
highest  elevations  of  the  mountains. 
The  squirrel  has  been  affected  by  a 
variety  of  human  activities  and  natu- 
ral events  that  have  altered  its  habi- 
tat. Disturbances  included  comple- 
tion of  a  road  to  the  mountain  top  in 
1933,  introduction  of  the  tassel-eared 
squirrel  (Sciurus  aberti)  in  1941  to 
1943,  extensive  logging  activities  in 
subalpine  coniferous  forests  from 
1946  to  1973,  a  major  fire  in  1956,  and 
extensive  windthrows  in  the  1960's 
(Spicer  et  al.  1985).  The  squirrel  was 
first  collected  from  the  Pinaleno 
Mountains  in  1894  and  was  consid- 
ered "connmon"  in  the  spruce-fir 


zone  above  2,590  m  in  1914.  Since  the 
early  1950's  it  has  been  considered 
"uncommon"  throughout  the  conifer- 
ous tree  zone  of  this  mountain  range 
(Spicer  et  al.  1985). 

THE  PROCESS 

The  problem-solving  process  used  in 
our  analysis  contains  three  principal 
steps  (fig.  1).  Problem  description, 
the  first  step,  defines  the  ecological 
problem  and  identifies  the  spedes, 
study  area,  and  time  frame  of  con- 
cern. 

Problem  analysis,  the  second  step, 
develops  biological  information  nec- 
essary to  achieve  a  solution.  An  ini- 
tial effort  is  to  describe  a  manage- 
ment goal  for  the  species  of  concern 
in  terms  of  a  specific  population  level 
to  be  achieved  and  maintained.  This 
numerical  target  is  not  a  vague  state- 
ment to  "maintain"  or  "enhance"  be- 
cause such  terms  cannot  be  used  to 
measure  the  results  of  management 
actions.  The  management  goal 
should  be  collaboratively  develof)ed 
so  that  all  interested  parties  reach  a 
consensus  on  the  desirability  for  per- 
petuating the  spedes  and  on  a  popu- 
lation level  to  be  achieved  by  man- 
agement. It  is  understood  that  mutu- 
ally agreed  upon  goals  represent 
compromise  and  that  compromises 
are  rarely  satisfactory  to  all  con- 
cerned parties. 

It  is  then  necessary  to  determine 
the  quality  and  quantity  of  habitat 


276 


required  to  achieve  the  management 
goal.  This  requires  building  a  model 
describing  habitat  requirements  for 
the  species.  An  understanding  of 
how  human  activities  and  natural 
events  impact  habitat  quality  and 
quantity  is  also  desirable  because  the 
management  of  these  restricting  ac- 
tions may  help  achieve  the  manage- 
ment goal  for  the  species.  The  identi- 
fication of  causes  contributing  to 


habitat  deficiencies  can  be  made  by 
interviewing  persons  familiar  with 
the  sf>ecies  and  the  particular  habitat 
conditions  within  the  study  area. 

The  third  step  in  the  process,  solv- 
ing the  problem  (fig.  1),  is  accom- 
plished after:  (1)  the  amount  and 
quality  of  habitat  necessary  to  fulfill 
the  management  goal  has  been  deter- 
mined, (2)  the  quantity  of  suitable 
habitat  presently  available  has  been 


I.  Describe  the  Problon 

II.  Analyze  the  Problem 

1.  Determine  the  Management  Goal  for  the  Species. 

2.  Describe  Inportant  Habitat  Cticxiitions  for  the  Species. 

3.  Determine  how  ftmn  Activities  Affect  Habitat  Conditions 
Important  to  the  Species. 

III.  Solve  the  Problem 


1.  Determine  Acceptable  Strategies  for  Managing  Habitats  Required 
by  the  Species. 


Figure  1  .—Steps  of  ttie  problem-solving  process. 


Abounds  in 
mountains  of 
lh«  Graham 
range 


Common 

Infir-spfuce 

fcrests 


Not 

abundant 


Uncommon 
to  rare 


Possibly 
extirpated 


Begun 
recovery  but 
probably  has 
not  acWevod 
former 
levels 


1889 


1914 


1929 


1951-2  1963-7 


Present 


Figure  2.— Possible  trends  in  abundance  of  the  Mount  Graham  red  squirrel.  (Population  de- 
scriptions are  those  of  Spicer  et  al.  1 985.) 


documented,  and  (3)  the  quantity  of 
suitable  habitat  that  would  be  avail- 
able under  different  land-use  options 
has  been  projected. 

Describe  the  Problem 

The  Mount  Graham  red  squirrel  has 
probably  declined  during  this  cen- 
tury (fig.  2)  in  part  because  of  the 
piecemeal  degradation  of  isolated 
forest  habitat.  The  variety  of  human 
activities  and  natural  events  causing 
this  decline  might  soon  be  aug- 
mented by  the  development  of  the 
astrophysics  observatory  on  the  Pi- 
naleno  Mountains.  Can  this  and  re- 
lated developments  occur  in  a  man- 
ner that  does  not  further  jeopardize 
the  existence  of  the  endangered  red 
squirrel  during  the  foreseeable  fu- 
ture? 


Analyze  the  Problem 

Determine  the  Management  Goal 
for  the  Species 

The  management  goal  is  described  in 
terms  of  a  population  to  be  attained 
and  maintained.  Ideally,  population 
goals  should  be  based  on  quantita- 
tive historical  levels  of  abundance. 
Population  goals  are  more  difficult  to 
establish  if  historical  information 
about  population  levels  are  fragmen- 
tary and  descriptive,  as  for  the 
Mount  Graham  red  squirrel.  In  such 
cases,  criteria  for  establishing  desired 
population  levels  should  consider: 
(1)  estimates  of  present  populations 
and  trends,  (2)  threshold  values  nec- 
essary to  ensure  the  survival  of  the 
species,  and  (3)  estimates  of  the  po- 
tential p>opulation  level  that  could  be 
attained  if  management  of  an  area 
was  accomplished  solely  to  benefit 
the  species. 

Estimates  of  population  trends  for 
the  Mount  Graham  red  squirrel  are 
largely  qualitative  (fig.  2).  The  results 
of  field  work  suggest  that  the  au- 
tumn 1987  population  of  red  squir- 


277 


rels  on  the  Pinaleno  Mountains  might 
be  246  (206-286),  (U.S.  Forest  Service 
1988:37).  Computer  simulations  of 
population  dynamics  of  the  red 
squirrel  (U.S.  Forest  Service  1988:74) 
are  only  minimally  helpful  because 
data  such  as  natality  and  mortality 
for  the  Mount  Graham  red  squirrel 
are  unknown.  The  computer  simula- 
tions suggest  probability  levels  for 
extinction  under  different  combina- 
tions of  mortality  and  reproduction. 
The  predicted  carrying  capacity  for 
the  squirrel  under  current  habitat 
conditions  has  been  estimated  at  502 
squirrels.  The  potential  future  carry- 
ing capacity,  based  on  the  quantity 
and  present  age  structure  of  mixed 
conifer  and  spruce-fir  stands,  is  725 
squirrels  (U.S.  Forest  Service  1988:72- 
73).  Thus,  the  current  population  of 
red  squirrels  might  be  somewhat 
higher  than  that  in  the  early  1960's 
when  the  species  was  reported  as 
possibly  extirpated  (fig.  2),  but  less 
than  one-half  the  present  carrying 
capacity  for  the  species.  The  collabo- 
ratively developed  management  goal 
might  state,  for  example,  that  the 
management  goal  for  the  species  is  to 
develop  and  perpetuate  a  red  squir- 
rel population  equal  to  the  present 
carrying  capacity  of  the  habitat  for 
the  squirrel  which  is  estimated  at  502 
squirrels  (U.S.  Forest  Service 
1988:73). 


Describe  Important  Habitat 
Conditions  for  \he  Species 

A  species-habitat  model  for  the  red 
squirrel  can  be  based  on  the  squir- 
rel's dependency  on  seed  cones  and 
trees  that  produce  those  cones.  Coni- 
fer seeds  are  the  primary  food  of  the 
red  squirrel,  which  cuts  cones  in 
summer  and  caches  then  in  middens 
in  dense  needle  litter  at  stumps, 
downed  timber,  and  on  the  base  of 
snags  or  live  trees  in  forests  with 
dense  overstory  canopies  (Spicer  et 
al.  1985). 

We  constructed  a  species-habitat 
model  for  Mount  Graham  red  squir- 


rels using  data  given  in  the  U.S.  For- 
est Service  (1987)  report.  The  struc- 
tural stage,  tree  species,  and  canopy 
density  that  compose  red  squirrel 
habitats  are  classified  as  excellent, 
good,  fair,  poor,  very  poor,  and  no 
value  (fig.  3).  These  data  were  devel- 
oped by  U.S.  Department  of  Agricul-  - 
ture  Forest  Service  personnel  and 
others  familiar  with  the  habitat  re- 
quirements of  the  squirrel  and  were 
based  on  vegetation  type  and  struc- 
tural stage,  the  number  of  snags  and 
downed  logs  per  hectare,  aspect,  and 
slope  (U.S.  Forest  Service  1987:44). 
Midden  complexes  are  a  focal  point 
of  territories  and  the  number  of  ac- 
tive middens  is  supposedly  associ- 
ated with  the  number  of  red  squirrels 
in  a  stand  (Spicer  et  al.  1985).  The 
data  for  middens  per  hectare  have 
been  adjusted  so  that  a  score  of  1.0  is 
listed  for  excellent  habitats,  0.0  for  no 
value  habitats,  and  intermediate  val- 
ues are  listed  for  habitats  of  interme- 
diate quality  (fig.  3). 

The  species-habitat  model  de- 
scribes conditions  in  habitats  of  dif- 
ferent quality.  A  simple  word  model 
was  then  developed  to  describe  a 
unit  of  good  or  excellent  habitat  for  a 
red  squirrel  (fig.  4).  The  model  devel- 
oped from  information  in  figure  3 
and  U.S.  Department  of  Agriculture 
(1987:33-37)  defines  suitable  habitat 
for  a  red  squirrel  as  a  1-ha  forested 
block  that:  (1)  is  contiguous  to  other 
similar  forested  blocks,  (2)  provides  a 
dense  overstory  canopy  of  spruce-fir 
or  mixed  conifers,  and  (3)  contains 
about  15  "good"  seed-bearing  trees 
per  hectare. 

Such  species-habitat  models  are 
general  and  approximate.  Still,  they 
provide  an  estimate  of  what  com- 
prises a  unit  of  habitat  area  and  con- 
dition that  might  be  required  by  a 
squirrel.  If  a  management  goal  is  to 
provide  habitat  for  X  red  squirrels 
then  that  goal  can  possibly  be 
achieved  by  providing  X  units  of 
good  to  excellent  habitat  (fig.  4).  The 
need  to  provide  this  quantity  of  a 
specific  habitat  condition  should 
drive  management  plans  for  the  sub- 


Trco  spccloi, 
ftlnirlurnl 
fct>(|t.f  and 
canopy  dcosi  ty 


All  titlicr  vcgetatiun 
coMLiin  i  I.  i  es 


Mixed  conircr  and  tuicn, 
polot, 

canopy  covar 


Spriice-'ir,  poles, 
'lUl  cMnnpy  tovyr 

Mixiyt  conirer  and  atpen, 
poles, 

71X  canopy  ccvar 


Sfinico*^ir  ard  aspon, 
iMture, 
10S  canopy  cover 
Nixad  conlfar.  anturo, 

7<^  canufjy  cover 


Spn*C0*rir,  acturtt, 

41S  canopy  cover 
Spnii:K*rlr  uimJ  bIwU 
coiifcr,  old  grovUi, 

^OX  citnupy  cover 
NIxuJ  cuKlrer,  BHUiru, 

ViX  canopy  cover 


Spnx:e-rir  acd  ailxud 
conircr,  old  qrowih, 
b<it  canopy  cov«r 


TT  1  1  1 


Figure  3.— Habitat  quality  for  ttie  Mount 
Grahiam  red  squirrel  can  be  described  in 
terms  of  tree  species,  structured  stage,  and 
canopy  density  wittiin  forest  tKibitats.  Habi- 
tat quality  is  measured  in  ternrts  of  middens 
per  tiectare  for  different  habitats  scaled  on 
ttie  basis  of  0.0  for  no  value  hiabitats  and  1 .0 
for  excellent  tiabitats.  Data  from  U.S.  Forest 
Service  (1988:112). 

A  habitat  block  suitable  for  a  Mount  Grahan  red 
squirrel  Is  at  least  1  ha  In  area  aid  is 
contlgjous  to  other  blocks  of  suitable  habitat 
(fragTBitation  or  isolaftioa  totally  raroves  a 
habitat  block  fran  suitability  oonsideraticn) . 


Yes 
i 

2.    The  tree  cxnposltitn  of  the  habitat  block  Is  (fran 
Figure  3): 

a.  spcuce-fir,  neture,  ^40%  canopy  coror 

b.  spcuce-fir,  old  growth 

c.  mixed  ocnlfer,  netuie,  ^70%  caociy  cover 

d.  mixed  cxnlfer,  old  growth 


yes 

3.    The  habitat  block  contains  at  least  15  "good- 
bearing  spnioe-fir  or  mixed  conifer  seed  trees  (a 
saTB*at  arbitrary  niiier). 


Yes 
i 

Ihe  habitat  block  provides  good  to  exoeUeot 
habitat  for  one  Mount  Graham  red  squirrel. 


Figure  4.— A  word  rTX>del  describing  good 
to  excellent  tiabitat  condition  for  a  red 
squirrel. 


278 


alpine  coniferous  forests  of  the  Pi- 
naleno  Mountains. 

Determine  How  HunfKin  Activities 
Affect  Habitat  Conditions 
Important  to  ttie  Species 

Several  human  activities  and  natural 
events  may  adversely  affect  habitats 
of  the  Mount  Graham  red  squirrel 
and  reduce  the  opportunity  to 
achieve  the  management  goal  for  the 
species.  A  listing  of  possible  impacts 
on  the  Mount  Graham  red  squirrel 
and  the  probable  resulting  habitat 
changes  is  in  figure  5. 

The  cells  in  a  cause-effect  matrix 
(table  1)  list  estimates  of  the  direction 
and  relative  importance  of  each  fac- 
tor affecting  a  habitat  criterion.  The 
cells  within  the  cause-effect  matrix 
can  be  completed  after  synthesizing 
information  from  the  literature,  from 


best  professional  judgments  elicited 
from  selected  personnel  or  preferably 
from  analyzing  results  of  appropriate 
research.  Information  within  the 
cause-effect  matrix  can  indicate  the 
relative  importance  of  different  hu- 
man activities  on  squirrel  habitats 
and  identify  actions  to  be  favored  or 
avoided  to  help  achieve  the  manage- 
ment goal.  For  example,  habitat  frag- 
mentation, clearcutting,  selective  har- 
vest, and  forest  management  favor- 
ing early  vegetation  successional 
stages  are  important  negative  factors 
to  red  squirrel  habitats  whereas  man- 
agement favoring  dense,  mature  or 
old-growth  stands  of  mixed  conifer 
and  spruce-fir  forests  are  important 
positive  actions,  favorable  to  red 
squirrels.  Causes  of  negative  and 
positive  impacts  to  species  or  habi- 
tats of  concern  are  factors  that 
should  be  considered  when  formulat- 
ing and  evaluating  plans  for  modify- 


cwsts 

Habitat  fra^MntAtlm  (stand  size  — 
reduction,  road  ccnstiuctlm,  otc.) 


Sal«ctlv«  harvwt  (t.MnrUng)  of  trm  (innj  

thtOH  1>  analogous  to  s*Lactlv«  harvest) 

CXasrcuttUig  6oc  dw^elopisnt  (foc««t  firee   

are  aralcgcxas  to  clearcutting) 


Mnagannt  favAsring  early  SLXxessianal 
wtegta 


HBDB^erant  favociog  dense,  nature,  or  - 
old  grcHtti  spcuoe-flr 

Moa^anwit  favoring  dacse,  mature,  or  - 
old  giouth  of  mlxsd  cxxiifeccus  gpeclw 

RBwal  of  anaga  and  Moody  deiarLi   


ConpetiUm  £ron  IntitxlDed  tassel-eeivd 


VXal  ctaogta  In  tim 
gjantity  and  quality  oC 
suitable  habitat  la  the 
Plnalwp  Mountains  tvm 
adversely  affected  the  Mount 
Gattm  red  agjirr*!  populaticn. 


->  - 


•  aiitablllty  ce  habitat  toe  squirrel 
territociee  (large  areas  with  aense 
tree  canqpy  ccwer)  is  nxliflad 

-  rjcgi  (a  food  source}  pzodLjcticn 
in  the  naedle  litt«-  is  redifled 

'  Storage  effectlvaoBss  of  middens 
is  changed 

'  Maindanca  of  preferred  cones  Is 
ncdified 

Abundance  of  tot^l  cans  is 

■  QiiBivtity  of  cavltiee  available 
as  nest  sltas  Is  ciianged 

-  Quiatity  of  atrjctural  niatArlAls 
available  for  rurways  and  niddea 
sites  is  changed 


Figure  5.— A  cause/effect  model  identifying  causes  thaf  affect  the  quantity  and  quaiity  of 
habitat  suitable  for  the  Mount  Grahann  red  squirrel. 


Quantity  of 

suitablfl 

habH^ 


Currtnl  habKat  corxJ'rtlon 


(1a)  Habitat  condition  exceods  that 
n«cessary  for  management  goal 

(2)  Required  habitat  condition  for 
management  goal 


(1b3)  Future  positive  impacts 
(1b2)  No  additional  future  Impacts 
(1b1)  Future  r^egative  impacts 


Time 

Figure  6.— Habitat  conditions  under  a  variety  of  nnanagennent  strategies. 


279 


ing  habitats  important  to  selected 
wildlife  species. 

Solve  the  Problem 

Determine  Acceptable  Strategies 
for  Managing  Habitats  Required 
by  ttie  Species 

A  way  to  evaluate  the  diversity  of 
different  land-use  scenarios  is  listed 
in  figure  6.  Threshold  values  describ- 
ing the  quantity  of  suitable  habitat 
necessary  for  achieving  the  manage- 
ment goal  for  the  red  squirrel  can  be 
represented  as  habitat  condition  2  in 
figure  6.  If  the  quantity  of  suitable 
habitat  presently  available  had  ex- 
ceeded this  threshold  value  (condi- 
tion la)  then  changes  to  the  quantity 
of  available  habitat  could  be  toler- 
ated and  that  fact  could  be  consid- 
ered in  making  a  decision  about  a 
potential  land  use. 

The  present  quantity  of  good  to 
excellent  habitat  for  the  red  squirrel 
in  the  Pinaleno  Mountains,  however, 
is  probably  more  closely  approxi- 
mated by  condition  lb  in  figure  6.  A 
variety  of  conditions  like  those  item- 
ized in  table  1  have  reduced  habitat 
quality  and  quantity  resulting  in  a 
diminished  squirrel  population  with 
an  endangered  species  listing.  A 
land-use  plan  that  continued  impacts 
(like  those  listed  in  table  1)  would 
further  reduce  the  area  and  quality  of 
contiguous  blocks  of  forest  habitat 
important  to  the  squirrel.  Any  fur- 
ther fragmentation  or  degradation  of 
habitat  would  be  exjjected  to  further 
diminish  the  population  (Ibl  in  fig, 
6)  and  perhaps  threaten  extinction  of 
the  subspecies.  A  land-use  plan  that 
neither  allowed  further  degradation 
of  habitat  nor  actively  improved 
habitat  conditions  for  the  squirrel 
might  result  in  maintaining  present 
population  levels  (lb2  in  fig.  6).  The 
most  desirable  land-use  scenarios  are 
those  likely  to  produce  trend  lines 
such  as  lb3  (fig.  6).  These  land-use 
plans  would  minimize  fragmentation 
of  habitats  and  would  actively  man- 


age  habitats  to  develop  large  contigu- 
ous blocks  of  old-growth  mixed  coni- 
fers and  spruce-fir  on  the  Pinaleno 
Mountains  to  help  attain  the  desired 
population  level  of  red  squirrels. 


CONCLUSIONS 

We  emphasize  that  potential  land- 
use  change  can  be  evaluated  in  a  ra- 


tional manner  if  management  goals 
for  wildlife  resources  have  been  pre- 
viously established  and  agreed  upon. 
The  merit  of  this  approach  is  that 
planning  becomes  an  active  rather 
than  a  reactive  exercise.  Too  often  we 
evaluate  proposed  land-use  changes 
in  terms  of  how  they  might  affect 
present  habitats  and  present  popula- 
tions without  considering  how  pres- 
ent conditions  compare  to  desired 


populations  and  necessary  habitats. 
Without  establishing  a  management 
goal  and  determining  the  habitat 
conditions  necessary  to  achieve  that 
goal,  we  could  accept  the  wrong 
baseline  for  developing  our  manage- 
ment strategy  (perhaps  something 
analogous  to  line  lb2  in  fig.  6).  If  this 
occurs,  we  might  have  little  success 
in  maintaining  viable  f)opulations 
because  we  frequently  strive  only  to 


Table  1  .—A  cause-effect  matrix  that  lists  the  relative  importance  of  causal  agents  (causes  listed  in  fig.  4)  that  change 
the  quantity  arKi  quality  of  habitat  features  (effects  listed  in  fig.  4)  for  the  Mount  Graham  red  squirrel.  A  {+)  value  indi- 
cates a  positive  impact  and  a  (-)  value  indicates  a  negative  impact.  Numerical  values  indicate  the  magnitude  of  an 
impact:  (0)  =■  negligible;  (1)  =  minor;  (2)  =  important;  and  (3)  =  very  important. 


-  0 

« «) 

k 

M 

I-  c 

0 

c 

c 

0 

c 

u 

-  (. 

0 

*t 

u 

L 

err 

TJ 

« *i 

u 

V 

(A 

i) 

3 

3 

0) 

a 

«-) 

E 

c 

A) 

0 

0 

0 

M 

w  n 

L 

0 

0 

«  w 

-0 

a. 

0 

0 

c 

>  01 

0 

*»  0 

V 

c 

m  ** 

L 

C  i- 

«^ 

19 

« 

0 

0 

ti  ii 

0  L 

ca 

1. 

0 

« 

M  l/t 

0 

C  0 

1. 

*i 

U  it 

c 

« 

«  C 

«  9(- 

Z  *i 
0- 

0 

a 

-  0 

A> 

£  L  0 

> 

tt 

*» 

u 

«  > 

V)  - 

b 

0 

« 

(9 

ii 

a 

Ai  — 

a 

OwO 

V  0 

0 

n  e 

E 

0- 

9 

V. 

e 

E 

>»«  N 

0X3 

c 

0 

0 

> 

*«» 

<«-  0 

ft 

« 

a 

-  c 

4)  >> 

0 

0 

0 

>  «. 

-  u  c 

(B  C 

u 

u 

«) 

g  (B 

—  CB 

—  0  (8 

0 

c 

c 

0 

3  M 

n*i  0 

0) 

K 

a 

«^>> 

n  E 

«e  — 

-r 

« 

T3 

0  « 

o>*» 

L 

c 

c 

3  > 

3  M 

c 

z  c 

0 

3 

> 

«-  C 

!. 

SOL 

(/)  4J  «J 

3 

■A.  — 

W 

< 

< 

(A  (. 

Habitat  fragmentation  (stand  size  -3 
reduction,  road  construction,  etc.) 

Selective  harvest  (thinning)  of  trees  -2 
(windthrow  is  analogous  to  selective 
harvest) 

Clearcutting  for  development  (forest  -3 
fires  are  analogous  to  clearcutting) 

Management  favoring  early  -2 
succession  stages 

Management  favoring  dense,  mature,  +3 
or  old  growth  spruce-fir 

Management  favoring  dense,  mature,  +3 
or  old  growth  of  mixed  coniferous  species 

Removal  of  snags  and  woody  debris  0 

Presence  of  tassel-eared  squirrels  0 


-1 


0 


-3 

0 

+3 

+3 

0 
-1 


-3 

0 

+3 

+3 

-2 
0 


-1 
-2 

-3 

-2 

+3 

+1 

0 
-1 


-1 


-3 

-2- 

+3 

+3 

0 
-1 


-2 

-3 

3 

+3 

+3 

-2 
0 


0 
+2 

-3 

-1 

+3 

+3 

-3 
0 


-3 
-2 

-3 
-2 

-1-3 

+2 

-1 
-1 


280 


maintain  marginal  populations  in 
marginal  habitats.  A  rule  for  judging 
the  suitability  of  a  proposed  land-use 
change  might  be  that  land-use 
change  that  can  be  accomplished 
while  promoting  trend  lines  like  lb3 
(with  strong  positive  slopes)  or 
i     which  produce  conditions  like  line  2 
in  figure  6  are  environmentally  ac- 
ceptable and  can  be  accomplished  if 
they  are  socially  and  economically 
I  desirable. 


LITERATURE  CITED 

Spicer,  R.  B.,  J.  C.  deVos,  Jr.,  and  R. 
L.  Glinski.  1985.  Status  of  the 
Mount  Graham  red  squirrel,  Tami- 
asciurus  hudsonicus  grahamensis 
(Allen),  of  southeastern  Arizona. 
Unpublished  report  by  Arizona 
Game  and  Fish  Dep.  for  U.S.  Fish 
Wildl.  Serv.,  Office  of  Endangered 
Species.  48  p. 

U.S.  Forest  Service.  1987.  Mount  Gra- 
ham red  squirrel.  A  biological  as- 
sessment of  impacts  proposed  Mt. 
Graham  Astrophysical  Project. 
Coronado  National  Forest, 
Tucson,  Ariz.  92  p. 

U.S.  Forest  Service.  1988.  Mount  Gra- 
ham red  squirrel.  An  expanded 
biological  assessment.  Coronado 
National  Forest,  Tucson,  Ariz. 
130  p. 


Comparative  Effectiveness  of 
Pitfalls  and  Live-Traps  in 
IVIeasuring  Small  Mammal 
Community  Structure^ 

Robert  C.  Szaro,^  Lee  H.  Simons,^  and  Scott 
C.  Beifit^ 


Abstract.— The  effectiveness  of  pitfalls  and  live- 
traps  for  assessing  small  mammal  community 
structure  was  compared  in  burned  and  unburned 
upland  Sonoran  Desert  and  in  an  elevational  series 
of  Sycamore  riparian  and  adjacent  habitats  in 
Arizona.  Although,  live-traps  v/ere  more  effective  in 
recapturing  previously  captured  small  mammals 
and  usually  resulted  in  more  total  captures  of  new 
individuals,  neither  method  gave  a  complete 
assessment  of  small  mammal  community  structure. 


Several  studies  that  compared  vari- 
ous types  of  pitfalls  and  live-traps 
(also  called  box-  or  cage-traps)  in  the 
field  (Chelkowska  1967,  Boonstra 
and  Krebs  1978,  Peterson  1980,  Boon- 
stra and  Rodd  1984,  Mengak  and 
Guynn  1987)  found  the  sampling  effi- 
ciency of  the  two  methods  varied 
considerably  (Andrzejewski  and 
Rajska  1972,  Briese  and  Smith  1974, 
Cockbum  et  al.  1979,  Williams  and 
Braun  1983).  Pitfall  cone  traps  were 
more  effective  than  live-traps  in  sam- 
pling small  mammals,  particularly 
shrews  in  southern  Finland  (Pankak- 
oski  1979).  In  contrast,  pitfalls  were 
less  effective  than  live-traps  in  cap- 
turing small-bodied  mice  in 
Durango,  Mexico,  although  more 
shrews  (Notiosorex  crawfordi)  were 
taken  in  pitfalls  (Peterson  1976).  Pit- 
falls of  various  materials,  shapes,  and 
sizes,  with  and  without  drift  fences, 

'Paper  presented  of  symposium,  Mon- 
ogement  of  Amphibions,  Repfiles.  ortd 
Small  Mammals  in  Norfhi  America.  (Flag- 
sf off,  AZ.  July  19-21  1988.) 

'Roberf  C.  Szaro  is  Research)  Wildlife  Bi- 
ologist, USDA  Forest  Service.  Roclcy  Moun- 
tain Forest  and  Range  Experiment  Station, 
Arizona  State  University  Campus.  Tempo,  AZ 
85287-1304. 

^Lee  hi.  Simorts.  formerly  a  graduate  stu- 
dent. Arizona  State  University.  Department 
of  Zoology,  Tempo,  Arizona  is  currently  a 
graduate  student.  Graduate  Group  in  Ecol- 
ogy, University  of  California,  Davis,  CA 
95616. 

"Scott  C.  Bel  fit  is  Wildlife  Biologist.  De- 
partment of  the  Army,  Wildlife  Manage- 
ment Section.  Fort  Huachuca.  AZ  856 13- 
6000.  Belfit's  current  address  is  P.O.  Box  336. 
Fort  Belvoir,  VA  22060-0336. 


have  been  used  for  capturing  small 
mammals  (Howard  and  Brock  1961, 
Andrzejewski  and  Wroclawek  1963, 
Pucek  1969,  Boonstra  and  Krebs 
1978,  Pankakoski  1979).  This  lack  of 
standardization  makes  it  difficult  to 
assess  the  relative  effectiveness  of 
pitfalls  versus  live-traps  in  sampling 
small  mammals  by  comparing  data 
between  studies.  Conflicting  results 
from  these  studies  argue  for  more 
comparisons  using  controls  for  as 
many  extraneous  factors  as  possible. 

Small  mammals  respond  dramati- 
cally to  many  environmental  factors, 
thus  confounding  attempts  to  assess 
species  or  community  relationships. 
Sampling  biases  caused  by  climate 
and  differences  in  activity  and  loco- 
motor adaptations  of  various  species 
further  compound  this  problem.  Still, 
trapping  remains  the  most  practical 
method  for  assessing  small  mammal 
populations  (Williams  and  Braun 
1983).  Because  responses  to  trapping 
methods  may  differ,  even  within  the 
same  species  (Andrzejewski  and 
Rajska  1972),  diverse  sampling 
schemes  might  reveal  population 
dynannics  and  community  structure 
more  completely  than  any  single 
method  (Weiner  and  Smith  1972, 
Boonstra  and  Krebs  1978). 

We  compared  the  effectiveness  of 
live-traps  versus  pitfalls  in  riparian 
and  desert  habitats  in  Arizona  to  an- 
swer the  following  questions:  (1) 
Does  sampling  method  influence  es- 
timates of  sf)ecies  com{X)sition  and 
abundance?  (2)  Are  various  species 


captured  or  recaptured  differen- 
tially? (3)  Are  individuals  within  a 
species  captured  differentially?  (4) 
Does  habitat  structure  influence  the 
effectiveness  of  these  methods? 


Study  Areas  and  Methods 

Riparian  and  Adjacent 
Communities 

The  riparian  and  adjacent  communi- 
ties (referred  to  in  general  as  the  ri- 
parian area)  were  located  at  Garden 
Canyon,  Fort  Huachuca  Military  Res- 
ervation, Arizona;  elevations  ranged 
from  1500  to  1630  m.  Riparian  com- 
munities sampled,  from  lowest  to 
highest  elevation,  were  sycamore 
(Platanus  mrightii),  sycamore/juniper 
(Juniperus  monosperma),  and  syca- 
more/juniper (/.  deppeam)/odk 
(Quercus  arizonica,  Q.  emoryi,  and  Q. 
hypoluecoides)  (Szaro  1988).  Plant 
communities  sampled  adjacent  to  the 
riparian  corridor,  from  lowest  to 
highest  elevation,  were  composite 
{Heterotheca  spp.) /grassland  {Poa 
spp.),  junifjer  (/.  monosperma)  wood- 
land, and  oak  (Quercus  emoryi)  wood- 
land. 

Six  trap  stations  were  set  in  each 
of  six  habitats:  composite /grassland, 
sycamore  riparian,  juniper  wood- 
land, sycamore/ juniper  riparian,  oak 
woodland,  and  sycamore/juniper/ 
oak  riparian  forest  (figs.  1-6)  (36  sta- 
tions in  all).  Trap  stations  consisted 
of  two  unbaited  pitfalls  (18.9  L  or  5 


282 


■      ...>   -^^      '  ^ 

.V               .■•               v'.'-''-" ■-■■'  *•'■■■■"  ■ 

■.''"..■.■■.■>.■..■ 
•    ■  ■                             .  .<  ^            <  "" 

"      /    "           '           -      4  •■      "  - 

Figure  1  .—Arizona  sycarDore  (Piafanus  wrightil)  study  site,  Garden  Figure  4.— Composite  (Heterotheca  spp.)/grassiand  (Poa  spp.) 

Canyon,  Fort  Huactiuca  Military  Reservation,  Arizona;  eievation  ca.  study  site,  Garden  Canyon,  Fort  Huactiuca  Military  Reservation, 
1500  m.  Arizona;  elevations  ca.  1510  m. 


Figure  2.— Arizona  sycarnore  (Platanus  wrightii)/one-seed  juniper  Figure  5.— One-seed  juniper  (J.  monosperma)  woodland  study  site, 

(Juniperus  monosperma)  study  site.  Garden  Canyon,  Fort  Garden  Canyon,  Fort  Huactiuca  Military  Reservation,  Arizona;  ele- 

Huact)uca  Military  Reservation,  Arizona;  elevation  ca.  1565  m.  vations  ranged  from  1570  m. 


Figure  3.— Arizona  sycamore  (Platanus  wrighfii)/a\\\ga\of  juniper  (J. 
deppeana)/m\xBd  oak  (Quercus  arizonica,  Q.  emoryl,  and  Q.  hy- 
poluecoldes)  study  site,  Garden  Canyon,  Fort  Huactiuca  Military 
Reservation,  Arizona;  elevation  ca.  1610  m. 


Figure  6.— ErTK>ry  oalc  (Quercus  emoryi)  woodland  study  site,  Gar- 
den Canyon,  Fort  Huactiuca  Military  Reservation,  Arizona;  eleva- 
tions ca.  1590  m. 


283 


gal.;  29  cm  in  diameter  by  36  cm 
deep)  with  a  7.6-m-long  by  20-cm- 
high  drift  fence  between  buckets. 
Covers  were  propped  2.5-5  cm  above 
openings  mouths.  Pitfalls  were  open 
from  16  April  through  28  May  and 
from  20  July  through  5  September 
1986  (6408  trap-nights)  and  were 
checked  three  times  each  week.  Sher- 
man live-traps  (8  by  9  by  23  cm) 
baited  with  rolled  oats  were  set 
around  each  pitfall  station  in  an  8- 
trap  pattern  with  at  least  5  m  be- 
tween traps  and  pitfalls.  Live-traps 
were  set  from  12  to  16  May  and  from 
17  to  21  August  1986  (2304  trap- 
nights)  and  were  checked  each  morn- 
ing. Most  live-trap  captures  were  re- 
leased after  being  ear-tagged.  Except 
for  some  Notiosorex,  all  pitfall  cap- 
tures were  collected.  Identification  of 
all  mammals  follows  Hoffmeister 
(1986).  Thomomys  species  include 
pure  and  hybrid  T.  utnbrinus  and  T. 
bottae. 


Desert  Community 

The  desert  study  area  was  in  the 
Tonto  National  Forest,  Maricopa 
County,  30  km  east  of  Phoenix,  Ari- 
zona. The  site  was  rocky  desert  dis- 
sected by  sandy  washes;  elevations 
ranged  from  450  to  550  m.  Vegetation 
was  typical  of  the  Arizona  upland 
subdivision  of  the  Sonoran  Desert 
biome  (Brown  1982),  with  mesquite 
(Prosopsis  juliflora)  along  wash  banks 
and  palo  verde  (Cercidium  micro- 
phyllum),  bursage  (Ambrosia  deltoides), 
and  cholla  (Opuntia  acanthocarpa)  on 
slopes. 

Two  grids  were  established  90  m 
apart,  each  with  100  sampling  sta- 
tions placed  in  a  10  by  10  pattern 
with  10-m  intervals  between  stations. 
Grid  1  was  in  mature  desert  and  grid 
2  had  50%  of  vegetative  cover 
burned  on  7  June  1985,  immediately 
before  the  start  of  trapping  (figs.  7-8). 
Interspaced  between  live-traps  (10  by 
10  by  25  cm)  on  each  grid,  but  no 
closer  than  10-m  intervals,  were  20 
single  pitfalls  (37.9  L  or  10  gal.,  34  cm 


diameter  by  40  cm  deep)  buried  to  set  and  baited  with  rolled  oats  for 

the  rim  with  a  cover  propped  5-10  two  consecutive  nights  on  19  occa- 

cm  over  the  opening.  Live- traps  were      sions  between  10  June  1985  and  3 


Figure  7.— Unburned  desert  study  area,  Tonto  National  Forest.  Maricopa  County,  30  km  east 
of  Phoenix,  Arizona;  elevation  ranged  from  450  to  550  m. 


Figure  8.— Burned  desert  study  area,  Tonto  National  Forest,  Maricopa  County,  30  Icm  east  of 
Ptioenix,  Arizona;  elevation  ranged  from  450  to  550  m. 


284 


August  1986 — weekly  in  spring  and 
early  summer,  biweekly  from  middle 
to  late  summer,  and  monthly  in  fall 
and  winter  (Simons  1986).  Unbaited 
pitfalls  were  always  open  during 
live-trapping  and  often  in  between 
when  live-trapping  occurred  weekly 
or  biweekly  (March-September).  All 
captures  except  for  casualties  were 
marked  and  released.  Each  method 
was  matched  with  an  approximately 
equal  sampling  effort  (about  3800 
trap-nights  per  grid). 

Results  and  Discussion 

Species  Composition  and 
Abundance 

Live-traps  and  pitfalls  provided  dif- 
ferent estimates  of  species  composi- 
tion and  relative  abundance  at  both 
study  areas.  In  the  riparian  area  we 
observed  no  consistent  pattern  be- 
tween trapping  method  and  number 
of  species  captured  (table  1).  Live- 
traps  caught  more  species  in  two 


habitats,  pitfalls,  in  three  habitats, 
and  in  the  sycamore/ juniper/ oak 
both  methods  captured  two  species. 
Neither  method  captured  all  species 
in  a  given  habitat  except  in  oak 
woodland  where  only  two  species 
were  encountered  and  pitfalls  cap- 
tured both.  However,  live-trapping 
was  significantly  more  successful 
than  pitfalls  in  number  of  new  cap- 
tures per  trap-night  (chi-square,  P  < 
0.05)  in  all  habitats  except  juniper 
woodland,  where  both  methods 
yielded  equal  numbers. 

In  the  desert,  live-traps  caught 
more  species  than  pitfalls  (table  2). 
Moreover,  significantly  more  new 
captures  and  total  captures  (chi- 
square,  P  <  0.05)  occurred  in  live- 
traps  than  in  bucket-traps  in  both 
burned  and  unburned  plots  (table  2). 
These  results  differ  from  those  of 
Williams  and  Braun  (1983)  who  re- 
ported that  number  of  species  and 
total  number  of  captures  were 
greater  in  pitfalls  than  in  the  com- 
bined catch  of  snap-  and  live-traps. 
They  recorded  six  species  in  pitfalls 


and  four  in  snap-  and  live-traps. 
Their  success  with  pitfalls  was  no 
doubt  increased  because  each  trap 
was  one-third  filled  with  water, 
drowning  all  captures.  Trapping  suc- 
cess for  voles  (Clethrionomys  glareo- 
lus)  was  also  reported  to  be  higher  in 
pitfalls  versus  live-traps  but  may 
vary  with  social  level,  age,  and  re- 
productive period  (Andrzejewski 
and  Rajska  1972,  Andrzejewski  and 
Wroclawek  1963,  Chelkowska  1967). 

New  individuals  represented  only 
31.5%  and  26.2  %  of  total  captures  in 
live-traps  on  the  burned  and  on  the 
unburned  plots,  respectively.  In  con- 
trast, 95.8%  and  92.7%  of  all  captures 
in  pitfalls  on  the  burned  and  un- 
burned areas,  respectively,  represent 
new  individuals.  The  lack  of  recap- 
tures in  pitfalls  is  not  explained  by 
differential  mortality  between  meth- 
ods because  sampling  with  both 
methods  occurred  simultaneously, 
and  most  animals  were  marked  and 
released.  These  differences  maybe  at 
least  partially  due  to  increased  at- 
tractiveness of  live-traps  with  bait 


Table  1.— Total  number  of  new  Individuals  captured  in  riparian  and  associated  habitats  using  live-traps  (384  trap- 
nights/habftat)  and  pitfalls  (1068  trap-nlghts/habitat)  during  spring  and  late  summer  1986. 


Composite/  Sycarrtore  Juniper        Sycarnore/  Oak  Sycamore/  Total 

grass  woodland         juniper  woodland  Juniper/oal< 

Species                         Live-     Pit-  Live-     Pit-  Live-     Pit-  Live-     Pit-  Live-     Pit-  Live-     Pit-  Live-  Pit- 
trap     fall  trap     fall  trap     fall  trap     fall  trap     fall  trap     fall  trap  fall 


Neotoma  albigula  ] 

Notiosorex  crawfordi  2  31  1        17  22  2 

Onychomys  forridus  8         2  ]         4  4 

Perognafhus  fkivus  1  1 

Perognathus  hispidcs  2  1 

Perognafhus  pencillatus  ^ 

Peromyscus  boylei  12  .1  13  13         3  12 

Peromyscus  leucopus  3  3 

Peromyscus  maniculatus         5  2 

Reithrodontomys  fulvescens    4         2         5         2  1  1 

Reithrodontomys  megahfis  1 
Sigmodon  ochrognofhus  1 
Sorex  arizonae 

Jhomomysspp.  2  13 

Total  captures  21  8 

Species  richiness  6  4 

Overall  species  richness  8 
New  coptures/trap-night     6.07  0.75 
X  100 

All  captures/trcp-night  7.26 
X  100 


1 

76 

12 

-7 
1 

1 

1 

3  . 

1 

51 

3 

6 

5 

2 

9 

6 

1 
1 

2 

6 

19 

37 

8 

23 

16 

26 

13 

5 

15 

4 

92 

103 

4 

5 

5 

4 

2 

3 

1 

2 

2 

2 

12 

8 

8 

7 

5 

2 

4 

14 

4,95 

3.46 

2.08 

2.15 

4.17 

2,43 

3.38 

0.47 

3.90 

0.37 

3.99 

1.61 

10.38 

2.10 

9.84 

4.92 

6.00 

6.76 

285 


and  with  concentrated  odors  from 
previous  captures  (Boonstra  and 
Krebs  1978,  Daly  and  Behrends 
1984).  Our  results  show  that  pitfalls 
provide  very  different  estimates  of 
species  composition  and  abundance 
than  live-traps.  We  therefore  ques- 
tion basic  assumptions  of  the  popular 
methods  of  population  estimation 
that  assume  either  equal  catchability 
of  all  members  in  the  population 
(Jolly  1965)  or  nearly  complete  caf>- 
ture  and  enumeration  of  a  popula- 
tion (Krebs  1966,  Hilborn  et  al.  1976). 


Differential  Trapping  Effectiveness 
Between  Species 

In  the  riparian  area,  80  of  81  shrews 
(Notiosorex  crawfordi  and  Sorex  arizo- 
me)  and  all  gophers  {Thomomys  spp.) 
were  captured  in  pitfalls.  In  contrast, 
only  5  of  67  captures  of  Peromyscus  (3 
species)  were  in  pitfalls  (table  1). 
Peromyscus  spp.  were  also  recaptured 
most  frequently  (57  of  64  recaptures). 
Similar  results  were  found  in  the  Si- 
erra Nevada  where  species  such  as 
shrews  (Sorex  trowbridgii  and  S.  mon- 
Hcolus)  and  gophers  (Thomomys  bot- 
tae),  which  tend  to  travel  in  burrows 
or  runways  or  along  obstacles,  were 
usually  captured  in  pitfalls  (Williams 
and  Braun  1983).  Williams  and  Braun 
(1983)  reported  in  their  first  test  that 
pitfalls  were  particularly  poor  for 
capturing  white-footed  mice  (Pero- 
myscus). In  a  subsequent  test  they 
implied  these  mice  might  be  taken  in 
pitfalls  after  losing  their  caution  for 
strange  objects.  This  did  not  happen 
in  our  study  because  very  few  Pero- 
myscus were  captured  in  pitfalls  over 
an  extended  period  even  though  live- 
trapping  showed  them  to  be  com- 
mon. More  likely  Peromyscus  may 
easily  escape  pitfalls  by  jumping  out, 
but  more  are  recorded  after  drown- 
ing in  water-filled  pitfalls  (Williams 
and  Braun  1983),  especially  when 
other  traps,  such  as  snap-  or  live- 
traps,  are  missing. 

In  the  desert  habitat,  a  single 
shrew  (Notiosorex  crawfordi)  was 


caught  in  a  pitfall  whereas  two 
species  (Dipodomys  merriami  and 
Peromyscus  eremicus)  were  caught 
only  in  live-traps.  Only  1  of  181  cap- 
tures (50  different  individuals)  of 
Neotoma  albigula  was  in  a  pitfall 
whereas  only  1  of  9  Onychomys  tor- 
ridus  was  not  captured  in  a  pitfall. 
Onychomys  was  probably  unable  to 
jump  out  of  the  buckets  used  in  this 
habitat.  Noted  accumulations  of 
Neotoma  feces  overnight  in  many  pit- 
falls indicated  these  rodents  had 
been  present  but  left.  Apparently 
larger  species  either  avoid  pitfalls  or 
simply  jump  out  of  them  (Clockburn 
et  al.  1979,  Williams  and  Braun  1983). 


Differential  Trapping  Effectiveness 
Wittiin  Species 

Few  significant  differences  in 
weights  of  small  mammals  caught 
with  the  two  methods  were  ob- 
served, but  weights  tended  to  be 
lower  in  pitfalls.  In  the  riparian  area, 
mean  weights  of  Reithrodontomys  ful- 
vescens  were  significantly  higher  in 
live-traps  (14.3  +  0.65  (S.E.)  g  versus 
5.1  +  0.56,  t-test,  P  <  0.001,  N  =  12).  In 


the  desert,  weight  differences  be- 
tween trap  methods  were  not  signifi- 
cant for  animals  less  than  about  20  g. 
However,  a  significant  difference  oc- 
curred in  the  mean  weight  of  Per- 
ognathus  baileyi  in  live-traps  (25.7  + 
0.97  g)  versus  pitfalls  (20.8  ±  1.61  g;  t- 
test,  P  =  0.014).  Similarly,  the  mean 
weight  of  Neotoma  albigula  caught  in 
live-traps  was  109.0  +  8.93  g,  whereas 
the  single  capture  in  a  pitfall 
weighed  31.0  g. 

Likewise  in  Canada  and  Poland, 
voles  (Microtus  townsendii  and  Cle- 
thrionomys  glareolus)  captured  in  pit- 
falls were  smaller  than  conspecifics 
taken  in  live-traps  (Andrzejewski 
and  Rajska  1972,  Boonstra  and  Krebs 
1978).  This  apparent  relationship  be- 
tween size  and  susceptibility  to  pit- 
falls is  likely  related  to  jumping  abil- 
ity which  tends  to  increase  with  age. 
For  some  sf>ecies,  pregnant  females 
may  be  more  susceptible  to  pitfalls. 

Effects  of  Habitat  on  Trapping 
Effectiveness 

Trapping  results  for  Onychomys  tor- 
ridus  varied  substantially  between 


Table  2.--Total  number  of  new  individuals  captured  in  burned  and  un- 
burned  desert  habitats  using  live-traps  cmd  pitfalls  (3800trap-nights/habl- 
tat/trap  type). 


Burned  Unbumed  Total 


Species  Live-trap  Pitfall  Uve-trap  Pitfall  Live-trap  Pitfall 


AmmospermopNIus  harrisii      2  14  1  6  1 

Dipodomys  memami  11  3  14 

Neofomo  albigula  11  1        38  49  1 

Notiosorex  crawfordi  1  1 

Onychomys  forridus  17.  1  1  8 

Peromyscus  eremicus  2  3  5 

Perognathus  amplus  91  76        85  25       176  101 

Perognathus  baileyi  18  30        21  10        39  40 

Total  captures  136  115       154  38      290  152 

Species  richness  7  5          6  5          7  6 

Overall  species  richness  7                  8  8 

New  captures/trap-night  3.58  3,02      4.05  1.00      3.81  2.00 
X  100 

All  captures/trap-night       11.36  3.16     15.45  1.08     13.41  2.11 
X  100 


286 


vegetative  communities.  On  the  des- 
ert sites,  8  of  9  captures  v^ere  in  pit- 
falls whereas  in  composite/grass 
habitat  in  Garden  Canyon,  8  of  10 
captures  were  in  live-traps.  Four  cap- 
tures were  made  with  each  method 
in  the  juniper  woodland.  Differences 
in  trapability  of  Oncychomys  may  be 
due  to  different  depths  of  pitfalls  in 
desert  (40  cm)  versus  riparian  (36 
cm)  habitats. 

Except  for  Perogmthus  spp.,  ro- 
dents were  about  equally  susceptible 
to  pitfalls  relative  to  live-traps  in 
both  burned  and  unburned  desert 
habitats.  Differences  in  total  number 
of  individuals  captured  by  both 
methods  in  the  desert  areas  may  be 
due  to  (1)  difference  in  abundance  of 
species  on  burned  and  unburned 
plots  (Simons  1986);  or  (2)  differences 
in  activity  patterns  related  to  the 
drastic  difference  in  shrub  cover.  Per- 
ogmthus spp.  typically  prefer  brush 
or  "cover"  microhabitats  (Price  1978) 
and  raised  pitfall  covers  may  have 
attracted  these  mice  more  on  the 
burned  area  where  natural  cover  was 
scare  than  on  the  unburned  area 
where  natural  cover  was  dense  (Si- 
mons 1986).  Whatever  the  cause,  the 
results  are  similar  to  those  found  in 
desert-shrub  and  mesquite-grassland 
habitats  in  Durango,  Mexico,  where 
significantly  more  small-bodied 
mammals  were  captured  with  live- 
traps  than  with  pitfalls  (5.4  L  tin  can 
pitfalls  with  a  depth  of  25.4  cm)  (Pe- 
terson 1980).  Possibly  a  greater  num- 
ber of  captures  (i.e.,  sample  size)  may 
be  needed  to  fully  reveal  the  impact 
of  habitat  on  trapping  methodology. 

Conclusions 

Neither  method  alone  was  able  to 
fully  assess  small  mammal  communi- 
ties in  the  desert-scrub  and  riparian 
communities  we  investigated.  We 
recommend  the  use  of  both  methods, 
particularly  when  it  is  imjx)rtant  to 
include  species  such  as  shrews  that 
are  not  easily  caught  in  live-traps  in 
investigations  of  small  mammal  com- 


munity structure  and  habitat  rela- 
tionships. 

Acknowledgments 

We  thank  T.  J.  O'Shea,  M.  G.  Ryan, 
D.  W.  Uresk,  and  D.  F.  Williams  for 
their  critical  reviews  of  this  manu- 
script. C.  Munns  helped  check  pit- 
falls at  Garden  Canyon.  The  Depart- 
ment of  Zoology,  Arizona  State  Uni- 
versity provided  support  for  L.  Si- 
mons. 


Literature  Cited 

Andrzejewski,  Roman,  and  Ewa 
Rajska.  1972.  Trapability  of  bank 
vole  in  pitfalls  and  live  traps.  Acta 
Theriologica  17:41-56. 

Andrzejewski,  Roman,  and  Henryka 
Wroclawek.  1963.  Metal  cylinder 
as  a  live  trap  with  a  bait.  Acta  Th- 
eriologica 6:297-300. 

Boonstra,  Rudy,  and  Charles  J. 
Krebs.  1978.  Pitfall  trapping  of 
Microtus  townsendii.  Journal  of 
Mammalogy  59:136-148. 

Boonstra,  Rudy,  and  F.  Helen  Rodd. 
1984.  Efficiency  of  pitfalls  versus 
live  traps  in  enumeration  of  popu- 
lations of  Microtus  pennsylvanicus. 
Canadian  Journal  of  Zoology 
62:758-765. 

Briese,  Linda  A.,  and  Michael  H. 
Smith.  1974.  Seasonal  abundance 
and  movement  of  nine  species  of 
small  mammals.  Journal  of  Mam- 
malogy 55:615-629. 

Brown,  David  E.  1982.  Biotic  commu- 
nities of  the  American  South- 
west— United  States  and  Mexico. 
Desert  Plants  4:1-342. 

Chelkowska,  Henryka.  1967.  An  at- 
tempt at  comparing  two  methods 
of  trapping  small  rodents  (in  pit- 
falls and  live  traps).  Ekologia 
Polska  Seria  A  15:779-785. 

Cockburn,  A.,  M.  Fleming,  and  J. 
Wainer.  1979.  The  comparative 
effectiveness  of  drift  fence  pitfall 
trapping  and  conventional  cage 
trapping  of  vertebrates  in  the  Big 


Desert,  North-Western  Victoria. 
Victorian  Naturalist  96:92-95. 

Daly,  Martin,  and  Philip  Behrends. 
1984.  Effect  of  moving  traps  be- 
tween trapping  stations  upon  ro- 
dent retrapping  data.  American 
Midland  Naturalist  112:205-207. 

Hilbom,  R.,  J.  A.  Redfield,  and  C.  J. 
Krebs.  1976.  On  the  reliability  of 
enumeration  for  mark  and  recap- 
ture census  of  voles.  Canadian 
Journal  of  Zoology  54:1019-1024. 

Hoffmeister,  Donald  F.  1986.  Mam- 
mals of  Arizona.  602  p.  University 
of  Arizona  Press  and  Arizona 
Game  and  Fish  Department, 
Tucson. 

Howard,  Walter  E.,  and  Elbert  M. 
Brock.  1961.  A  drift-fence  pit  trap 
that  preserves  captured  rodents. 
Journal  of  Mammalogy  42:386-391. 

Jolly,  G.  M.  1965.  Explicit  estimates 
from  capture-recapture  data  with 
both  death  and  immigration — sto- 
chastic model.  Biometrica  52:225- 
247. 

Krebs,  Charles  J.  1966.  Demographic 
changes  in  fluctuating  populations 
of  Microtus  californicus.  Ecological 
Monographs  36:239-273. 

Mengak,  Michael  T.,  and  David  C. 
Guynn,  Jr.  1987.  Pitfalls  and  snap 
traps  for  sampling  small  mammals 
and  herpetofauna.  American  Mid- 
land Naturalist  118:284-288. 

Pankakoski,  Erkki.  1979.  The  cone 
trap — a  useful  tool  for  index  trap- 
ping of  small  mammals.  Annales 
Zoologici  Fennici  16:144-150. 

Peterson,  Michael  K.  1976.  Notewor- 
thy range  extensions  of  some 
mammals  in  Durango,  Mexico. 
Southwestern  Naturalist  21:139- 
142. 

Peterson,  Michael  K.  1980.  A  com- 
parison of  small  mammal  popula- 
tions sampled  by  pit-fall  and  live- 
traps  in  Durango,  Mexico.  South- 
western Naturalist  25:122-124. 

Price,  Mary  V.  1978.  The  role  of  mi- 
crohabitat  in  structuring  desert 
rodent  communities.  Ecology 
59:910-921. 

Pucek,  Z.  1969.  Trap  response  and 
estimation  of  numbers  of  shrews 


287 


in  removal  catches.  Acta  Theriol- 
ogica  14:403-426. 

Simons,  Lee  Hervey.  1986.  Rodent 
dynamics  in  burned  and  unburned 
desert  habitat  with  tests  of  the  ef- 
fects of  fire.  M.S.  Thesis.  57  p.  Ari- 
zona State  University,  Tempe. 

Szaro,  Robert  C.  1988.  Riparian  forest 
and  scrubland  community  types 
of  Arizona  and  New  Mexico.  Des- 
ert Plants  9  [In  press]. 

Weiner,  James  G.,  and  Michael  H. 
Smith.  1972.  Relative  efficiencies 
of  four  small  mammal  traps.  Jour- 
nal of  Mammalogy  53:868-873. 

Williams,  Daniel  F.,  and  Suzanne  E. 
Braun.  1983.  Comparison  of  pitfall 
and  conventional  traps  for  sam- 
pling small  mammal  populations. 
Journal  of  Wildlife  Management 
47:841-845. 


The  Role  of  Habitat  Structure 
in  Organizing  Smali  IVIammal 
Populations  and 
Comnnunities^ 

Gregory  H.  Adler^ 


Abstract.— Microhabitat  structure  influences 
population  density  more  than  other  demographic 
variables  such  as  age  and  sex  composition. 
Microhabitat  heterogeneity,  or  quantitative 
variation  in  microhabitat  structure,  apparently  has 
little  influence  on  phenomena  such  as  population 
stability.  Scale  mediates  effects  of  habitat  structure 
and  heterogeneity  on  population  and  community 
organization.  I  suggest  that  microhabitat  structure 
influences  density  more  than  other  aspects  of 
demography,  whereas  mocrohobitat  structure  and 
heterogeneity  are  more  important  in  influencing 
population  stability,  demography,  and  community 
structure. 


Environmental  heterogeneity  has 
maintained  a  position  of  prominence 
in  theoretical  population  and  com- 
munity ecology  (reviewed  by  Levin 
1976,  Wiens  1976,  and  Wiens  et  al. 
1986).  Heterogeneity  allows  organ- 
isms to  select  different  habitats, 
which  subsequently  can  have  pro- 
found consequences  for  the  organiza- 
tion of  populations  and  connmunities. 
Environmental  heterogeneity  can  be 
studied,  both  theoretically  and  em- 
pirically, at  different  scales.  Conclu- 
sions based  on  the  study  of  habitat 
structure  may  differ  widely  depend- 
ing upon  the  scale  of  structure  exam- 
ined. The  scale  of  environmental  sub- 
division can  be  viewed  as  occurring 
along  various  continua,  e.g.,  from  the 
area  occupied  by  a  single  individual 
to  a  biogeographic  or  continental 
area  (Wiens  et  al.  1986),  or  from  mi- 
crohabitat to  macrohabitat. 

In  this  paper,  I  concentrate  on  the 
microhabitat  to  macrohabitat  scale.  I 
define  microhabitat  as  physical  habi- 
tat characteristics  likely  to  vary  over 
the  home  range  of  a  single  individual 
(e.g.,  the  number  of  herbaceous 
stems  within  a  circumscribed  area) 
and  macrohabitat  as  the  major  habi- 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibiarts.  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff. AZ.  July  19-21.  1988.) 

'Gregory  H.  Adier  is  Research)  Fellow  in 
Population  Sciences.  Department  of  Popu- 
lation Sciences.  Schiool  of  Public  l-lealthi. 
Harvard  University.  665  Huntington  Avenue. 
Boston.  MA  021 15. 


tat  type  where  an  entire  population 
may  be  found  (e.g.,  grassy  field  or 
deciduous  woodland  in  the  case  of 
small  mammals).  Microhabitat  struc- 
ture therefore  can  vary  substantially 
within  a  single  macrohabitat. 

I  summarize  results  from  a  series 
of  long-term  studies  on  the  role  of 
habitat  structure  in  organizing  small 
mammal  populations  and  communi- 
ties that  I  conducted  in  eastern  Mas- 
sachusetts. These  studies  were  de- 
signed to  examine  (1)  habitat  associa- 
tions and  habitat  selection  and  the 
roles  of  intra-  and  interspecific  inter- 
actions in  affecting  habitat  utiliza- 
tion, and  (2)  the  influence  of  habitat 
structure  on  density  and  demogra- 
phy. In  these  studies,  I  focus  primar- 
ily on  microhabitat  structure,  and  I 
develop  a  conceptual  scheme  which 
shows  how  microhabitat  and  mac- 
rohabitat structure  organize  small 
mammal  populations  and  communi- 
ties. 


STUDY  SITES  AND  GENERAL 
METHODS 

Study  Sites 

The  long-term  studies  were  con- 
ducted at  three  sites  in  eastern  Mas- 
sachusetts: Broad  moor /Little  Pond 
Audubon  Sanctuary,  South  Natick; 
Great  Island,  near  West  Yarmouth; 
and  the  University  of  Massachusetts 
Nantucket  Field  Station,  Nantucket. 
Sampling  areas  within  each  study 


site  were  confined  to  a  300-ha  area 
and  were  exposed  to  the  same  cli- 
mate and  the  same  predators,  com- 
petitors, and  parasites. 

Broadmoor  consists  of  a  mosaic  of 
grassy  fields  separated  by  nnixed  de- 
ciduous-coniferous woodland.  Sam- 
pling at  Broadmoor  was  confined  to 
the  fields,  which  were  dominated  by 
the  grasses  Agropyron  repens  and  Poa 
pratensis.  Other  herbaceous  and 
woody  plants,  including  goldenrod 
(Solidago  spp.),  milkweed  (Asclepias 
syriaca),  poison  ivy  (Rhus  radicans), 
and  several  species  of  deciduous  tree 
saplings,  were  much  less  prevalent. 

Great  Island  is  a  240-ha  island 
connected  to  mainland  Cape  Cod  by 
a  causeway.  The  island  is  dominated 
by  deciduous  and  coniferous  wood- 
land but  has  structurally  simpler 
habitat  along  the  shore.  This  shore- 
line habitat  consists  primarily  of 
beach  grass  (Ammophila  breviligulata), 
with  patches  of  poison  ivy,  Virginia 
creeper  (Parthenocissus  quinquefolia), 
bayberry  (Myrica  pensylvanica),  rose 
(Rosa  Carolina),  and  juniper  (Juniperus 
virginiana). 

Nantucket  Island  (ca.  12,300  ha 
and  lying  approximately  30  km  off 
the  coast  of  Cape  Cod)  has  large  ar- 
eas of  low,  dense  woody  growth 
(heath)  where  small  mammals  were 
sampled.  Heath  at  the  study  site  was 
composed  primarily  of  rose  and  bay- 
berry,  with  patches  of  goldenrod  and 
other  herbaceous  plants  and  grasses 
interspersed  within  the  brush.  Scat- 
tered juniper  trees  also  were  present. 


289 


Sampling  Procedures 

I  sampled  small  mammals  at  each 
study  site  by  monthly  live-trapping 
with  Longworth  live-traps  for  ap- 
proximately 4  to  5  years.  At  each 
study  site,  I  monitored  two  0.4-ha 
grids  located  in  grassy  or  brushy 
habitat.  One  grid  served  as  a  control 
in  which  all  small  mammals  were 
individually  marked  by  ear-tags  (ro- 
dents) or  toe-clips  (insectivores).  The 
other  grid,  located  30.4  m  from  the 
control  and  situated  in  contiguous 
habitat,  served  as  an  experimental 
grid  from  which  all  small  mammals 
were  removed  permanently  upon 
first  capture  (Adler  1985).  All  small 
mammals  captured  on  this  grid  after 
the  initial  removal  period  were  con- 
sidered colonists.  I  also  sampled 
small  mammals  on  4  nearby  trapping 
plots  which  also  were  located  in 
similar  macrohabitat  but  covered  a 
range  of  microhabitats.  Each  plot 
consisted  of  two  parallel  traplines 
located  30.4  m  apart.  Each  trapline 
was  15  stations  long  at  Broadmoor 
(except  on  one  plot  where  both  tra- 
plines were  12  stations  long)  and  on 
Nantucket  and  20  stations  long  on 
Great  Island.  These  plots  were 
trapped  on  a  rotation  basis  (Adler 
1987).  On  Great  Island,  an  additional 
4  control  grids  were  monitored 
monthly  from  April  through  Septem- 
ber for  five  years  (Adler  and  Wilson 
1987).  These  grids  were  not  confined 
to  structurally  simple  macrohabitats 
but  ranged  from  grassland  to  mature 
woodland  habitats. 

Grid  1  was  located  at  the  edge  of  a 
stand  of  pitch  pine.  (Pinus  rigida), 
white  oak  (Quercus  alba),  and  black 
oak  (Q.  velutim).  Dense  brushy 
understory  covering  a  large  portion 
of  the  grid  consisted  of  bayberry, 
huckleberry  (Gaylussacia  baccata),  and 
inkberry  holly  (Ilex  glabra).  Low-lying 
areas  of  the  grid  were  damp  and  har- 
bored large  cranberry  (Vaccinium 
macrocarpon)  and  sundew  {Drosera 
spp.).  Very  little  herbaceous  vegeta- 
tion was  present.  Grid  2  also  was  lo- 
cated at  the  edge  of  a  pitch  pine. 


white  oak,  and  black  oak  woodland 
but  was  more  elevated  and  conse- 
quently drier.  A  dense  brushy  under- 
story consisted  of  bayberry,  poison 
ivy,  and  common  greenbrier  (Smilax 
rotundifolia).  Grass  was  present  in  the 


brushy,  treeless  portions  of  the  grid. 
Grid  3  was  located  within  a  white 
oak  and  black  oak  woodland.  A 
dense  shrub  cover  of  blueberry, 
bearberry  (Arctostaphylos  uva-ursi), 
common  greenbrier,  and  bullbrier 


Table  1  .—Description  of  the  habitat  variables  measured  at  each  trap  sta- 
tion. All  variables  measured  as  proporHons  was  arcsin  square  root  trans- 
formed. 


Name 


Description 


WOOD  Density  of  woody  stems  within  a  1  -m^  circle  at  ground 

level. 

HERB  Number  of  Inerbaceous  stems  (excluding  grasses  and 

sedges)  within  a  1-m^  circle  at  ground  level. 

WDSPEC  Number  of  woody  species  within  a  1  -m^  circle  at  ground 

level. 

HBSPEC  Number  of  herbaceous  species  (excluding  grasses  and 

sedges)  within  a  1-m^  circle  at  ground  level. 
HB50  Number  of  herbaceous  sterris  (excluding  grasses  and 

sedges)  within  a  1-m^  circle  at  50  cm  above  ground  level. 
HBICX)  Number  of  herbaceous  stems  (excluding  grasses  and 

sedges)  within  a  1-m^  circle  at  1  m  above  ground  level. 
VHBDEN  Mean  of  HERB,  HB50,  and  HBIOO. 

WD50  Number  of  woody  stems  within  a  1  -m^  circle  at  50  cm 

above  ground  level, 
WDIOO  Number  of  woody  stems  within  a  1-m^  circle  at  1  m 

above  ground  level. 
VWDEN  Mean  of  WOOD,  WD50,  and  WDIOO. 

OVER  Number  of  overstory  species  within  a  15-m^  circle. 

UNDER  Number  of  shrub  level  species  within  a  15-m^  circle. 

FORB  Number  of  f orb  species  within  a  15-m^  circle. 

GRASPEC  Number  of  grass  and  sedge  species  within  a  1 5-m^  circle. 
GRNDSPEC       Number  of  woody  ground-dwelling  vine  species  within  a 

IS-m^circie, 

SPECIES  Total  number  of  angiosperm  and  gymnosperm  species 

within  a  15-m^  cirde. 
TPSHRUB  Trar^formed  proportion  of  a  15-m^  circle  dominated  by 

woody  shrub-level  vegetation. 
TPHERB  Transformed  proportion  of  a  ]S-m'^  circle  dominated  by 

herbaceous  vegetation  (excluding  grasses  and  sedges). 
TPGRND  Trar^sformed  proportion  of  ]5-m^  circle  dominated  by 

woody  ground-dwelling  vines. 
TPGRASS  Trar>sformed  proportion  of  IS-m^  circle  dominated  by 

grasses  and  sedges. 
TPVEG  Trortsformed  proportion  of  a  15-m^  circle  covered  by 

vegetation. 

HBGRND  An  index  of  herbaceous  ground  cover  (excluding  grasses 

and  sedges),  calculated  as  HERB-HB50, 
WDGRND         An  index  of  woodv  ground  cover  calculated  as  WOOD- 

WD50. 

TPCANOPY       Trar^sformed  proportion  of  a  canopy  cover  measured  only 

on  the  five  Great  Island  control  grids. 
TPGREEN         Trar^sformed  proportion  of  evergreen  canopy  cover, 

measured  only  on  the  five  Great  Island  control  grids. 


290 


greenbrier  (S.  bom-nox)  was  present, 
along  with  bracken  fern  (Pteridium 
aquilinium).  Grid  4  was  located  on 
Pine  Island,  a  7-ha  islet  37  m  from 
Great  Island  and  connected  to  the 
latter  by  a  narrow  sandy  spit.  White 
oak  and  black  oak  formed  a  canopy 
over  much  of  the  grid,  and  a  dense 
woody  understory  of  bayberry  and 
other  shrubs  also  was  present.  Dense 
beach  grass  was  present  in  the  tree- 
less portions  of  the  grid.  Grid  5  was 


the  companion  control  for  the  experi- 
mental grid  and  was  located  in  dense 
beach  grass  containing  scattered 
patches  of  bayberry,  juniper,  and  poi- 
son ivy. 

I  sampled  vegetation  structure  at 
every  trap  station  on  all  grids  and 
plots  by  measuring  23  habitat  vari- 
ables related  to  plant  structure  and 
species  richness  (table  1).  Two  addi- 
tional habitat  variables  describing 
canopy  structure  were  included  in 


Tcrt>le  2.— Summary  of  the  sampling  design  and  statistical  approach  em- 
ployed in  this  study. 


Sampling  area 


Control  grid 


Experimental 
grid 


Trapping  plots 


Additional  con- 
trol grids  (4  at 
Great  Island) 


Topic 


Memods 


Habitat  associotions. 


Temporal  dynamics  of 
habitat  use. 


Habitat  selection. 


Relationship  between 
demography  and  mi- 
crohabitat  structure 
within  a  macrohabitat. 

Relationship  between 
demography  and  mi- 
crohabitat  structure  of 
a  habitat  generalist 
across  macrohabitat 
boundaries. 


Multiple  linear  regression 
of  numbers  of  captures  at 
a  trap  station  on  mi 
crohabitat  variables  de- 
rived from  PCA, 

DFA  to  derive  a  quantita- 
tive measure  of  seasonal 
habitat  use  (the  distinction 
between  favorable  and 
unfavorable  microhabi- 
tafs,  or  habitat  discrimina- 
tion). Regression  of  dis- 
crimination values  on 
population  densities  to  de- 
termine the  relatior^ship 
between  microhabitat  use 
and  intra- and  interspeci- 
fic population  der^ities. 

Multiple  linear  regression 
of  numbers  of  captures  at 
a  trap  station  (in  a  per- 
turbed area)  on  mi- 
crohabitat variables  de- 
rived from  PCA/ compared 
with  control  grid. 

Regression  (and  residual 
analysis)  of  demographic 
variables  on  plot  means  of 
microhabitat  gradients 
and  heterogeneity. 

Regression  (and  residua! 
analysis)  of  demographic 
variables  on  grid  means  of 
microhabitat  gradients 
and  heterogeneity. 


the  analysis  on  Great  Island  control 
grids  (table  1).  Measurement  proce- 
dures were  given  by  Adler  (1985) 
and  Adler  and  Wilson  ( 1987). 


Data  Analysis 

I  relied  extensively  upon  principal 
components  analysis  (PCA)  and  dis- 
criminant function  analysis  (DFA)  in 
order  to  uncover  the  structure  of 
complex  and  temporally  variable 
small  mammal  populations  and  their 
relationships  to  habitat  structure. 
Specifically,  my  aims  were  to  (1)  re- 
duce the  number  of  habitat  dimen- 
sions, (2)  derive  a  quantitative  meas- 
ure of  habitat  heterogeneity,  (3) 
quantify  patterns  of  habitat  utiliza- 
tion, (4)  combine  covarying  demo- 
graphic traits  into  single  variables, 
and  (5)  derive  indices  of  demo- 
graphic variability. 

In  these  studies,  I  recognized  two 
related  descriptors  of  microhabitat 
structure.  I  defined  a  microhabitat 
structure-diversity  variable  or  gradi- 
ent as  a  characteristic  that  described 
the  physical  structure  of  the  mi- 
crohabitat and  that  varied  in  magni- 
tude along  a  continuum.  I  defined 
microhabitat  heterogeneity  as  a 
quantitative  measure  of  horizontal 
variation  in  microhabitat  characteris- 
tics (August  1983,  Adler  1987). 

I  subjected  the  habitat  data  meas- 
ured at  each  trap  station  to  PCA  to 
reduce  the  number  of  habitat  vari- 
ables. At  each  site,  I  conducted  two 
PC  As  of  the  23  variables,  one  with 
control  and  experimental  grids  com- 
bined and  one  with  the  4  trapping 
plots  combined.  I  also  conducted  a 
PCA  of  24  habitat  variables  for  all 
five  control  grids  on  Great  Island 
combined.  HBIOO  was  eliminated 
from  this  analysis  because  only  one 
nonzero  value  was  recorded  on  the 
five  grids. 

Each  principal  component  (PC) 
with  an  eigenvalue  greater  than  1.0 
was  retained  for  further  analysis  as  a 
new  habitat  variable.  Principal  com- 
ponents derived  from  PCAs  of  grid 


291 


and  plot  data  were  quite  similar 
within  each  site,  based  upon  factor 
loadings  on  the  original  habitat  vari- 
ables (Adler  1985, 1987).  At  Broad- 
moor, five  PCs  were  retained  for 
analysis  from  both  grid  and  plot 
data,  whereas  six  were  retained  from 
analysis  of  grid  data;  four  PCs  were 
interpreted  similarly  in  both  data 
sets.  PCAs  of  Nantucket  grid  and 
plot  data  both  yielded  seven  retain- 
able PCs,  three  of  which  could  be  in- 
terpreted similarly  between  the  two 
data  sets.  The  PCA  of  habitat  data 
from  the  five  control  grids  on  Great 
Island  yielded  seven  PCs. 

I  computed  a  microhabitat  hetero- 
geneity index  for  each  of  the  four 
trapping  plots  at  the  three  study  sites 
and  for  each  of  the  five  control  grids 
on  Great  Island  (Adler  1987,  Adler 
and  Wilson  1987).  This  index  was 
based  on  the  supposition  that  the 
standard  deviation  of  the  within-plot 
or  within-grid  mean  vector  of  a  PC 
described  the  variability  of  a  mi- 
crohabitat gradient  on  a  given  plot  or 
grid.  Since  each  successive  PC  con- 
tributed less  to  the  total  variance  in 


habitat  data,  I  adjusted  for  each  PC's 
contribution  to  the  total  variance  by 
multiplying  the  factor  scores  by  the 
square  root  of  that  PC's  eigenvalue. 

I  examined  capture  data  in  rela- 
tion to  habitat  structure  at  both  the 
level  of  individual  trap  stations 
(habitat  association  and  selection) 
and  at  the  level  of  a  grid  or  plot  (de- 
mography). I  used  multiple  linear 
regression  and  residuals  analysis  to 
relate  these  small  nnammal  (depend- 
ent) variables  to  habitat  (independ- 
ent) variables.  More  complete  de- 
scriptions of  analytical  techniques  are 
given  in  each  section  below,  and  a 
brief  outline  of  the  sampling  design 
is  given  in  table  2. 


SPECIES  COMPOSITION 

I  recorded  9,170  captures  of  10  small 
mammal  species  in  42,773  trapnights 
at  the  3  study  sites  (table  3).  Each 
study  site  generally  had  an  abundant 
herbivore  (Microtus  pennsylvanicus), 
an  abundant  granivore  (Peromyscus 
leucopus,  except  at  Broadmoor  where 


it  was  rare  in  the  grassland  trapping 
areas),  a  common  insectivore  (Blarina 
brevicauda  or  Sorex  cinereus),  and  any 
of  several  rarer  granivores,  omni- 
vores,  or  insectivores. 


HABITAT  STRUCTURE  AND 
POPULATION  STATISTICS 


Habif  at  Associations  and 
Selection 


Study  Purpose 

I  exan\ined  both  small  mammal  mi- 
crohabitat associations  and  selection 
at  all  three  study  sites  (Adler  1985). 
Density-dependent  effects  of  con- 
specifics  and  other  species  may  re- 
strict access  to  certain  habitat  types, 
thereby  resulting  in  different  patterns 
of  habitat  utilization.  I  therefore  re- 
served the  term  habitat  selection  for 
situations  where  individuals  had 
more  or  less  unrestricted  access  to  a 
variety  of  habitat  types. 


Table  3.— Trapping  effort  and  numbers  of  captures  of  small  mammals  at  three  study  sites  In  eastem  Massachusetts. 
Species  designations  are  MP  (Microtus  pennsytvanhus),  PL  (Peromyscus  leucopus),     (Bfarina  brevicauda),  SC 
(Sorex  cinereus),  ZH  (Zapus  hudsonlus),  RN  (Rattus  norvegicus),  SA  (Scalopus  aquaticus),  TS  (Tamias  striatus),  CG 
(Clethrionomys  gapperl),  and  CC  (Condylura  crisfata). 


Site 

Trap 
periods 

Trap- 
nights 

MP 

PL 

BB 

SC 

ZH 

RN 

SA 

TS 

CG 

cc 

TOTAL 

Broadmoor 

Control 

32 

3332 

416 

1 

76 

7 

6 

0 

0 

0 

0 

0 

507 

Exptl. 

32 

3136 

225 

11 

61 

5 

23 

0 

0 

2 

0 

1 

327 

Plots 

32 

1824 

212 

19 

48 

7 

8 

1 

0 

0 

0 

0 

295 

Great  Islar^d 

Ctl(l) 

43 

4157 

86 

400 

0 

22 

18 

0 

0 

15 

7 

0 

548 

Cti(2) 

31 

2940 

146 

359 

25 

12 

7 

0 

0 

39 

0 

0 

588 

Ctl  (3) 

31 

2989 

12 

358 

13 

10 

0 

0 

0 

71 

0 

0 

464 

Ctl  (4) 

30 

2934 

551 

404 

0 

41 

0 

0 

0 

0 

0 

0 

996 

Ctl  (5) 

43 

4193 

603 

349 

11 

61 

21 

0 

0 

3 

0 

0 

1048 

Exptl. 

35 

3381 

219 

111 

15 

74 

18 

0 

0 

1 

0 

0 

438 

Plots 

35 

2686 

336 

221 

3 

54 

12 

0 

0 

0 

0 

0 

626 

Nantucket 

Control 

35 

5782 

1364 

255 

83 

0 

3 

2 

2 

0 

0 

0 

1709 

Exptl. 

35 

3621 

420 

270 

75 

2 

0 

3 

0 

0 

0 

0 

770 

Plots 

35 

1798 

400 

420 

28 

0 

0 

6 

0 

0 

0 

0 

854 

TOTAL 

449 

42773 

4990 

3178 

438 

295 

116 

12 

2 

131 

7 

1 

9170 

292 


Analytical  Approach 

I  defined  an  association  as  a  staristi- 
cal  relationship  between  the  numbers 
of  captures  of  a  species  at  trap  sta- 
tions and  a  quantitative  measure  of 
microhabitat  structure.  To  determine 
these  relationships,  I  regressed  the 
total  number  of  captures  of  a  species 
at  each  control  grid  trap  station  on 
factor  scores  of  each  PC.  The  experi- 
mental grid  represented  an  area 
where  densities  were  continually 
being  reduced  and  vacant  microhabi- 
tats  were  more  often  available  to 
colonizing  individuals. 

To  determine  differences  in  mi- 
crohabitat associations  between  con- 
trol and  experimental  grids,  I  in- 
cluded a  dummy  variable  coding  for 
grid  (control  or  exjjerimental)  and 
habitat  variable  x  grid  interaction 
tern\s  (Adler  1985). 


Inferences 

Most  small  mammals  (8  of  11  fx)pu- 
lations  examined)  demonstrated  af- 
finities for  specific  microhabitat  types 
on  either  control  or  experimental 
grids  (table  4).  These  affinities  gener- 
ally were  consistent  with  other  pub- 
lished reports  of  habitat  associations 
of  these  species.  For  instance,  P.  leu- 
copus  generally  were  associated  posi- 
tively with  woody  microhabitats  or 
negatively  associated  with  herba- 
ceous microhabitats.  M.  pennsylvani- 
cus  generally  showed  the  opposite 
associations.  Microhabitats  selected 
by  small  mammals,  as  determined 
from  capture  data  on  experimental 
removal  grids,  sometimes  differed 
from  associations  determined  from 
capture  data  on  the  adjacent  control 
grids  (table  4).  Differences  in  habitat 
selection  and  association  were  attrib- 


Table  4.— Habitat  associations  of  small  mammals  In  eastern  Massachusetts 
determined  from  regressions  of  numbers  of  captures  at  trap  stations  on 
habitat  variables  derived  from  principal  components  analysis.  The  direc- 
tion of  Vne  regression  slope  Is  given  by  +  or  -  ond  tti«  strength  of  the  rela- 
tionship Is  given  by  the  number  of  signs  (1 ,  P<0.05;  2,  P<0.01 ;  3,  P<0.001). 
Species  designations  are  as  In  table  3. 


Site  and 
Species 


Description  of  habitat  variable 


Control 
grid 


Experimental 


Broadmoor 
BB 
MP 


ZH 

Great  Island 
SC 
BB 
PL 

MP 


ZH 

Nantucket 
BB 
PL 

MP 


Alt  variables 

Herbaceous  density  and  height 
Shrubby  vegetation 
Ground-level  herbaceous  density 
Herbaceous  density  and  height 

All  variables 

Vertical  woody  vegetation  density 
Total  vegetation  cover,  primarily 
herbaceous 
Habitat  complexity 
Habitat  structure,  reflecting 
increasing  herbaceousness 
and  decreasing  woodiness 
Ground-level  woody  vine  density 
All  variables 

Herbaceous  species  richness 
Number  of  overstory  trees 
Vertical  vegetation  structure 
Herb  species  richness 


N$ 
+++ 


NS 

NS 
+ 

NS 


NS 

+ 

+ 

NS 


NS 

NS 
+ 

+ 

NS 


NS 
NS 


NS 

NS 
NS 
+ 

NS 


utable  to  opportunistic  responses  of 
small  mammals  to  between-grid  dif- 
ferences in  microhabitat  structure 
and  to  differences  in  the  level  of  in- 
traspecific  interactions  brought  about 
through  density  reductions  on  the 
experimental  grids  (Adler  1985). 


Tennporol  Patterns  of  Habitat  Use 
Study  Purpose 

I  examined  temporal  patterns  of  mi- 
crohabitat use  by  M.  pennsylvanicus 
at  the  three  study  sites  and  by  P.  leu- 
copus  on  Great  Island  and  Nantucket. 

Analytical  Approach 

Monthly  trapping  periods  were 
grouped  into  winter  (Dec.-Feb.), 
spring  (Mar.-May),  summer  (Jun.- 
Aug.),  and  fall  (Sep.-Nov.)  seasons 
each  year.  I  divided  trap  stations  on 
control  grids  into  favorable  and  unfa- 
vorable nnicrohabitats  each  season 
depending  upon  whether  the  total 
number  of  captures  in  a  season  was 
above  (favorable)  or  below  (unfavor- 
able) the  seasonal  mean  (Van  Home 
1982;  Adler  1985).  I  then  used  a  two- 
group  DFA,  with  favorable  and  unfa- 
vorable trap  stations  defining  the 
two  groups,  to  develop  a  discrimina- 
tion index  of  habitat  use  (Rice  et  al. 
1983;  Adler  1985).  This  index  was  the 
percentage  of  trap  stations  classified 
correctly  as  either  favorable  or  unfa- 
vorable. High  discrimination  values 
indicated  a  sharp  distinction  between 
favorable  and  unfavorable  micro- 
habitats; low  values  indicated  little 
difference  between  favorable  and  un- 
favorable areas. 

To  determine  the  importance  of 
intra-  and  interspecific  population 
densities  on  temporal  patterns  of 
habitat  discrimination  by  P.  leucopus 
and  M.  pennsylvanicus,  I  regressed 
the  seasonal  discrimination  values  on 
the  mean  seasonal  densities  of  each 
of  the  major  small  mammal  species 
present  at  each  study  site. 


293 


Inferences 

In  the  case  of  M.  pennsylvanicus,  den- 
sity and  discrimination  were  nega- 
tively related  at  Broadnnoor  and 
positively  related  on  Great  Island. 
The  unexpected  positive  relationship 
on  Great  Island  could  be  explained 
by  the  distribution  of  captures  over 
the  grid;  17  capture  stations  had  less 
than  two  captures  during  the  entire 
study  and  were  in  a  sparsely  vege- 
tated area.  As  density  increased,  the 
reniaining  32  trap  stations  became 
increasingly  utilized.  The  distinction 
between  favorable  and  unfavorable 
microhabitats  increasingly  became  a 
distinction  between  unoccupied, 
sparsely  vegetated  stations  and  occu- 
pied, densely  vegetated  stations. 

On  Nantucket,  discrimination  fol- 
lowed a  pattern  similar  to  density 
but  was  not  linearly  related  to  the 
latter.  For  P.  leucopus  on  both  Great 
Island  and  Nantucket,  habitat  dis- 
crimination was  related  negatively  to 
density  (fig.  1),  indicating  that  the 
distinction  between  favorable  and 
unfavorable  microhabitats  decreased 
with  increasing  density.  Densities  of 
other  species  were  not  related  to 
temporal  variation  in  habitat  use 
(Adler  1985). 

Therefore,  intraspecific  competi- 
tion appeared  to  be  more  important 
than  interspecific  interactions  in  de- 
termining microhabitat  use  by  the 
species  I  examined.  As  intraspecific 
density  increased,  the  range  of  mi- 
crohabitat types  utilized  also  in- 
creased, as  predicted  by  early  theo- 
ries of  habitat  selection  (e.g., 
Svardson  1949). 


Microhabitat  Structure  and 
Dennography 

Study  Purpose 

I  examined  the  relationship  between 
demography  of  M,  pennsylvanicus 
and  microhabitat  structure  from  data 
collected  on  the  four  trapping  plots 
at  each  study  site  (Adler  1987). 


Analytical  Approacti 

I  calculated  density  (log^^  number 
per  100  trapnights),  sex  composition 
(proportion  males,  arcsin  square  root 
transformed),  age  structure  (propor- 
tion of  adults  captured  during  sam- 
pling periods  from  April  through 
September,  arcsin  square  root  trans- 
formed), and  breeding  intensity  (pro- 
p>ortion  of  adults  in  breeding  condi- 
tion captured  in  sampling  periods 
from  April  through  September, 
arcsin  square  root  transformed)  each 
trapping  period. 

I  also  computed  variability  meas- 
ures for  each  of  these  demographic 


variables  as  squared  distances  from 
plot  means.  I  divided  the  estimates 
for  density  variability  on  each  plot  by 
the  mean  density  of  the  respective 
plot  in  order  to  adjust  for  population 
size. 

I  regressed  the  estimates  for  each 
of  the  eight  demographic  variables 
separately  on  plot  means  for  each 
microhabitat  variable  derived  from 
PC  A  and  the  index  of  heterogeneity. 
I  then  regressed  the  unstandardized 
residuals  from  each  of  these  regres- 
sions separately  on  each  habitat  PC 
and  the  heterogeneity  index  to  search 
for  nonlinearities  and  missing  vari- 
ables (Framstad  et  al.  1985). 


GREAT  ISLAND 


80-1 


4  6 

DENSITY 
NANTUCKET 


o 
I— 
< 


100 


90- 


80- 


o 

00 


^  70- 


60-1 


10 


DENSITY 


Figure  1.— Relationships  between  seasonal  habitat  discrimination  and  population  density  In 
Peromyscus  leucopus  at  two  study  sites  In  eastern  Massachusetts. 


294 


Inferences 

Densities  of  M.  pennsylvanicus  and  P. 
leucopus  were  ordered  linearly  along 
microhabitat  gradients  (Adler  1987), 
consistent  with  patterns  of  mi- 
crohabitat associations  and  selection 
in  these  two  species  (table  5).  In 
general,  M.  pennsylvanicus  densities 
were  higher  on  plots  with  more  her- 
baceous and  grassy  cover  or  less 
woody  cover.  Nantucket  was  excep- 
tional, however,  with  M.  pennsylvani- 
cus densities  increasing  along  gradi- 
ents of  increasing  woody  growth  and 
shrub  species  richness.  I  captured 
large  numbers  of  this  vole  in  dense 
heath  with  little  or  no  herbaceous 
vegetation. 

Peromyscus  leucopus  densities  on 
Great  Island  could  not  be  related  to 
microhabitat  structure,  probably  be- 
cause of  the  generalist  nature  of  this 
mouse  relative  to  the  breadth  of  mi- 
crohabitats  sampled.  Indeed,  when 
sampling  areas  included  other  mi- 
crohabitats,  density  could  be  related 
to  overall  microhabitat  structure  (see 
below).  P.  leucopus  densities  on  Nan- 
tucket increased  with  increasing 
shrub  species  richness  (table  5).  Den- 
sities of  both  species  were  more  vari- 


able in  poorer  habitats.  Microhabitat 
structure  was  a  poor  predictor  of 
other  aspects  of  demography  such  as 
age  and  sex  composition.  However, 
variability  in  demographic  structure 
often  was  greater  in  low-density 
habitats.  While  some  of  the  variabil- 
ity in  density  and  demography  may 
have  been  due  to  statistical  depend- 
ence on  population  size  (i.e.,  greater 
sampling  error  at  small  population 
sizes),  biological  effects 
(e.g.,response  to  environmental  fluc- 
tuations) also  must  have  been  impor- 
tant. More  favorable  microhabitats 
should  have  maintained  a  more 
stable  composition  over  time  due  to 
greater  intraspecific  interactions, 
whereas  poorer  microhabitats  should 
have  contained  a  more  unstable  as- 
semblage of  predominantly  transient 
and  subordinate  individuals  due  to 
spillover  during  periods  of  high  den- 
sity (Adler  1987).  In  contrast  to  the 
importance  of  microhabitat  gradi- 
ents, the  quantitative  measure  of  mi- 
crohabitat heterogeneity  generally 
was  unrelated  to  demographic  phe- 
nomena. In  only  one  case  did  mi- 
crohabitat heterogeneity  explain  vari- 
ation in  demography  better  than  any 
structure-diversity  variable. 


Macrohabrtdt  Structure  and 
Dennography 

Study  Purpose 

I  further  examined  the  relationship 
between  demography  of  P.  leucopus 
and  microhabitat  structure  across 
macrohabitats.  P.  leucopus  is  a  habitat 
generalist  which  occurs  in  habitats 
ranging  from  grassland  to  mature 
deciduous  and  coniferous  forests  in 
southeastern  Massachusetts. 


Analytical  Approacti 

For  this  purpose,  data  from  the  five 
control  grids  on  Great  Island  were 
analyzed  (Adler  and  Wilson  1987). 
Monthly  trapping  data  were  ana- 
lyzed with  respect  to  10  demo- 
graphic variables.  Grid  means  of 
density  (log^^  minimum  number 
known  alive),  adult  male  body  mass, 
and  observed  range  length  (ORL,  the 
maximum  linear  distance  between 
capture  points  of  an  individual, 
Stickel  1954)  were  compared  using 
Tukey's  multiple  comparisons  test. 
Mean  male  and  female  ORLs  were 
compared  on  each  grid  using  t-tests. 

Contingency  table  analysis  was 
used  to  compare  age  structure  (pro- 
portion adult),  adult  survival  (stan- 
dardized 14-day  rates),  sex  composi- 
tion (proportion  of  mice  tagged  that 
were  males),  adult  residence  rates 
(proportions  of  adults  captured  in  at 
least  two  trapping  periods),  overwin- 
ter residence  (proportions  of  mice 
present  during  Sep.  and  surviving  to 
the  subsequent  Apr.),  the  propor- 
tions of  adults  that  were  reproduc- 
tively  active,  and  the  proportions  of 
young  mice  (mice  with  some  grey 
pelage  remaining)  that  were  repro- 
ductively  active.  These  10  variables 
were  examdned  for  intersex  differ- 
ences within  a  grid  (except  sex  com- 
position) and  for  intergrid  differ- 
ences. 

To  examine  temporal  dynamics  of 
demography,  monthly  trapping  data 
were  grouped  into  early  summer 


r 


ldb\e  5.— Relationships  between  density  of  Peromyscus  leucopus  (PL)  and 
Microtus  pennsylvanicus  (MP)  and  microhabitat  variables  (derived  from 
principal  components  analysis)  at  three  study  sites  In  eastern  Massachu- 
setts. Signs  of  correlation  are  as  Indicated  In  table  4. 


Site 


Species 


Habitat  variable 


Correlation 


Broadmoor 
Great  Island 


Nantucket 


MP   Decreasing  vertical  woody  stem  + 

der^ities  and  shrub  cover. 
PL    All  gradients.  NS 
MP   Increasing  woody  and  herbaceous  — 

stem  der^ties,  cover  and  species 

richness;  decreasing  grassiness 

!ncreasir>g  woody  ground  vine  — 

species  richness  and  cover, 

increasing  total  vegetation  cover  ++ 

decreasing  overstory  species  richness. 
PL    Increasing  shrub  species  richness,  + 
MP   Increasing  herbaceous  growth; 

decreasing  woody  growth. 

Increasing  plant  species  richness.  + 

Increasing  shrub  species  richness.  + 


295 


(Apr.-Jun.)  and  late  summer  (Jul.- 
Sep.)  seasons.  The  following  demo- 
graphic variables  were  estimated  on 
each  grid  during  each  season:  density 
(mean  log^g  minimum  number 
known  alive),  proportions  of  males 
and  of  females  that  were  adults,  pro- 
portion of  males,  mean  adult  male 
body  mass,  proportions  of  adult 
males  and  of  adult  females  breeding, 
and  survival  rates  of  adult  males  and 
of  adult  females  (weighted  mean  14- 
day  rates).  Variables  expressed  as 
proportions  were  arcsin  square  root 
transformed. 

Many  rodent  population  parame- 
ters are  known  to  covary  (e.g.,  Schaf- 
fer  and  Tamarin  1973).  Accordingly, 
a  PCA  of  the  eight  variables  was  exe- 
cuted in  order  to  include  covarying 
parameters  as  single  demographic 
variables;  four  PCs  with  eigenvalues 
greater  than  1.0  were  retained  for 
further  analysis. 

These  PCs  were  correlated  with 

(1)  density  and  adult  survival, 

(2)  adult  female  breeding  activity, 

(3)  adult  male  breeding  activity,  and 

(4)  the  proportion  of  males.  Variabil- 
ity indices  of  each  of  these  PCs  were 
calculated  each  season  for  each  grid 
as  squared  distances  from  grid 
means  (Adler  and  Wilson  1987).  A 
measure  of  overall  demographic 
variability  was  calculated  for  each 
grid  each  season  as  squared  dis- 
tances of  the  factor  scores  from  the 
mean  factor  score,  summed  over  the 
four  PCs. 

Factor  scores  within  each  PC  were 
multiplied  by  the  square  root  of  that 
PC's  eigenvalue  in  order  to  account 
for  the  unequal  contributions  to 
overall  variance  of  each  PC  (Adler 
and  Wilson  1987).  This  method  al- 
lowed variables  with  different  scales 
of  measurement  to  be  included  to- 
gether without  further  scaling  or 
weighting.  Seasonal  estimates  of  each 
of  the  PCA-derived  demographic 
variables  and  their  variability  esti- 
mates were  regressed  separately  on 
each  of  the  PCA-derived  microhabi- 
tat  variables  and  the  index  of  hetero- 
geneity. 


inferences 

Statistical  tests  which  were  signifi- 
cant at  P<0.05  are  qualitatively  sum- 
marized in  table  6.  Grid  means  of  the 
first  three  demographic  PCs  revealed 
three  demographic  groups.  Grids  1 
and  5  were  located  farthest  from  any 
adjacent  grid  in  three-dimensional 
space,  whereas  grids  2, 3,  and  4  were 
clustered  more  tightly  together  with 
respect  to  demographic  structure 
(table  7).  Grid  1  was  characterized  by 
low  density  and  survival,  a  low  pro- 
portion of  females,  low  breeding  in- 
tensity, and  high  demographic  vari- 
ability. Grid  5  was  characterized  by 
low  density  and  survival,  a  high  pro- 
portion of  females,  moderate  breed- 
ing intensity,  and  high  demographic 
variability.  Grids  2, 3,  and  4  were 
characterized  by  high  density  and 
survival,  low  to  moderate  proportion 
of  females,  moderate  to  high  breed- 
ing intensity,  and  low  demographic 
variability.  Two  low-density  groups 
(represented  by  grid  1  and  grid  5) 
and  one  high-density  group  (repre- 
sented by  grids  1, 2,  and  3)  therefore 


Variable  Comment 


Density 

Adu't  male  body  mass 
Observed  range  length 

Proporfion  male 


were  evident.  The  low-density 
groups  were  more  variable  in  terms 
of  each  of  the  demographic  PCs  and 
in  overall  demographic  structure.  In 
general,  density,  survival,  and  breed- 
ing activity  increased  along  gradients 
of  increasing  woodiness  or  decreas- 
ing herbaceousness,  whereas  demo- 
graphic variability  decreased  along 
these  gradients  (table  8). 

SYNTHESIS 

I  found  microhabitat  structure  to  be  a 
potentially  important  force  in  organ- 
izing small  mammal  fX)pulations, 
particularly  in  relation  to  associa- 
tions and  densities.  Small  mammals 
generally  were  associated  with  par-  ! 
ticular  microhabitats,  as  revealed  by  \ 
analysis  of  single  trap  stations.  How-  j 
ever,  associations  often  differed  be-  | 
tween  control  and  experimental  \ 
grids.  I  suggest  that  the  small  mam-  j 
mals  I  studied  selected  specific  mi-  i 
crohabitats  and  were  opportunistic  in  } 
their  responses  to  habitat  not  occu- 
pied by  other  individuals  (as  on  the  | 

i 
i 


Adult  breeding  activity 


Young  breeding  activity 
Adult  residence 

Overwinter  residence 
Adult  survival 


Grids  1  and  5  had  lower  densities  than  grids  2. 

3,  and  4. 

No  differences. 

Males  had  a  greater  ORL  than  females  on  grid 

2. 

Grids  2. 3.  and  4  had  a  higher  proportion  of 
adult  males.  Grid  3  had  a  higher  proportion  of 
adult  males  than  females. 
Grids  1  and  5  had  a  lower  proportion  of  males 
breeding  than  grids  2, 3.and  4.  Grid  1  had  a 
lower  proportion  of  females  breeding  than  did 
the  other  grids.  A  higher  proportion  of  females 
was  breeding  on  grids  1,3,  and  4  than  were 
males. 

No  differences. 

Grids  1  and  5  had  lower  residence  rates  of 
adult  males  than  grids  2, 3,  and  4, 
No  differences. 

Males  on  grids  1  and  5  had  poorer  survival 
rates  than  on  grids  2,3,  and  4. 


Table  6.— Summary  of  differences  In  demography  of  Peromyscus  leucopus 
on  Greal  Island,  determined  from  monthly  trapping  data  on  five  grids. 


296 


experimental  grids).  Since  most  small 
mammals  that  I  studied  were  mi- 
crohabitat  selectors,  microhabitat 
structure  therefore  was  a  crucial  de- 
terminant of  local  community  com- 
position. Furthermore,  microhabitat 
structure  also  should  have  affected 
temporal  variability  of  community 
structure  since  populations  in  low- 
density  areas  were  more  variable. 

Affinities  of  each  small  mammal 
species  for  particular  microhabitats 
resulted  in  density-habitat  relation- 
ships when  averaged  over  a  larger 
sampling  area  (grids  or  plots).  Thus, 
small  mammal  densities  generally 
could  be  related  to  nnicrohabitat 
structure.  Survival  and  breeding  ac- 
tivity, which  generally  covary  with 
density,  also  could  be  related  to  mi- 
crohabitat structure  when  sampling 
areas  spanned  macrohabitat  bounda- 
ries. The  importance  of  microhabitat 
structure  in  affecting  other  demo- 


graphic characteristics  such  as  sex 
composition  and  age  structure  was 
not  as  pronounced.  Gradients  of  n\i- 
crohabitat  structure  can  be  envi- 
sioned as  comprising  an  environ- 
mental suitability  gradient,  with  the 
endpoints  being  uninhabitable  and 
optimal  (where  individual  fitness  is 
highest).  Demographic  characteris- 
tics then  vary  along  this  gradient  of 
suitability  and  along  other  gradients. 
The  gradient  of  suitability  is  com- 
posed of  factors  related  not  only  to 
habitat  structure  but  also  to  food  re- 
sources and  release  from  predation, 
competition,  and  parasitism.  Density 
alone  may  not  be  a  strong  correlate 
of  suitability  (Van  Home  1983),  but 
density  in  concert  with  survival  and 
breeding  activity  should  increase 
along  the  gradient  of  suitability.  By 
contrast,  demographic  variability 
should  decrease  along  this  gradient. 
Several  habitat  types  may  represent 


Table  7.— Distances  between  grid  means  of  the  first  three  principal  compo- 
nents derived  from  an  analysis  of  demographic  data  of  Peromyscus  leu- 
copus. 


Grid 


1 
2 
3 
4 


1.14 


L38 
0.46 


1.36 
0.45 
0.22 


1.09 
1.00 
1.26 


Table  8.— Relationships  between  Peromyscus  /et/copt/s  demographic  vari- 
alDles  arKl  habitat  variables  derived  from  PCA  on  Great  Island.  Signs  of  re- 
lationships are  as  Indicated  in  table  4. 


Demographic  variables  Habitat  variables 


Correlation 


Derisity  and  survival 
Adult  male  breeding 
activity 

Variability  of  der^ity 
and  survival 
Variability  of  male 
breeding  activity 
Variability  in  sex 
composition 


Herbaceous  ground-level  vegetation 
Herbaceous  ground-level  vegetation 
Woody  ground  vine  species  richness 
Herbaceous  cover  and  species  richness  + 

Herbaceous  ground-level  vegetation 

Woody  vegetation  der>sity  and  richness  — 
Herbaceous  cover  and  species  richness  -»-++ 
Canopy  cover  — 


similar  conditions  of  environmental 
suitability,  particularly  for  habitat 
generalists  such  as  Peromyscus  leu- 
copus.  Therefore,  it  may  be  difficult  to 
relate  demography  to  microhabitat 
structure  because  similar  demo- 
graphic structure  may  be  found  in 
different  habitats  (Adler  and  Wilson 
1987). 

Quantitative  measures  of  habitat 
heterogeneity  generally  were  unre- 
lated to  demographic  variables,  in 
contrast  to  the  mass  of  theory  pre- 
dicting that  heterogeneity  promotes 
population  stability  (e.g.,  den  Boer 
1968,  Levins  1969,  Smith  1972,  Mayn- 
ard  Smith  1974,  Steele  1974,  Tanner 
1975,  Stenseth  1977, 1980,  Lomnicki 
1978, 1980,  de  Jong  1979,  Hassell 
1980).  The  contrast  between  my  re- 
sults and  theoretical  predictions  may 
be  reconciled  by  introducing  scale. 
My  measures  of  heterogeneity  were 
at  the  microhabitat  level,  whereas 
many  models  have  implied  mac- 
rohabitat heterogeneity  so  that  or- 
ganisms may  disperse  into  a  patch 
and  establish  a  resident  population 
(e.g..  Levins  1969).  Increasing  the 
numb>er  of  such  patches  increases  the 
spatial  heterogeneity  of  an  area, 
which  then  promotes  population  sta- 
bility. I  suggest  that  microhabitat 
structure  will  affect  density  more 
than  it  will  other  demographic  char- 
acteristics, whereas  macrohabitat 
structure  and  heterogeneity  will  be 
more  important  in  stabilizing  popu- 
lations and  in  influencing  demo- 
graphic structure  (e.g.,  sex  composi- 
tion and  age  structure). 

My  conclusions  concerning  the 
importance  of  habitat  structure  in 
organizing  small  mammal  popula- 
tions and  communities  can  be  shown 
schematically  (fig.  2).  According  to 
this  scheme,  microhabitat  structure 
primarily  affects  habitat  selection, 
density,  and  density  variability  (since 
density  generally  is  related  inversely 
to  variability).  Macrohabitat  struc- 
ture primarily  affects  population  sta- 
bility (stability  being  enhanced  by 
macrohabitat  heterogeneity)  and 
demographic  structure.  Habitat  se- 


297 


lection  and  demography  then  deter- 
mine local  community  composition 
and  variability,  respectively.  While 
habitat  structure  ultimately  deter- 
mines community  composition,  it 
does  so  at  the  population  or  individ- 
ual level.  Therefore,  I  added  no  di- 
rect links  between  habitat  structure 
and  the  community  variables. 

Additional  links  may  be  added; 
factors  such  as  random  events,  com- 
petition, predation,  parasitism,  and 
infection  manifest  their  effects  at 
various  levels.  For  instance,  a  preda- 
tor may  selectively  feed  on  a  particu- 
lar species,  thereby  depressing  its 
density  and  affecting  community 
composition  and  structure.  Competi- 
tion between  species  also  may  affect 
species  densities  in  certain  small 
mammal  communities.  The  structur- 
ally simple  habitats  that  I  have  stud- 
ied generally  contain  an  abundant 
herbivore,  an  abundant  granivore,  a 
common  insectivore,  and  any  of  sev- 
eral rarer  omnivores.  These  poorly 
diversified  communities  are  quite 
different  from  other  systems  such  as 
deserts  or  tropical  forests  where 
communities  are  comprised  of  regu- 
larly structured  guilds  containing 
several  species.  Competition,  which 
apparently  is  important  in  structur- 
ing communities  in  other  areas  (e.g.. 
Brown  and  Bowers  1984),  should  not 
be  very  important.  The  opportunity 
for  competition  between  guilds 
would  be  expected  to  be  quite  low. 
The  occurrence  of  several  easily  stud- 
ied genera  with  interesting  life-his- 
tory traits  (e.g.,  Microtus,  Peromyscus, 
and  Tamias)  made  these  sites  ideal 
for  population-level  studies,  but  be- 
cause of  poorly  diversified  guilds  or 
even  guild  singularity  (only  one  spe- 
cies per  guild)  at  my  study  sites, 
these  same  areas  were  far  less  suit- 
able for  community-level  studies. 

Habitats  with  which  different  spe- 
cies of  small  mammals  are  associated 
are  well  known,  but  the  effects  of 
relevant  scales  of  habitat  structure 
are  only  now  becoming  apparent.  I 
suggest  that  future  studies  shift  from 
repetitious  descriptions  of  habitats 


with  which  well-studied  species  as- 
sociate to  innovative  experimental 
approaches  that  test  hypothesized 
effects  of  habitat  structure  on  popu- 
lation and  community  organization 
and  that  identify  relevant  scales  of 
such  structure. 


ACKNOWLEDGMENTS 

Bruce  Lund,  Wesley  N.  Tiffney,  Jr., 
and  the  Chace  family  granted  per- 
mission to  trap.  John  W.  Classer, 
Thomas  W.  Schoener,  and  Robert  H. 
Tamarin  read  an  earlier  draft  of  the 
manuscript.  Mark  L.  Wilson  shared 
in  much  of  the  field  work  on  Great 
Island,  and  Marita  Sheridan  shared 
in  field  work  on  Nantucket.  I  thank 
Robert  H.  Tamarin  for  providing  re- 
search supplies  and  facilities  at  Bos- 
ton University  while  I  conducted  the 
field  research.  I  also  thank  Thomas 
W.  Schoener  for  providing  facilities 
and  resources  at  the  University  of 
California  (Davis),  where  I  wrote  the 
manuscript.  This  research  was  sup- 
ported by  NSF  grants  DEB-8103483 
to  Robert  H.  Tamarin  and  BSR- 
8700130  to  G.H.A.  and  from  Sigma  Xi 
and  the  Boston  University  chapter  of 
Sigma  Xi  to  GHA.  This  study  is  a 


contribution  of  the  University  of 
Massachusetts  Nantucket  Field 
Station. 


LITERATURE  CITED 

Adler,  Gregory  H.  1985.  Habitat  se- 
lection and  species  interactions:  an 
experimental  analysis  with  small 
mammal  populations.  Oikos 
45:380-390. 

Adler,  Gregory  H.  1987.  Influence  of 
habitat  structure  on  demography 
of  two  rodent  species  in  eastern 
Massachusetts.  Canadian  Journal 
of  Zoology  65:903-912. 

Adler,  Gregory  H.,  and  Mark  L. 
Wilson.  1987.  Demography  of  a 
habitat  generalist,  the  white- 
footed  mouse,  in  a  heterogeneous 
environment.  Ecology  68:1785- 
1796. 

August,  Peter  V.  1983.  The  role  of 
habitat  complexity  and  heteroge- 
neity in  structuring  tropical  mam- 
mal communities.  Ecology 
66:1495-1507. 

Brown,  James  H.,  and  Michael  A. 
Bowers.  1984.  Patterns  and  proc- 
esses in  three  guilds  of  terrestrial 
vertebrates,  p.  282-296.  In  Ecologi- 
cal communities:  conceptual  issues 


Density 


Density 
Variability 


Macrohabitat 
Structure 


Microhabitat 
Structure 


Habitat 
Selection 


Population 
Performance 


z 


Demographic 
Structure 


Commimity 
Composition 


Community 
Structure 


Temporal  Variability 
of  Commvmities 


Figure  2.— Conceptual  scheme  showing  the  effects  of  habitat  structure  on  population  and 
community  processes. 


298 


and  the  evidence.  Donald  R. 
Strong,  Jr.,  Daniel  Simberloff, 
Lawrence  G.  Abele,  and  Anne  B. 
Thistle,  editors.  Princeton  Univer- 
sity Press,  Princeton,  NJ. 

de  Jong,  G.  1979.  The  influence  of  the 
distribution  of  juveniles  over 
patches  of  food  on  the  dynamics 
of  a  population.  Netherlands  Jour- 
nal of  Zoology  29:33-51. 

den  Boer,  P.  J.  1968.  Spreading  of  risk 
and  stabilization  of  animal  num- 
bers and  the  heterogeneity  of  the 
environment:  the  problem  of  the 
persistence  of  sparse  populations, 
p.  77-97.  In  Dynamics  of  Popula- 
tions. Proceedings  of  the  Ad- 
vanced Study  Institute  on  Dynam- 
ics and  Numbers  of  Populations, 
Oosterbeek,  7-18  September  1970. 
P.J.  den  Boer  and  G.  R.  Gradwell, 
editors.  Centre  for  Agricultural 
Publishing  and  Documentation 
(PUDOC),  Wageningen,  The  Neth- 
erlands. 

Framstad,  E.,  S.  Engen,  and  N. 
Stenseth.  1985.  Regression  analy- 
sis, residual  analysis,  and  missing 
variables  in  regression  models. 
Oikos  44:319-323. 

Hassell,  Michael  P.  1980.  Some  con- 
sequences of  habitat  heterogeneity 
for  population  dynamics.  Oikos 
35:150-160. 

Levin,  Simon  A.  1976.  Population 
dynamic  models  in  heterogeneous 
environments.  Annual  Review  of 
Ecology  and  Systematics  7:287- 
310. 

Levins,  Richard.  1969.  Some  demo- 
graphic and  genetic  consequences 
of  environmental  heterogeneity  for 
biological  control.  Bulletin  of  the 
Entomological  Society  of  America 
15:237-240. 

Lomnicki,  Adam.  1978.  Individual 
differences  between  animals  and 
the  natural  regulation  of  their 
numbers.  Journal  of  Animal  Ecol- 
ogy 47:461-475. 

Lomnicki,  Adam.  1980.  Regulation  of 
population  density  due  to  individ- 
ual differences  and  patchy  envi- 
ronment. Oikos  35:185-193. 


Maynard  Smith,  John.  1974.  Models 
in  ecology.  Cambridge  University 
Press,  Cambridge,  United  King- 
dom. 

Schaffer,  William  M.,  and  Robert  H. 
Tamarin.  1973.  Changing  repro- 
ductive rates  and  population 
cycles  in  lemmings  and  voles. 
EvoluHon  27:111-124. 

Smith,  Frank  E.  1972.  Spatial  hetero- 
geneity, stability,  and  diversity  in 
ecosystems.  Transactions  of  the 
Connecticut  Academy  of  Arts  and 
Sciences  44:309-335. 

Steele,  John  H.  1974.  Spatial  hetero- 
geneity and  population  stability. 
Nature  248:83. 

Stenseth,  N.  C.  1977.  On  the  impor- 
tance of  spatio-temporal  heteroge- 
neity for  the  population  dynamics 
of  rodents:  towards  a  theoretical 
foundation  of  rodent  control. 
Oikos  29:545-552. 

Stenseth,  N.  C.  1980.  Spatial  hetero- 
geneity and  population  stability: 
some  evolutionary  consequences. 
Oikos  35:165-184. 

Stickel,  Lucille  F.  1954.  A  comparison 
of  certain  methods  of  measuring 
ranges  of  small  mammals.  Journal 
of  Mammalogy  35:1-15. 

Svardson,  Gunnar.  1949.  Competi- 
tion and  habitat  selection  in  birds. 
Oikos  1:157-174. 

Tanner,  James  T.  1975.  The  stability 
and  the  intrinsic  growth  rates  of 
prey  and  predator  populations. 
Ecology  56:855-867. 

Van  Horne,  Beatrice.  1982.  Niches  of 
adult  and  juvenile  deer  mice  (Pero- 
myscus  maniculatus)  in  serai  stages 
of  coniferous  forest.  Ecology 
63:992-1003. 

Van  Home,  Beatrice.  1983.  Density  as 
a  misleading  indicator  of  habitat 
quality.  Journal  of  Wildlife  Man- 
agement 47:893-901. 

Wiens,  John  A.  1976.  Population  re- 
sponses to  patchy  environments. 
Annual  Review  of  Ecology  and 
Systematics  7:81-120. 


Wiens,  John  A,  John  F.  Addicott,  Ted 
J.  Case,  and  Jared  Diamond.  1986. 
Overview:  the  importance  of  spa- 
tial and  temporal  scale  in  ecology 
investigations,  p.  145-153.  In  Com- 
munity ecology.  Jared  Diamond 
and  Ted.  J.  Case,  editors.  Harper 
and  Row  Publishers,  New  York, 
NY. 


299 


Microhabitat  as  a  Template 
tor  the  Organization  of  a 
Desert  Rodent  Community^ 

Michael  A.  Bowers^  and  Christine  A. 
Flanagan^ 


Abstract.— We  used  20  0.25-ha  fenced  plots  to 
experimentally  study  microhabitat  use  by  1 1  desert 
rodent  species  in  southeastern  Arizona.  Removal  of 
the  largest  granivore,  Dipodomys  spectabilis. 
produced  the  most  pervasive  shifts  in  the  use  of 
microhobitats  while  adding  food  or  removing  ants 
produced  few  responses.  These  results  support  the 
idea  that  this  community  is  organized  around 
competitive  interactions  involving  aggression, 
preemption,  and  relegation. 


It  is  generally  believed  that  species 
have  different  fitnesses  in  different 
habitats,  that  most  communities  are 
comprised  of  sufficient  habitat  vari- 
ation over  which  fitness  differentials 
can  be  expressed,  and  that  species 
select  habitats  that  maximize  their 
fitness  (e.g..  Levins  1962,  Schoener 
1971).  The  manner  and  degree  to 
which  species  respond  to  the  habitat 
template  involves  elements  of  selec- 
tion in  its  purest  form  (i.e.,  choice), 
relegation,  and  correlation. 

At  the  community  level  rarely  do 
species  occupy  habitats  in  an  ideal  or 
cost-free  fashion.  By  occupying  space 
or  using  resources  in  a  habitat  spe- 
cific manner  organisms  alter  habitat 
suitability  and  thereby  change  the 
basis  over  which  habitats  are  selected 
(Fretwell  and  Lucas  1970).  Species 
that  use  limited  resources  in  an  effi- 
cient manner  or  are  behaviorally 
dominant  can  monof)olize  the  choic- 
est habitats  and  relegate,  directly  or 
indirectly,  subordinate  or  competi- 
tively inferior  species  to  secondary 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff. AZ,  July  19-21.  1988.) 

'Michiael  A.  Bowers  is  Assistant  Professor 
in  ttie  Department  of  Environmental  Sci- 
ences and  Researchi  Coordinator  at  the 
Blandy  Experimental  Farm.  University  of  Vir- 
ginia. Clark  Hall.  Charlottesville,  VA  22903. 

^Christine  A.  Flanagan  is  Assistant  Cura- 
tor of  the  Orland  E.  White  Arboretum.  Uni- 
versity of  Virginia.  P.O.  Box  175.  Boyce.  VA 
22620. 


habitats  (Col well  and  Fuentes  1973, 
Bowers  et  al.  1987).  If  the  capture 
success  rates  of  predatory  species 
varies  among  habitats  this  can  also 
affect  the  absolute  and  relative  fit- 
ness of  prey  species  and  their  distri- 
bution among  habitats  (Kotler  1984, 
Bowers  1988). 

Marked  patterns  of  habitat  occu- 
pancy and  segregation  are  often  cited 
as  evidence  that  ecological  communi- 
ties are  structured.  The  general  pat- 
tern is  that  some  (if  not  most)  species 
in  a  community  utilize  habitats  dif- 
ferently from  random  and  differently 
than  if  each  species  occurred  by  it- 
self. Observational  and  manipulative 
experiments  have  shown  that  dy- 
namical properties  of  populations 
(including  patterns  of  growth,  demo- 
graphics, and  interaction)  often  be- 
come expressed  as  spatial  phenom- 
ena, thereby  establishing  a  connec- 
tion between  habitat  occupancy  and 
population  dynamics  (see  Connor 
and  Bowers  1987). 

Many  communities  are  comprised 
of  an  array  of  microhabitats  which 
represent  discrete,  exploitable  re- 
sources which  occur  with  suffident 
variability  so  as  to  be  parti tionable 
among  spedes.  The  availability  and 
distribution  of  microhabitats  have 
been  shown  to  limit  the  growth  and 
density  of  many  populations  and, 
thereby  provide  an  ecologically  rele- 
vant and  readily  identifiable  context 
over  which  species  interactions  and 
population  growth  can  be  studied 
(Price  1978,  Rosenzweig  1981). 


Desert  rodents  have  long  pro- 
vided ecologists  with  a  model  system 
for  examining  the  role  of  microhabi- 
tat in  structuring  communities.  The 
basic  pattern  throughout  the  major 
North  American  deserts  is  that  lo- 
cally co-occurring  species  character- 
istically forage  in  microhabitats  that 
are  structurally  distinctive  with  re- 
spect to  perennial  vegetation  and  soil 
type  (Rosenzweig  and  Winakur  1969, 
Price  1978;  for  reviews  see  Brown  et 
al.  1979,  Munger  et  al.  1983,  Price 
and  Brown  1983). 

Three  mechanisms,  alone  or  in 
combination,  apparently  account  for 
the  general  pattern.  First,  because  of 
differences  in  body  size,  mode  of  lo- 
comotion and  behavior,  rodents  dif- 
fer in  their  abilities  to  exploit  particu- 
lar distributions  of  food  (i.e.,  seed) 
resources  that  are  created  by  struc- 
tural features  of  the  microhabitat 
(Bowers  1982,  Harris  1984,  Price 
1983,  Reichman  1981).  Second,  ro- 
dents may  differ  in  their  ability  to 
escape  visually  oriented  predators  so 
that  the  most  susceptible  rodents  are 
limited  to  the  safest  microhabitats 
(i.e.,  under  vegetative  cover)  while 
more  vagile  rodents  show  more  un- 
restricted use  of  alternate  microsites 
(Kotler  1984).  Third,  the  ability  of 
some  species  to  aggressively  defend 
areas  from  other  rodents  may  be 
high  in  some  habitats  and  low  in  oth- 
ers resulting  in  habitat  dependent 
segregation  involving  domination/ 
relegarion  (Hutto  1978,  Frye  1983, 
Bowers  et  al.  1987). 


300 


Desert  rodent  populations  are  re- 
markable in  their  ability  to  respond 
to  short-term  changes  in  the  abun- 
dance and  distribution  of  food  re- 
sources; primarily  seeds.  Some  of  the 
more  marked  responses  involve 
changes  in  use  of  microhabitats.  For 
example,  enriching  microhabitats 
with  supplemental  seeds  increases 
the  use  of  these  by  desert  rodents 
(Harris  1984,  Kotler  1984,  Price  and 
Waser  1985).  Such  shifts  are  particu- 
larly noteworthy  for  microhabitats 
where  the  risk  of  being  preyed  upon 
is  high,  and  suggests  that  both  ener- 
getic profits  and  predatory  risk  play 
a  role  in  determining  which  mi- 
crosites  are  used  (Hay  and  Fuller 
1981,  Price  and  Waser  1985,  Bowers 
1988).  Food  availability  also  can 
change  the  manner  in  which  some 
rodent  species  interact:  from  com- 
1    petitive  exploitative  interactions  un- 
I    der  low  levels  of  food  to  aggressive 
interference  interactions  under  high 
levels  of  food  (e.g.,  Congdon  1974). 
i       In  complex  communities  mi- 
'    crohabitat  use  originates  with  prefer- 
ences of  individual  species  for  certain 
microhabitats,  but  these  basic  re- 
||    sponses  may  become  altered,  directly 

or  indirectly,  by  interactions  with 
I    other  species.  Moreover,  at  the  com- 
munity level  it  is  not  clear  how 
changes  in  the  resource  base  are 
manifest  in  patterns  of  spatial  usage. 
Some  important  questions  are:  Does 
interspecific  competition  become 
more  or  less  important  with  increas- 
ing food  availability?  Does  the  mode 
of  competition  change?  How  does 
food  availability  change  the  relative 
roles  of  preference  and  relegation  in 
determining  habitat  occupancy? 
Thus,  detailing  the  interplay  between 
population  and  community-level  re- 
sponses to  changes  in  resource 
availability  should  reveal  much 
about  the  processes  influencing  mi- 
crohabitat  use  and,  thereby,  the  fac- 
tors responsible  for  the  organization 
of  these  communities. 

In  this  paper  we  describe  patterns 
of  microhabitat  use  of  11  Chihua- 
huan  Desert  rodents  over  a  span  of 


more  than  eight  years.  We  exp)eri- 
mentally  manipulated  both  species 
composition  and  food  supply  and 
measured  resulting  shifts  in  mi- 
crohabitat use.  By  detailing  shifts  in 
microhabitat  use  in  response  to  our 
manipulations  we  were  able  to  iden- 
tify the  most  important  interactions 
among  species,  estimate  their  relative 
strengths,  and  say  something  about 
the  mode  of  interaction  promoting 
the  shifts. 

Our  results  suggest  that  the  or- 
ganization of  this  community  re- 
volves more  around  differences  in 
the  ability  of  species  to  occupy  and 
defend  certain  key  microhabitats 
than  changes  in  food  availability. 

Study  Site  and  Methods 

The  present  paper  details  changes  in 
microhabitat  use  in  response  to  long- 
term  experimental  manipulation  of 
rodent  composition  and  food  supply. 
Our  study  site  was  located  at  an  ele- 
vation of  1330  m  in  a  relatively 
homogeneous  desert  shrub  habitat 
on  the  Cave  Creek  Bajada  6.5  km  east 
and  2  km  north  of  Portal,  in  Arizona, 
USA.  Manipulations  were  performed 
in  twenty  0.25-ha  plots.  Each  plot 
was  fenced  with  0.64-cm  mesh  hard- 
ware cloth,  extending  0.7-m  above 
and  buried  0.2-m  below  ground.  In 
addition  to  an  unmanipulated  fenced 
control  (see  below),  the  remaining 
treatments  consisted  of  two  general 
classes:  treatments  where  one  or 
more  rodent  species  were  removed, 
potentially  changing  both  food 
availability  and  the  ix)tential  for  di- 
rect behavioral  interactions;  and  food 
alteration  treatments  where  supple- 
mental millet  seeds  were  added  at  a 
rate  of  96  kg  per  year  or  seed-eating 
ants  were  removed.  Experimental 
treatments  were  assigned  to  plots  at 
random. 

Fourteen  rodent  species  of  which 
11  were  commonly  captured,  inhab- 
ited the  study  site,  all  except  those 
mentioned  above  had  equal  access  to 
all  plots  (fig.  1).  Because  of  problems 


in  consistently  identifying  the  two 
Onychomys  species  (as  either  O.  tor- 
ridus  or  O.  leucogaster)  we  group 
these  together  under  the  designation, 
Onychomys  spp. 

Sixteen  equally-spaced  gates  in 
each  plot  allowed  the  selective  exclu- 
sion of  rodent  species  above  a  thresh- 
old body  size  while  allowing  all 
other  species  access.  Access  gates 
varied  in  size  among  the  treatments. 
Large  gates  (3.7  x  5.7-cm)  allowed  all 
rodent  species  free  access  to  control 
(2  plots),  ant  removal  (4  plots),  and 
the  seed  addition  plots  (8  plots;  see 
below);  medium-sized  gates  (2.6  x 
3.0-cm)  were  used  to  exclude  only 
the  largest  granivore,  Dipodomys 
spectabilis  (2  plots);  and  small  gates 
(1.9  x  1.9-cm)  were  used  to  exclude 
all  Dipodomys  species  (4  plots).  The 
seed  addition  treatments  included 
six  plots  where  supplemental  seeds 
were  applied  in  12  monthly  applica- 
tions (hereafter  referred  to  as  "con- 
stant seed  additions");  two  plots  re- 
ceived the  total  allotment  of  seeds  in 
three  applications  during  the  fall 
(September-November;  referred  to  as 
"pulsed  seed  additions").  Seeds  were 
uniformly  scattered  by  hand  over 
each  plot. 

It  was  estimated  from  productiv- 
ity measurements  at  the  site  that  the 
addition  of  96  kg  of  seeds  per  year 
should  have  approximately  doubled 
the  total  biomass  of  seeds  produced 
annually  (our  estimate  of  seed  pro- 
duction was  ca.  400  kg/ha/yr).  The 
constant  seed  additions  included  two 
plots  where  whole  millet  (Panicum 
miliaceam)  was  added  (mean  seed 
mass  =  6  mg);  two  plots  where 
cracked  millet  was  added  (mean 
mass  =  1  mg);  and  two  plots  where 
an  equal  mixture  of  whole  and 
cracked  millet  was  added.  The 
pulsed  seed  treatment  was  designed 
to  represent  a  doubling  of  the  seed 
production  of  summer  annual  plants, 
a  particularly  important  food  source 
for  the  rodents  in  this  community 
(Davidson  et  al.  1985).  Brown  and 
Munger  (1985)  found  no  differences 
in  responses  of  rodents  to  addition  of 


301 


seeds  of  different  size,  so  the  four 
constant  seed  addition  treatments 
will  be  lumped  together  here  (6 
plots). 

Rodents  were  censused  monthly 
during  the  week  of  the  new  moon 
(moonlight  has  been  shown  to  effect 
the  microhabitats  used  by  desert  ro- 
dents; Bowers  1988)  using  live  traps 
placed  in  each  plot  in  7  x  7  grids  with 
6.5-m  between  trap  stations.  Traps 
were  baited  with  millet  and  opened 
for  one  night  per  month  with  plot 
gates  closed  so  that  only  plot  resi- 
dents would  be  captured.  For  more 
details  concerning  the  experimental 
design,  see  Bowers  et  al.  1987,  Brown 
and  Munger  (1985),  and  Brown  et  al. 
(1986). 

Following  the  lead  of  many  previ- 
ous studies  on  desert  rodent  commu- 
nities we  used  the  percent  cover  of 
perennial  plants  to  characterize  the 
microhabitat  at  each  of  the  980  trap 
stations.  Percent  cover  within  a  2-m 


radius  of  each  trap  station  was  meas- 
ured by  ocular  estimation  using  ref- 
erence disks  of  known  percent  cover- 
age. Cover  was  measured  in  1978, 
1981,  and  1983.  There  was  no  signifi- 
cant changes  in  perennial  cover  over 
this  five  year  period  (Mann-Whitney 
U-test;  P  >  0.05),  so  we  used  data 
from  1983  to  characterize  microhabi- 
tats. Table  1  summarizes  vegetation 
cover  data  over  the  entire  study  site. 

Fence  installation  was  completed 
in  June,  1977;  premanipulative  trap- 
ping was  conducted  from  July-Sep- 
tember, 1977;  and  the  manipulations 
were  initiated  in  October,  1977.  We 
restrict  our  analyses  to  include  post- 
manipulation  data  compiled  from 
October  1977  to  December  1984  and 
to  only  those  20  plots  to  which  ro- 
dents had  access. 

Analyses  were  designed  to  answer 
two  questions:  first,  what  are  the  pat- 
terns of  species  associations  occur- 
ring at  the  community  level;  and  sec- 


ond, what  role  does  microhabitat 
play  in  the  distribution  of  individual 
species.  In  this  study  patterns  of  as- 
sociation (including  the  association 
of  rodent  species  with  each  other  and 
with  structural  microhabitats)  are 
analyzed  at  the  level  of  individual 
trap  stations  (980  total).  Hence,  we 

^able  l,— Mean  percent  cover  of 
the  seven  most  common  perennial 
plant  species  over  the  study  site 
(standard  deviations  In  parenthe- 
ses). 


Species 


%  cover 


Acacia  constricfa  1.5(4.4) 

Ephedra  forreyana  2.7  (3.6) 

Florensia  cemva  2  7  (4.7) 

Gufierrezla  lucida  3.3(3.9) 

Lycium  andersonii  0.4  (2,2) 

Mimosa  biuncifera  0.2  (1 .4) 

Prosopis  Juliflora  0.6(4.5) 

i^otal  cover  (all  species)  127  (3.7)  ^ 


Re  i  thro- 


Ueotoma 


Olnedomyt       Dipodomft      Dipodomys    Perognathus  PtrognalhuM    PeromyMcus    Peiomyscut  donto»,yt 

'  penicillatus  mtnlculmtut    •ramleua        megalotis  albigula 


Mpaclablllt       mmrrlami  ordU 


120g 


Control  184  529  51 

(2  plots) 

Ant  removal  539  787  248 

(4  plots) 

Seed-pulse  278        390  116 

(2  plots) 

Seed-constant  821  1005  217 
(6  plots) 

Dipodomys-  0  0  0 

removal 

(4  plots) 

D.  spectabills-         0  465  50 

removal 

(2  plots) 


^  ^  ^  ^ 


45g       52g  17g 


25 


33 


27 


42 


49 


7g 


25 


82 


22 


121 


243 


13 


8 


9 

10 
74 

21 


Figure  1  .—Rodent  species  on  study  site  (Including  their  body  sizes)  along  with  their  capture 
frequencies  In  each  of  ttie  experimental  treatments.  Included  are  the  number  of  plots  in 
each  treatment. 


10 


22 


75 


65 


42 


24g      25g      11g  156g 


24 
48 
21 
32 
175 

33 


60 


94 


56 


145 


61 


66 


Onyehomyt 
tpp. 


29g 


49 

106 

47 

132 

104 

53 


302 


were  interested  in  measuring  re- 
sponses of  rodents  to  microhabitat 
variation  occurring  at  a  scale  of  a  me- 
ter or  two.  However,  we  acknowl- 
edge that  habitats  may  also  be  se- 
lected at  larger  spatial  scales  (Morris 
1987).  For  example,  rodents  may  also 
select  areas  on  the  basis  of  mi- 
crohabitat composites  (e.g.,  at  the 
level  of  the  home  range)  which  might 
be  best  examined  by  considering 
structural  microhabitats  over  trap 
station  aggregates.  However,  there  is 
reason  to  believe  that  even  if  selec- 
tion does  occur  at  these  larger  scales 
it  is  still  oriented  towards  excluding 
or  including  certain  key  microhabi- 
tats. Hence,  we  were  confident  our 
analyses  would  detect  patterns  at 
both  scales. 

Indices  of  species  association  were 
calculated  by  using  the  frequency 
that  species  were  captured  at  the 
same  trap  station  using  trap  data  for 
the  eight  year  period.  This  involved 
several  steps:  (i)  tabulating  the  pro- 
portion of  trap  stations  where  each 
species  was  captured  over  the  eight 
year  study;  (ii)  tallying  the  number  of 
trap  stations  where  each  pair  of  spe- 
cies co-occurred;  and  (iii)  comparing 
the  observed  frequency  of  co-cap- 
tures to  that  expected  if  species  cap- 
tures were  distributed  independently 
and  randomly  among  trap  stations. 
The  expected  frequency  of  species 
co-capture  was  calculated  by  multi- 
plying together  the  proportion  of  sta- 
tions capturing  species  individually 
to  generate  a  probability  of  joint  oc- 
currence. A  modified  chi-square  sta- 
tistic, including  the  sign  of  associa- 
tion, was  then  used  as  an  index  of 
association:  i.e.,  a  measure  of  the  dif- 
ference between  the  observed  and 
expected  values.  The  null  hypothesis 
was  that  there  would  be  an  equal 
number  of  positive  and  negative  as- 
sociations with  less  than  5%  of  the 
association  values  being  statistically 
significant  at  a  P  =  0.05. 

The  analysis  described  above  can 
also  be  used  to  examine  the  associa- 
tion of  all  species  in  the  community 
at  individual  trap  stations.  Specifi- 


cally, instead  of  asking  how  fre- 
quently species  pairs  associate  we 
can  use  the  maximum  likelihood  esti- 
mation technique  to  estimate  how 
many  trap  stations  should  have  cap- 
tured 0, 1,  2, . .  n  species  (where  n  is 
the  number  of  species  in  the  commu- 
nity) over  the  eight  year  period.  As 
in  the  above  analysis,  this  uses  the 
proportion  of  stations  capturing  each 
species,  multiplies  these  together  in 
all  possible  combinations  that  might 
produce  co-captures  of  from  0  to  n 
species,  and  sums  these  probabilities 
for  each  number  of  possible  co-cap- 
tures to  give  an  expected  distribution 
over  the  population  of  trap  stations. 
The  null  expectation  here  is  that  spe- 
cies captures  are  independently  and 
randomly  distributed  among  trap 
stations. 

Analyses  were  also  performed  to 
examine  the  individualistic  responses 
of  species  to  variation  in  microhabi- 
tat and,  particularly,  how  these 
change  when  manipulations  are  ap- 
plied at  the  level  of  the  entire  com- 
munity. We  used  percent  cover  by 
perennial  plants  at  trap  stations  as  a 
general  descriptor  of  microhabitat 
type.  Our  goal  was  not  to  use  a  series 
of  variables  to  explain  the  largest 
amount  of  variation  in  nnicrohabitats 
where  species  were  captured  but 
rather  we  were  interested  in  identify- 
ing a  major  resource  axis  over  which 
both  species  distributions  and  com- 
munity-level responses  could  be  ana- 
lyzed. Past  work  justified  using  cover 
as  such  a  variable  (Brown  et  al.  1979, 
Munger  et  al.  1983,  Price  and  Brown 
1983).  Our  scheme  of  categorizing 
microhabitats  was  simple:  trap  sta- 
tions were  grouped  into  those  with 
greater-than-median  and  those  with 
less-than-median  cover.  This  was 
performed  separately  for  stations  in 
each  of  the  six  treatments.  Hence, 
each  microhabitat  category  was  rep- 
resented by  an  equal  number  of  trap 
stations  in  each  treatment  type.  The 
null  hypothesis  for  analyzing  the  trap 
data  was  if  rodents  use  microhabitats 
randomly,  and  without  regard  to 
vegetative  cover,  they  should  be 


trapped  in  equal  frequencies  at  sta- 
tions in  the  two  microhabitat  catego- 
ries. Avoidance  or  preference  for  mi- 
crohabitats would  be  indicated  by  a 
disproportionate  number  of  captures 
in  one  or  the  other  category. 

We  were  also  interested  in  exam- 
ining (1)  the  microhabitat  affinities  of 
species  in  the  different  treatments, 
and  (2)  shifts  in  types  of  microhabi- 
tats used  by  the  same  species  over 
the  different  seasons  of  the  year  and 
over  the  six  experimental  treatments. 
In  the  first  case  we  used  the  Fisher 
Exact  Probability  procedure  in  a  two- 
tailed  test  of  the  null  hypothesis  that 
captures  in  the  two  microhabitats  did 
not  differ  from  a  1:1  ratio  (Siegel 
1956);  in  the  second  we  subjected  the 
proportion  of  species'  captures  in  the 
two  microhabitats  to  a  2-way 
ANOVA  where  season  and  treatment 
represented  treatment  factors. 

Results 

Results  are  based  on  8,019  captures 
of  the  11  most  common  rodent  spe- 
cies. Figure  1  lists  the  frequency  of 
capture  for  each  species  in  the  six 
treatments  summed  over  the  eight 
year  study  period. 

Community-Wide  Patterns  of 
Microtiabitot  Use 

What  are  the  patterns  of  species  asso- 
ciation at  the  level  of  the  entire  com- 
munity? In  answering  this  we  consid- 
ered the  frequency  that  species  were 
captured  at  the  same  trap  station.  We 
performed  two  tests.  We  first  calcu- 
lated species  associations  for  all  pos- 
sible pairings  of  the  11  species  occur- 
ring in  plots  with  intact  rodent  as- 
semblages (i.e.,  those  14  plots  with 
large  gates)  resulting  in  a  total  of  45 
values  of  species  association.  Plotting 
all  association  values  show  that  most 
species  in  this  community  are  cap- 
tured at  the  same  station  much  less 
frequently  than  predicted  by  chance 
(fig.  2;  the  null  hypothesis  is  that 


303 


there  would  be  an  equal  number  of 
positive  and  negative  associations 
and  that  only  5%  of  these  would  be 
statistically  significant  at  P  <  0.05). 
The  deviation  from  what  is  expected 
is  particularly  striking  considering 
that  27  of  the  association  values  ex- 
ceeded the  cutoff  value  for  signifi- 
cance (3.84  for  p  <0.05  and  d.f.=l) 
and  all  of  these  were  in  the  direction 
of  negative  species  associations;  there 
was  not  a  single  significant  positive 
association.  This  suggests  a  high 
level  of  organization  revolves  around 
the  spatial  segregation  of  species. 

Among  those  factors  that  could  be 
responsible  for  this  marked  segrega- 
tion are  unique  habitat  preferences  of 
species.  These  could  work  alone  or  in 
conjunction  with  habitat  segregation 
that  is  mediated  through  interactions 
with  other  rodent  species.  The  design 
of  our  experiment  allows  a  further 
examination  of  the  role  of  species  in- 
teractions in  producing  the  pattern. 
Specifically,  our  experiment  includes 
treatments  with  an  intact  rodent  as- 
semblage (14  plots;  686  stations)  as 
well  as  treatments  where  either  D, 
spectabilis  (2  plots;  98  stations)  or  all 
Dipodomys  (4  plots;  196  stations) 
were  selectively  removed  and  ex- 
cluded. Because  previous  studies 
have  shown  Dipodomys  (and 
especially  D.  spectabilis)  to  be  behav- 
iorally  dominant  over  many  of  the 
species  they  co-occur  with  (Blaustein 
and  Riser  1974,  Frye  1983,  Bowers  et 
al.  1987)  there  is  reason  to  think  that 
by  their  removal  the  patterns  of  asso- 
ciation of  the  remaining  species  may 
change.  To  evaluate  this  possibility 
we  restricted  the  analyses  to  include 
just  those  eight  non-Dipodomys  spe- 
cies that  occurred  in  all  three  treat- 
ments (number  of  pairwise  associa- 
tion values  for  this  group  =  21).  The 
degree  to  which  these  species  were 
associated  with  each  other  at  trap 
stations  in  each  of  the  three  treat- 
ments was  calculated  as  before,  and 
then  compared  across  the  three  treat- 
ments (fig.  3).  The  results  show  that 
removing  either  all  Dipodomys  or  just 
D.  spectabilis  significantly  alters  the 


degree  to  which  the  remaining  spe- 
cies are  spatially  segregated  (X^  = 
17.33,  df  =  2;  P  <  0.000).  While  the 
trend  is  clearly  towards  more  posi- 
tive and  fewer  negative  associations 
when  competitors  are  removed,  most 
of  the  species  are  still  negatively  as- 
sociated with  each  other. 

The  previous  analysis  can  be  ex- 
tended from  the  two-species  case  to 
one  considering  the  association  of  all 
11  species.  Specifically,  instead  of 
asking  how  frequently  species  pairs 
associate  we  can  use  the  maximum 


likelihood  estimation  technique  to 
estimate  how  many  trap  stations 
should  have  captured  0, 1,  2  ...  11 
species  over  the  eight  year  period. 
Comparing  the  actual  number  of  spe- 
cies captured  per  station  with  that 
expected  (fig.  4)  shows  that  the  ob- 
served distribution  is  shifted  to  the 
left  of  that  expected  (significantly  dif- 
ferent at  P  <  0.05  using  Kolmogorov- 
Smimov  one  sample  test),  that  there 
are  significant  differences  in  the 
mode  of  species  co-captured  per  sta- 
tion (expected=4;  observed=3),  and 


18  T 


Chi  Square 

Figure  2.— Estimates  of  species  associations  for  plots  witti  intact  rodent  assemblages  (i.e., 
thiose  with  large  gates).  Association  values  represent  modified  ctii-squares  (witti  ttie  sign  of 
association)  and  wt^ere  calculated  according  to  wtiettier  species  were  captured  at  ttie 
sanr»e  trap  station  more  or  less  frequently  tt>an  expected  by  chiance.  See  text  for  rrjore  de- 
tail. 


Figure  3.— Histogram  of  ttie  number  of  positive  and  negative  species  associations  for  non- 
Dipodomys  species  broiten  into  thiree  treatrT>ent  categories:  (i)  treatments  witti  intact  rodent 
assemblages;  (ii)  D.  spectabilis  renrwval  plots;  and  (iii)  Dipodomys  removal  plots. 


304 


that  there  are  large  differences  in  the 
proportion  of  stations  capturing  two 
species  (ca.  5%  for  the  expected  com- 
pared to  23%  for  the  observed).  The 
main  result  is  that  trap  stations  cap- 
tured fewer  species  than  expected  if 
species  captures  were  random,  which 
further  evidence  that  species  in  this 
community  are  spatially  segregated. 


Use  of  Space  by  Individual 
Species:  The  Role  of  Cover 

In  this  section  we  are  interested  in 
the  individualistic  responses  of  spe- 
cies to  microhabitat  variation  and, 
particularly,  how  these  change  when 
manipulations  are  applied  at  the 
level  of  the  entire  community. 


Observed 
Expected 


#  species  per  Station 

Figure  4.— Histogram  of  expected  and  observed  number  of  species  captured  at  individuai 
trap  stations. 


r 


Table  2,— Results  of  analyses  testing  for  <l)  microhabitat  associations  of 
species  In  control  plots  and  (II)  for  shifts  In  microhabltats  between  control 
and  experimental  plots.  Microhabitat  associations  of  species  In  the  unma- 
nipulated  community  are  indicated  under  the  "control"  treatment:  "c"  If 
they  were  trapped  significantly  more  often  in  grecrter-than-median  cover; 
and  "o"  If  more  often  In  lesser-than-median  cover.  Significant  shifts  In  mi- 
crotxibitat  use  relative  to  that  on  "controls"  are  Indicated  by  a  "+"  if  the 
shift  was  towards  high  cover  and     It  towards  low  cover  (more  open) 
sites.  "R"  Is  used  to  Indicate  which  species  were  removed  from  treatments; 

Indicates  the  level  of  statistical  significance  C  for  P  <0.05;  "  for  P 
<0.01). 

Treatments 


Seed 


Seed 


Ant 


D.s. 


Species 

Control    pulsed  constant  removal 

removal 

removal 

D,  specfabilis 

o 

R 

R 

D.  merriami 

c 

R 

D.  ordii 

R 

P.  pencillafus 

• 

• 

+• 

P.  flavus 

o 

R.  megalofis 

c 

+• 

P.  maniculafus 

+• 

P.  eremicus 

c 

+* 

N.  albigula 

c 

•» 

O.  spp. 

There  was  marked  variability  both 
within  and  between  species  in  the 
usage  of  microhabitats  (table  2  and 
figs.  5  and  6).  On  control  plots  Pero- 
myscus  eremicus,  Neotorm  albigula, 
Reithrodontomys  megalotis,  and  Dipod- 
omys  merriami  (in  all  treatments  but 
the  D.  spectabilis  removals)  all 
showed  positive  associations  for  trap 
stations  with  greater-than-median 
cover. 


Treatment 


CO 
3 

0 
CC 


Figure  5.— Distribution  of  captures  in 
greater-ttian  and  less-than  rrtedian  cover 
for  the  five  heteromyid  species  listed  ac- 
cording to  treatnr»ent  and  season.  Capture 
data  is  graphed  relative  to  what  the  null 
hypothesis  predicts  (i.e.,  an  equal  number 
of  captures  in  both  microhabitat  types;  the 
zero  line).  Preference  for  higher-than -me- 
dian sites  is  represented  by  positive  values; 
less-than -nr»edian  cover  by  negative  val- 
ues. Bars  within  treatment  categories  indi- 
cate season:  from  left  to  right  Spring 
(March-May),  Summer  (June-August),  Fall 
(September-November),  Winter  (Decem- 
ber-February). Treatment  designation  Is  as 
follows:  "-DS",  Dipodomys  spectabilis  re- 
rTX>val;  "C",  control;  "SC",  corwtant  seed 
addition;  "-A",  ant  removal;  "-D",  Dipod- 
omys rerrxjval;  "SP",  pulsed  seed  additions. 


305 


i 


Those  species  associated  with 
more  open  microhabitats  included 
the  large  kangaroo  rat,  Dipodomys 
spedabilis,  and  the  smallest  species, 
Perogmthus  flavus.  The  remaining 
species  used  the  two  microhabitats 
more  indiscriminantly  with  the  ex- 
ception that  Peromyscus  maniculatus 
was  captured  more  frequently  in 
high-cover  microsites  in  the  D.  specta- 
bilis  removal  treatment. 

Figures  5  and  6  and  table  2  show 
our  experiments  were  of  the  kind 
and  were  of  sufficient  intensity  to 
promote  community-wide  changes  in 
the  use  of  microhabitats  by  all  spe- 
cies; only  the  Onychomys  showed  sig- 
nificant seasonal  shifts  in  microhabi- 
tat  use  (captured  more  frequently  in 
higher-cover  areas  during  the  fall 
than  in  the  other  seasons).  Using  the 
control  treatment  as  a  reference  point 
showed  that  the  majority  of  species 
shifted  their  use  of  microhabitats  on 
plots  where  D.  spectabilis  was  experi- 
mentally removed.  These  shifts,  in- 
volving eight  of  the  nine  species 
present,  included  an  increase  in  the 
use  of  microsites  with  less-than-me- 
dian  cover  by  D.  merriami,  P.  pencilla- 
tus,  P.  flavus,  and  N.  albigula,  and  an 
increase  in  the  use  of  high-cover  sites 
by  P.  maniculatus,  P.  eremicus,  R. 
megalotis,  and  D,  ordii. 

The  remaining  manipulations  reg- 
istered fewer  and  less  dramatic 
shifts:  i.e.,  increased  use  of  open  mi- 
crohabitats by  P.  pencillatus  and  P, 
maniculatus  on  constant  seed  addi- 
tion plots;  and  shifts  towards  higher- 
cover  microsites  by  R.  megalotis  and 
P.  pencillatus  in  ant  removal  and  Di- 
podomys removal  treatments,  respec- 
tively. 

The  role  of  microhabitat  in  the  or- 
ganization of  this  community  can  be 
further  evaluated  by  comparing  the 
distribution  of  trap  captures  for  all 
species  with  what  is  available  at  trap 
stations  (fig.  7).  The  objective  was  to 
determine  whether  certain  types  of 
microhabitats  are  used  by  the  rodent 
community  more  frequently  than 
others.  This  analysis  shows  that  the 
distribution  of  captures  in  control,  D. 


spectabilis  removal,  and  Dipodomys 
removal  plots  all  differ  significantly 
from  that  expected  if  the  use  of  mi- 
crohabitats was  random  with  respect 
to  vegetative  cover  (Kolmogorov- 
Smimov  two  sample  test;  P  <  0.05). 
However,  there  are  characteristic 
ways  these  differ  from  expected.  On 
control  plots  there  were  fewer  than 
expected  rodent  captures  in  traps 
having  <  5%  cover;  on  D.  spectabilis 
removal  plots  there  were  a  greater- 
than-expected  number  of  captures 
for  this  same  cover  category;  and  on 
Dipodomys  removal  plots  most  ro- 
dents were  captured  at  trap  stations 
with  >  10  %  cover. 


Discussion 

Out  results  identify  species  interac- 
tions as  the  principal  factor  produc- 
ing structure  in  this  community.  It  is 
significant  that,  by  adding  supple- 
mental seeds  or  removing  ants,  we 
were  able  to  change  microhabitats 
used  by  only  a  few  of  the  species  but 
removing  a  large,  potentially  domi- 
nant competitor  produced  many 
shifts.  This  suggests  that  the  primary 
mode  of  interaction,  as  it  effects  the 
patterns  of  microhabitat  use  in  this 
community,  involves  the  direct  re- 
sponses of  rodent  species  to  each 
other  rather  than  interactions  medi- 
ated through  the  exploitation  of  food 
resources,  or  the  individualistic  re- 
sponses of  rodents  to  particular  mi- 
crohabitat types. 

The  results  point  to  the  impor- 
tance of  one  dominant  species,  D. 
spectabilis,  whose  presence  in  the 
community  plays  a  disproportionate 
role  in  determining  which  microhabi- 
tats are  utilized  by  the  other  species, 
and  thus  the  organization  of  the  com- 
munity as  a  whole.  Whenever  it  is 
present,  regardless  of  how  much 
food  is  available,  it  appears  to  rele- 
gate the  majority  of  other  rodent  spe- 
cies to  higher-than-median  cover 
habitats,  thereby  reducing  the  den- 
sity of  potential  competitors  in  the 
open  habitats  it  prefers.  A  notable 


exception  is  Perogmthus  flavus  which 
was  captured  in  open  sites  along 
with  D.  spectabilis.  Because  of  its 
small  size  (ca.  7  g)  and  low  popula- 
tion density,  P.  flavus  may  have  only 
a  negligible  impact  on  the  food  re- 
sources that  can  be  harvested  by  D, 
spectablilis  and,  therefore,  may  not 
compete  directly  with  or  be  subjected 
to  its  aggressive  behavior.  The  im- 
portance of  such  size-ratio  thresholds 
in  allowing  species  to  coexist  has 
been  discussed  (Bowers  and  Brown 
1982).  Defending  open  areas  from 
other  rodents  may  be  a  mechanism 
by  which  D.  spectabilis  is  able  to 
preempt  food  resources  for  its  exclu- 
sive use.  Supporting  evidence  for  this 
comes  from  other  research  at  our 
study  site  where  it  was  found  that 

Treatment 


CO 
=3 

tr 


Figure  6.— Distribution  of  captures  in  thie  two 
microtiabitat  categories  for  the  six  Cricetid 
rodents  listed  by  treatment  and  season. 
See  legend  to  figure  5  for  nriore  details. 


306 


experimental  seeds  placed  in  open 
microhabitats  remained  largely  un- 
harvested  when  D.  spectabilis  was 
present  but  quickly  disappeared  in 
plots  where  it  was  removed  (see 
Bowers  et  al.  1987). 

Our  results  also  infer  something 
about  the  mechanism  by  which  D. 


spectabilis  affects  the  use  of  space  by 
other  rodent  species  in  the  commu- 
nity. Competition  can  be  mediated 
through  two  processes:  (i)  exploita- 
tive interactions  where  species  inter- 
act through  a  shared  resource  base; 
or  (ii)  contest  interactions  involving 
aggressive  dominance  and  relegation 


to  suboptimal  areas  and  resources. 
For  exploitation  alone  to  account  for 
the  patterns  of  microhabitat  use,  D, 
spectabilis,  through  its  foraging, 
would  have  to  significantly  alter  the 
distribution  of  food  (seed)  resources 
among  the  microhabitats  in  ways 
that  are  ecologically  significant  for 
the  other  species.  This  is  unlikely  for 
several  reasons.  First,  many  of  the 
seeds  utilized  by  the  smaller  species 
appear  to  be  too  small  to  be  economi- 
cally harvestable  by  D.  spectabilis  (see 
Bowers  et  al.  1987).  Second,  many  of 
the  species  showing  significant  mi- 
crohabitat shifts  were  non-granivores 
(i.e.,  Neotoma),  and  hence,  should  be 
relatively  insensitive  to  changes  in 
the  resource  base  attributable  to  the 
foraging  of  D.  spectabilis.  Third,  add- 
ing seeds  should  have  made  food 
more  available  to  all  species  and  re- 
duced the  degree  to  which  D.  specta- 
bilis was  able  to  alter  the  distribution 
of  food  resources,  so  that  shifts  by 
the  other  species  would  have  been 
expected  in  response  to  this  treat- 
ment. Moreover,  significant  changes 
in  the  distribution  of  food  resources 
were  more  likely  to  have  been  caused 
by  D.  merriami  that  occurs  at  higher 
densities  than  D.  spectabilis.  Our  re- 
sults show  that  adding  supplemental 
seeds  or  removing  D,  merriami  pro- 
duced fewer  shifts  than  removing 
just  D.  spectabilis. 

As  an  alternative  to  exploitation, 
competitors  of  large  body  size  may 
directly  restrict  the  foraging  activities 
of  smaller  species  through  interfer- 
ence. Under  an  interference  mode  of 
competition  adding  seeds  may  not 
alter  the  intensity  or  outcome  of  the 
interaction.  Because  most  significant 
shifts  in  microhabitat  use  occurred  in 
the  D.  spectabilis  removal  treatment — 
coupled  with  the  fact  that  adding 
seeds  had  litUe  effect  on  the  patterns 
of  microhabitat — leads  us  to  the  con- 
clusion that  aggressive  interference 
by  D.  spectabilis  is  the  mechanism 
most  consistent  with  our  results. 

Our  study  also  indicates  that  the 
majority  of  shifts  in  microhabitat  use 
originate  with  the  D,  spectabilis-D. 


UJ 
> 

< 

-I 
UJ 

cr 


50-1 


25  - 


CONTROL 


-1 
I 


>- 

y  50 


UJ 

o 

UJ 

cr 

U- 


25  - 


D.  spectabilis 
REMOVAL 


i 

I. 


50  n 


25  - 


Dipodomys 
REMOVAL 


I 
I 
I 


I 


0.0  2.5 


— I  J  1  1  1 — 

5.0    10.0    20.0   40.0  80.0 

PERCENT  COVER 


Figure  7.— Distribution  of  trap  captures  (broken  line;  all  species  combined)  and  available 
trap  sites  (sold  line)  relative  to  vegetative  cover  on  (i)  control;  (Ii)  D.  spectobiWs  removal; 
and  (iii)  Dipodomys  rerrtoval  plots. 


307 


merriami  interaction  and,  at  the  com- 
n\unity  level,  this  one  interaction  af- 
fects the  microhabitat  utilization  of 
the  majority  of  rodent  species 
through  a  complex  network  of  direct 
and  indirect  interactions.  Perhaps  the 
most  striking  shift  (not  in  the  magni- 
tude of  response  but  in  the  number 
of  individuals  involved)  was  the  in- 
creased use  of  open  areas  by  the  nu- 
merically dominant  D.  merriami 
when  D.  spedabilis,  which  had  for- 
mally used  these  sites  was  removed. 
Most  other  shifts  by  the  smaller  ro- 
dents, including  the  increased  use  of 
open  microhabitats  by  Perognathus 
flavus,  Peromyscus  maniculatus  and 
Reithrodontomys  megaloHs  when  all 
Dipodomys  were  removed,  suggest 
that  these  species  responded  directly 
to  D.  merriami  and  only  indirectly  to 
D.  spectabilis.  Hence,  there  appears  to 
be  a  hierarchy  of  interactions.  The 
primary  one  is  between  the 
behavioral  (D.  spectabilis)  and 
numerical  CD.  merriami)  dominants 
and  it  is  this  interaction  around 
which  the  community  is  organized. 
Other  studies  have  noted  the  poten- 
tial for  interference  betv/een  desert 
rodents  (Blaustein  and  Riser  1974, 
Hutto  1978,  Rebar  and  Conley  1983), 
especially  between  D.  spectabilis  and 
D.  merriami  (Frye  1983),  and  our 
study  shows  how  this  one  interaction 
can  resound  throughout  the  commu- 
nity to  affect  many  other  species. 

A  primary  motivation  for  our 
study — and  most  studies  focusing  on 
the  role  of  habitat — is  that  microhabi- 
tats represent  a  limited  and  exploit- 
able resource  and  the  manner  in 
which  they  are  used  directly  im- 
pinges on  population  growth  and 
density.  Many  of  the  experimentally 
induced  microhabitat  shifts  we  have 
reported  were  accompanied  by 
changes  in  local  species  density 
(Brown  and  Munger  1985,  Brown  et 
al.  1986)  that  support  the  contention 
that  D.  spectabilis  controls  the  dynam- 
ics of  this  community  through  a  com- 
bination of  direct  and  indirect  effects. 
For  example,  increasing  food  levels 
by  adding  seeds  resulted  in  an  in- 


crease of  D.  spectabilis  and  a  decrease 
in  D.  merriami  densities.  Removal  of 
D.  spectabilis  resulted  in  positive  den- 
sity compensation  of  D.  merriami  but 
no  changes  in  densities  of  the  smaller 
seed-eaters;  removal  of  all  Dipod- 
omys,  however,  resulted  in  large  den- 
sity increases  in  several  of  the 
smaller  rodents.  Taken  together,  the 
microhabitat  and  density  responses 
to  our  manipulations  indicate  that 
interference  competition  for  certain 
foraging  sites  not  only  determines 
the  spatial  organization  of  this  com- 
munity but  that  it  is  directly  involved 
in  the  regulation  of  rodent  densities. 

There  are  several  aspects  that  war- 
rant further  comment.  First,  our  re- 
sults show  that  when  D.  spectabilis  is 
present  open  sites  are  underutilized 
by  the  community  as  a  whole;  when 
D.  spectabilis  is  removed  the 
remaining  Dipodomys  shift  to  use 
these  open  sites;  but  when  all  Dipod- 
omys are  removed  the  remaining  sp)e- 
cies  are  unable  to  fully  utilize  the  va- 
cated microhabitats  (fig.  7).  Hence, 
there  appears  to  be  a  limit  to  how  far 
the  community  can  compensate  for 
the  absence  of  certain  species. 
Among  the  possible  explanations  for 
this  might  be  that  assemblages  of 
desert  rodents  have  been  associating 
together  for  a  sufficient  time  to  have 
lost  the  flexibility  to  respond  to  situ- 
ations where  one  or  more  of  the  spe- 
cies are  absent  (Schroder  and 
Rosenzweig  1975).  Another  is  that 
quadrupedal  species  may  have  a  lim- 
ited ability  to  avoid  predators  in 
open  microhabitats  and  this  limits 
the  degree  to  which  they  can  com- 
pensate when  the  bipeds  are  re- 
moved. In  either  case  the  relaxation 
of  one  factor  (in  this  case  the  removal 
of  dominant  competitors)  appears  to 
be  accompanied  by  the  increased  im- 
portance of  others. 

Second,  the  effects  of  interference 
competition  by  D.  spectabilis  appear 
to  be  effective  in  excluding  inter- 
specifics  primarily  in  open  areas  al- 
though this  dominant  does  occur  in 
greater-than-median  cover  habitats. 
It  may  be  that  aggression  is  of  lim- 


ited value  in  bushy  microsites  where 
subdominant  species  may  readily 
find  refugia.  As  a  result,  D,  spectabilis 
may  be  involved  in  two  kinds  of 
interactions  with  each  of  its  competi- 
tors; exploitatively  for  seeds  in  bushy 
sites  and  through  interference  in 
open  microhabitats.  As  a  result,  the 
highly  asymmetrical  interactions  be- 
tween the  dominant/subordinates  in 
open  sites  may  become  more  nearly 
symmetrical  in  bushy  sites  where 
premiums  are  on  foraging  efficiency. 

Third,  the  existence  of  strong,  ag- 
gressive interactions  among  sp>ecies 
increases  the  potential  for  indirect 
and  high-order  interactions  that  in- 
volve species  that  overlap  very  little 
in  resource  utilization.  For  example, 
the  large  herbivore,  Neotoma  albigula 
was  as  likely  to  shift  its  microhabitat 
use  as  the  granivorous  species.  How- 
ever, it  is  interesting  to  note  that  al- 
though the  non-granivores  shifted 
microhabitat  use  when  granivorous 
species  were  removed,  significant 
density  changes  were  limited  to  just 
other  granivores  (Brown  and  Munger 
1985).  Hence,  while  interference  may 
play  a  role  in  determining  use  of  mi- 
crohabitats by  rodents  in  several  for- 
aging guilds,  its  effects  appear  to  be 
most  significant  for  ecologically  simi- 
lar species. 

The  goal  of  experimental  pro- 
grams is  to  hold  most  variables  con- 
stant while  manipulating  others,  and 
then  to  measure  for  shifts  in  response 
variables.  In  this  paper  we  have  used 
patterns  of  microhabitat  use  in  con- 
trol plots  as  a  reference  point  for 
interpreting  our  experimental  results. 
The  assumption  in  doing  this  is  that 
the  degree  to  which  the  community 
responds  to  a  particular  manipula- 
tion provides  an  estimate  of  its  im- 
portance in  producing  the  basic  pat- 
tern. In  our  particular  case  we 
wanted  to  know  how  the  baseline 
patterns  of  microhabitat  use  (i.e., 
those  in  control  plots)  change  when 
supplemental  food  is  added  or  spe- 
cies are  removed.  While  some  of  our 
patterns  are  easy  to  interpret,  others 
are  very  complex  and  appear  to  in- 


308 


volve  a  hierarchy  of  responses  that 
operate  over  different  scales  in  time 
and  space.  The  existence  of  such  a 
dynamic  and  diverse  set  of  responses 
shows  the  limitations  of  most  two- 
species  models  of  interspecific  inter- 
actions upon  which  past  theories  of 
community  organization  have  largely 
been  based;  they  also  call  into  ques- 
tion the  value  of  studies  seeking  to 
understand  the  mechanistic  proc- 
esses that  determine  community 
composition  through  comparative, 
nonexperimental  methods. 

Implications  for  Management 

While  the  spatial  association  of  small 
mammals  with  particular  microhabi- 
tats  has  been  rigorously  and  repeat- 
edly documented,  and  the  patterns 
suggest  almost  a  universal  role  of 
microhabitat  in  ''structuring"  small 
mammal  communities,  the  processes 
responsible  for  producing  these  asso- 
ciations are  poorly  understood  (Price 
and  Brown  1983,  Bowers  1986).  To 

(successfully  manage/ manipulate 
such  communities  there  is  a  clear 
need  to  better  understand  the  proc- 
esses that  determine  which  mi- 
crohabitats  are  used  and  which  are 
I     not.  Towards  this  end  we  identify 
two  particularly  relevant  areas  for 
our  discussion:  (1)  the  scales  in  time 
j     and  space  over  which  microhabitat 
'     use  occurs;  and  (2)  the  roles  of  corre- 
lation, and  selection/ relegation  in  the 
occupancy  of  microhabitats. 

Vagile  organisms,  e.g.  small  mam- 
mals, can  potentially  respond  to  fea- 
tures of  the  habitat  at  several  differ- 
ent scales.  At  the  macro-end  of  the 
habitat  spectrum  animals  choose  ar- 
eas in  which  to  establish  home 
ranges.  Microhabitat  selection,  in 
contrast,  usually  involves  the  use/ 
disuse  of  small  areas  within  the 
home  range.  There  are  also  temporal 
differences  in  schedules  of  usage: 
macrohabitat  selection  occurs  over  a 
much  longer  timescale  (weeks- 
I      months)  while  microhabitat  use  oc- 
curs more  immediately  (seconds- 


minutes).  While  it  was  assumed  for 
years  that  macrohabitat  selection  oc- 
curred through  the  selection  of  com- 
posite microhabitats,  recent  work  on 
small  mammals  suggests  that  the  two 
may  be  largely  separate  (Morris 
1987). 

Most  factors  that  are  demonstra- 
bly important  to  the  structure  of 
small  mammal  communities,  i.e.,  pri- 
mary productivity,  plant  species  and 
foliage  height  diversity,  vegetation 
cover,  substrate  type,  competitor  di- 
versity and  abundance,  and  preda- 
tory pressure,  vary  more  between 
macrohabitats  than  among  mi- 
crohabitats within  particular  locales. 
For  example,  primary  productivity 
and  plant  cover  are  determined  by 
plant  species  composition  and  gen- 
eral conditions  for  growth  that  vary 
over  large  environmental  gradients 
at  the  macrohabitat  scale.  These  large 
scale  gradients  influence  patterns  of 
microhabitat  use  by  determining 
which  rodent  species  are  present, 
their  densities,  the  distribution  and 
abundance  of  food  resources,  and  the 
types  of  microhabitats  that  are  avail- 
able for  selection.  As  a  consequence, 
the  composition,  densities  and  demo- 
graphical  behavior  of  small  mammal 
populations  and  communities  may 
more  closely  reflect  habitat  variabil- 
ity at  the  macro — rather  than  the  mi- 
cro— scale.  On  the  other  hand,  mi- 
crohabitat usage  is  a  phenomena  in- 
volving choices  of  individuals.  Mi- 
crohabitats that,  by  definition,  vary 
over  scales  smaller  than  individual 
home  ranges,  have  significance  for 
the  survivorship  or  reproduction  of 
foraging  individuals,  but  may  have 
little  relevance  when  integrated  over 
the  population  as  a  whole. 

Most  experimental  studies  exam- 
ining the  role  of  microhabitat  in 
structuring  small  mammal  communi- 
ties tend  to  confound  micro-  and 
macrohabitat  effects.  Typically,  ma- 
nipulations (e.g.,  food  addition,  spe- 
cies removal,  tailoring  of  vegetation) 
are  applied  at  the  level  of  the  mac- 
rohabitat with  microhabitat  usage  by 
individuals  measured  as  a  response 


variable.  The  research  reported  here 
suffers  from  such  a  confounding. 
Other  field  experiments  that  examine 
the  allocation  of  foraging  time  among 
patches  restrict  manipulations  to  the 
level  of  microhabitats  (Kotler  1984, 
Price  and  Waser  1985),  and  are  not 
confused  by  responses  of  entire 
populations.  Clearly,  the  time  has 
come  to  utilize  the  information  we 
now  have  to  design  comprehensive 
studies  that  distinguish  between  mi- 
cro- and  macrohabitat  selection:  i.e, 
studies  that  manipulate  certain  mi- 
crohabitats on  a  scale  over  which 
populations  might  respond. 

Correctly  gauging  the  scale  over 
which  species  respond  to  the  envi- 
ronmental mosaic  is  critical  to  the 
successful  management  of  that  spe- 
cies. Programs  aimed  at  managing 
species  by  manipulating  microhabi- 
tats may  or  may  not  be  successful 
depending  on  the  scale  at  which  the 
manipulation  is  applied.  If  the  goal  is 
to  manage  pK)pulations  then  mac- 
rohabitat may  be  the  correct  context 
for  the  program.  This  is  not  to  sug- 
gest that  microhabitat  is  an  inappro- 
priate context  for  management  pro- 
grams. What  it  does  suggest  is  that 
management  oriented  programs 
should  be  directed  towards  popula- 
tions rather  than  the  behavior  of  indi- 
viduals. In  many  cases  this  may  in- 
volve changing  the  focus  from  the 
micro  to  macro  level. 

Our  second  point  for  discussion 
involves  habitat  correlation  versus 
selection/ relegation.  Habitat  usage  is 
determined  by  the  habitats  available, 
the  tolerances/preferences  of  organ- 
isms for  these  habitats,  and  the 
among-habitat  variability  in  fitness. 
Clearly,  there  must  be  some  variabil- 
ity in  the  structure  of  the  habitat  in 
order  for  selection  to  occur.  Habitats 
that  are  relatively  homogeneous  at 
the  smaller  scales  may  not  exhibit 
habitat  associations  even  by  highly 
selective  species.  Conversely,  show- 
ing that  a  habitat  has  a  significant 
degree  of  microhabitat  variability 
does  not  imply  that  organisms  have 
the  ability  or  inclination  to  respond 


309 


to  that  variability.  In  order  to  apply 
the  patterns  of  microhabitat  use  from 
one  site  to  predict  what  is  occurring 
at  another  requires  an  understanding 
of  the  biological  factors  underlying 
microhabitat  use.  Achieving  this  has 
proved  difficult  because  of  several 
problems.  First,  it  is  clear  from  a 
growing  body  of  experimental  work 
(including  the  present  study)  that 
habitat  association  does  not  necessar- 
ily imply  habitat  selection.  Because 
microhabitats  are  rarely  discrete, 
usually  grade  from  one  type  to  an- 
other, and  involve  a  suite  of  factors 
that  either  characterize  or  are  corre- 
lated with  specific  microhabitats,  it  is 
rare  that  habitat  occupancy  can  be 
tied  to  a  single  factor.  As  a  result  it  is 
difficult  to  conclude  that  an  animal  is 
selecting  a  habitat  per  se,  some  fea- 
ture of  that  habitat,  or  some  factor 
that  is  only  correlated  with  that  mi- 
crohabitat. As  a  complicating  factor 
habitat  selection  probably  reflects 
integrated  responses  of  organisms  to 
maximize  fitness  relative  to  several 
largely  independent  processes.  For 
example,  animals  might  select  mi- 
crohabitats so  as  to  minimize  preda- 
tory risk,  or  food  encounter  rates,  or 
to  jointly  maximize  food  intake  while 
minimizing  predatory  risk  (Bowers 
1987). 

Second,  the  present  results  and 
those  of  others  (Price  1978, 
M'Closkey  1978,  Wondolleck  1978, 
Bowers  et  al.  1987)  show  that  mi- 
crohabitat provides  a  template  over 
which  species  interactions  and  com- 
petitive hierarchies  become  ex- 
pressed. The  pattern  is  one  of  selec- 
tion/relegation— the  competitive 
dominant  selecting  its  preferred  mi- 
crohabitat and  through  exploitative 
or  interference  competition  relegat- 
ing other  species  to  less  preferred 
sites.  The  more  ecologically  similar 
two  species — and  hence,  the  greater 
the  intensity  of  competition  between 
them — the  greater  the  potential  role 
of  interspecific  competition  in  deter- 
mining microhabitat  usage. 

Competitive  interactions  represent 
dynamical  processes  impinging  on 


microhabitat  association  and  usage. 
Seasonal  or  year-to-year  fluxes  in  re- 
source availability  or  changes  in  the 
distribution  of  resources  among  mi- 
crohabitats can  alter  the  economical 
basis  underlying  competitive  interac- 
tions, and  thereby  promote  shifts  in 
microhabitat  usage.  For  example, 
Congdon  (1974)  found  during  peri- 
ods of  low  resource  availability  that 
the  large  D.  deserti  and  the  smaller, 
D.  merriami,  coexisted  in  the  same 
microhabitats  but  that  the  former  be- 
came aggressive  and  excluded  the 
latter  from  these  sites  when  food  lev- 
els increased.  Similarly,  Frye  (1983) 
found  that  D.  spectabilis  excluded  D, 
merriami  from  areas  around  its  bur- 
rows just  in  the  fall  when  seeds  from 
summer  annuals  were  abundant. 

Competitively  based  selection/ 
relegation  has  the  effect  of  increasing 
usage  of  secondary  habitats  while 
decreasing  usage  of  the  most  pre- 
ferred ones.  The  result  is  that  compe- 
tition promotes  the  segregation  of 
species  among  microhabitats  and  the 
degree  to  which  the  community  is 
spatially  organized.  Thus  it  is  no  ac- 
cident that  the  most  striking  patterns 
of  microhabitat  use  and  segregation 
are  in  communities  that  are  highly 
competitive  (Connor  and  Bowers 
1987).  As  the  present  study  has  dem- 
onstrated even  one  strong  interaction 
involving  just  two  species  (in  this 
case  the  behavioral  and  numerical 
dominants)  can  affect  microhabitat 
usage  by  all  species  in  the  commu- 
nity through  direct  and  indirect  path- 
ways of  interaction. 

Care  must  be  taken  when  examin- 
ing the  spatial  organization  of  com- 
munities where  competition  might  be 
occurring.  Efforts  to  understand  mi- 
crohabitat utilization  through  recon- 
stitution  studies  that  measure  indi- 
vidual species  preferences  for  mi- 
crohabitats, then  combines  these  in  a 
general  model  of  microhabitat  asso- 
ciation, will  miss  higher-order  com- 
petitive effects  that  may  be  the  main 
determinants  of  microhabitat  use. 
Further,  since  competition  can  be  in- 
determinate, work  over  complex 


pathways,  and  operate  over  widely 
varying  scales  in  time  and  space  it  is 
doubtful  that  any  one  model  can  be 
used  to  predict  microhabitat  use  over 
all  communities.  As  a  first  step  to- 
wards using  microhabitat  utilization 
as  a  tool  for  management  programs 
we  need  to  know  which  communities 
are  interactive  (i.e.,  structured 
around  selection/ relegation 
schemes),  which  are  non-interactive, 
and  something  about  ecological  at- 
tributes of  each.  It  may  be  that  in 
some  communities  microhabitat  is 
the  correct  context  for  management 
programs  while  in  other  communi- 
ties the  focus  should  be  on  species 
interactions.  Species  removal  experi- 
ments such  as  the  one  described  here 
provide  a  straightforward  test  of 
these  models. 

What  we  are  suggesting  here  is 
that  microhabitat  use  be  viewed  as  a 
manifestation  of  process  and  that 
these  processes  provide  the  basis  for 
management.  We  feel  that  the  most 
important  question  is  not  which 
habitats  are  being  used  by  a  particu- 
lar species  but  why  it  is  using  that 
microhabitat  and  not  others.  Recent 
work  has  shown  that  the  pathways 
by  which  species  interact  at  the  level 
of  ecological  communities  can  be 
very  complex  and  that  similar  pat- 
terns of  microhabitat  usage  need  not 
share  a  common  sequence  of  causa- 
tion (see  papers  in  Diamond  and 
Case  1986). 

Without  knowing  something 
about  which  processes  are  locally  im- 
portant it  is  risky  to  extrapolate  find- 
ings from  one  site  in  managing  an- 
other. For  example.  Bowers  (1986) 
found  in  rarefaction  studies  of  the 
same  three  species  rodent  commu- 
nity that  microhabitat  use  at  one  site 
was  affected  by  intersp)ecific  compe- 
tition but  not  at  two  others.  Such  re- 
sults underscore  the  fact  that  mi- 
crohabitat use  involves  multidimen- 
sional responses  of  organisms  to 
their  environment.  Understanding 
the  basics  of  such  relationships 
should  be  the  goal  of  community 
ecologists  and  managers  alike. 


310 


Acknowledgments 

We  thank  J.H.  Brown  and  D.B.  Th- 
ompson for  help  in  the  field  and  for 
discussion.  J.H.  Brown  and  R.T. 
M'Closkey  provided  critical  reviews 
of  the  n[\anuscript. 

Literature  Cited 

Blaustein,  Andrew  R.,  and  Arthur  C. 
Riser.  1974.  Dominance  relation- 
ships of  the  dark  kangaroo  mouse 
(Microdipodops  megacephalus)  and 
the  little  pocket  mouse  (Per- 
ogmthus  longimembris)  in  captivity. 
Great  Basin  Naturalist  34:312-314. 

Bowers,  Michael  A.  1982.  Foraging 
behavior  of  heteromyid  rodents: 
field  evidence  of  resource  parti- 
tioning. Journal  of  Mammalogy 
63:361-367. 

Bowers,  Michael  A.  1986.  Geographic 
comparison  of  microhabitats  used 
by  three  heteromyid  in  response  to 
rarefaction.  Journal  of  Mam- 
malogy 67:46-52. 

Bowers,  Michael  A.  1987.  Seed  re- 
moval experiments  on  desert  ro- 
dents: the  microhabitat  by  moon- 
light effect.  Journal  of  Mam- 
malogy 69:201-204. 

Bowers,  Michael  A.,  and  James  H. 
Brown.  1982.  Coexistence  of  desert 
rodents:  Chance  or  community 
struchire?  Ecology  63:391-400. 

Brown,  James  H.,  Diane  W. 

Davidson,  James  C.  Munger,  and 
Richard  S.  Inouye.  1986.  Experi- 
mental community  ecology:  the 
desert  granivore  system,  p.  41-61. 
In  (J.  Diamond  and  T.J.  Case,  eds.) 
Community  ecology.  Harper  and 
Row,  New  York. 

Brown,  James  H.,  and  James  C. 
Munger.  1985.  Experimental  ma- 
nipulation of  a  desert  rodent  com- 
munity: food  addition  and  species 
removal.  Ecology  66:1545-1563. 

Brown,  James  H.,  O.J.  Reichman,  and 
Diane  W.  Davidson.  1979.  Gra- 
nivory  in  desert  ecosystems.  An- 
nual Review  of  Ecology  and  Sys- 
tematics  10:201-227. 


Bowers,  Michael  A.,  Daniel  B.  Th- 
ompson, and  James  H.  Brown. 
1987.  Spatial  organization  of  a  des- 
ert rodent  community:  food  addi- 
tion and  species  removal.  Oecolo- 
gia  72:77-82. 

Colwell,  Robert  K.,  and  Eduardo  R. 
Fuentes.  1975.  Experimental  stud- 
ies of  the  niche.  Annual  Review  of 
Ecology  and  Systematics  6:281- 
310. 

Congdon,  Justin.  1974.  Effect  of  habi- 
tat quality  on  distributions  of 
three  sympatric  species  of  desert 
rodents.  Journal  of  Mammalogy 
55:659-662. 

Connor,  Edward  F.,  and  Michael  A. 
Bowers.  1987.  The  spatial  conse- 
quences of  interspecific  competi- 
tion. Annales  Zoologici  Fennici 
24:213-226. 

Davidson,  Diane  W.,  Douglas  A. 
Sampson,  and  Richard  S.  Inouye. 
1985.  Granivory  in  the  Chihua- 
huan  Desert:  interactions  within 
and  between  trophic  levels.  Ecol- 
ogy 66:486-502. 

Diamond,  Jared,  and  T.J.  Case.  1986. 
Community  Ecology.  Harper  and 
Row,  New  York,  New  York.  663  p. 

Fretwell,  Steven  D.,  and  H.L.  Lucas, 
Jr.  1970.  On  territorial  behavior 
and  other  factors  influencing  habi- 
tat distribution  in  birds.  I.  Theo- 
retical development.  Acta  Biotheo- 
redca  19:16-36. 

Frye,  Robert  J.  1983.  Experimental 
field  evidence  of  interspecific  ag- 
gression between  two  species  of 
kangaroo  rats  (Dipodomys). 
Oecologia  59:74-78. 

Harris,  John  H.  1984.  An  experimen- 
tal analysis  of  desert  rodent  forag- 
ing ecology.  Ecology  65:1579-1584. 

Hay,  Mark  E.,  and  P.J.  Fuller.  1981. 
Seed  escape  from  heteromyid  ro- 
dents: the  importance  of  mi- 
crohabitat and  seed  preference. 
Ecology  62:1395-1399. 

Hutto,  Richard  L.  1978.  A  mecha- 
nism for  resource  allocation 
among  sympatric  heteromyid  ro- 
dent species.  Oecologia  33:115-126. 

Kotler,  Burt  P.  1984.  Risk  of  preda- 
tion  and  the  structure  of  desert 


rodent  communities.  Ecology 
65:689-701. 

Lemen,  Cliff  A.,  and  Patricia  W.  Free- 
man. 1986.  Interference  competi- 
tion in  a  heteromyid  community 
in  the  Great  Basin  of  Nevada, 
USA.  Okios  46:390-396. 

Levins,  Richard.  1962.  Theory  of  fit- 
ness in  a  heterogeneous  environ- 
ment. I.  The  fitness  set  and  adap- 
tive function.  American  Naturalist 
96:361-373. 

M'Closkey,  Robert  T.  1978.  Niche 
separation  and  assembly  in  four 
species  of  Sonoran  Desert  rodents. 
American  Naturalist  112:683-694. 

Morris,  Douglas  W.  1987.  Ecological 
scale  and  habitat  use.  Ecology 
68:362-369. 

Munger,  James  C,  Michael  A.  Bow- 
ers, and  W.  Thomas  Jones.  1983. 
Desert  rodent  populations:  factors 
affecting  abundance,  distribution, 
and  genetic  structure.  Great  Basin 
Naturalist  Memiors  7:91-116. 

Price,  Mary  V.  1978.  The  role  of  mi- 
crohabitat in  structuring  desert 
rodent  communities.  Ecology 
59:910-921. 

Price,  Mary  V.  1983.  Ecological  con- 
sequences of  body  size:  a  model 
for  patch  choice  in  desert  rodents. 
Oecologia  59:384-392. 

Price,  Mary  V.,  and  James  H.  Brown. 
1983.  Patterns  of  morphology  and 
resource  use  in  North  American 
desert  rodent  communities.  Great 
Basin  Natur.  Memiors  7:117-134. 

Price,  Mary  V.,  and  Nickolas  M. 
Waser.  1985.  Microhabitat  use  by 
heteromyid  rodents:  effects  of  arti- 
ficial seed  patches.  Ecology  66:211- 
219. 

Rebar,  Cindy,  and  Walt  Conley.  1983. 
Interactions  in  microhabitat  use 
between  Dipodomys  ordii  and 
Onychomys  leucogaster.  Ecology 
64:894-899. 

Reichman,  O.J.  1981.  Factors  influ- 
encing foraging  in  desert  rodents, 
p.  195-213.  In  (A.C.  Kamil  and 
T.D.  Sargent,  eds.)  Foraging  be- 
havior: ecological,  ethological  and 
psychological  approaches.  Gar- 
land STPM,  New  York. 


311 


Rosenzweig,  Michael  L.  1981.  A  the- 
ory of  habitat  selection.  Ecology 
62:327-335. 

Rosenzweig,  Michael  L.,  and  Jerald 
Winakur.  1969.  Population  ecology 
of  desert  rodent  communities: 
habitats  and  environmental  com- 
plexity. Ecology  50:558-572. 

Schoener,  Thomas  W.  1971.  The  the- 
ory of  foraging  strategies.  Annual 
Review  of  Ecology  and  Sys- 
tematics  2:369-404. 

Schroder,  Gene,  and  Michael  L. 
Rosenzweig.  1975.  Perturbation 
analysis  of  competition  and  over- 
lap in  habitat  utilization  between 
Dipodomys  ordii  and  Dipodomys 
merriami.  Oecologia  19:9-28. 

Siegel,  Sidney  1956.  Nonparametric 
statistics  for  the  behavioral  sci- 
ences. McGraw-Hill,  New  York. 

Wondolleck,  John  T.  1978.  Forage- 
area  separation  and  overlap  in 
heteromyid  rodents.  Journal  of 
Mammalogy  59:510-518. 


Response  of  Small  Mammal 
Communities  to  Silvicultural 
Treatments  in  Eastern 
Hardwood  Forests  of  West 
Virginia  and  Massachusetts^ 

Robert  T.  Brooks  and  William  M.  Healy^ 


Abstract.— We  studied  small  mammal 
communities  and  associated  habitats  in  West 
Virginia  and  Massachusetts  hardwood  forests  with 
different  silvicultural  treatments.  In  Massachusetts, 
white-tailed  deer  (Odocoileus  virginianus)  density 
was  a  second  interactive  treatment.  Total  capture 
rates  were  relatively  stable  across  all  treatment 
classes.  Small  mammal  community  composition  and 
individual  species  capture  rates  varied  according  to 
treatment.  White-toiled  deer  density  had  a  greater 
effect  on  the  small  mammal  community  than  did 
silvicultural  practices. 


Small  mammals  (i.e.  New  World 
mice,  voles  and  jumping  mice  [Crice- 
tidae  and  Zapodidae],  shrews  [Sori- 
cidae],  and  squirrels  [Sciuridae])  are 
an  important  component  of  north- 
eastern forest  ecosystems.  Their  posi- 
tions in  the  food  web  are  broad, 
functioning  as  foragers  on  plant  and 
faunal  biomass  and  as  prey  to  nu- 
merous predators.  Small  mammals 
play  an  important  role  in  forest  dy- 
namics by  dispersing  seeds  and  my- 
corrhizal  fungal  sjx)res  and  by  en- 
hancing organic  matter  decomposi- 
tion and  mineral  cycling  (Spurr  and 
Barnes  1980). 

Relatively  little  is  known  of  the 
response  of  small  mammals,  by  spe- 
cies and  as  a  community,  to  silvicul- 
tural treatments  of  northeastern 
hardwood  forests.  Several  studies 
have  shown  that  the  response  varies 
by  species  but  that  the  small  mam- 
mal community  is  generally  resilient 
to  forest  harvesting  (Healy  and 
Brooks  1988,  Kirkland  1977,  Lovejoy 
1975,  Clough  1987,  Monthey  and 
Soutiere  1985).  These  studies  report 
the  predominant  effect  of  silvicultu- 
ral treatments  on  small  mammal 
habitat  is  the  enhancement  of  the 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America  (Flagstaff, 
AZ.July  19-21,  1988). 

'Robert  T.  Brooks  and  William  M.  Healy 
are  Researcti  Wildlife  Biologists,  USDA  Forest 
Service,  Norttieastern  Forest  Experiment 
Station,  University  of  Massact^usetfs, 
Holdswortt)  Hall,  Amherst,  MA  01003. 


ground  cover  and  lesser  woody 
vegetation.  Stenotopic  species  sensi- 
tive to  understory  plant  cover  and  its 
influence  on  microclimate  seem  to  be 
encouraged,  at  least  temporarily,  by 
most  forest  harvesting,  while  eu- 
ry topic  sp)ecies  seem  unaffected. 

The  study  began  in  West  Virginia 
(WV),  where  one  field  season  was 
completed,  and  was  continued  in 
Massachusetts  (MA).  Our  objective 
was  to  investigate  the  response  of  the 
small  mammal  community,  as  char- 
acterized by  live-trapping  statistics, 
to  standard  eastern  hardwood 
silvicultural  treatment  (Marquis  et  al. 
1975,  Hibbs  and  Bentley  1983).  In 
WV,  we  studied  the  effects  of  even- 
aged  regeneration  clearcutting  and 
subsequent  succession  on  small 
mammal  trapping  data.  In  MA,  the 
silvicultural  treatment  was  interme- 
diate thinnings,  with  a  second  inter- 
active treatment  of  differential  white- 
tailed  deer  density. 

STUDY  AREAS 

The  WV  study  sites  were  on  the 
Cheat  Ranger  District,  Monongahela 
National  Forest.  Three  randomly  lo- 
cated stands  in  each  of  four  stand- 
age  classes  were  studied  to  evaluate 
the  small  mammal  community  over  a 
silvicultural  rotation  for  even-aged 
management  of  a  northern  hard- 
wood forest.  The  four  age  classes 
were  seedling  (8-9  years),  sapling 
(12-14  years),  sawtimber  (61-76 


years),  and  mature  (>100  years).  The 
12  stands  averaged  19.4  ha  and 
ranged  in  area  from  6.1  to  38.8  ha. 
The  study  area  is  described  in  Healy 
and  Brooks  (1988). 

The  MA  study  sites  were  on  the 
Quabbin  Reservation  in  Franklin 
county.  This  watershed  is  managed 
by  Boston's  Metropolitan  District 
Commission  for  water  production. 
Four  randomly  selected  stands  in 
each  of  four  treatment  classes  were 
studied  to  evaluate  the  interactive 
effects  of  intermediate  thinning  and 
white-tailed  deer  density  on  a  south- 
em  New  England  oak  forest's  flora 
and  fauna.  The  treatments  were  com- 
binations of  thinned  vs.  unthinned 
and  low  (6-8/mi2)  vs.  high  (34-59/ 
mi^)  deer  density.  The  16  stands  av- 
eraged 19.1  ha  and  ranged  in  area 
from  4.9  to  57.5  ha.  The  MA  study 
site  is  described  in  Healy  et  al.  (1987). 

METHODS 

SnrKili  Mammals 

Small  mammals  were  live-trapped  at 
10  systematically  located  stations 
along  a  transect  in  each  of  the  28 
stands.  Transects  were  located  along 
the  long  axis  of  each  stand.  Trap  sta- 
tions were  no  less  than  80  m  apart  in 
any  stand.  At  each  station  three  Sher- 
man-type box  traps  (7.6  X  7.6  X  30.5 
cm)  were  baited  with  a  mixture  of 
peanut  butter,  rolled  oats,  and  bacon 
fat  and  set  within  1  m  of  each  station. 


313 


Traps  were  set  for  three  successive 
nights,  left  closed  for  one  (WV)  or 
four  (MA)  nights,  and  then  set  for 
three  additional  successive  nights. 
Sprung  traps  were  noted  and  their 
numbers  subtracted  from  the  total 
number  of  trap  nights  per  station  (18) 
to  calculate  the  number  of  effective 
trap  nights.  Each  forest  stand  was 
trapped  once  per  year.  Mammals 
were  trapped  from  mid-September  to 
early  October  1981  in  WV.  Mammals 
were  trapped  during  June  and  July  of 
1985-87  in  MA.  Captured  mammals 
were  marked  for  individual  identifi- 
cation and  released. 


Vegetation 

Vegetation  sampling  techniques  var- 
ied between  states.  Vegetation  plots 
were  systematically  located  along  the 
same  transects  as  were  the  small 
mammal  trapping  stations.  In  WV, 
trees  (>  2.5  cm)  were  sampled  using 
point-centered-quarter  method  (Cot- 
tam  and  Curtis  1956),  while  in  MA, 
trees  were  sampled  using  fixed-ra- 
dius plots.  Herbaceous  and  woody- 
stemmed  understory,  including  trees 
<  2.5  cm,  were  sampled  in  WV  using 
the  line  intercept  method  (Eberhardt 
1978)  and  in  MA,  these  flora  were 
sampled  using  fixed-radius  plots. 
Tree  and  understory  sampling  oc- 
curred at  the  same  locations  along 


the  transects.  These  data  were  used 
to  estimate  tree  density,  dominance, 
and  average  diameter  and  under- 
story cover  by  major  plant  life  form 
(i.e.  forb,  fern,  graminoid,  and 
woody-stem  species). 


Analysis 

Small  mammal  trapping  results  and 
vegetation  samples  were  summa- 
rized by  treatment  class  and  forest 
stand.  Treatment  effects  on  small 
mammal  capture  rates,  standardized 
as  captures  per  100  trap  nights  (TN), 
were  analyzed  by  one-way  (WV)  or 
two-way  (MA)  analysis  of  variance 
in  a  balanced,  nested  design  with 
stand  sum-of-squares  the  error  term 
for  treatment  effect.  Treatment  ef- 
fects on  species  composition  of  small 
mammal  capture  rates  were  analyzed 
using  multivariate  analysis  of  vari- 
ance. Testing  of  treatment  effects  was 
done  using  the  SPSS  MANOVA  pro- 
cedure (Hull  and  Nie  1981). 


RESULTS 
Vegetation  Structure 
West  Virginia 

Tree  density  declined  and  both  basal 
area  and  average  tree  diameter  in- 


creased as  the  forest  stands  matured 
from  an  even-aged  regeneration  har- 
vest (table  1). 

The  understory  changed  more  in 
life  form  composition  than  in  total 
cover.  Forb  cover  increased  in  per- 
centage of  cover  with  stand-age  as 
did  ferns  while  shrub  cover  declined 
(table  1).  All  forest  stands  supported 
a  luxuriant  understory  regardless  of 
age. 


Massachusetts 

Tree  density  and  basal  area  de- 
creased with  thinning  while  average 
tree  diameter  changed  little  (table  1). 
The  effect  of  deer  density  is  under- 
standable if  one  considers  the  low 
deer-unthinned  treatment  to  be  a 
"control"  condition. 

From  this  perspective,  high  deer- 
density  stands  had  lower  tree  density 
and  basal  area,  and  a  larger  average 
diameter  because  of  poor  regenera- 
tion resulting  from  browse  damage 
(table  1). 

Forb  cover  declined  with  higher 
deer  densities,  while  graminoid 
cover  increased  (table  1).  Shrub  and 
fern  cover  responded  irregularly  to 
the  treatments  except  for  a  dramatic 
increase  in  fern  cover  in  high  deer- 
thinned  stands,  an  effect  reported 
elsewhere  (Marquis  1987). 


Table  1  .—Average  structural  characteristics  of  scmnpied  forest  stands  by  state  and  treatment  class. 


West  Virginia 


Massachusetts 


Low  Deer 

High  Deer 

Characteristic 

Seedling 

Sapling 

Sdwtlmber 

Mature 

Unthinned 

Thinned 

Unthinned 

Thinned 

Tree  stems  >  2.5  cm 

Stems/ha 

1970 

2482 

969 

772 

1334 

876 

974 

645 

Basal  area  (mVha) 

5.3 

12.3 

41.7 

35.9 

24.5 

15.7 

22.8 

15.7 

Average  diameter  (cm) 

5.2 

6.8 

17.9 

17.6 

12.2 

10.7 

13.5 

13.5 

Percent  understory  cover 

Forb  species 

17 

18 

18 

36 

18 

16 

7 

14 

Fern  species 

5 

11 

21 

14 

15 

13 

12 

32 

Graminolds 

2 

2 

<1 

<1 

1 

2 

5 

17 

Shrubs  and  trees  <  2.5  cm 

32 

17 

9 

9 

15 

31 

26 

26 

314 


SrTKill  Mammals 
West  Virginia 

In  the  one  trapping  season,  662  indi- 
viduals of  15  sf)edes  were  captured. 
Total  capture  rate  averaged  33.2  indi- 
viduals/100  TN.  Average  total  cap- 
ture rate  declined  with  stand-age, 
from  42.4  individuals/100  TN  in 
seedling  stands  to  27.4/100  TN  in 
sawtimber  stands,  and  then  in- 
creased to  31.0/100  TN  in  mature 
stands  (table  2).  The  effect  of  stand- 


age  class  on  total  capture  rate  was 
not  statistically  significant  (F  =  3.16, 
P  =  0.086,  d.f.  =  3,8). 

Six  species  were  captured  in  all 
four  forest  age  classes,  eight  addi- 
tional species  were  captured  in  three 
or  fewer  treatment  classes  (table  2). 
Species  richness  was  greatest  in  the 
sawtimber  stands,  intermediate  in 
the  younger  stands,  and  least  in  the 
mature  stands. 

The  southern  red-backed  vole  (see 
table  2  for  small  mammal  scientific 
nomenclature)  was  the  most  com- 


mon species,  averaging  12.7  indi- 
viduals/100 TN.  Capture  rate  for  this 
species  declined  with  stand-age 
through  sawtimber  stands  (table  2), 
but  treatment  effect  was  not  signifi- 
cant (F  =  2.37,  P  =  0.146,  d.f.  =  3,8). 
Deer  mice  were  the  second  most 
common  species,  with  an  average 
capture  rate  of  10.0  individuals/ 100 
TN.  Capture  rates  for  this  species 
were  similar  across  treatment  class 
except  for  a  lower  rate  in  the  seedling 
stands.  No  significant  differences 
were  found  between  stand-age  class 


Table  2.— Average  number  of  individual  small  mammals  captured  per  100  trap  nights  by  species,  state,  and  treat- 
ment class. 


West  Virginia 


Massachusetts 


Low  deer 


High  deer 


Characteristic 


Seedling    Sapling  SawtimlDer   Mature    Unthinned  Thinned  Unthlnned  Thinned 


S.  red-backed  vole 

19.8 

12.7 

7,3 

n.i 

15.0 

12,3 

2.8 

3.8 

(y^i&iiiiiufiofTiys  yufjf^&fi/ 

Short-tailed  shrew 

9  J 

3.5 

3.6 

2.1 

1.1 

1.4 

0.2 

0,9 

(Blarina  brevicauda) 

E.  chipmunk 

0,6 

0.6 

1.2 

1.8 

0.7 

0.4 

0.4 

1.2 

(Tamias  sfriafus) 

White-footed  mouse 

0.6 

17,2 

17.8 

30.9 

23.4 

(Peromyscus  leucopus) 

Deer  mouse 

7.1 

ID.O 

11.7 

11.3 

(P.  maniculafus) 

Woodland  jumping  mouse 

2.1 

2.3 

0.6 

3.0 

0.1 

(Napaeozapus  insignis) 

Rock  vole 

1.6 

1.4 

0.4 

1.2 

(Microfus  chroforrhinus) 

S.  flying  squirrel 

1.1 

1.4 

0.4 

(Glaucomys  volans) 

Smoky  shrew 

0.2 

0.2 

0.4 

(Sorex  fumeus) 

Meadow  vole 

1.0 

0.1 

(M.  pennsyivanicus) 

Red  squirrel 

0.7 

<0.1 

(Tamiasciurus  hudsonicus) 

Masked  shrew 

0,2 

<0.1 

(S,  cinereus) 

Long-tailed  shrew 

0.2 

(S.  dispar) 

Woodland  vole 

0.2 

0.1 

0.1 

(M.  pineforum) 

Total  all  species 

42.4 

32.1 

27.4 

31.0 

34.2 

32.2 

34.3 

29.4 

Total  number  trap  nights^ 

497 

485 

510 

608 

2036 

2055 

2021 

2008 

'Scientific  names  from  Jones  et  al.  1975. 

^Total  number  of  possible  trap  nighits  (WV=540;  MA=2160)  minus  sprung  traps. 


315 


(F  =  0.29,  P  =  0.766).  Short-tailed 
shrews  were  the  only  other  species 
frequently  caught  in  all  stand-age 
classes.  Shrews  were  most  common 
in  the  seedling  stands  but  no  signifi- 
cant treatment  effect  was  found  (F  = 
0.96,  P  =  0.459).  No  significant  treat- 
ment effect  was  found  for  eastern 
chipmunks  (F  =  1.26,  P  =  0.351), 
woodland  jumping  mice  (F  =  0.21,  P 
=  0.885),  and  rock  voles  (F  =  0.41,  P  = 
0.749),  which  were  caught  infre- 
quently in  all  stand-age  classes  (table 
2).  The  remaining  eight  species  were 
caught  with  less  regularity.  No  fur- 
ther analysis  was  completed  for  these 
species.  No  significant  treatment  ef- 
fect was  found  in  the  simultaneous 
capture  rates  of  the  six  most  com- 
monly trapped  species  (i.e.,  red- 
backed  and  rock  voles,  short-tailed 
shrews,  chipmunks,  and  deer  and 
jumping  mice)  (Wilks  lambda  = 
0.046,  Rao's  F  =  0.979,  P  =  0.54). 

Massachusetts 

Over  3  years,  2,630  individual  small 
mammals  of  nine  species  were  cap- 
tured. Average  total  capture  rate  was 
32.6  individuals/ 100  TN.  There  was 
a  significant  decline  in  capture  rate 
across  the  years  (F  =  30.02,  P  <  0.001, 
d.f.  =  2,24).  The  capture  rate  of  43.7 
individuals/100  TN  in  1985  declined 
to  33.7  in  1986  and  20.3  in  1987.  The 
decline  was  observed  across  all  treat- 
ments and  all  stands. 

We  found  no  significant  full  model 
treatment  effect  on  total  capture  rate 
(F  =  1.78,  P  =  0.204,  d.f.  =  3,12).  Total 
capture  rate  for  all  species  was  high- 
est in  the  unthinned  stands  and  low- 
est in  the  thinned  stands,  especially 
in  the  high  deer-density  stands  (table 
2).  Neither  thinning  (F  =  3.99,  P  = 
0.069,  d.f.  =  1,12)  nor  deer  density  (F 
=  0.60,  P  =  0.453)  had  a  significant 
effect  on  total  capture  rates. 

Species  richness  was  highest  in  the 
high  deer-thinned  treatment  class, 
intermediate  in  the  two  low  deer- 
density  classes,  and  lowest  in  the 
high  deer-unthinned  treatment  (table 


1).  White-footed  mouse  was  the  most 
commonly  captured  species,  fol- 
lowed by  southern  red -backed  voles 
(table  2).  Capture  rates  for  both  spe- 
cies differed  by  treatment  class  (F  = 
9.01,  P  =  0.002,  d.f.  =  3,12  for  mice;  F 
=  6.06,  P  =  0.009  for  voles),  with  deer 
density  a  significant  effect  (F  =  20.7, 
P  =  0.0007.  d.f.  =  1.12  for  mice;  F  = 
17.5,  P  -  0.01  for  voles),  and  thinning 
effect  nonsignificant  (F  =  2.72,  P  = 
0.125  for  mice;  F  =  0.11,  P  =  0.74  for 
voles).  Voles  were  most  commonly 
captured  in  stands  of  low  deer-den- 
sity, and  mice  most  commonly  cap- 
tured in  stands  of  high  deer-density. 

Short-tailed  shrews  and  eastern 
chipmunks  were  the  only  other  spe- 
cies captured  in  each  of  the  four 
treatments.  Shrew  captures,  like 
those  for  red-backed  voles,  declined 
with  increasing  deer  density  (F  =  6.2, 
P  =  0.028)  but  showed  no  significant 
response  to  thinning  (F  =  3.1,  P  = 
0.1).  Chipmunk  captures  showed  no 
significant  response  to  either  deer 
density  (F  =  0.95,  P  =  0.35)  or  thin- 
ning (F  =  1.52,  P  =  0.24).  The  remain- 
ing five  species  were  infrequently 
caught  in  three  or  fewer  treatment 
classes,  and  no  further  analysis  was 
performed. 

Relative  capture  abundance  of  the 
four  most  commonly  captured  spe- 
cies (i.e.,  white-footed  mice,  red- 
backed  voles,  short-tailed  shrews, 
and  chipmunks)  differed  between  the 
two  levels  of  deer  density  (Wilks 
lambda  =  0.237,  Rao's  F  =  7.25,  P  = 
0.007).  No  difference  in  relative  cap- 
ture abundance  was  found  between 
the  two  thinning  classes  (Wilks  lamb- 
da =  0.496,  Rao's  F  =  2.28,  P  =  0.14). 

DISCUSSION 

Silvicultural  treatments  had  no  sig- 
nificant effect  on  total  small  mammal 
captures.  Total  capture  rates  were 
stable  across  the  range  of  treatments 
in  both  WV  (clear-cutting  and  subse- 
quent regrowth)  and  MA  (intermedi- 
ate thinning)  with  the  exception  of 
WV  seedling  stands  (table  2).  In 


those  stands,  where  regenerating 
trees,  shrubs,  and  herbaceous  plants 
flourish  in  the  sunlight  afforded  by 
the  removal  of  the  overstory,  total 
capture  rates  increased.  Otherwise, 
treatment  effects  on  habitat  structure 
were  insufficient  to  alter  total  cap- 
ture rates,  as  changes  in  the  species 
composition  of  small  mammal  cap- 
tures were  compensatory. 

Six  of  the  14  small  mammal  spe- 
cies captured  in  WV  were  captured 
in  all  four  stand-age  classes.  Of  the 
other  species:  red  squirrels  were  ob- 
served in  all  stands  but  poorly  cap- 
tured in  our  traps;  white-footed 
mice,  woodland  and  meadow  voles, 
and  masked  and  long-tailed  shrews 
were  each  captured  in  one  stand; 
four  smoky  shrews  were  captured  in 
three  stands;  and  southern  flying 
squirrels  were  captured  in  sapling 
and  older  stands.  McKeever  (1955) 
generally  concurs  that  these  species 
are  uncommon  in  WV  (woodland 
vole,  masked  and  long-tailed  shrew), 
or  are  common  in  forests  not 
sampled  in  this  study  (white-footed 
mouse  in  lower  elevation  forests)  or 
other  habitats  (meadow  vole).  Smoky 
shrews  and  southern  flying  squirrels 
are  more  common  WV  small  mam- 
mals but  were  poorly  represented  in 
our  sample.  Capture  data  for  these 
species  are  insufficient  for  drawing 
any  conclusions  regarding  species 
response  to  clearcutting. 

West  Virginia  red-backed  vole  and 
short-tailed  shrew  captures  increased 
concurrent  with  a  decline  in  deer 
mouse  captures  (table  2).  Vole  and 
shrew  capture  rates  were  highest  in 
seedling  stands.  Kirkland  (1977)  and 
Lovejoy  (1975)  reported  a  similar  re- 
sponse in  vole  captures  but  not  for 
shrew  captures.  The  increase  in  vole 
captures  could  be  a  response  to  the 
flush  in  vegetation  associated  with 
overstory  removal  and  to  the  volume 
of  slash  occurring  immediately  sub- 
sequent to  harvest.  These  factors  al- 
ter ground  level  microclimate,  in- 
creasing humidity  and  improving 
conditions  for  red-backed  voles 
(Lovejoy  1975,  Merrit  1981). 


316 


Vole  and  shrew  captures  declined 
and  deer  mice  captures  increased  as 
the  forest  stands  matured.  Forb  cover 
remained  stable  with  increasing 
stand-age  while  fern  cover  increased 
and  shrub  cover  declined  (table  1). 
These  changes  presumably  altered 
microhabitat  conditions  to  the  detri- 
ment of  red-backed  voles.  In  mature 
forest  stands,  vole  and  mouse  cap- 
tures were  equal.  In  these  stands, 
forb  cover  increased  dramatically 
from  conditions  observed  in  sawtim- 
ber  stands,  fern  cover  declined,  and 
shrub  cover  remained  stable  (table  1). 
These  habitat  conditions  resulted  in 
an  increase  in  red-backed  vole  cap- 
tures in  mature  stands  over  capture 
rates  for  the  species  in  sawtimber 
stands. 

Less  frequently  trapped  rock  voles 
were  captured  in  stands  with  rock 
outcrops.  Eastern  chipmunks  cap- 
tures increased  with  stand  age,  and 
woodland  jumping  mice  captures 
showed  no  clear  response  to  stand 
age.  Capture  rates  for  these  two  spe- 
cies were  not  related  to  measured 
habitat  variables  (Healy  and  Brooks 
1988). 

Species  composition  of  WV  small 
mammal  captures  and  individual 
species  capture  rates  were  not  signifi- 
cantly different  between  treatment 
classes.  No  major  small  mammal  spe- 
cies was  eliminated  by  clearcutting 
and  the  subsequent  maturing  of  the 
regeneration  of  the  hardwood 
stands.  These  species  either  survived 
within  clearcut  stands  or  recolonized 
harvested  stands  from  adjacent  un- 
cut stands.  Within  maturing  stands, 
habitat  conditions  were  sufficiently 
diverse  to  support  all  major  species. 

These  results  demonstrate  that 
clearcutting  of  WV  northern  hard- 
wood forests  allowed  for  the  contin- 
ued maintenance  of  the  small  mam- 
mal community.  Out  data  showed 
the  small  mammal  community  to  be 
relatively  stable  across  a  silviculhiral 
rotation,  with  no  major  changes  in 
composition  or  capture  rates  that 
could  alter  forest  ecosystem  function- 
ing or  character. 


Total  capture  rates  were  stable 
across  treatment  classes  in  MA. 
Treatment  effects  upon  habitat  struc- 
ture in  these  stands  were  insufficient 
to  alter  total  capture  rates.  However, 
capture  rates  for  individual  small 
mammal  species  varied  among  forest 
treatments.  Deer-density  had  a 
greater  influence  on  both  individual 
species  capture  rates  and  species 
composition  than  did  silvicultural 
treatment.  There  was  a  reciprocal 
change  in  the  relative  abundance  of 
red-backed  voles  and  white-footed 
mice  with  changes  in  deer  density 
(table  2). 

During  the  3  years  of  this  study, 
fall  deer  density  averaged  18/km^  in 
the  high  deer-density  stands  and  3/ 
km^  in  the  low  deer-density  stands 
(Healy  et  al.  1987).  Red-backed  voles 
were  scarce  in  high  deer-density 
stands.  Ferns  and  ericaceous  shrubs 
dominated  the  understory  of  these 
stands  while  the  understory  of  low 
deer-density  stands  contained  a 
greater  overall  number  of  plant  spe- 
cies and  forb  species  were  more 
abundant  (Healy  et  al.  1987).  Red- 
backed  voles  prefer  mesic  to  hydric 
sites,  especially  in  the  southern  por- 
tion of  their  New  England  range 
(Miller  and  Getz  1972, 1973).  It  seems 
that  foraging  by  deer  may  have  suffi- 
ciently altered  the  understory  vegeta- 
tion to  depress  vole  populations. 

The  response  of  white-footed  mice 
to  deer  density  in  MA  was  less  clear. 
Although  capture  rates  in  low  deer- 
density  stands  were  fewer  than  in 
high  deer-density  stands,  they  never- 
theless exceeded  capture  rates  for 
red-backed  voles  in  all  treatment 
classes  (table  2).  White-footed  mice 
are  ubiquitous  in  habitat  preference 
within  the  forest  ecosystem  (King 
1968,  Godin  1977,  Hamilton  and 
Whitaker  1979).  Whereas  Wolff  and 
Dueser  (1986)  suggest  that  the  these 
two  species  can  coexist  noncompeti- 
tively  through  microhabitat  and  food 
habit  differences,  our  data  suggest 
that  mice  capture  rates  are  sup- 
pressed in  stands  with  high  vole  cap- 
ture rates.  Our  stand  data  are  at  too 


coarse  a  scale  to  address  microhabi- 
tat separation.  One  would  need  to 
manipulate  vole  jx)pulations  experi- 
mentally to  evaluate  whether  the 
abundance  of  voles  is  competitively 
suppressing  mice  populations  in  low 
deer-density  stands  with  better  qual- 
ity vole  habitat. 

Short-tailed  shrews  captures  were 
more  common  in  low  deer-density 
stands,  a  possible  response  to  the 
greater  forb  cover  observed  in  these 
stands  and  probable  increase  in 
ground  level  humidity.  Eastern  chip- 
munk captures  offer  no  ready  inter- 
pretation in  regard  to  response  to 
treatment  effect  or  habitat  structure. 
The  remaining  five  species  were  cap- 
tured so  infrequently  that  it  is  impos- 
sible to  draw  any  conclusions  as  to 
the  effects  of  either  thinning  or  deer 
density  on  capture  rates. 

Thinning  MA  oak  forests  had  no 
significant  effect  on  capture  rates  of 
the  four  major  small  mammal  species 
or  species  composition  of  the  cap- 
tures. From  a  management  perspec- 
tive, intermediate  thinning  of  these 
forests  did  not  alter  the  continuation 
of  the  pretreatment  small  mammal 
community.  For  those  situations 
where  white-tailed  deer  have  been 
allowed  to  reach  population  levels 
where  vegetation  is  altered,  signifi- 
cant changes  in  the  small  mammal 
community  are  found.  Silvicultural 
treatment  effects  on  small  mammal 
habitat  are  temporary  and  ecosystem 
resources  (i.e.  nutrients,  energy)  re- 
main available  to  small  mammals. 
Long-term,  high  populations  of  deer, 
a  large,  possibly  competing  herbi- 
vore, alter  the  structure  and  compo- 
sition of  small  mammal  habitat  to  the 
detriment  of  some  species. 

CONCLUSIONS 

The  small  mammal  community  is  an 
important  component  of  northeast- 
em  forested  ecosystems,  functioning 
both  as  a  consumer  of  plant  and  ani- 
mal biomass  and  as  prey  to  numer- 
ous predators.  Intermediate  thinning 


317 


and  clearcutting  treatments,  which 
are  common  silvicultural  practices, 
have  minimal  or  ephemeral  effects 
on  the  numbers  of  small  manunals 
and  the  composition  of  the  small 
mammal  community  found  in  these 
forests.  Long-term,  high  deer  popula- 
tions may  permanently  alter  habitat 
structure  to  the  extent  that  changes 
occur  in  small  mammal  community 
composition. 

LITERATURE  CITED 

Clough,  Garrett  C.  1987.  Relations  of 
small  mammals  to  forest  manage- 
ment in  northern  Maine.  Canadian 
Field-Naturalist  101:40-48. 

Cottam,  Grant  and  J.  T.  Curtis.  1956. 
The  use  of  distance  measures  in 
phytosociological  sampling.  Ecol- 
ogy 37:451-460. 

Eberhardt,  L.  L.  1978.  Transect  meth- 
ods for  population  studies.  Journal 
of  Wildlife  Management  42:1-31. 

Godin,  Alfred  J.  1977.  Wild  mammals 
of  New  England.  304  p.  The  John 
Hopkins  Press,  Baltimore,  Md. 

Hamilton,  William  J.  and  John  O. 
Whitaker.  1979.  Mammals  of  the 
eastern  United  States.  346  p.  Cor- 
nell University  Press,  Ithaca,  N.Y. 

Healy,  William  M.  and  Robert  T. 
Brooks.  1988.  Small  mammal 
abundance  in  northern  hardwood 
stands  of  different  ages  in  West 
Virginia.  Journal  of  Wildlife  Man- 
agement 52(3)  :[in  press]. 

Healy,  William  M.,  Robert  T.  Brooks, 
and  Paul  J.  Lyons.  1987.  Deer  and 
forests  on  Boston's  municipal  wa- 
tershed after  50  years  as  a  wildlife 
sanctuary,  p.  3-21.  In  Deer,  for- 
estry, and  agriculture:  Interactions 
and  strategies  for  management. 
David  Marquis,  editor.  Sympo- 
sium Proceedings,  Plateau  and 
Northern  Hardwood  Chapter,  Al- 
legheny Society  American  Forest- 
ers, Warren,  Pa. 

Hibbs,  David  E.  and  William  R. 
Bentley.  1983.  A  Management 
Guide  for  Oak  in  New  England.  12 
p.  The  University  of  Connecticut, 


Cooperative  Extension  Service, 
Storrs. 

Hull,  C.  Hadlai  and  Norman  H.  Nie. 
1981.  SPSS  Update  7-9:  New  pro- 
cedures and  facilities  for  releases 
7-9.  402  p.  McGraw-Hill  Book 
Company,  NY. 

Jones,  J.  Knox,  Dilford  C.  Carter,  and 
Hugh  H.  Genoways.  1975.  Revised 
checklist  of  North  American  mam- 
mals north  of  Mexico.  14  p.  The 
Museum,  Texas  Tech.  Univ.,  Lub- 
bock. Occasional  Paper  No.  28. 

King,  John  A.  1968.  Biology  of  Pero- 
myscus  (Rodentia).  593  p.  The 
American  Society  of  Mammalo- 
gists.  Special  Publication  No.  2. 

Kirkland,  Gordon  L.  1977.  Responses 
of  small  mammals  to  the  clearcut- 
ting of  northern  Appalachian  for- 
ests. Journal  of  Mammalogy 
58:600-609. 

Lovejoy,  David  A.  1975.  The  effect  of 
logging  on  small  mammal  popula- 
tions in  New  England  northern 
hardwoods.  University  of  Con- 
necticut Occasional  Papers,  Bio- 
logical Science  Series  2:269-287. 

Marquis,  David  A.,  editor.  1987. 
Deer,  forestry,  and  agriculture: 
Interactions  and  strategies  for 
management.  183  p.  Symposium 
Proceedings,  Plateau  and  North- 
ern Hardwood  Chapter,  Al- 
legheny Society  American  Forest- 
ers. Warren,  Pa. 

Marquis,  David  A.,  Ted  J.  Grisez, 
John  C.  Bjorkbom,  and  Benjamin 
A.  Roach.  1975.  Interim  Guide  to 
Regeneration  of  Allegheny  Hard- 
woods. USDA  Forest  Service  Gen- 
eral Technical  Report  NE-19. 14  p. 
Northeastern  Forest  Experiment 
Station,  Broomall,  PA. 

McKeever,  Sturgis.  1955.  Ecology 
and  distribution  of  the  mammals 
of  West  Virginia.  Ph.D.  Disserta- 
tion. 335  p.  North  Carolina  State 
College,  Raleigh. 

Merrit,  Joseph  F.  1981.  Clethrionomys 
gapperi.  9  p.  The  American  Society 
Mammalogists  Mammalian  Spe- 
cies No.  146. 

Miller,  Donald  H.  and  Lowell  L. 
Getz.  1972.  Factors  influencing  the 


local  distribution  of  the  redback 
vole,  Clethrionomys  gapperi,  in  New 
England.  University  of  Connecti- 
cut Occasional  Papers,  Biological 
Science  Series  2:115-138. 

Miller,  Donald  H.  and  Lowell  L. 
Getz.  1973.  Factors  influencing  the 
local  distribution  of  the  redback 
vole,  Clethrionomys  gapperi,  in  New 
England.  II.  Vegetation  cover,  soil 
moisture,  and  debris  cover.  Uni- 
versity of  Connecticut  Occasional 
Papers,  Biological  Science  Series 
2:159-180. 

Monthey,  Roger  W.  and  E.  C.  Souti- 
ere.  1985.  Responses  of  small 
mammals  to  forest  harvesting  in 
northern  Maine.  Canadian  Field- 
Naturalist  99:13-18. 

Spurr,  Stephen  H.  and  Burton  V.  Bar- 
nes. 1980.  Forest  Ecology.  687  p. 
3rd  edition.  John  Wiley  and  Sons, 
N.Y. 

Wolff,  Jerry  O.  and  Raymond  D. 
Dueser.  1986.  Noncompetitive  co- 
existence between  Peromyscus  spe- 
cies and  Clethrionomys  gapperi.  Ca- 
nadian Field-Naturalist  100:186- 
191. 


318 


Habitat  Structure  and  the 
Distribution  of  Small 
Mamnnals  in  a  Northern 
Hardwoods  Forest^ 

Jeffery  A.  Gore^ 


Abstract.— The  influence  of  habitat  structure  on 

the  distribution  of  small  mammals  was  studied  in  an 
old-growth  northern  hardwoods  forest  in  New 
Hampshire.  Logistic  regression  equations  developed 
with  data  from  three  live-trapping  grids  were  able  to 
classify  locations  of  just  three  of  eight  small  mammal 
species  better  than  expected  by  chance.  For  all 
species  the  regression  models  failed  to  correctly 
predict  presence  in  an  independent  grid.  At  the 
scale  tested,  habitat  structure  hod  little  effect  on  the 
distribution  of  small  mammals  within  this  forest  type. 


In  northern  temperate  forests  small 
mammals  are  distributed  unevenly 
across  available  habitat,  even  within 
a  single  forest  type  or  age  class 
(Dueser  and  Shugart  1978,  Vickery 
1981,  Parren  1981,  Seagle  1985a).  Dif- 
ferential use  of  certain  segments  or 
microhabitats  within  a  broader  habi- 
tat type  has  most  often  been  reported 
for  sympatric  species  of  small  mam- 
mals, but  intraspecific  variation  in 
microhabitat  use  has  also  been  noted 
(Kitchings  and  Levy  1981,  Vickery 
1981,  Seagle  1985a). 

Differential  use  of  microhabitats 
by  small  mammals  may  be  a  conse- 
quence of  the  ecological  require- 
ments of  each  species  (i.e.  habitat  se- 
lection) or  it  may  be  the  result  of  par- 
titioning of  habitat  by  competing  sf)e- 
cies  (Crowell  and  Pimm  1976,  Porter 
and  Dueser  1982).  Another  hypothe- 
sis is  that  the  observed  use  of  mi- 
crohabitats by  small  mammals  is  pri- 
marily a  function  of  the  density  of 
small  mammal  populations. 

Under  this  hypothesis,  use  of  a 
certain  microhabitat  is  determined 
more  by  the  availability  of  animals  to 
occupy  the  area  than  by  structural 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  reptiles,  and  Small 
Mammals  in  Nortti  America.  (Flagstaff,  AZ, 
July  19-21,  1988.) 

^Jeffery  A.  Gore,  formerly  a  graduate 
student.  University  of  Massachiusetts,  De- 
partment of  Forestry  and  Wildlife  Manage- 
ment is  currently  Nongame  Wildlife  Biolo- 
gist, Florida  Game  and  Fresh  Water  Fish 
Commission,  f?t.  4.  Box  759.  Panama  City. 
Florida  32405. 


characteristics  of  the  microhabitat. 
Extrinsic  factors,  such  as  food  availa- 
bility, disease,  or  predation,  could 
alter  population  levels  and  thus  indi- 
rectly influence  the  distribution  of 
small  mammals  among  microhabi- 
tats. Observed  microhabitat  use 
might  also  be  a  function  of  some 
combination  of  habitat  selection, 
competitive  partitioning,  and  factors 
affecting  population  density. 

The  question  of  which  mechanism 
most  influences  the  distribution  of 
small  mammals  is  of  more  than  aca- 
demic impK)rtance.  If  small  mammals 
select  among  microhabitats  based  on 
structural  features,  then  disturbance 
such  as  timber  harvesting  may  have 
a  considerable  impact  on  population 
density  or  species  composition.  Con- 
versely, if  the  distribution  of  small 
mammals  is  primarily  a  function  of 
population  density,  then  habitat  dis- 
turbance is  likely  to'have  less  effect, 
or  at  least  a  less  direct  effect,  on  local 
populations.  Furthermore,  if  mi- 
crohabitat requirements  are  known, 
it  might  be  possible  to  manipulate 
population  levels  by  altering  struc- 
tural components  of  the  habitat. 

I  measured  use  of  microhabitat  by 
small  mammals  in  an  old-growth 
northern  hardwoods  (Acer-Fagus-Bet- 
ula)  forest,  a  habitat  that  contains  a 
variety  of  microhabitats  (Bormann 
and  Likens  1979)  and  supports  sev- 
eral species  of  small  mammals  (Love- 
joy  1970).  In  this  paper  I  identify  the 
small  mammal-microhabitat  associa- 
tions observed,  compare  them  to  re- 


sults of  previous  studies,  and  suggest 
that  the  distribution  of  small  mam- 
mals among  microhabitats  in  the 
northern  hardwoods  forest  is  influ- 
enced little  by  structural  features  of 
the  forest. 


Methods 

The  study  area  was  located  in  the 
White  Mountain  National  Forest, 
New  Hampshire  in  a  topographically 
isolated  site  known  as  the  Bowl 
(Martin  1977).  All  fieldwork  was  con- 
fined to  the  uncut,  old-growth  north- 
em  hardwoods  forest  that  comprises 
about  210  ha  in  the  lower  (600-750  m) 
elevations  of  the  Bowl.  The  old- 
growth  forest  is  structurally  hetero- 
geneous; numerous  treefalls  and 
gaps  in  the  canopy  are  present  and 
the  portion  of  the  forest  floor  cov- 
ered by  rock,  soil,  water,  or  vegeta- 
tion varies  greatly  across  the  stand 
(Gore  1986). 


Trapping 

In  1983,  small  mammals  were  live- 
trapped  on  three  60  x  105-m  grids, 
each  consisting  of  40  trapping  sta- 
tions spaced  at  15-m  intervals  along 
five  rows.  Two  stations  were  added 
to  each  grid  in  1984  to  increase  sam- 
pling at  seeps  and  along  streanns.  A 
fourth  grid  of  42  trapping  stations 
was  also  established  in  1984.  This 
grid  was  used  to  evaluate  the  robust- 


319 


ness  of  models  of  microhabitat  use 
that  were  developed  with  data  from 
the  initial  three  grids. 

One  Sherman  (5  x  6  x  17  cm), 
Pymatuning  (Tyron  and  Snyder 
1973),  and  pitfall  trap  were  located  at 
each  trapping  station  (Gore  1986). 
Traps  were  baited  with  sunflower 
seeds  and  set  simultaneously  for  four 
consecutive  days  each  month  from 
August  to  October,  1983  and  June  to 
October,  1984.  Each  trapping  period 
began  3-5  days  after  a  new  moon  (to 
minimize  variation  in  ambient  light) 
and  continued  regardless  of  weather 
conditions.  Captured  animals  were 
marked  and  released  at  the  capture 
site.  In  1983, 4,320  trap-nights  (12 
nights  X  120  stations  x  3  traps)  were 
recorded  on  three  grids.  In  1984  four 
grids  provided  10,080  trap-nights  (20 
nights  X  168  stations  x  3  traps). 


combinations  of  variables  to  be  incor- 
porated, but  at  the  risk  of  complicat- 
ing the  interpretation  of  results. 


Data  Analysis 

Chi-square  tests  (Snedecor  and  Co- 
chran 1980)  were  used  to  identify  sta- 
tistically significant  associations  be- 
tween capture  locations  of  all  pos- 
sible pairs  of  species.  Within  species, 
differences  between  capture  locations 
in  different  years  and  seasons  were 
analyzed.  The  relative  strength  of  as- 
sociations was  measured  via  the  con- 
tingency correlation  coefficient.  Phi 
(Brown  1983). 

Habitat  variables  from  plots  with 
and  without  captures  in  1984  were 
compared  for  each  small  mammal 
Sf)ecies.  Capture  locations  from  the 


entire  trapping  period  were  grouped 
even  though  some  values,  such  as  for 
vegetative  cover,  varied  between  sea- 
sons. If  the  habitat  values  within  the 
two  groups  were  normally  distrib- 
uted and  had  equal  variances,  signifi- 
cant differences  between  groups 
w6re  determined  via  t-tests;  if  not, 
Mann-Whitney  tests  were  used 
(Snedecor  and  Cochran  1980). 

Logistic  regression  (Bishop  et  al. 
1975,  Engelman  1983)  was  used  to 
identify,  for  each  mammal  species, 
the  microhabitat  variables  that  ac- 
counted for  statistically  significant 
portions  of  the  variation  in  capture 
success.  The  product  of  the  analysis 
is  a  set  of  regression  equations  for 
predicting  presence  of  each  species  at 
a  station  based  upon  quantitative 
measures  of  the  station's  habitat 
characteristics.  Logistic  regression 


Microhabitat 

At  each  trapping  station  microhabitat 
was  quantified  by  measures  from 
within  the  15  x  15-m  plot  in  which 
the  station  was  centered.  The  vari- 
ables used  to  quantify  the  habitat  are 
defined  in  table  1  and  the  methods 
for  measuring  them  are  described  in 
detail  by  Gore  (1986). 

The  26  habitat  variables  selected 
for  analysis  were  a  subset  of  a  larger 
group  of  variables  on  which  meas- 
ures were  made.  The  number  was 
reduced  in  order  to  facilitate  inter- 
pretation of  results  and  to  increase 
the  ratio  of  sample  cases  to  variables 
(Morrison  1985).  The  initial  group 
was  condensed  by  deleting  one  of 
each  pair  of  highly  correlated 
(r>0.50)  variables  that  were  similar  in 
ecological  form  or  function.  Some 
highly  correlated  variables,  such  as 
the  number  of  dead  trees  and  the 
number  of  logs,  were  retained  be- 
cause I  felt  they  represented  distinct 
ecological  features.  Variables  were 
included  regardless  of  whether  their 
means  or  distributions  varied  be- 
tween capture  and  no-capture  sta- 
tions. This  allowed  significant  linear 


Table  1.— Names  and  definitions  of  26  habitat  variables  measured  at  each 
trapping  station  in  1984. 


Name 


Definition 


SLOPE  angle  of  ground  from  horizontal  (%  of  90 ) 

NTREES  number  of  trees  (stems  >  1 0  cm  diameter) 

BATREE  basal  area  (m^  of  trees 

DEDTREE  number  of  dead  trees 

NSTUMPS  number  of  stumps  (dead  trees  <  1.5  m  high) 

TREEDIS  mean  distance  (m)  from  trees  to  station  center 

N$HRBL3  number  of  shrubs/saplings  (i.e.  stems  <10  cm  diameter) 

with  diameter  <3  cm  (measured  10  cm  above  ground) 

NSHRB36  number  of  shrubs/saplings  with  diameter  3-6  cm 

NSHRBG6  number  of  shrubs/saplings  with  diameter  6-1 0  cm 

NLOGS  number  of  logs  (stems  >10  cm  diameter) 

LOGDIST  mean  distance  (m)  from  logs  to  station  center 

AVGVOLA  mean  volume  (m^)  of  logs  in  class  of  least  decay 

AVGVOLB  mean  volume  of  logs  in  class  of  moderate  decay 

AVGVOLC  mean  volume  of  logs  in  class  of  advanced  decay 

DVEG  relative  cover  (%)  by  vegetation  <0.5  m  above  ground 

CVEG  relative  cover  by  coniferous  vegetation  <0.5  m  above 

ground 

ROCK  relative  cover  by  rocks  >0.5  m^ 

SOIL  relative  cover  by  exposed  soil 

WATER  relative  cover  by  water 

VEG52  relative  cover  by  vegetation  0.5-2  m  above  ground 

CVEG52  relative  cover  by  coniferous  vegetation  0.5-2  m 

VEGT2  relative  cover  by  vegetation  >2  m  above  ground 

CVEGT2  relative  cover  by  coniferous  vegetation  >2  m 

AVGLTR  mean  depth  (cm)  of  leaf  litter 

AVGHUMS  mean  depth  (cm)  organic  soil  (humus) 

AVGSHR  mean  horizontal  sheer  strength  (kg/m)  of  soil 


320 


was  used  instead  of  discriminant 
analysis  because  it  does  not  assume 
that  independent  or  explanatory 
variables  are  normally  distributed  or 
have  homogeneous  variances  (Press 
and  Wilson  1978). 

I  used  the  BMDP-LR  computer 
program  (Engelman  1983)  to  con- 
struct regression  equations,  or  mod- 
els, for  predicting  microhabitat  use 
as  defined  by  captures.  The  program 
selected,  in  a  stepwise  manner,  habi- 
tat variables  that  distinguished  sta- 
tions where  animals  were  captured 
from  those  where  they  were  not. 
Variables  were  entered  into  an  equa- 
tion if  their  F  value  was  significant  at 
P<0.10  and  removed  if  P  subse- 
quently exceeded  0.15. 

Initial  regression  models  were 
formed  using  data  obtained  in  1984 
from  Grids  1-3.  The  regression  mod- 
els for  each  mammal  species  were 
used  to  classify  stations  within  the 
three  grids  as  locations  where  the 
species  was  either  present  or  absent. 
The  models  were  then  used  to  pre- 
dict the  presence  of  each  sp)ecies  at 
stations  in  Grid  4.  Finally,  new  re- 
gression models  were  developed  for 
each  species  using  data  from  all  four 
grids.  These  were  compared  to  the 
models  from  the  initial  three  grids  in 
order  to  assess  the  effect  of  different 
sites  on  the  models.  The  Kappa  sta- 


tisHc  (Reiss  1973,  Engelman  1983) 
determined  the  significance  of  agree- 
ment between  locations  where  a  spe- 
cies was  observed  and  the  species- 
present  locations  predicted  by  the 
regression  models. 


Results 


Trapping 


Thirteen  species  of  small  mammals 
were  captured  during  the  study,  but 
only  those  captured  more  than  ten 
times  are  considered.  The  number  of 
captures  varied  widely  among  spe- 
cies and,  for  some  species,  between 
years  (table  2).  The  smokey  shrew 
(Sorex  fumeus),  pygmy  shrew  (S. 
hoyi),  and  eastern  chipmunk  (Tamias 
striatus)  were  captured  infrequently 
in  both  1983  and  1984,  while  the 
masked  shrew  (S.  cinereus)  and  the 
southern  red-backed  vole  (Clethriono- 
mys  gapperi)  were  common  in  both 
years.  The  northern  short-tailed 
shrew  (Blarina  brevicauda),  deer 
mouse  (Peromyscus  maniculatus)  and 
woodland  jumping  mouse  (Na- 
paeozapus  insignis)  were  captured 
more  often  in  1984  than  in  1983. 

Captures  from  August  through 
October  at  the  120  stations  trapped  in 
both  years  were  compared  for  each 


Table  2.— Number  of  individuals  captured,  total  captures,  and  individuals 
captured  per  TOO  trap-nights  (TN)  for  eight  small  mammal  species  in  1983 
and  1984. 


Number  of  captures 

_  1984 

Indivi-  /TOO       Indlvl-  /1 00 

Species  duals      Total        TN         duals  Total  TN 


Short-tailed  Shrew 

5 

5 

0.2 

160 

170 

2.4 

Masked  Shrew 

32 

32 

1.1 

69 

73 

1.0 

Smokey  shrew 

3 

3 

0.1 

8 

9 

0.1 

Pygmy  shrew 

6 

6 

0.2 

4 

4 

<0.1 

Eastern  chipmunk 

6 

7 

0.2 

22 

26 

0.3 

Deer  mouse 

41 

61 

1.4 

303 

671 

4.5 

Red-backed  vole 

25 

30 

0.9 

64 

87 

1.0 

Jumping  mouse 

71 

92 

2.5 

305 

494 

4.5 

species  to  determine  whether  1983 
and  1984  capture  locations  were  as- 
sociated. For  all  but  two  species,  in- 
dividuals were  captured  in  both 
years  at  few  (0-8  percent)  of  the  sta- 
tions with  captures.  No  species 
showed  a  significant  association  in 
capture  locations  between  years  (X^ 
tests,  P>0.05).  Even  for  deer  mice  and 
jumping  mice,  which  were  abundant 
in  both  years,  stations  with  captures 
in  both  years  comprised  only  32-41 
percent  of  all  stations  with  captures 
of  these  species. 

A  similar  comparison  was  made 
between  the  capture  locations  of  each 
species  in  summer  (June- August)  and 
fall  (September-October)  of  1984  on 
all  four  grids.  No  species  exhibited  a 
significant  (P>0.05)  association  be- 
tween summer  and  fall  locations. 

For  all  trapping  periods  and  spe- 
cies combined,  each  trapping  station 
had  at  least  two  captures.  In  1984,  all 
168  stations  recorded  at  least  one 
capture  and  85  percent  had  more 
than  four  captures.  The  maximum 
number  of  captures  at  a  single  station 
was  22  for  all  species  combined  and 
15  for  a  single  species,  the  jumping 
mouse.  Only  for  deer  mice  and  jump- 
ing mice  did  stations  with  multiple 
captures  outnumber  stations  with 
single  captures.  All  other  species 
were  taken  only  once  or  not  at  all  at 
the  majority  of  trapping  stations. 

The  only  pairs  of  species  captured 
at  the  same  locations  more  often  than 
expected  by  chance  (X^  tests,  P<0.05) 
were  jumping  mouseired-backed 
vole,  short-tailed  shrew:masked 
shrew,  and  short-tailed 
shrew:smokey  shrew.  The  associa- 
tion between  each  pair  was  positive 
and  weak  (0.15<Phi<0.30). 

Some  animals  died  after  capture, 
but  the  effect  upon  local  populations 
of  each  species  was  unknown.  For 
abundant  species  that  experienced 
low  mortality,  such  as  the  deer 
mouse  and  jumping  mouse,  the  effect 
was  probably  negligible.  For  the 
shrews,  which  had  high  mortality 
rates  during  capture,  the  effect  may 
have  been  substantial.  However,  cap- 


321 


hire  rates  indicated  no  adverse  ef- 
fects on  shrew  abundance.  In  1984 
more  shrews  of  each  species  were 
captured  in  September  than  in  the 
previous  three  months.  Furthermore, 
of  the  four  shrew  species,  only  cap- 
tures of  pygmy  shrews  decUned  be- 
tween years  (table  2). 


Microhabitat  Use 

Differences  between  habitat  values 
for  the  capture  and  no-capture  sta- 
tions of  Grids  1-3  from  1984  were 
compared.  For  each  species,  at  least 
one  habitat  variable  differed  signifi- 
cantly between  stations  with  and 
without  captures  (table  3). 

Logistic  regression  is  not  useful  if 
the  number  of  cases  of  either  of  the 
dependent  variable  values  (species 
presence  or  absence)  is  less  than  five 
percent  of  the  total  number  of  cases 
(D.  Hosmer,  University  of  Massachu- 
setts, personal  communication).  Be- 
cause the  pygmy  shrew  was  found  at 
only  two  percent  of  the  stations,  it 
was  deleted  from  the  analysis.  Deer 
mice  and  smokey  shrews  each  also 
had  widely  disparate  group  sizes,  95 
percent  present  and  95  percent  ab- 
sent respectively;  therefore,  results 
for  these  species  should  be  consid- 
ered cautiously. 

The  first  set  of  logistic  regression 
models  of  microhabitat  use  were 
based  on  captures  in  1984  from  the 
126  stations  in  Grids  1-3.  The  number 
of  significant  variables  included  in 
each  model  ranged  from  one,  when 
presence  of  red-backed  voles  was  the 
dependent  variable,  to  12,  when 
presence  of  eastern  chipmunks  was 
used  (table  4).  For  most  species  the 
variables  included  in  the  regression 
models  were  not  the  same  as  those 
whose  means  differed  between  cap- 
ture and  no-capture  stations  (table  3). 
This  suggests  that  some  linear  combi- 
nation of  habitat  variables  was  im- 
portant in  defining  the  microhabitat 
where  a  species  was  captured,  even 
though  individual  variables  alone 
were  not. 


The  habitat  variables  included  in 
the  logistic  regression  models  (table 
4)  were  selected  because  each  was 
associated  with  a  significant  (P<0. 10) 
portion  of  the  variance  in  the  capture 
data  of  a  species.  However,  if  these 
variables  and  their  regression  coeffi- 
cients cannot  be  used  to  correctly 
predict  the  capture  success  of  a  spe- 
cies at  a  station,  they  are  of  limited 
practical  value  regardless  of  their  sta- 
tistical significance.  To  assess  the 
utility  of  regression  models  as  de- 
scriptors of  microhabitat,  they  were 
independently  used  to  classify  each 
trapping  station,  based  on  habitat 
pararneters,  as  one  with  the  species 
present  or  absent  (table  5). 

Tests  of  the  agreement  between 
predicted  and  observed  capture  suc- 
cess were  not  possible  for  the 
smokey  shrew  because  no  sites  were 
classified  as  having  the  species  pres- 
ent. The  regression  model  was  not 
able  to  identify,  based  upon  the  habi- 
tat variables  measured,  the  eight  sta- 


tions that  captured  smokey  shrews. 
Conversely,  nearly  all  stations  were 
predicted  to  capture  deer  mice  and 
jumping  mice  (table  5).  The  regres- 
sion models  for  those  two  species 
were  unable  to  distinguish  those  sta- 
tions where  the  animals  were  not 
captured. 

For  the  other  four  species  the 
numbers  of  stations  with  and  with- 
out captures  were  more  similar  and, 
consequently,  so  were  the  number  of 
species-present  and  species-absent 
classifications.  For  red-backed  voles, 
however,  only  47  percent  of  the  clas- 
sifications of  present  were  correct 
and  this  was  not  significantly  better 
than  chance  (table  5).  (Dnly  for  the 
eastern  chipmunk,  short-tailed 
shrew,  and  masked  shrew  were  the 
regression  models  able  to  classify 
capture  success  at  a  level  better  than 
chance  agreement.  For  these  three 
species,  the  logistic  regression  mod- 
els may  be  useful  descriptors  of  the 
microhabitat  used  within  Grids  1-3. 


Table  3.— Means  (SE)  of  habitat  variables  that  differed  signlficqntly  be- 
tween stations  with  and  without  captures  of  each  species  in  Grids  1-3  In 
1984. 


Habitat 

Stations  with 

Stations  without 

Species 

Variable' 

captures 

captures 

Short-tailed  shrew 

NSHRBL3 

188(10) 

158(15) 

<0.05 

CVEGT2 

0.2(0.1) 

0.0 

<0.05 

VEG52 

30.9  (2.0) 

24.7  (2.2) 

<0.05 

Masked  shrew 

LOGDIST 

5.1  (0.1) 

5.5  (0.1) 

<0.05 

AVGVOLA 

0.20  (0.06) 

0.06  (0.02) 

<0.001 

DVEG 

38.2  (2.4) 

27.6(1.3) 

<0.001 

AVGLTR 

3.4(0,1) 

3.0  (0.1) 

<0.05 

DEDTREE 

0.63  (0.12) 

1.0  (0.1) 

<0.05 

Smokey  shrew 

AVGLTR 

3.8  (0.3) 

3.1  (0.1) 

<0.05 

Eastern  chipmunk 

BATREE 

0.93  (0.08) 

0.77  (0.02) 

<0.05 

AVGVOLA 

0.23(0.10) 

0.08  (0.02) 

<0.05 

VEG52 

37.1  (4.5) 

26.9(1.6) 

<0.05 

NSHRBL3 

220  (28) 

168(9) 

<0.05 

Deer  mouse 

DEDTREE 

0.85  (0.0) 

1.6  (0.42) 

<0.05 

NSHRBG6 

7.8  (0.4) 

10.8(1.2) 

<0.05 

AVGVOLB 

0,23  (0.03) 

0.51  (0.14) 

<0.05 

Red-backed  vole 

DVEG 

36.9  (2.3) 

27.7(1.4) 

<0.001 

Jumping  mouse 

VEGT2 

75.4  (0.9) 

70.2  (2.6) 

<0.05 

'Definitions  ofhobifaf  variables  are  given  in  table  h 
'Probability  that  capture  groups  have  equal  means. 


322 


Because  of  the  large  number  of 
variables  included  in  the  regression 
model  for  the  eastern  chipmunk 
(table  4),  it  was  difficult  to  concisely 
describe  the  microhabitat  of  this  spe- 
cies. Briefly,  the  eastern  chipmunk 
was  associated  with  large  trees 
(+BATREE),  downed  wood 
(+NLOGS,  -LOGDIST,  +AVGVOLB, 
+AVGVOLC),  and  dense  vegetation 
taller  than  0.5  m  (+VEG52,  +VEGT2, 
+NSHRB36). 

The  microhabitat  of  the  short- 
tailed  shrew  has  fewer  variables  but 
is  also  difficult  to  characterize.  Cap- 
tures were  negatively  associated  with 


numbers  of  medium-sized  shrubs 
and  with  vegetative  cover  between 
0.5  and  2  m  above  ground.  The 
masked  shrew  was  found  at  stations 
with  numerous  logs  and  dense  vege- 
tation <0.5  m  tall.  The  pxDsitive  asso- 
ciation with  slightly  decayed  logs 
and  dense  ground  cover  and  the 
negative  association  with  standing 
dead  trees  suggest  that  recent 
treefalls  may  provide  good  habitat 
for  masked  shrews. 

To  be  useful  predictors  of  species 
microhabitat,  regression  models 
should  be  successful  with  data  that 
are  indef>endent  of  those  from  which 


the  models  were  formed.  To  test  site- 
specificity  of  regression  models,  they 
were  applied  to  data  from  42  stations 
in  Grid  4.  Unlike  the  other  three 
grids.  Grid  4  had  a  perennial  stream 
running  through  it,  two  extensive 
canopy  gaps  from  recent  treefalls, 
and  highly  variable  soil  conditions. 

Regression  models  from  each  of 
the  seven  species  were  used  to  clas- 
sify the  stations  in  Grid  4  according 
to  capture  success.  None  of  the  clas- 
sifications, even  those  of  the  eastern 
chipmunk,  short-tailed  shrew,  and 
masked  shrew  were  correct  more  of- 
ten than  expected  due  to  chance  (for 
all  tests  Kappa  <0.48,  P>0.05). 

Because  none  of  the  regression 
models  were  useful  in  predicting 
capture  locations  in  Grid  4,  data 
from  all  four  grids  were  combined 
and  new  logistic  regression  models 
for  predicting  species  presence  were 
developed  to  determine  the  influence 
of  data  from  Grid  4  (table  7).  Deer 
nnice,  jumping  mice,  and  smokey 
shrews  again  had  widely  disparate 
group  sizes  and  the  models  could  not 
correctly  classify  the  stations  in  the 
less  common  group  (table  7). 

Agreement  between  observed  and 
predicted  locations  was  significant 
for  red-backed  voles  as  well  as  east- 
em  chipmunks,  short-tailed  shrews, 
and  masked  shrews  (table  7),  which 
had  significant  models  earlier.  Some 
of  the  variables  included  in  the  mod- 
els of  each  species  (table  6)  were  dif- 
ferent from  those  included  when 
only  data  from  Grids  1-3  were  used 
(table  4).  For  the  masked  shrew,  east- 
ern chipmunk,  and  red-backed  vole 
the  regression  models  created  with 
and  without  the  data  from  Grid  4 
were  similar,  even  though  the  predic- 
tions from  Grids  1-3  for  the  red- 
backed  vole  were  not  better  than  ex- 
pected by  chance.  The  coefficients 
changed  but  most  variables  were  the 
same.  This  suggests  that  for  these 
three  species  the  microhabitats  in 
Grid  4  were  similar  to  those  identi- 
fied in  the  other  three  grids.  The  rela- 
tionship of  species  to  microhabitat 
parameters  may  not  be  as  sensitive 


Table  4.— Logistic  regression  models  for  predicting  presence  of  small 
mommcd  species  based  on  data  collected  In  Grids  1  -3  in  1 984. 


independent 

Regression 

Coefficient/ 

Species 

Variable' 

Coefficient 

Standard  Error 

Short-tailed  shrew 

N$HRB36 

-0.062 

-1J84 

VEG52 

-0.063 

2.059 

CONSTANT 

0.487 

0.907 

Masked  shrew 

DEDTREE 

-0.704 

-2.538 

LOGDIST 

-O.660 

-2.577 

AVGVOLA 

2.621 

2.212 

DVEG 

0J37 

3.273 

CONSTANT 

0.784 

0.566 

Smokey  shrew 

AVGLTR 

1.514 

2.631 

CONSTANT 

-7.869 

-3.651 

Eastern  chipmunk 

NTREES 

-0.210 

-1.431 

BATREE 

2.523 

1.957 

NSHRB36 

0.123 

1.457 

NSHRBG6 

-0.447 

-2.831 

NLOGS 

0.188 

2.171 

LOGDIST 

-0,702 

-1.651 

NSTUMPS 

-1,706 

-1.272 

AVGVOLB 

3.031 

2.584 

AVGVOLC 

2.305 

2.560 

VEG52 

0.052 

1.187 

VEGT2 

0.086 

1.772 

AVGSHR 

-0.102 

-2.369 

CONSTANT 

-3.238 

-0.965 

Deer  mouse 

NLOGS 

-0.132 

-2.033 

AVGVOLB 

-1.155 

-1.586 

CONSTANT 

4.543 

4.804 

Red-backed  vole 

DVEG 

0.112 

3.230 

CONSTANT 

-2.450 

-4.415 

Jumping  mouse 

CVEG52 

-0.512 

-1.833 

VEGT2 

0.070 

2.120 

CONSTANT 

-2.544 

-1.205 

'See  table  1  for  definition  oftiabifat  variables. 


323 


as  the  regression  models  suggest. 
This  would  account  for  the  poor  per- 
formance of  the  models  from  Grids 
1-3  in  predicting  captures  on  Grid  4. 

For  short-tailed  shrews  the  regres- 
sion model  changed  greatly  when 
data  from  Grid  4  were  included. 
Four  new  variables  were  added,  and 
the  sign  of  the  coefficient  was  re- 
versed on  the  only  variable,  vegeta- 
tion between  0.5  and  2  m,  that  was 
retained.  This  suggests  that  captures 
of  short-tailed  shrews  or  the  meas- 
ured habitat  parameters  poorly  re- 
flect the  microhabitat  requirements 
of  the  species,  or  that  short- tailed 
shrews  are  not  restricted  by  mi- 
crohabitat within  this  forest. 


Discussion 

In  an  environment  of  limited  re- 
sources, sympatric  species  are  ex- 
pected to  partition  resources  as  a 
means  of  coexisting,  i.e.  avoiding 
competitive  exclusion  (Schoener 
1974).  Since  Brown  (1973)  first  sug- 
gested that  temperate  forest  rodents 
would  be  likely  to  partition  habitat 
rather  than  seasonally  variable  food 
supplies,  numerous  studies  in  north- 
ern temperate  forests  have  identified 
statistically  significant  associations 
between  habitat  structure  and  small 
mammal  distributions  (Dueser  and 
Shugart  1978,  Kitchings  and  Levy 
1981,  Parren  1981,  Vickery  1981, 
Schloyer  1983,  Seagle  1985a). 

Statistical  significance,  however, 
does  not  necessarily  impart  biologi- 
cal meaning  to  observed  patterns  of 
species  distributions.  Few  authors 
have  tested  the  biological  relevance 
of  their  models  of  microhabitat  use 
by  using  them  to  predict  microhabi- 
tat use  at  independent  locations  or 
times.  Parren  and  Capen  (1985) 
found  that  capture  locations  of  deer 
mice  could  not  be  accurately  pre- 
dicted using  discriminant  functions 
of  microhabitat  use  developed  with 
data  from  similar  habitats  the  previ- 
ous year.  Similarly,  none  of  the  logis- 
tic regression  models  I  developed 


were  useful  in  predicting  capture  lo- 
cations at  stations  other  than  those 
from  which  the  models  were  devel- 
oped. 

One  reason  for  the  poor  predictive 
capabilities  of  the  multivariate  mod- 
els may  be  that  trapping  does  not  ac- 
curately portray  the  relationship  be- 
tween species  presence  and  habitat 
requirements.  In  addition,  the  way  in 
which  habitat  features  are  measured 
may  not  depict  the  variability  per- 
ceived by  small  mammals  or  the 
variation  in  microhabitat  structure 
may  be  small  relative  to  the  niche 
breadth  of  each  species.  Unfortu- 
nately, these  problems  are  not  easily 
identified  or  solved.  Ideally,  the  ac- 
tivity of  many  individual  animals 
would  be  intensively  monitored,  but 
that  is  very  difficult  to  accomplish. 

Another  reason  for  the  poor  per- 
formance of  the  models  is  that  prob- 
lems in  applying  the  multivariate 
analyses,  such  as  disparate  sizes  of 
presence  and  absence  groups  and 
multicollinearity  of  variables,  make  it 
difficult  to  interpret  the  results  of 
habitat  models  (Noon  1984).  The 
scale  at  which  habitat  and  small 
mammals  are  sampled  also  greatly 


affects  the  relationship  that  can  be 
defined  (Morris  1984). 

Despite  these  potential  limitations, 
I  believe  the  inability  of  my  models 
to  predict  species  presence  on  a  inde- 
pendent grid  in  the  same  forest  stand 
suggests  that  structural  features 
alone,  at  least  at  the  microhabitat 
level,  are  not  important  to  the  distri- 
bution of  small  mammals.  Compari- 
sons of  capture  locations  and  review 
of  habitat  requirements  for  each  spe- 
cies supports  my  argument. 

The  locations  where  species  were 
captured  suggest  that  no  interspecific 
segregation  of  microhabitats  oc- 
curred. Overlap  in  capture  sites  was 
high  among  species  and  no  inverse 
relationships  were  observed,  even 
when  data  were  examined  by  season. 
This  suggests  that  habitat  partition- 
ing or  microhabitat  selection  is  ab- 
sent or  operating  at  a  finer  scale  than 
my  trap  stations. 

The  weak  association  I  found 
among  capture  locations  of  each  spe- 
cies between  years  and  seasons  sug- 
gests that  individual  species  were  not 
selecting  particular  trapping  stations. 
It  is  possible  that  subtle  shifts  in  the 
microhabitat  used  would  not  be  per- 


Table  5.— Classification  of  126  trapping  stations  in  Grids  1  -3  as  locations 
where  each  of  eight  small  mammal  species  Is  present  or  absent  based  on 
logistic  regression  models,  and  agreement  between  predicted  and  ob- 
sen^ed  classifications. 


No.  of  stations 


classified^ 

%  Correct^ 

Agreement^ 

Present 

Absent 

Present 

Absent 

K 

ASE 

P 

Short-tailed  shrew 

104 

22 

64 

64 

0.185 

0.079 

<0.025 

Masked  shrew 

23 

103 

65 

81 

0.381 

0.093 

<0.001 

Smokey  shrew 

0 

126 

0 

94 

Eastern  chipmunk 

12 

114 

75 

92 

0.648 

0.114 

<0.001 

Deer  mouse 

125 

1 

94 

0 

0.014 

0.013 

NS 

Red-backed  vole 

15 

111 

47 

72 

0J13 

0.084 

NS 

Jumping  mouse 

124 

2 

87 

50 

0.079 

0.090 

NS 

'Prior  probability  of  presence  =  0.05. 

'Percent  ofstatioris  wt)ere  present/absent  classificotion  agreed  withi  observations 
^om  trapping  in  1984, 

=  Kappa  statistic  (Fleiss  1973).  ASE  =  asymptotic  standard  error.  P  =  ProbabiHty 
tttat  agreement  is  due  to  ctiance.  i.e.  K=0.  NS  =  not  significant  >0.05. 


324 


ceived  by  examination  of  trapping 
locations  alone.  The  niicrohabitat  oc- 
cupied by  sn-iall  mammals  has  been 
reported  to  shift  with  season  (Kitch- 
ings  and  Levy  1981,  Vickery  1981), 
population  density  (M'Closkey  1981, 
Adler  1985),  and  species  composition 
(Seagle  1985b).  This  suggests  that  mi- 
crohabitat  use  is  dynamic,  regardless 
of  whether  the  shifting  is  determinis- 
tic or  stochastic. 

Another  argument  against  differ- 
ential use  of  microhabitats  by  the 
small  mammals  I  observed  is  the  va- 
riety of  habitats  they  occupy.  The 


eight  species  I  found  in  the  old- 
growth  northern  hardwoods  forest 
have  been  found  in  other  age-  and 
size-classes  of  northern  hardwoods 
forests  as  well  as  in  other  forest  types 
(Lovejoy  1970,  Richens  1974, 
Kirkland  1977,  Miller  and  Getz  1977, 
Hill  1981,  and  others).  Except  for 
smokey  shrews  and  pygmy  shrews, 
the  sf)ecies  are  common  in  a  variety 
of  habitats  comprising  a  wide  range 
of  structural  characteristics.  In  fact, 
descriptions  of  the  important  habitat 
features  associated  with  each  species 
do  not  always  agree  [e.g.  see  Hamil- 


Table  6.— Logistic  regression  models  for  predicting  presence  of  small 
mammal  species  based  on  data  collected  In  Grids  1  -4  In  1 984. 


Species 


Independent 
Variable' 


Regression 
Coefficient 


Coefficient/ 
Standard  Error 


Short-tailed  shrew 


Masked  shrew 


Eastern  chipmunk 


Deer  mouse 


Red-backed  vole 


Jumping  mouse 


DEDTREE 

DVEG 

ROCK 

VEG52 

AVGLTR 

CONSTANT 

DEDTREE 

LOGDIST 

DVEG 

AVGHUMS 

CONSTANT 

NSHRBG6 

NLOGS 

NSTUMPS 

AVGVOLB 

AVGVOLC 

VEGT2 

AVGLTR 

CONSTANT 

DEDTREE 

NSHRB36 

AVGVOLB 

AVGHUMS 

CONSTANT 

DEDTREE 

DVEG 

CONSTANT 

LOGDIST 

AVGVOLB 

SOIL 

CONSTANT 


-0.355 
0.053 
0,207 
0,043 
0.809 
-3,463 
-0.401 
-0.810 
0.159 
0.499 
-0.475 
-0.280 
0,125 
-1.761 
2.085 
1.042 
0.072 
0.675 
-9.267 
-0.934 
-0.170 
-1.064 
-0.794 
9763 
-0.321 
0.083 
-1.841 
0.475 
-0.869 
-0.192 
-0.140 


-2.157 
2.275 
2,721 
1.826 
-3.198 
-3.392 
-2.029 
-3.415 
5.348 
2.467 
-0,336 
-2.825 
1.979 
-1.639 
2.762 
1.874 
2.088 
1.867 
-3.506 
-2.511 
-2.366 
-1.599 
-2.165 
4.248 
-1.849 
3.778 
-4.365 
1.742 
-1,815 
-2.351 
-0.098 


'See  table  1  for  deftiifion  of  habitat  variables. 


ton  (1941),  Brower  and  Cade  (1966), 
Lovejoy  (1970),  Vickery  (1981),  and 
Parren  (1981)  for  descriptions  of 
jumping  mouse  habitat].  If  each  spe- 
cies is  common  under  a  wide  range 
of  habitat  conditions,  it  seems  un- 
likely that  they  would  partition  or 
select  habitat  based  on  the  advan- 
tages of  structural  features  alone. 

Fine  discrimination  of  the  forest 
habitat  seems  more  improbable  when 
the  temjx)ral  variability  of  mi- 
crohabitats is  considered.  Within  the 
northern  hardwoods  forest  of  New 
Hampshire  microhabitats  are  greatly 
modified  in  winter  by  deep  snow 
cover,  in  summer  by  closed  canopies 
and  sparse  ground  cover,  and  in  fall 
by  deep  leaf  litter.  Therefore,  resi- 
dent species  must  accommodate  sea- 
sonally variable  microhabitats  as  well 
as  seasonally  variable  food  supplies. 
The  reasoning  Brown  (1973)  used  to 
suggest  that  temperate  forest  rodents 
could  not  specialize  on  seasonally 
variable  food  resources  seems  appli- 
cable also  to  seasonally  variable  mi- 
crohabitats. 

In  the  forest  I  sampled,  presence 
of  most  species  at  individual  trap- 
ping stations  could  not  be  accurately 
predicted  based  on  structural  fea- 
tures of  the  habitat.  If  microhabitat 
structure  does  not  greatly  influence 
the  distribution  of  small  mammals 
within  this  forest  type,  disturbance  of 
the  habitat  should  not  directly  affect 
population  levels.  However,  the  scale 
at  which  the  disturbance  occurs  may 
determine  to  what  extent  local  popu- 
lations are  affected.  Small  scale  dis- 
turbance of  the  habitat,  such  as  har- 
vesting by  single-tree  or  small-group 
selection,  would  likely  not  affect  Sf)e- 
cies  composition  or  density  of  the 
resident  small  mammals.  More  wide 
scale  disturbance,  such  as  clear-cut- 
ting of  the  entire  forest  stand,  might 
alter  the  habitat  so  greatly  that  spe- 
cies abundance  and  distribution  is 
affected  (Kirkland  1977).  Given  the 
apparent  wide  range  of  habitat  con- 
ditions in  which  these  small  mam- 
mals occur,  even  a  large  scale  distur- 
bance of  the  northern  hardwoods 


325 


forest  would  likely  cause  only  tem- 
porary changes  in  species  composi- 
tion or  population  levels  of  small 
mammals. 

The  relationship  between  small 
mammals  and  habitat  structure 
within  the  northern  hardwoods  for- 
est remains  poorly  understood. 
However,  the  data  presented  here,  as 
well  as  comparisons  at  different 
scales  (Morris  1984, 1987),  suggest 
that  microhabitat  features  play  only  a 
minor  role  in  the  distribution  of 
small  mammals  within  the  forest.  A 
more  important  determinant  of  small 
mammal  distribution  may  be  popula- 
tion size  and  the  factors  that  affect  it, 
such  as  food,  weather,  and  preda- 
tors. Consequently,  models  for  pre- 
dicting the  distribution  of  small 
mammals  within  the  northern  hard- 
woods forest  will  likely  remain  un- 
successful until  factors  that  affect 
population  size  are  included.  The 
temporal  and  spatial  scales  at  which 
these  factors  influence  distribution 
must  also  be  addressed. 


Acknowledgments 

I  thank  C.  D.  Warren,  J.  M.  Prince, 
and  S.  L.  Crane  for  assistance  with 
fieldwork.  Staff  of  the  U.S.  Depart- 
ment of  Agriculture,  Forest  Service 
(USPS)  facilitated  work  in  the  White 
Mountain  National  Forest.  W.  A.  Pat- 
terson III,  R.  M.  DeGraaf,  S.  L.  Gar- 
man,  S.  W.  Seagle,  M.  W.  Sayre,  D.  P. 
Snyder,  and  an  anonymous  reviewer 
provided  helpful  comments  on  ear- 
lier drafts  of  the  manuscript.  This  re- 
search was  supported  by  grants  from 
the  USPS  and  the  Mclntire-Stennis 
Cooperative  Forestry  Research  Pro- 
gram (Grant  No.  MS-47). 


Literature  Cited 

Adler,  Gregory  H.  1985.  Habitat  se- 
lection and  species  interactions:  an 
experimental  analysis  with  small 
mammal  populations.  Oikos 
45:380-390. 


Bishop,  Yvonne  M.  M.,  Stephen  E. 
Feinberg,  and  Paul  W.  Holland. 

1975.  Discrete  multivariate  analy- 
sis: theory  and  practice.  557  p. 
MIT  Press,  Cambridge,  MA. 

Bormann,  F.  Herbert,  and  Gene  E. 
Likens.  1979.  Pattern  and  process 
in  a  forested  ecosystem.  253  p. 
Springer- Verlag,  New  York,  New 
York. 

Brower,  J.  E.,  and  T.  J.  Cade.  1966. 
Ecology  and  physiology  of  Na- 
peozapus  insignis  (Miller)  and  other 
woodland  mice.  Ecology  47:46-63. 

Brown,  James  H.  1973.  Species  diver- 
sity of  seed-eating  desert  rodents 
in  sand  dune  habitats.  Ecology 
54:788-797. 

Brown,  M.  B.  1983.  Frequency  tables, 
p.  143-206.  In  BMDP  Statistical 
Software  1983.  W.  L.  Dixon,  edi- 
tor. University  of  California  Press, 
Berkeley. 

Crowell,  Kenneth  L.,  and  S.  L.  Pimm. 

1976.  Competition  and  niche  shifts 
of  mice  introduced  on  small  is- 
lands. Oikos  27:251-258. 

Dueser,  Raymond  D.,  and  H.H. 
Shugart,  Jr.  1978.  Microhabitats  in 
a  forest-floor  small  mammal 
fauna.  Ecology  59:89-98. 


Engelman,  Laszlo.  1983.  Stepwise  lo- 
gistic regression,  p.  330-344.  In 
BMDP  statistical  software  1983. 
W.L.  Dixon,  editor.  University  of 
California  Press,  Berkeley. 

Heiss,  J.  L.  1973.  Statistical  methods 
for  rates  and  proportions.  223  p. 
John  Wiley  and  Sons,  New  York, 
New  York. 

Gore,  Jeffery  A.  1986.  Small  mam- 
mals and  habitat  structure  in  an 
old-growth  northern  hardwoods 
forest.  Ph.D.  Dissertation.  135  p. 
University  of  Massachusetts, 
Amherst. 

Hamilton,  W.J.,  Jr.,  and  David  B. 
Cooke.  1941.  Small  mammals  and 
the  forest.  Journal  of  Forestry 
38:468-473. 

Hill,  Barbara  J.  1981.  Small  mammal 
habitat  associations  in  selected 
timber  community  types  on  the 
White  Mountain  National  Forest, 
New  Hampshire.  M.S.  Thesis.  140 
p.  University  of  Massachusetts, 
Amherst. 

Kirkland,  Gordon  L.,  Jr.  1977.  Re- 
sponses of  small  mammals  to  the 
clearcutting  of  northern  Appala- 
chian forests.  Journal  of  Mam- 
malogy 58:600-609. 


Table  7.— Classification  of  1 68  trapping  stations  in  Grids  ]  -4  as  locations 
where  each  of  seven  small  mammal  species  Is  present  or  absent  based  on 
logistic  regression  models,  and  agreement  between  predicted  and  ob- 
sen^ed  classifications. 


No.  of  stations 


Species 

classified^ 

%  Correct* 

Agreement' 

Present  Absent 

Present  Absent 

K 

ASE  P 

Short-tailed  shrew 

106 

62 

71 

68 

0.371 

0.072  <0.001 

Masked  shrew 

45 

123 

67 

78 

0.412 

0.075  <0.001 

Smokey  shrew 

2 

166 

100 

96 

0.351 

0.183  NS 

Eastern  chipmunk 

7 

161 

71 

90 

0.314 

0.116  <0.01 

Deer  mouse 

167 

1 

95 

0 

0.011 

0.010  NS 

Red-backed  vole 

27 

141 

57 

72 

0.223 

0.076  <0.01 

Jumping  mouse 

165 

3 

88 

62 

0.133 

0.093  NS 

'Prior  probability  of  presence  =  0.05. 

'Percent  of  stations  wtiere  present/ absent  classificotion  agreed  withi  observations 
frorr)  trapping  in  1984. 

=  Kappa  statistic  (Fleiss  1973).  ASE  =  asymptotic  standard  error.  P  =  Probability 
tt>at  agreement  is  due  to  ctiance,  i.e.  K=0.  NS  =  not  significant 


326 


Kitchings,  J.  Thomas,  and  Douglas  J. 
Levy.  1981.  Habitat  patterns  in  a 
small  mammal  community.  Jour- 
nal of  Mammalogy  62:814-820. 

Lovejoy,  David  A.  1970.  The  effect  of 
logging  on  small  mammal  popula- 
tions in  the  White  Mountain  Na- 
tional Forest  of  New  Hampshire. 
Ph.D.  Dissertation.  173  p.  Univer- 
sity of  Connecticut,  Storrs. 

Martin,  C.  Wayne.  1977.  Distribution 
of  tree  species  in  an  undisturbed 
northern  hardwood-spruce-fir  for- 
est, the  Bowl,  New  Hampshire. 
USDA  Forest  Service  Research 
Note  NE-244,  Northeastern  Forest 
Experiment  Station,  Broomall, 
Pennsylvania. 

M'Closkey,  Robert  T.  1981.  Mi- 
crohabitat  use  in  coexisting  desert 
rodents — the  role  of  population 
density.  Oecologia  50:310-315. 

Miller,  Donald  H.,  and  Lowell  L. 
Getz.  1977.  Factors  influencing  lo- 
cal distribution  and  species  diver- 
sity of  forest  small  mammals  in 
New  England.  Canadian  Journal 
of  Zoology  55:806-814. 

Morris,  Douglas  W.  1984.  Patterns 
and  scale  of  habitat  use  in  two 
temperate-zone,  small  mammal 
faunas.  Canadian  Journal  of  Zool- 
ogy 62:1540-1547. 

Morris,  Douglas  W.  1987.  Tests  of 
density-dependent  habitat  selec- 
tion in  a  patchy  environment.  Eco- 
logical Monographs  57:269-281. 

Morrison,  Michael  L.  1985.  Influence 
of  sample  size  on  discrinninant 
function  analysis  of  habitat  use  by 
birds.  Journal  of  Field  Ornithology 
55:330-335. 

Noon,  Barry  R.  1984.  Summary:  Bi- 
ometric  approaches  to  modeling — 
the  researcher's  viewpoint,  p.  197- 
201.  In  Wildlife  2000:  Modeling 
habitat  relationships  of  terrestrial 
vertebrates.  Jared  Verner,  Michael 
L.  Morrison,  and  C.  John  Ralph, 
editors.  The  University  of  Wiscon- 
sin Press,  Madison. 


Parren,  Steven  G.  1981.  Habitat  selec- 
tion by  small  mammals  in  a  north- 
ern hardwoods  forest.  M.S.  Thesis. 
77  p.  University  of  Vermont,  Burl- 
ington. 

Parren,  Steven  G.,  and  David  E. 
Capen.  1985.  Local  distribution 
and  coexistence  of  Peromyscus  in 
Vermont.  Journal  of  Mammalogy 
66:36-44. 

Porter,  John  H.,  and  Raymond  D. 
Dueser.  1982.  Niche  overlap  and 
competition  in  an  insular  mammal 
fauna:  a  test  of  the  niche  overlap 
hypothesis.  Oikos  39:228-236. 

Press,  S.  James,  and  Sandra  Wilson. 
1978.  Choosing  between  logistic 
regression  and  discriminant  analy- 
sis. Journal  of  the  American  Statis- 
tical Association  73:699-705. 

Richens,  Voit  B.  1974.  Numbers  and 
habitat  affinities  of  small  mam- 
mals in  northwestern  Maine.  Ca- 
nadian Field-Naturalist  88:191-196. 

Schloyer,  Carroll  R.  1983.  Ecological 
relationships  of  sympatric  Pero- 
myscus in  western  Massachusetts. 
Ph.D.  Dissertation.  149  p.  Univer- 
sity of  Massachusetts,  Amherst. 

Schoener,  Thomas  W.  1974.  Resource 
partitioning  in  ecological  commu- 
nities. Science  185:27-39. 

Seagle,  Steven  W.  1985a.  Patterns  of 
small  mammal  microhabitat  utili- 
zation in  cedar  glade  and  decidu- 
ous forest  habitats.  Journal  of 
Mammalogy  66:22-35. 

Seagle,  Steven  W.  1985b.  Competi- 
tion and  coexistence  of  small 
mammals  in  an  east  Tennessee 
pine  plantation.  American  Mid- 
land Naturalist  114:272-282. 

Snedecor,  George  W.,  and  William  G. 
Cochran.  1980.  Statistical  methods. 
507  p.  Iowa  State  University  Press, 
Ames. 


Tyron,  C.  A.,  and  D.  P.  Snyder.  1973. 
Biology  of  the  eastern  chipmunk, 
Tamias  striatus:  life  tables,  age  dis- 
tributions, and  trends  in  popula- 
tion numbers.  Journal  of  Mam- 
malogy 54:145-168. 

Vickery,  W.  L.  1981.  Habitat  use  by 
northeastern  forest  rodents. 
American  Midland  Naturalist 
106:111-118. 


327 


The  Value  of  Rocky  Mountain 
Juniper  (Juniperus 
scopulorum)  Woodlands  in 
South  Dakota  as  Small 
Mammal  Habitat' 


Abstract.— Small  mammals  and  vegetation  were 
sampled  over  two  years  in  Rocky  Mountain  juniper 
woodlands  and  adjacent  grasslands  in  South 
Dakota.  Juniper  woodlands  provided  specialized 
habitat  for  two  woodland  species,  white-footed 
mice  and  bushy-tailed  woodrats,  and  attracted  a 
number  of  species  generally  associated  with 
grasslands. 


Carolyn  Hull  Sieg^ 


Native  woodlands  constitute  only  a 
small  percentage  of  the  total  land 
area  in  the  Northern  Great  Plains,  yet 
they  provide  critical  habitat  for  many 
wildlife  species.  Isolated  woodlands 
provide  a  sharp  contrast  with  adja- 
cent grasslands,  increasing  available 
cover,  vertical  structure,  and  habitat 
interspersion,  and,  hence,  the  num- 
ber of  potential  niches  available  for 
wildlife.  Research  on  the  value  of  na- 
tive woodlands  as  wildlife  habitat 
has  focused  mainly  on  wildlife  use  of 
deciduous  woodlands  (Faanes  1984, 
Gaines  and  Kohn  1982,  Hopkins  et  al. 
1986,  Uresk  1982),  although  the  im- 
portance of  Rocky  Mountain  juniper 
woodlands  for  mule  deer  (Odocoileus 
hemionus)  has  been  documented  (Sev- 
erson  1981,  Severson  and  Carter 
1978).  Information  on  small  mam- 
mals associated  with  Rocky  Moun- 
tain juniper  stands  is  limited  to  brief 
studies  conducted  in  North  Dakota 
(Hansen  et  al.  1980,  Hopkins  1983, 
Seabloom  et  al.  1978). 

Native  woodlands  in  the  Northern 
Great  Plains  are  limited  to  areas  of 
increased  moisture,  such  as  along 
streams  and  rivers,  and  to  areas  with 

'Paper  presented  af  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ,  July  19-21.  1988.) 

'Carolyn  Hull  Sieg  is  Research!  Wildlife 
Biologist,  USDA  Forest  Service,  Rocky  Moun- 
tain Forest  and  Range  Experiment  Station, 
located  at  Rapid  City,  Southi  Dakota. 
Headquarters  is  in  Fort  Collins,  in  associa- 
tion with)  Colorado  State  University. 


increased  topographic  variation. 
Rocky  Mountain  juniper  is  restricted 
to  areas  of  steep  topography,  such  as 
the  "Badlands"  of  North  and  South 
Dakota,  the  Black  Hills,  areas  along 
drainageways  of  major  rivers,  and 
areas  on  high  limestone  plateaus  in 
South  Dakota  and  Wyoming.  It  is 
more  likely  to  occur  on  steep,  north- 
facing  slopes,  and  is  often  associated 
with  soils  that  are  calcareous,  poorly 
developed,  and  shallow  (Powells 
1965). 

The  purpose  of  this  study  was  to 
characterize  small  mammal  species 
composition  and  distribution  in 
Rocky  Mountain  juniper  woodlands 
and  in  adjacent  mixed-grass  range- 
lands  in  the  Badlands  National  Park, 
southwestern  South  Dakota.  The  ob- 
jectives were  to  determine  if  the  pres- 
ence of  isolated  juniper  woodlands 
increased  mammal  species  richness 
of  the  area,  and  to  form  preliminary 
hypotheses  as  to  how  these  wood- 
lands function  as  small  mammal 
habitat. 


Study  Area  and  Methods 

The  study  area  is  in  Pennington 
County,  South  Dakota,  approxi- 
mately 15  km  south  of  the  town  of 
Wall,  in  Sage  Creek  Basin,  Badlands 
National  Park.  Elevation  ranges  from 
950  to  1000  m  above  sea  level.  An- 
nual precipitation  averages  36  cm, 
most  of  which  is  received  in  May, 
June,  and  July.  The  terrain  in  Bad- 


lands National  Park  is  typically 
rough  and  irregular,  with  steep 
bluffs  rising  above  floodplains  onto 
upland  grasslands.  Dense  stands  of 
Rocky  Mountain  juniper  occur  on 
steep,  north-facing  slopes  and  in 
draws.  Upland  grasslands  are  domi- 
nated by  western  wheatgrass 
(Agropyron  smithii),  green 
needlegrass  (Stipa  viridula), 
buffalograss  (Buchloe  dactyloides),  and 
blue  grama  (Bouteloua  gracilis). 

Eight  study  sites  were  established, 
four  in  Rocky  Mountain  junijjer 
woodlands  on  north-facing  slopes  in 
draws,  and  four  on  adjacent  grass- 
lands. Vegetation  and  relative  abun- 
dance of  small  mammals  were 
sampled  on  a  regular  basis  for  2 
years.  Plant  canopy  cover  on  grass- 
lands and  understory  cover  in  the 
juniper  woodlands  were  sampled  in 
June  and  August  of  both  sampling 
years.  Plant  canopy  cover,  by  species, 
was  estimated  in  150,  0.1-m^  quad- 
rats spaced  at  1-m  intervals  along 
three  permanent  50-m  transects  on 
each  site  (Daubenmire  1959).  Over- 
story  vegetation  in  Rocky  Mountain 
juniper  study  sites  was  sampled  in 
eight,  7-  by  7-m  macroplots  spaced  at 
30-m  intervals  on  each  site.  Tree  den- 
sities, heights,  diameters  (d.b.h.),  and 
crown  heights  of  all  trees  were  meas- 
ured. 

Small  mammal  abundance  was 
sampled  monthly  from  June  through 
October  in  both  years.  Forty  Sherman 
live  traps,  spaced  at  10-m  intervals 
along  two  permanent  200-m 


328 


transects,  were  set  on  each  study  site 
for  four  consecutive  nights,  after  one 
night  of  prebaiting.  Total  trap  effort 
was  6400  trap  nights  per  vegetation 
type  per  year.  Rolled  oats  mixed 


-with  peanut  butter  were  used  as  bait. 
Captured  animals  were  identified  by 
species  and  assigned  a  unique  4-digit 
number  by  toe  clipping  (Taber  and 
Cowan  1971). 


Differences  in  small  mammal 
numbers  and  vegetation  between  the 
two  vegetation  types  were  tested 
with  ref>eated  measures  analyses  of 
variance  (SPSS  1986).  Both  years 
were  combined  for  analyses.  Total 
unique  small  mammals  and  numbers 
of  each  species  were  analyzed  sepa- 
rately: trap  session  and  year  were 
within-subject  factors;  vegetation 
type  was  the  between-subject  factor. 
Total  plant  canopy  cover  was  the 
vegetation  parameter  analyzed:  sam- 
pling session  and  year  were  within- 
subject  factors;  vegetation  type  was 
the  between-subject  factor.  Homoge- 
neity of  variances  was  tested  with 
Bartlett's  Box  F  test;  variables  with 
heterogeneous  variances  were  log- 
transformed. 


Results 

Vegetation 

Overstory  vegetation  in  juniper 
woodlands  was  nearly  a  monocul- 
ture of  Rocky  Mountain  junip>er,  al- 
though an  occasional  green  ash  (Frax- 
inus  pennsylvanica)  tree  was  ob- 
served. Tree  density  averaged  260 
trees/ha  (+117  SD),  and  ranged  from 
an  average  of  160  to  380  trees/ha  on 
the  four  sites.  Tree  heights  ranged 
from  a  mean  of  2.8  to  3.1  m,  and  the 
crowns  extended  nearly  to  the 
ground,  averaging  approximately  2.3 
m  in  height.  The  diameters  of  the  ju- 
niper trees  were  small,  ranging  from 
a  mean  of  4.8  cm  to  7.6  cm. 

Total  plant  canopy  cover  of  under- 
story  vegetation  in  the  juniper  wood- 
lands was  lower  (P  <  0.01)  than  on 
grasslands.  Total  cover  in  the  juniper 
woodlands  averaged  25%  (table  1). 
Yellow  sweetclover  (Melilotus  offici- 
nales) was  the  most  common  under- 
story  plant,  then  stonyhills  muhly 
(Muhlenbergia  cuspidata)  and  littleseed 
ricegrass  (Oiyzopsis  micrantha). 
Shrubs  were  uncommon  in  juniper 
woodlands;  chokecherry  (Prunus  vir- 
giniam),  western  wild  rose  (Rosa 
luoodsii),  western  snowberry  (Sym- 


Table  1.— Two-year  average  percent  (±  SD)  canopy  cover  of  dominant 
species  in  four  Rocky  Mountain  juniper  woodlands  and  four  grassland 
sites,  Badlands  National  Park,  South  Dakota. 


Category 

Juniper 

Grassland 

Total  cover 

24.5 

5.5 

52.8  +  4.9 

Litter  cover 

44.6 

±  11.3 

39.5+  10.2 

Bareground 

36.7 

±11.8 

15.8  +  8.3 

Forbs 

Yellow  sweetclover  (Melilotus  officinalis) 

8.7 

+  4.9 

1.2+  1.0 

Russian  thistle  (Salsola  kali) 

<  1 

±<  1 

2.4  +  2.8 

Scarlet  globemallow  (Sphaeralcea  coccinea) 

<  1 

+  <  1 

2.9+1.8 

Grasses 

Western  wheatgrass  (Agropyron  smifhii) 

1.4 

+  1.4 

9.7  +  7.5 

Blue  grama  (Boufeloua  gracilis) 

<  1 

+  <  1 

5.4  +  4.0 

Cheoigross  (Bromus  fecforum) 

<  1 

±<  1 

2.5+1.5 

Buffalogross  (Buchloe  dacfyloides) 

<  1 

+  <  1 

3.4  +  3,5 

Threadleof  sedge  (Carex  filifolia) 

1.3 

+  2.1 

3.3  +  4,2 

Sun  sedge  (Carex  heliophila) 

<  1 

+  <  1 

3.0+1.5 

Stonyhills  muhly  (Muhlenbergia  cuspidata) 

3.4 

+  3.2 

0 

Littleseed  ricegrass  (Oryzopsis  micrantha) 

1.5 

+  1.4 

0 

Needleondthread  grass  (Sfipa  comafa) 

<  1 

+  <  1 

6.5  +  9.0 

Table  2— Two-year  average  number  (+ SD)  of  small  mammals  captured 
per  site  in  four  Rocky  Mountain  juniper  woodlands  and  four  adjacent 
grassland  sites,  BadlarKis  National  Park,  South  Dakota. 


Species 

Juniper 

Grassland 

Meadow  vole  (Microtus  pennsylvanicus) 

<l±l.la 

14.8  ±17.2^ 

House  mouse  (Mus  musculus) 

<  1  +  0.4° 

0° 

Bushy- tailed  woodrat  (Neotoma  cinerea) 

1.0±  ].2° 

0° 

Northern  grasshopper  mouse 

(Onychomys  leucogaster) 

1.0+1.1° 

4.9  +  3.3'^ 

Plains  pocket  mouse  (Perognafhus  flavescens) 

1.0+  1.8° 

1.0+  1.8° 

Hispid  pocket  mouse  (Perognafhus  h'ispidus) 

1.0  ±0.9° 

2.1  ±2.0° 

White-footed  mouse  (Peromyscus  leucopus) 

20.4+12.7'^ 

1.6  +  2.2° 

Deer  mouse  (Peromyscus  maniculafus) 

44.5  +  24.9° 

35.9  ±27,4° 

Western  harvest  mouse 

(Reithrodontomys  megalotis) 

<  1±1° 

2.9  ±3.5*^ 

Thirteen-lined  ground  squirrel 

(Spermophilus  tridecemlineafus) 

<  1  +  1° 

11.0±6.9^ 

Total 

68.5  +  27.4° 

74.1+21.9° 

'Means  in  a  row  followed  by  the  same  superscript  were  not  significantly  (P  >  0. 1) 
different 


329 


phoricarpos  occidentalis),  and 
skunkbush  sumac  (Rhus  aromatica) 
each  comprised  less  than  1%  of  the 
total  canopy  cover.  Litter  cover  in  the 
juniper  woodlands  averaged  45% 
and  bare  ground  30%. 

Total  plant  canopy  cover  on  grass- 
lands averaged  53%  (table  1).  West- 
ern wheatgrass  was  the  most  com- 
mon plant  species,  then 
needleandthread  grass  (Stipa  comata), 
blue  grama,  and  buffalograss.  Scarlet 
globemallow  (Sphaeralcea  coccinea) 
was  the  most  common  forb.  Shrub 
species  were  limited  to  fringed  sage 
(Artemisia  frigida)  and  dwarf 
sagebrush  (A.  cam),  each  comprising 
a  small  percentage  of  the  total  cover 
on  grasslands.  Mean  litter  cover  was 
40%  and  bare  ground  16%  over  the 
two  sampling  years. 

Small  Mammals 

Average  numbers  of  small  mammals 
were  similar  (P  =  0.4)  on  the  two 
vegetation  types;  however,  species 
composition  differed  between  juni- 
per woodlands  and  adjacent  grass- 
lands (table  2).  Deer  mice  (Pero- 
myscus  maniculatus)  were  the  most 
common  species  captured  in  both 
juniper  woodlands  and  on  grass- 
lands, constituting  66%  of  the  total 
capture  in  juniper  woodlands  and 
48%  on  grasslands.  Number  of  deer 
mice  captured  was  similar  (P  =  0.4) 
in  both  vegetation  types,  averaging 
42  and  36  individuals  per  site  in  juni- 
per woodlands  and  grasslands,  re- 
spectively. White-footed  mice  (P.  leu- 
copus)  were  the  next  most  abundant 
small  mammal  species  captured  in 
juniper  woodlands,  constituting  ap- 
proximately 29%  of  the  total  cap- 
tures; their  numbers  were  much 
lower  (P  =  0.04)  on  grassland  sites. 
Bushy-tailed  woodrats  (Neotoma  cin- 
erea)  were  captured  in  small  numbers 
in  the  juniper  woodlands  but  were 
absent  from  grasslands.  Average 
numbers  of  meadow  voles  (Microtus 
pennsylvanicus)  (P  =  0.03),  thirteen- 
lined  ground  squirrels  (Spermophilus 


tridecemlineatus)  (P  =  0.03),  northern 
grasshopper  mice  (Onychomys  leu- 
cogaster)  (P  =  0.06),  and  western  har- 
vest mice  (Reithrodontomys  megalotis) 
(P  =  0.08)  were  higher  on  grasslands 
than  in  juniper  woodlands.  Small 
numbers  of  plains  pocket  mice  (Per- 
ognathus  flavescens)  and  hispid  pocket 
mice  (P.  hispidus)  were  captured  in 
both  vegetation  types.  One  house 
mouse  (Mus  musculus)  was  captured 
in  a  juniper  woodland. 

Discussion 

Rocky  Mountain  juniper  stands  did 
not  support  significantly  higher  num- 
bers of  small  mammals  than  did  ad- 
jacent grasslands,  but  enhanced 
small  mammal  diversity  by  provid- 
ing specialized  habitat  for  white- 
footed  mice  and  bushy-tailed 
woodrats.  White-footed  mice  prefer 
and  are  commonly  restricted  to  ri- 
parian forests  and  shrubby  habitats 
in  this  region  (Armstrong  1972,  Sea- 
bloom  et  al.  1978),  and  were  a  com- 
mon species  in  Rocky  Mountain  juni- 
per woodlands  in  North  Dakota 
(Hopkins  1983).  White-footed  mice 
forage  (M'Closkey  1975)  and  nest 
(Wolff  and  Hurlbutt  1982)  in  trees 
and  show  a  tendency  to  use  woody 
vegetation  as  escape  routes  (Barry 
and  Francq  1980).  Their  preferred 
habitat  is  often  characterized  by 
dense  woody  understory  (Yahner 
1982).  Rocky  Mountain  juniper 
woodlands  lack  vertical  layering  pro- 
vided by  shrubs,  but  the  dense  tree 
canopy  and  presence  of  branches 
nearly  to  the  ground  may  substitute 
for  shrub  layers  found  in  other 
woodlands.  Further,  juniper  wood- 
lands may  function  as  dispersal  path- 
ways for  woodland  species  such  as 
white-footed  mice.  Turner  (1974) 
postulated  that  riparian  habitats 
along  major  drainageways  allowed 
the  western  expansion  of  the  white- 
footed  mouse. 

Bushy-tailed  woodrats  are  often 
restricted  to  rocky  areas  in  this  re- 
gion (Jones  et  al.  1983),  and  their 


presence  has  been  documented  in 
deciduous  woodlands  in  northwest- 
ern South  Dakota  (Hodorff  et  al.  In 
Press).  Bushy-tailed  woodrats  were 
captured  in  ponderosa  pine  (Pinus 
ponderosa)  stands,  toe  slopes,  hilly 
scoria,  and  upland  breaks  in  western 
North  Dakota  (Seabloom  et  al.  1978). 
Juniper  stands  likely  provide  den 
sites,  which  grasslands  lacked.  Mid- 
dens constructed  of  juniper  branches 
were  observed  in  three  of  four  Rocky 
Mountain  juniper  sites  in  this  study. 

Three  species — deer  mice,  plains 
pocket  mice,  and  hispid  pxKket 
mice — apparently  showed  no  prefer- 
ence between  grasslands  or  juniper 
woodlands.  The  high  proportion  of 
deer  mice  in  the  total  capture  on  both 
grasslands  and  in  juniper  woodlands 
is  not  uncommon  on  the  Northern 
Great  Plains.  Deer  mice  are  a  ubiqui- 
tous species,  occurring  in  nearly  ev- 
ery habitat  in  this  region  (Jones  et  at. 
1983).  Deer  mice  were  the  most  com- 
monly captured  species  in  green  ash 
woodlands  in  northwestern  South 
Dakota  (Hodorff  et  al.  In  Press),  and 
were  abundant  in  both  green  ash  and 
Rocky  Mountain  juniper  woodlands 
in  western  North  Dakota  (Hopkins 
1983).  Rocky  Mountain  junif)er 
woodlands  in  South  Dakota  are 
probably  not  critical  habitat  for  deer 
mice,  but  when  available,  will  be  ex- 
ploited by  this  adaptive  species. 

Hispid  pocket  mice  apparentiy 
prefer  rocky  areas,  where  a  variety  of 
shrubs,  forbs,  and  yucca  (Yucca  spp.) 
grow  (Jones  et  al.  1983).  Plains  pocket 
mice  are  considered  rare  mammals  in 
South  Dakota  (Houtcooper  et  al. 
1985);  hence  little  is  known  about  the 
distribution  and  habitat  preferences 
of  this  species  in  the  state.  Hodorff  et 
al.  (In  Press)  captured  low  numbers 
of  both  plains  and  hispid  pocket  mice 
in  green  ash  woodlands  in  north- 
western South  Dakota.  Haufler  and 
Nagy  (1984)  captured  plains  pocket 
mice  in  pinyon  pine  (Pinus  edulis)- 
Utah  juniper  (/.  osteosperma)  wood- 
lands in  Colorado,  and  reported  that 
juniper  comprised  17%  of  the  pocket 
mouse's  diet.  The  small  captures  of 


330 


both  species  of  pocket  mice  make 
generalizations  about  habitat  prefer- 

'  ence  suspect,  but  Rocky  Mountain 
juniper  woodlands  likely  provided 
habitat  interspersion  and  food  re- 

!     sources  for  these  species. 

;        Juniper  woodlands,  with  sparse 
understory  cover,  are  atypical  habitat 
for  grassland  inhabitants  such  as 
meadow  voles,  thirteen-lined  ground 
squirrels,  northern  grasshopper  mice, 
and  western  harvest  mice.  Meadow 
voles,  in  particular,  are  generally  as- 
sociated with  dense  stands  of  grass 
(Birney  et  al.  1976).  However,  Rocky 
Mountain  juniper  woodlands  in 
southwestern  North  Dakota  sup- 
ported meadow  voles  in  some  areas 
(Seabloom  et  al.  1978)  and  prairie 
voles  (M.  ochrogaster)  on  other  sites 
(Hopkins  1983). 

The  ability  of  North  Dakota  juni- 

1     per  woodlands  to  support  microtines 
was  attributed  to  differences  in  plant 
community  attributes.  Littleseed 
ricegrass  and  mosses  dominated  the 
understory  and  total  plant  cover  av- 
eraged over  60%  (vs.  25%  in  South 
Dakota)  in  most  juniper  stands 
sampled  by  Hopkins  (1983)  (Hansen 
et  al.l984).  The  more  dense  under- 
story of  the  North  Dakota  wood- 

I    lands,  which  South  Dakota  wood- 
lands lacked,  apparently  provided 

(adequate  cover  for  microtines. 
Thirteen-lined  ground  squirrels 
were  most  frequently  captured  in 
northwestern  South  Dakota  in  road- 
ways and  fencerows  in  shortgrass 
prairies  (Andersen  and  Jones  1971). 
Northern  grasshopper  mice  are  gen- 
erally restricted  to  shortgrass  and 
desert  sites  (McCarty  1978),  in  areas 
with  adequate  dust-bathing  sites 
I)    (Egoscue  1960).  Western  harvest 
mice  were  occasionally  captured  in 
pinyon-juniper  woodlands  in  south- 
eastern Colorado,  but  were  associ- 
ated with  dense  herbaceous  cover 
lacking  tree  canopy  cover  (Ribble 
and  Samson  1987).  Rocky  Mountain 
juniper  woodlands  may  provide  sup>- 
plemental  food  resources  for  small 
mammals  generally  restricted  to 
grasslands. 


Conclusion 

Rocky  Mountain  juniper  woodlands 
enhance  small  mammal  richness  of 
the  generally  treeless  Northern  Great 
Plains  by  providing  specialized  habi- 
tat for  at  least  two  species,  bushy- 
tailed  woodrats  and  white-footed 
mice.  Juniper  woodlands  lack  well- 
developed  shrub  layers,  but  the 
dense  canopy  of  the  juniper  trees  and 
crowns  that  extend  nearly  to  the 
ground  may  provide  foraging  and 
nesting  substrates  for  woodland 
mammals.  Further,  Rocky  Mountain 
juniper  woodlands  may  function  as 
dispersal  pathways  for  these  two 
species.  Juniper  woodlands  lack 
dense  herbaceous  understories  neces- 
sary to  support  microtines  such  as 
meadow  voles,  but  likely  serve  as 
food  resource  supplemental  areas  for 
a  variety  of  mammals  associated 
with  grasslands.  Adaptable  species 
such  as  the  deer  mouse  may  not  re- 
quire juniper  woodlands,  but  will 
exploit  this  habitat  when  available. 
Finally,  Rocky  Mountain  juniper 
woodlands  may  figure  into  the  habi- 
tat needs  of  pocket  mice,  but  low 
captures  of  two  species  preclude 
clear  definition  of  preferred  habitat. 

Acknowledgments 

Critical  reviews  by  Dan  Uresk,  Deb 
Paulson,  Dick  Hansen,  and  Bill  Clark 
were  helpful  in  improving  this 
manuscript.  Personnel  at  Badlands 
National  Park  were  most  cooperative 
during  this  study.  Deb  Paulson  and 
Bob  Hodorff  helped  with  field  work. 

Literature  Cited 

Andersen,  Kenneth  W.,  and  J.  Knox 
Jones,  Jr.  1971.  Mammals  of  north- 
western South  Dakota.  University 
of  Kansas  Museum  of  Natural  His- 
tory 19:361-393. 

Armstrong,  David  M.  1972.  Distribu- 
tion of  mammals  in  Colorado. 
415  p.  Monograph  No.  3,  Museum 


of  Natural  History,  University  of 
Kansas,  Lawrence,  KS. 

Barry,  Ronald  E.,  Jr.,  and  Edward  N. 
Francq.  1980.  Orientation  to  land- 
marks within  the  preferred  habitat 
by  Peromyscus  leucopus.  Journal  of 
Mammalogy  61:292-303. 

Birney,  Elmer  C,  W.E.  Grant,  and 
Donna  Day  Baird.  1976.  Impor- 
tance of  vegetation  cover  to  cycles 
of  Microtus  populations.  Ecology 
57:1043-1051. 

Daubenmire,  R.  1959.  A  canopy-cov- 
erage method  of  vegetation  analy- 
sis. Northwest  Science  33:43-64. 

Egoscue,  H.  J.  1960.  Laboratory  and 
field  studies  of  the  northern  grass- 
hopper mouse.  Journal  of  Mam- 
malogy 51:622-623. 

Faanes,  Craig  A.  1984.  Wooded  is- 
lands in  a  sea  of  prairie.  American 
Birds  38:3-6. 

Powells,  H.A.  1965.  Rocky  Mountain 
Juniper  {Juniperus  scopulorum 
Sarg.).  In  silvics  of  Forest  trees  of 
the  United  States.  Agricultural 
Handbook  No.  271.  p.  217-222. 
Washington,  D.C. 

Gaines,  Ronald  C,  and  Stanley  C. 
Kohn.  1982.  Raptor  use  of  hard- 
wood draws  in  central  North  Da- 
kota. Prairie  Naturalist  14:55-58. 

Hansen,  Paul  L.,  George  R.  Hoffman, 
and  Ardell  J.  Bjugstad.  1984.  The 
vegetation  of  Theodore  Roosevelt 
National  Park,  North  Dakota:  a 
habitat  type  classification.  General 
Technical  Report  RM-113.  35  p. 
Fort  Collins,  CO:  U.S.  Department 
of  Agriculture,  Forest  Service, 
Rocky  Mountain  Forest  and  Range 
Experiment  Station. 

Hansen,  Paul  L.,  Ricky  B.  Hopkins, 
and  George  R.  Hoffman.  1980.  An 
ecological  study  of  the  habitat 
types  and  their  animal  compx)- 
nents  at  Theodore  Roosevelt  Na- 
tional Park,  North  Dakota.  182  p. 
Department  of  Biology,  Vermil- 
lion, University  of  South  Dakota. 

Haufler,  Jonathan  B.,  and  Julius  G. 
Nagy.  1984.  Summer  food  habits 
of  a  small  mammal  community  in 
the  pinyon-juniper  ecosystem. 
Great  Basin  Naturalist  44:145-150. 


331 


Hodorff,  Robert  A.,  Carolyn  Hull 
Sieg,  and  Raymond  L.  Linder.  In 
Press.  Wildlife  response  to  stand 
structure  of  deciduous  wood- 
lands. Journal  of  Wildlife  Manage- 
ment. 

Hopkins,  Rick  B.  1983.  Small  mam- 
mal component  of  habitat  types  in 
Theodore  Roosevelt  National 
Park.  Prairie  Naturalist  15:45-46. 

Hopkins,  Rick  B.,  Frank  J.  Cassel, 
and  Ardell  J.  Bjugstad.  1986.  Rela- 
tionships between  breeding  birds 
and  vegetation  in  four  woodland 
types  of  the  Little  Missouri  Grass- 
lands. Research  Paper  RM-270. 12 
p.  Fort  Collins,  CO:  U.S.  Depart- 
ment of  Agriculture,  Forest  Serv- 
ice, Rocky  Mountain  Forest  and 
Range  Experiment  Station. 

Houtcooper,  Wayne  C,  David  J. 
Ode,  John  A.  Pearson,  and  George 
M.  Vandel,  III.  1985.  Rare  animals 
and  plants  of  South  Dakota.  Prai- 
rie Naturalist  17:143-165. 

Jones,  J.  Knox,  Jr.,  David  M.  Arm- 
strong, Robert  S.  Hoffman,  and 
Clyde  Jones.  1983.  Mammals  of 
the  Northern  Great  Plains.  379  p. 
Lincoln,  NE:  University  Nebraska 
Press. 

McCarty,  Richard.  1978.  Onychomys 
leucogaster.  Mammal  Species  87:1- 
6.  American  Society  of  Mammalo- 
gists. 

M'Closkey,  Robert  T.  1975.  Habitat 
dimensions  of  white-footed  mice, 
Peromyscus  leucopus.  American 
Midland  Naturalist  93:158-167. 

Ribble,  David  O.,  and  Fred  B.  Sam- 
son. 1987.  Microhabitat  associa- 
tions of  small  mammals  in  south- 
eastern Colorado,  with  special 
emphasis  on  Peromyscus  (Roden- 
tia).  Southwestern  Naturalist 
32:291-303. 

Seabloom,  Robert  W.,  Richard  D. 
Crawford,  and  Michael  G.  McK- 
enna.  1978.  Vertebrates  of  south- 
western North  Dakota:  amphibi- 
ans, reptiles,  birds,  mammals.  Re- 
search Report  24, 549  p.  Grand 
Forks,  ND:  University  of  North 
Dakota;  Institute  of  Ecological 
Studies. 


Severson,  Kieth  E.  1981.  Plains  habi- 
tat. In  Wallmo  O.C.  Mule  and 
black-tailed  deer  in  North  Amer- 
ica, p.  459-468.  Lincoln,  NE:  Uni- 
versity of  Nebraska  Press  and 
Washington,  D.C.,  Wildlife  Man- 
agement Institute. 

Severson,  Kieth  E.,  and  Arthur  V. 
Carter.  1978.  Movements  and 
habitat  use  by  mule  deer  in  the 
Northern  Great  Plains,  South  Da- 
kota, p.  466-468.  In  D.N.  Hyder, 
editor.  Proceedings  of  the  First 
International  Rangeland  Congress. 
[Denver,  CO,  August  14-18, 1978] 
Society  for  Range  Management. 
Peerless  Printing,  Denver,  CO. 

SPSS.  1986.  SPSSX  user's  guide.  988 
p.  New  York,  N.Y.  McGraw-Hill 
Book  Company. 

Taber,  Richard  D.,  and  McT.  I. 
Cowan.  1971.  Capturing  and 
marking  wild  animals,  p.  277-317. 
In  Wildlife  Management  Tech- 
niques, 3rd  edition.  R.H.  Giles  Jr., 
editor.  Washington,  D.C.,  Wildlife 
Society. 

Turner,  Ronald  W.  1974.  Mammals  of 
the  Black  Hills  of  South  Dakota 
and  Wyoming.  Miscellaneous  Pub- 
lication No.  60, 178  p.  Lawrence, 
KS:  University  of  Kansas,  Museum 
of  Natural  History. 

Uresk,  Daniel  W.  1982.  Importance  of 
woodlands  to  wildlife  and  live- 
stock use  on  the  Northern  High 
Plains,  p.  7-12.  In  North  Platte, 
NE:  Proceedings  of  the  Great 
Plains  Agricultural  Council. 

Wolff,  Jerry  O.,  and  Bethia,  Hurlbutt. 
1982.  Day  refuges  of  Peromyscus 
leucopus  and  Peromyscus  manicula- 
tus.  Journal  of  Mammalogy  63:666- 
668. 

Yahner,  Richard  H.  1982.  Microhabi- 
tat use  by  small  mammals  in  farm- 
stead shelterbelts.  Journal  of 
Mammalogy  63:440-445. 


332 


Postfire  Rodent  Succession 
Following  Prescribed  Fire  In 
Southern  California 
ChaparraP 

William  O.  Wirtz,  11,^  David  Hoekman,^  John 
R.  Muhm/  and  Sherrie  L.  Souza^ 


Abstract.— This  paper  describes  species 
composition  and  density  changes  in  rodent 
populations  during  postfire  succession  following 
prescribed  fire  in  the  chaparral  community  of  the 
San  Gabriel  Mountains.  Conclusions  are  drawn  from 
a  4-year,  live-trap,  mark  and  release  study  of  postfire 
succession  in  two  watersheds  receiving  "hot"  burns 
and  two  receiving  "normal"  burns. 


The  chaparral  community  of  south- 
ern California  is  associated  with 
nearly  two  million  years  of  fire  his- 
tory (Hanes  1971).  In  recent  centuries 
major  fires  have  occurred  at  intervals 
of  20  to  40  years  (Byrne  et  al.  1977; 
Philpot  1977).  Postfire  plant  succes- 
sion (Patric  and  Hanes  1964,  Hanes 
and  Jones  1967,  Hanes  1971)  and  the 
fire  itself  have  varying  short  term 
effects  on  the  birds  and  small  mam- 
mals found  in  the  chaparral  (Law- 
rence 1966,  Quinn  1979,  Wirtz  1977, 
1979).  Wirtz  (1977)  summarized  the 
work  of  earlier  authors  concerning 
conditions  in  small  vertebrate  mi- 
crohabitats  during  fire,  vertebrate  be- 
havior during  fire,  and  survival  of 
small  vertebrates  exposed  to  fire. 
Both  Lawrence  (1966)  and  Quinn 
(1979)  studied  rodent  populations 
before  and  after  a  bum,  in  addition 
to  documenting  microhabitat  condi- 
tions during  the  fire.  Wirtz  (1977, 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff, AZ,  July  19-21.  1988.) 

'William  O.  Wirtz.  II  is  Professor  of  Biology. 
Department  of  Biology.  Pomona  College. 
Claremont.  CA  91711. 

^David  Hoekman  is  a  research  assistant. 
Department  of  Biology,  Pomona  College, 
Claremont.  CA9171I. 

"John  R.  Muhm  is  a  research  assistant. 
Department  of  Biology.  Pomona  College. 
Claremont.  C A  91711. 

^Sherrie  L.  Souza  is  a  research  assistant. 
Department  of  Biology.  Pomona  College. 
Claremont.  CA  91711. 


1982, 1984)  presents  preliminary 
analyses  of  data  collected  on  postfire 
rodent  succession  following  wildfire 
in  the  chaparral  community  of  south- 
ern California. 

Because  of  the  recently  recognized 
significance  of  the  use  of  prescribed 
fire  in  the  management  of  chaparral 
ecosystems,  the  Pacific  Southwest 
Forest  and  Range  Experiment  Sta- 
tion, USD  A  Forest  Ser  vice,  began  for- 
mulating plans  in  1983  for  a  series  of 
prescribed  fires  in  the  San  Dimas  Ex- 
perimental Forest,  located  in  the  San 
Gabriel  Mountains  of  southern  Cali- 
fornia, that  might  be  utilized  for  long 
range  studies  of  the  effects  of  pre- 
scribed fire  in  chaparral.  In  October, 
1984,  the  Forest  Service  burned  four 
chaparral  watersheds  of  approxi- 
mately 40  ha  each  in  the  San  Dimas 
Experimental  Forest.  This  paper  de- 
scribes the  changes  in  rodent  com- 
munity structure  for  the  4-year  pe- 
riod following  prescribed  burning. 

Methods 

In  October,  1984,  four  chaparral  wa- 
tersheds of  approximately  40  ha  each 
were  subjected  to  prescribed  bums  in 
the  San  Dimas  Experimental  Forest. 
The  vegetation  of  the  two  of  these 
watersheds  (874  and  775)  had  been 
hand  cut  in  the  spring  of  1984  to  pro- 
duce the  dried  fuel  for  an  exception- 
ally hot  fire.  Two  adjacent  water- 
sheds (804  and  776)  burned  normally 


for  climatic  conditions  at  the  time.  A 
fifth  watershed  (803),  which  has  been 
extensively  studied  since  1976  (see 
Wirtz  1977, 1979, 1982, 1984),  serves 
as  a  control  for  studies  on  the  pre- 
scribed bum  areas. 

Rodent  live-trap,  mark  and  re- 
lease, studies  were  conducted  on  all 
experimental  areas  prior  to  the  bums 
to  document  the  size  and  species 
composition  of  the  prefire  rodent 
community  on  all  watersheds,  and 
175  individuals  were  permanently 
marked  by  toe-clipping  to  provide  a 
prefire  pool  of  marked  rodents  from 
which  to  determine  survival  rates 
following  the  bum.  Following  the  fire 
grids  of  50  stations  at  15  m  intervals 
were  established  in  each  of  the  four 
watersheds  on  the  sites  of  the  prefire 
censusing,  and  a  live-trap,  mark  and 
release,  program  was  initiated  to  de- 
termine fire  survival  and  postfire  ro- 
dent succession  patterns. 

For  this  paper,  population  esti- 
mates were  done  by  the  Hayne  (1949) 
equation.  Area  sampled,  for  each 
species,  for  each  month,  was  esti- 
mated by  determining  the  mean  dis- 
tance traveled  for  each  species  be- 
tween captures,  for  each  month,  and 
then  adding  a  zone  equal  to  the  mean 
distance  travelled  to  the  perimeter  of 
the  grid.  Biomass  was  determined  by 
the  product  of  the  estimated  popula- 
tion times  the  mean  weight  for  each 
species  for  the  month,  and  these  val- 
ues are  then  summed  for  all  species 
taken  on  the  grid  for  the  month. 


333 


Results 

Postfire  trapping  was  initiated  in 
February  1985,  and  both  experimen- 
tal and  control  plots  were  sampled 
bi-monthly.  Hayne  equation  popula- 
tion estimates  for  rodent  populations 
on  each  study  plot  are  presented  in 
figures  1-5,  The  absence  of  data 
fX)ints  from  February  through  April 
or  May  means  that  no  rodents  were 
trapped,  except  for  watershed  803  in 
which  trapping  was  not  begun  until 
June  1985. 

Mice  of  the  genus  Peromyscus 
(deer  mouse,  P.  maniculatus;  brush 
mouse,  P.  boylii;  California  mouse,  P. 
calif ornicus),  and  California  pocket 
mice,  Perogmthus  californicus,  consti- 
tute the  bulk  of  the  postfire  rodent 
population.  Pacific  kangaroo  rat,  Di- 
podomys  agilis,  dusky-footed  wood 
rat,  Neotoma  fuscipes,  and  California 
vole,  Microtus  californicus,  are  present 
in  low  numbers,  and  a  few  Botta's 
pocket  gopher,  Thomomys  bottae,  and 


1985  1066  1987  1088 


Figure  1  .—Hayne  equation  estimates  of 
population  size  of  rodent  species  in  Bell 
803,  ttie  28-year-old  ctiaparral  control  plot. 


.or 

"  L 

20 
N  10 
0 


■e 

70 
«0 

so 

N  40 

30 

20 
10 
0 


p.  boylii 


p.  maniculslus 


D.  •gilit 


•T"i" 


Perognathus i 


A 
\  V 


o  jumjsnjmmjsnjmmjsnjmmj 
1985        .        1986        .        1987        •  1988 

Figure  2.— Hayne  equation  estimates  of 
population  size  of  rodent  species  in  Bell 
804,  normal  prescribed  burn.  Note  ttiat 
points  at  0  on  tt>e  x-axis  against  ttie  y-axis 
are  populations  estimates  prefire. 


'r    0.  agilit 

 I  -r-i-T 


O  JMMJSHJMMJ8NJMMJSNJWU 

198S       •        1986        •        1987        •  1988 

Figure  4.— Hayne  equation  estirrtates  of 
population  size  of  rodent  species  in  San 
Dimas  776,  nornnal  prescribed  burn.  Note 
that  points  at  0  on  the  x-axis  against  the  y- 
axis  are  populations  estimates  prefire. 


N  30 


P.cali'ornic  jg 


P.  boylii 


\   /  p\maniculatu»  v. 


O  JMMJ9NJMMJSNJMUJSNJMMJ 

1985        •        1986       •        1987        •  1988 

Figure  3.— Hayne  equation  estimates  of 
populations  size  of  rodent  species  in  Bell 
874,  hot  prescribed  burn.  Note  that  points  at 
0  on  the  x-axis  against  the  y-axis  are  popu- 
lations estimates  prefire. 


N  30 


0  »     ■  —l—t' 


h  II 

A  Parognathua       >  Ml 
!/       V  •- •'^       D.  aglris  ^ 


Naotoma 


10  r 

I  MIcrolua 


O  JMMJSNJMMJSNJMMJSHJMM 
1986        •        19B«        •        1987        •  1988 

Figure  5.— Hayne  equation  estimates  of 
population  size  of  rodent  species  in  San 
Dimas  775,  hot  prescribed  bum.  Note  thiot 
points  at  0  on  the  x-axis  against  the  y-axis 
are  populations  estirrKites  prefire. 


334 


western  harvest  mouse,  Reithrodonto- 
mys  megalotis,  have  also  been  taken. 
Larger  mammals  observed  in  burned 
watersheds,  for  which  no  quantita- 
tive data  are  available,  included 
Beechey  ground  squirrel,  Spermophi- 
lus  beecheyi,  Audubon's  cottontail, 
Sylvilagus  auduboni,  brush  rabbit,  S. 
bachmani,  coyote,  Canis  latrans,  black 
bear,  Ursus  americanus,  badger.  Tax- 
idea  taxus,  and  mule  deer,  Odocoileus 
hemionus. 


Fire  Survival 

No  marked  wood  rats  survived  the 
fires.  Nine  (12.5%)  Peromyscus  sur- 
vived normal  fires,  and  one  (1.4%) 
survived  hot  fires.  Tow  (12.5%) 
Pocket  mice  survived  normal  fires, 
and  two  (12.5%)  Survived  hot  fires. 
These  data  support  the  currently 
held  opinion  that  some  rodents  do 
survive  fires,  and  help  provide  the 
nucleus,  along  with  immigration 
from  unburned  areas,  for  rodent 
postfire  succession. 

Larger  mammals  seen  in  the 
burned  watersheds  in  the  first  month 
postfire  included  coyote,  black  bear, 
badger,  and  mule  deer. 

Early  Postfire  Succession 

Pocket  mice  and  all  three  Peromyscus 
species  were  present  on  one  hot  burn 
(874)  by  April  1985,  six  months 
postfire,  but  no  rodents  were  present 
on  the  other  hot  burn  (775).  Pocket 
mice  moved  into  this  hot  burn  (775) 
by  May,  and  two  Peromyscus  species, 
(P.  californicus,  P.  maniculatus)  were 
present  by  July. 

Pacific  kangaroo  rats  appeared  on 
some  burned  areas  by  June  or  July 
1985  (they  are  rare  in  mature  chapar- 
ral). Woodrats  appeared  on  one  nor- 
mal bum  (804)  by  June  1985,  and  an- 
other (776)  by  September  1985,  and 
on  one  hot  burn  (874)  by  August 
1985.  Single  pocket  gophers  and  har- 
vest mice  have  been  taken  on  one  hot 
bum  (775). 


Demography 

Sampling  was  not  begun  on  the  con- 
trol plot  (803)  until  June  1985.  The 
rodent  population  on  this  plot  con- 
sists chiefly  of  wood  rats,  California 
mice,  and  pocket  mice  (fig.  1).  The 
California  mouse  population  peaked 
during  the  fall,  winter,  and  spring  of 
1985-86,  and  again  in  the  winter  and 
spring  of  1986-1987.  Pocket  mice 
were  rare  on  the  control  until  the  fall 
of  1986  and  remained  common  until 
the  summer  of  1987  (fig.  1).  The 
wood  rat  population  has  peaked  in 
each  summer  studied  to  date. 

The  prefire  rodent  population  on 
the  normal  bum  in  Bell  (804)  was 
composed  primarily  of  woodrats, 
with  smaller  numbers  of  other  spe- 
cies (fig.  2)  (note  that  symbols  at  0  on 
the  X-axis  against  the  y-axis  represent 
prefire  density  estimates).  The 
postfire  rodent  population  on  this 
grid  has  been  composed  primarily  of 
brush  mice  and  pocket  mice,  with 
population  peaks  of  the  latter  in  each 
winter  (1985, 1986,  and  1987).  Wood 
rat  populations  did  not  show  signifi- 
cant increases  on  this  grid  until  the 
spring  of  1987,  about  30  months  after 
the  burn,  and  they  have  yet  (June 
1988)  to  reach  prefire  densities  (fig. 
2).  Pacific  kangaroo  rats  have  oc- 
curred on  this  bumed  area  in  num- 
bers above  prefire  densities  since  the 
summer  of  1985.  Brush  and  Califor- 
nia mouse  populations  have  oc- 
curred in  numbers  above  prefire  den- 
sities since  the  winter  of  1985-86  (fig. 
2). 

The  prefire  rodent  population  on 
the  hot  bum  in  Bell  (874)  was  com- 
posed largely  of  wood  rats,  Califor- 
nia mice,  and  pocket  mice  (fig.  3).  All 
species,  except  kangaroo  rats,  were 
present  again  on  this  grid  by  August 
1985, 10  months  postfire.  The  postfire 
rodent  community  on  this  hot  burn 
has  been  dominated  by  brush  mice 
and  pocket  mice  (fig.  3),  with  both 
species  reaching,  or  exceeding,  pre- 
fire densities  by  the  winter  of  1985, 
approximately  a  year  after  the  burn. 
Califomia  mouse  and  wood  rat 


populations  have  yet  (June  1988)  to 
reach  prefire  densities  (fig.  3). 

The  prefire  rodent  population  on 
the  normal  bum  in  San  Dimas  (776) 
was  composed  primarily  of  Califor- 
nia mice  and  wood  rats,  with  smaller 
numbers  of  pxxket  mice  and  no 
brush  mice  (fig.  4).  The  postfire  ro- 
dent community  has  been  dominated 
by  Califomia  mice  and  pocket  mice, 
with  both  species  exceeding  prefire 
densities  by  the  winter  of  1985,  ap- 
proximately one  year  postfire.  Wood 
rats  have  yet  (June  1988)  to  reach 
prefire  densities,  brush  mice  have  not 
appeared  on  this  grid,  and  Califomia 
voles  were  common  in  the  summer 
of  1987  and  the  spring  of  1988  (fig.  4). 

The  prefire  rodent  population  on 
the  hot  burn  in  San  Dimas  (775)  was 
very  similar  to  that  on  the  normal 
bum  here  (fig.  5).  And,  like  the  nor- 
mal bum,  the  postfire  community 
has  been  dominated  by  Califomia 
mice  and  p)ocket  mice,  with  pocket 
mice  exceeding  prefire  densities  by 
the  summer  of  1985  and  Califomia 
mice  exceeding  prefire  densities  by 
the  fall  of  1986  (fig.  5).  Pacific  kanga- 
roo rats  also  exceeded  prefire  densi- 
ties within  one  year  postfire  on  this 
grid. 

Comment  should  be  made  about 
the  presence  of  deer  mice  (P.  manicu- 
latus) and  Olifornia  voles  (Microtus) 
on  these  grids.  Neither  species  was 
present  on  any  grid  prefire,  and  nei- 
ther has  been  taken  on  the  control 
(figs.  1-5).  P.  maniculatus  has  been 
taken  on  all  burned  grids,  with  peaks 
of  abundance  by  the  second  year 
postfire  and  declining  abundance  by 
the  fourth  year  postfire  (figs.  4  and 
5). 

Effects  of  Hot  and  Normal  Fires 

The  effects  of  hot  and  normal  fires  on 
rodent  demography  were  examined 
by  (1)  comparing  pre  and  post  fire 
populations  in  areas  exposed  to  these 
two  fire  regimes  (figs.  6  and  7),  (2) 
comparing  the  number  of  captures  of 
each  species  postfire  under  each  fire 


335 


regime  (fig.  8),  and  (3)  comparing 
total  postfire  biomass  on  areas  ex- 
posed to  different  fire  regimes  (fig.  9) 
(note  again  that  points  at  0  on  the  x- 
axis  against  the  y-axis  are  prefire 
populations  estimates).  Only  species 
with  relatively  high  abundances  are 
considered  in  this  paper. 

Prefire  populations  of  brush  mice 
were  essentially  the  same  on  both 
areas  to  be  burned  in  Bell,  while  den- 
sities of  pocket  mice  and  California 
mice  were  greater  on  the  area  to  re- 
ceive the  hot  bum,  and  deer  mice 
were  not  present  on  either  grid  (fig. 
6).  All  prefire  populations  were  se- 
verely impacted  by  fire,  dropping  in 
most  instances  to  near  zero  for  sev- 
eral months  postfire.  Pocket  mice  in- 
creased to  twice  their  prefire  density 
on  the  hot  bum  and  25  times  prefire 
density  on  the  normal  burn  (fig.  6). 
Bmsh  mice  increased  to  14  times 
their  prefire  density  on  the  hot  burn 
and  six  times  prefire  density  on  the 
normal  bum  (fig.  6).  Califomia  mice 
returned  to  prefire  density  by  one 
year  postfire  on  the  normal  burn,  and 
numbers  have  remained  relatively 
constant  since  then.  Deer  mice  were 
present  on  both  burned  areas 
postfire,  but  have  been  more  abun- 
dant on  the  hot  bum  (fig.  6). 

Prefire  populations  of  Califomia 
mice  and  pocket  mice  were  similar 
on  both  areas  to  be  burned  in  San 
Dimas  (fig.  7).  Some  individuals  sur- 
vived the  normal  burn.  Pocket  mouse 
populations  exceeded  prefire  densi- 
ties on  both  normal  and  hot  bums  by 
eight  months  postfire  (fig.  7).  Califor- 
nia mouse  populations  exceeded  pre- 
fire densities  by  one  year  postfire  on 
the  normal  burn,  but  took  two  years 
to  reach  prefire  densities  on  the  hot 
bum  (fig.  7).  Two  species  not  present 
prefire.  Pacific  kangaroo  rats  and 
deer  mice,  colonized  both  burned 
areas  by  eight  months  postfire;  kan- 
garoo rats  have  remained  numerous 
on  the  hot  burn,  and  deer  mice  are 
more  numerous  on  the  hot  bum  than 
on  the  normal  bum  (fig.  9). 

Captures  of  Califomia  mice 
postfire  are  greater  on  normal  burns 


than  on  hot  burns,  and  exceed  cap- 
tures on  the  control  on  one  normal 
bum  (776)  (fig.  8).  Captures  of  bmsh 
mice  postfire  are  greater  on  both  hot 
bums  and  one  normal  burn  than  on 
the  control,  and  captures  on  hot 
bums  are  greater  than  on  normal 
bums  for  each  pair  of  watersheds 
bumed  (fig.  8).  Deer  mice  have  not 
been  captured  on  the  control;  cap- 
tures are  greater  postfire  on  hot 
bums  than  on  normal  burns  for  each 
pair  of  watersheds  burned  (fig.  8). 
Captures  of  wood  rats  are  less  on  all 
bumed  areas  than  on  the  control,  and 
they  are  less  on  hot  bums  than  on 
normal  burns  for  each  pair  of  water- 
sheds bumed  (fig.  8). 

California  voles  have  not  been 
taken  f)ostfire  on  the  control  nor  on 
one  normal  burn,  and  are  greater  on 
the  other  normal  burn  than  on  either 
hot  bum  (fig.  8).  Captures  of  Pacific 
kangaroo  rats  p>ostfire  are  greater  on 


"     I   ■    I    I  •  ■  -^r  I"  --  ■  - 

O  JMMJtNJMMJSNJMMJtNJMMJ 


1685       •        1»ae       •       1987       •  1888 

Figure  6.— Comparison  of  roder^t  postfire 
population  growthi  on  normal  (804)  and  hot 
(874  prescribed  fire  plots  in  Bell.  Note  thiat 
points  at  0  on  \he  x-axis  against  ttie  y-axis 
are  populations  estimates  prefire. 

336 


both  normal  and  one  hot  bum  than 
on  the  control,  while  captures  of 
pocket  mice  postfire  are  greater  on 
all  bumed  areas  than  on  the  control 
(fig.  8). 

Total  biomass  on  the  control,  not 
28  years  old,  has  fluctuated  during 
the  period  of  study,  but  shows  a 
slight  increasing  trend  (fig.  9).  Total 
biomass  on  both  bumed  plots  in  Bell, 
the  location  of  the  control,  has  also 
fluctuated,  with  a  slight  increasing 
trend,  in  a  fashion  similar  to  that  of 
the  control  (fig.  9).  Total  biomass  on 
the  bumed  plots  in  San  Dimas  has 
also  fluctuated,  with  slight  increasing 
trend,  but  with  two  dramatic  bio- 
mass increases,  one  in  the  Spring  of 

1987  and  the  other  in  the  spring  of 

1988  (fig.  9).  The  pattem  of  fluctua- 
tion, and  increase,  on  the  normal 
bum  in  San  Dimas  is  similar  to  that 
observed  for  the  control,  and  the  pat- 
tern of  fluctuation,  and  increase,  if 


1986       •        1988       •        1987       •  1986 


Figure  7.— Comparison  of  rodent  postfire 
population  growtti  on  normal  (776)  and  ho\ 
(775)  prescribed  fire  plots  in  San  Dioxis. 
Note  ttiat  points  at  0  on  \he  x-axis  against 


the  two  sharp  peaks  are  not  consid- 
ered, is  also  similar  to  the  control 
(fig.  9). 

Discussion 

General  patterns  of  rodent  postfire 
succession  following  these  prescribed 
bums  are  similar  to  those  reported 
by  Wirtz  (1977, 1982, 1984)  for  suc- 
cession following  wildfire  in  the 
chaparral  of  the  San  Gabriel  Moun- 
tains, but  lack  the  dramatic  increases 
in  density,  and  therefore  biomass, 
observed  in  these  earlier  studies.  He 
notes  (1984)  that  rodent  succession 
following  wildfire  takes  about  four 
years  before  populations  stabilize  at 
essentially  prefire  conditions  found 
in  older  chaparral  stands.  The  re- 
sponse of  species  to  these  prescribed 
fires  varied,  with  some  species  reach- 
ing prefire  densities  in  less  than  four 
years  and  others  having  not  yet 
reached  prefire  densities  at  essen- 
tially four  years  postfire. 

Only  slight  differences  are  noted 
between  rodent  postfire  succession 


on  normal  and  hot  bums,  and  these 
may  probably  be  attributed  to  differ- 
ences in  the  biology  of  individual 
species.  In  Bell,  both  normal  and  hot 
bums  were  dominated  postfire  by 
pocket  mice  and  brush  mice,  though 
pocket  mice  had  the  highest  density 
on  the  normal  burn  (804)  and  bmsh 
mice  had  the  highest  density  on  the 
hot  burn  (874)  (fig.  6).  Califomia 
mice  recovered  to  prefire  density  on 
the  normal  bum,  but  have  not  yet 
(June  1988)  recovered  on  the  hot 
burn,  and  wood  rats  have  not  recov- 
ered to  prefire  densities  on  either 
bumed  area  (fig.  6).  Deer  mice  have 
been  more  prevalent  on  the  hot  bum 
than  on  the  normal  burn  during  the 
period  of  the  study.  By  the  second 
year  postfire,  populations  of  all  spe- 
cies, except  wood  rats,  exceeded  pre- 
fire densities  on  the  normal  burn  (fig. 
2),  and  populations  of  brush  mice 
and  p>ocket  mice  had  exceeded  pre- 
fire densities  on  the  hot  bum  (fig.  3). 

In  San  Dimas,  where  considerable 
brush  was  left  alive  on  the  normal 
bum  (776),  both  normal  and  hot 
bums  were  dominated  postfire  by 


{XKket  mice  and  Califomia  mice  (fig. 
7).  Both  of  these  species  recovered  to 
prefire  densities  on  the  normal  burn 
by  one  year  postfire  (fig.  4),  as  did 
pocket  mice  on  the  hot  bum  (fig.  5), 
but  Califomia  mice  did  not  reach 
prefire  densities  on  the  hot  burn  until 
the  second  year  postfire  (fig.  5).  For 
reasons  not  immediately  apparent, 
but  probably  because  of  the  presence 
of  some  grass  prefire,  California 
voles  were  found  only  in  these  two 
watersheds  postfire.  The  greater  rela- 
tive abundance  of  Pacific  kangaroo 
rats  on  the  hot  burn  is  most  likely 
due  to  the  fact  that  more  open  space, 
necessary  for  kangaroo  rat  saltitorial 
locomotion,  was  left  by  the  hot  fire 
here. 

Pocket  mice  increase  rapidly  on 
bumed  areas,  there  being  essentially 
no  difference  between  normal  and 
hot  burns  (figs.  6  and  7).  Bmsh  mice, 
if  present  prefire,  recover  more  rap- 
idly p)ostfire  than  California  mice, 
and  the  latter  recover  more  rapidly 
on  normal  burns  than  on  hot  burns 
(figs.  6  and  7).  Deer  mice,  virtually 
nonexistent  in  mature  chaparral, 
colonize  both  normal  and  hot  bums, 
and  increase  more  rapidly  on  hot 
bums  (figs.  6  and  7). 

Data  on  captures  (fig.  8)  indicate 
that  increase  of  deer  mice  on  hot 
bums.  The  sf)ecies  is  known  to  colo- 
nize disturbed  areas,  whether  they  be 
caused  by  fire,  logging,  or  over- 


Figure  9.— Total  postfire  biorrxjss  (grams)  for 
control  and  burned  plots. 


800 


700 


eoo 


CO 

UJ  500 

QC 

3 

Q. 

<  400 

u 


300 


200 


100 


Peromyscus 
cahlornicus 


CONTROL 
NORMAL      '  » 
HOT  HW 


CNHNH       CNHNH  CNHNH 


Ptro-nyscut 


Pvromy acu* 
maniculaluS 


200 


100 


jQlL 


0  0 


CNHNH  CNHNH 


luselptt 


300 


200 


100 


OtpodoMyt 
■  gills 


Uicrotut 
oUlornleut 


CNHNH  CNHNH 


Parognathua 

ealliornicut 


Figure  8.— Connparison  of  postfire  captures  of  all  rodent  species  on  control  and  prescribed 
burn  plots. 


337 


grazing  (Williams  1955).  These  data 
also  illustrate  the  decline  of  Califor- 
nia mice  on  hot  bums  and  its  in- 
crease in  normal  burns,  and  the  in- 
crease of  brush  mice,  where  present 
prefire,  on  both  normal  and  hot 
bums.  Buming  favors  density  in- 
creases of  pocket  mice,  with  essen- 
tially no  difference  between  normal 
and  hot  burns.  Kangaroo  rats  exhibit 
variable  increases  in  response  to  fire, 
and  wood  rats  are  severely  impacted 
by  fire. 

Biomass  increases  in  response  to 
fire  are  variable,  and  in  this  study, 
were  similar  in  variability  to  those 
occurring  on  the  control  (fig.  9).  The 
sharp  peaks  in  biomass  observed  on 
one  hot  burn  (775)  are  due  to  large 
density  increases  in  pocket  mice  dur- 
ing these  periods. 

It  is  important  to  note,  when  com- 
paring data  for  normal  and  hot 
bums,  that  in  one  normal  burn  (776) 
a  lot  of  unburned  brush  remained, 
perhaps  more  accurately  simulating 
an  "island"  in  a  bum  rather  than  a 
burn  per  se.  So,  for  this  study,  the 
data  for  776  are  somewhat  atypical, 
and  804  represents  more  accurately 
the  situation  following  a  normal 
bum.  But  it  is  also  important  to  note 
that  "islands"  of  unburned  vegeta- 
tion are  frequently  left  by  wildfire, 
providing  refugia  for  both  plants  and 
animals  from  fire. 

Several  general  conclusions  may 
be  drawn  from  the  rodent  data:  (1) 
fire  may  impact  rodent  species  se- 
verely, probably  chiefly  through  loss 
of  habitat  resources,  especially  shel- 
ter and  food;  (2)  some  individuals 
survive  fire;  (3)  colonization  from 
adjacent  habitats  may  be  rapid;  (4) 
postfire  succession  is  somewhat  de- 
pendent on  prefire  species  composi- 
tion of  the  area;  (5)  in  southem  Cali- 
fornia chaparral,  at  least  two  species, 
deer  mouse  and  California  vole,  are 
fire  specialists,  entering  the  system 
only  for  relatively  short  periods  of 
the  postfire  succession;  (6)  species 
requiring  brush  for  cover  and /or 
food,  like  wood  rats  and  California 
mice,  are  most  severely  impacted  by 


fire,  and  require  the  longest  time  to 
recover  to  prefire  densities;  (7)  there 
is  no  clear-cut  difference  in  rodent 
postfire  succession  following  normal 
and  hot  fires;  (8)  rodent  postfire  suc- 
cession is  characterized  by  increases 
in  successionally-adapted  species, 
with  declines  in  those  species  for 
which  essential  habitat  features  are 
lacking;  and  (9)  recovery  of  the  ro- 
dent community  to  its  prefire  condi- 
tion probably  takes  four  to  six  years, 
with  the  exact  pattern  of  recovery 
being  dependent  on  prefire  species 
composition  and  features  of  the  pre- 
fire plan  community  and  postfire 
plant  succession  that  have  not  been 
delineated. 


Acknowledgments 

This  research  was  supported  by 
USDA  Forest  Service,  Pacific  South- 
west Forest  and  Range  Experiment 
Station  Grant  Number  PSW-85- 
0004CA  to  WOW  and  a  summer  re- 
search assistantship  from  Pomona 
College  to  JRM. 

We  are  indebted  to  Susan  Conard, 
Project  Leader,  Pacific  Southwest 
Forest  and  Range  Experiment  Sta- 
tion, Forest  Fire  Laboratory,  River- 
side, CA,  for  her  support  and  coop- 
eration during  this  study.  Many  biol- 
ogy students  at  Pomona  College  have 
assisted  with  field  work.  The  senior 
author  wants  to  acknowledge  5  years 
of  field  work  by  Sherrie  Souza  and 
David  Hoekman,  all  computer  pro- 
gramming by  David  Hoekman,  and 
all  data  analysis  by  John  R.  Muhm. 
We  are  grateful  to  Helen  Wirtz  for 
our  figures.  Preparation  of  this  paper 
was  greatly  assisted  by  a  summer 
research  assistantship  to  JRM. 

Literature  Cited 

Byrne,  Roger,  Joel  Michaelsen  and 
Andrew  Soutar.  1977.  Fossil  char- 
coal as  measure  of  wildfire  fre- 
quency in  southern  Califomia:  A 
preliminary  analysis,  p.  361-367.  In 

338 


Harold  A.  Mooney,  Eugene  C. 
Conrad,  eds.  Proceedings  of  the 
Symposium  on  the  Environmental 
Consequences  of  Fire  and  Fuel 
Management  in  Mediterranean 
Ecosystems.  USDA  Forest  Service, 
General  Technical  RepK3rt  WO-3. 

Hanes,  Ted  L.  1971.  Succession  after 
fire  in  the  chaparral  of  southern 
California.  Ecological  Monographs 
41(l):27-52. 

Hanes,  Ted  L.  and  Harold  W.  Jones. 
1967.  Postfire  chaparral  succession 
in  Southem  Califomia.  Ecology 
48:259-263. 

Hayne,  Don  W.  1949.  Two  methods 
of  estimating  population  from 
trapping  records.  Journal  of  Mam- 
malogy 30(4):399-411. 

Lawrence,  George  E.  1966.  Ecology 
of  vertebrate  animals  in  relation  to 
chaparral  fire  in  the  Sierra  Nevada 
foothills.  Ecology  47(2):278-291. 

Patric,  James  H.  and  Ted  L.  Hanes. 
1964.  Chaparral  succession  in  a 
San  Gabriel  Mountain  Area  of 
Califomia.  Ecology  45:353-360. 

Philpot,  Charles  W.  1977.  Vegetative 
features  as  determinants  of  fire 
frequency  and  intensity,  p.  12-16. 
In  Harold  A.  Mooney,  Eugene  C. 
Conrad,  eds.  Proceedings  of  the 
Symposium  on  the  Environmental 
Consequences  of  Fire  and  Fuel 
Management  in  Mediterranean 
Ecosystems.  USDA  Forest  Service, 
General  Technical  Report  WO-3. 

Quinn,  Ronald  D.  1979.  Effects  of  fire 
on  small  mammals  in  the  chapar- 
ral. Cal-Neva  Wildlife  Transac- 
tions, 125-133. 

Williams,  Olwen.  1955.  Distribution 
of  mice  and  shrews  in  a  Colorado 
montane  forest.  Journal  of  Mam- 
malogy 36(2):221 -231. 

Wirtz,  William  O.,  II.  1977.  Verte- 
brate postfire  succession,  p.  46-57. 
In  Harold  A.  Mooney,  Eugene  C. 
Conrad,  eds..  Proceedings  of  the 
Symposium  on  the  Environmental 
Consequences  of  Fire  and  Fuel 
Management  in  Mediterranean 
Ecosystems.  USDA  Forest  Service, 
General  Technical  Report  WO-3, 
Washington,  D.C.,  498  p. 


Wirtz,  William  O.,  II.  1979.  Effects  of 
fire  on  birds  in  chaparral.  Cal- 
Neva  Wildlife  Transactions,  114- 
124. 

Wirtz,  William  O.,  II.  1982.  Postfire 
community  structure  of  birds  and 
rodents  in  southern  California 
chaparral,  p.  241-246.  In  Eugene  C. 
Conrad,  Walter  C.  Oechel,  techni- 
cal coordinators.  Proceedings  of 
the  Symposium  on  Dynamics  and 
Management  of  Mediterranean- 
type  Ecosystems.  USDA  Forest 
Service,  General  Technical  Report 
PSW-58. 

Wirtz,  William  O.,  II.  1984.  Postfire 
rodent  and  bird  communities  in 
the  chaparral  of  southern  Califor- 
nia, p.  167-168.  In  MEDECOS  IV. 
Proceedings  of  the  4th  Interna- 
tional Conference  on  Mediterra- 
nean Ecosystems.  University  of 
Western  Australia,  Perth,  Western 
Australia. 


Douglas- Fir  Forests  in  the 
Cascade  IVIountains  of 
Oregon  and  Washington:  Is 
the  Abundance  of  Small 
Mammals  Related  to  Stand 
Age  and  Moisture?^ 

Paul  Stephen  Corn,^  R.  Bruce  Bury,^  and 
Thomas  A.  Sples^ 


Abstract.— Red  tree  voles  (Arborimus  longicaudus) 
were  the  only  small  mammal  strongly  associated 
with  old-growth  forests,  whereas  vagrant  shrews 
(Sorex  vagrans)  were  most  abundant  in  young 
forests.  Pacific  marsh  shrews  (S.  bendirii)  were  most 
abundant  in  wet  old-growth  forests,  but  abundance 
of  this  species  in  young  (wet)  forests  needs  further 
study.  Clearcuts  had  a  mammalian  fauna  distinct 
from  young  forest  stands.  Abundance  of  several 
species  was  correlated  to  habitat  features  unique  to 
naturally  regenerated  forests,  indicating  an  urgent 
need  to  study  the  long-term  effects  of  forest 
management  on  nongome  wildlife. 


Management  of  old-growth  Douglas- 
fir  (Pseudotsuga  menziesii)  forests  west 
of  the  Cascade  Mountains  in  the  Pa- 
cific Northwest  is  an  increasingly 
controversial  topic,  arising  from  a 
fundamental  conflict.  These  forests 
are  extremely  valuable  sources  of 
timber;  40  ha  of  old  growth  is  valued 
at  about  $1.6  million  (Meslow  et  al. 
1981).  At  the  same  time,  conserva- 
tionists view  old  growth  as  a  unique 
ecosystem  that  is  nonrenewable  un- 
der current  management  practices 
(Cutler  1984,  Schoen  et  al.  1981).  Old- 
growth  forests  are  disappearing;  dur- 
ing the  past  30  years,  removal  of 
Douglas-fir  saw  timber  from  western 
Oregon  and  Washington  has  ex- 
ceeded annual  growth  by  a  factor  of 
three  (Harris  1984).  Now,  less  than 
20%  of  the  original  old-growth  forest 
in  the  Pacific  Northwest  remains 
(Spies  and  Franklin  in  press). 

Historically,  old-growth  forests 
were  viewed  as  decadent  stands  of 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibiar^.  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff. AZ.  July  19-21.  1988.) 

'Paul  Stephien  Corn  is  Zoologist.  USDI  Rshi 
and  Wildlife  Service.  National  Ecology  Re- 
search) Center.  1300  Blue  Spruce  Drive.  Fort 
Collins.  CO  80524. 

^R.  Bruce  Bury  is  Zoologist  (Research)). 
USDI  Fish  and  Wildlife  Service.  National  Ecol- 
ogy Research  Center.  1300  Blue  Spruce 
Drive.  Fort  Collins.  CO  80524. 

^Thomas  A.  Spies  is  Research  Forester. 
USDA  Forest  Service.  Pacific  Northwest  Re- 
search Station.  Corvallis.  OR  9733 1. 


wasted  timber  that  provided  little 
wildlife  habitat.  For  example,  Tevis 
(1956)  stated: 

Virgin  forest  in  the  Douglas-fir 
(Pseudotsuga  taxifolia  [menzi- 
esii]) region  of  northwestern 
California  is  sterile  habitat  for 
wildlife.  Dense  shade  and 
competition  from  large  old 
trees  prevent  the  growth  of 
nearly  all  bushy  and  herba- 
ceous vegetation  except  a 
weak  understory  of  tan  oak 
(Lithocarpus  densiflora).  Food 
for  animals  is  scarce. 

The  value  of  old  growth  has  been 
rehabilitated.  Currently,  old-growth 
Douglas-fir  forests  are  considered 
excellent  wildlife  habitats,  dominated 
by  large  trees,  but  possessing  a  com- 
plex and  varied  structure  (Franklin  et 
al.  1981,  Franklin  and  Spies  1984), 
including  some  of  the  highest 
amounts  of  coarse  woody  debris 
(CWD)  reported  for  any  forest  eco- 
system (Spies  et  al.  in  press). 

Most  remaining  old  growth  in  the 
Pacific  Northwest  is  on  Federal  land 
managed  by  the  Forest  Service  and 
Bureau  of  Land  Management  (Harris 
1984).  The  policy  of  the  U.S.  Depart- 
ment of  Agriculture  is  to  "...maintain 
viable  populations  of  all  existing  na- 
tive vertebrate  populations..."  (Cut- 
ler 1980)  but,  until  recently,  the  infor- 
mation needed  to  achieve  this  goal 
did  not  exist.  Most  lists  of  species 
with  some  degree  of  dependence  on 


or  association  with  old  growth  are 
incomplete  or  inferential  (e.g.,  Harris 
and  Maser  1984,  Meslow  et  al.  1981). 
Recent  research  has  improved  this 
situation,  but  little  of  it  is  directed 
toward  nongame  species.  A  recent 
symposium  on  wildlife  and  old- 
growth  relations  (Meehan  et  al.  1984) 
included  27  papers.  Two-thirds  (17) 
of  the  papers  concerned  game  spe- 
cies, and  only  four  papers  discussed 
ecology  of  nongame  wildlife.  Re- 
maining pap>ers  discussed  either 
characteristics  of  old-growth  forests 
(three  papers)  or  management  objec- 
tives (three  papers). 

In  1981,  to  provide  the  informa- 
tion necessary  for  managing  wildlife 
in  the  national  forests  of  the  Pacific 
Northwest,  the  U.S.  Forest  Service 
chartered  the  Old-Growth  Wildlife 
Habitat  Program^  (OGWHP).  Its 
goals  (Ruggiero  and  Carey  1984) 
were  to:  (1)  identify  old-growth  for- 
ests were  unique  components  of  co- 
niferous forest  ecosystems,  (2)  iden- 
tify the  ecological  characteristics  of 
old  growth,  (3)  identify  any  wildlife 
species  dependent  on  old  growth  for 
survival  or  optimal  habitat,  and  (4) 
determine  the  amount  and  distribu- 
tion of  old  growth  necessary  to  meet 
the  needs  of  dependent  species. 

Vegetation  and  vertebrate  commu- 
nity studies  were  performed  on  a 
matrix  of  forest  conditions  in  natu- 
rally regenerated  stands.  Forest  de- 

^Now  the  Wildlife  Habitat  Relationships  in 
Western  Oregon  and  Washington  Project. 


340 


Figure  1  .—Maps  of  study  areas  where  pitfall  trapping  was  conducted  in  1983.  HJAEF  =  H.  J. 
Andrews  Experirr>ental  Forest;  WREF  =  Wind  River  Experimental  Forest.  Note  ttiat  \he  scale  for 
eachi  nr>ap  is  different. 


velopment  was  examined  across  a 
chronosequence,  and  a  moisture  gra- 
dient was  examined  for  the  old- 
growth  stands. 

Field  work  began  in  1983  with 
vegetation  and  vertebrate  commu- 
nity pilot  studies  at  30  stands  spread 
between  two  sites  in  the  Oregon  and 
Washington  Cascade  Mountains.  The 
primary  goal  of  the  first  year  was  to 
evaluate  and  recommend  sampling 
techniques.  The  pilot  studies  were 
successful  in  developing  and  refining 
sampling  methods  (e.g..  Bury  and 
Corn  1987,  Thomas  and  West  1984, 
West  1985).  In  1984  and  1985,  com- 
munity studies  expanded  to  more 
than  180  stands  in  the  Washington 
Cascades,  the  Oregon  Cascades,  the 
Oregon  Coast  Range,  and  the 
Siskiyou  and  Klamath  mountains  of 
southern  Oregon  and  northern  Cali- 
fornia. Since  1985,  species-specific 
studies  have  been  emphasized, 
largely  concerning  the  ecology  and 
management  of  the  spotted  owl 
(Strix  occidentalis)  and  its  prey  base. 

Our  paper  concerns  the  commu- 
nity ecology  of  small  mammals  as 
revealed  by  pitfall  trapping  in  1983. 
The  data  collected  in  1983  are  useful 
for  other  than  evaluating  techniques, 
but  these  data  are  difficult  to  inte- 
grate into  1984  and  1985  results,  be- 
cause the  sampling  methods  were 
changed  (Bury  and  Corn  1987). 
Therefore,  we  report  these  results 
with  the  caveat  that  variation  be- 
tween years  is  not  examined. 

Our  specific  objectives  are  to  ex- 
amine the  relations  of  the  abundance 
of  small  mammal  sf)ecies  to  the 
chronosequence  and  the  moisture 
gradient  and  to  identify  specific  habi- 
tat features  that  contribute  to  abun- 
dance. The  effects  of  forest  manage- 
ment are  also  discussed. 


METHODS 

Study  Areas 

Forest  stands  were  studied  in  two 
areas  on  the  western  slop)es  of  the 


341 


Cascade  Mountains  (fig.  1).  Twelve 
stands  were  in  the  Wind  River  Ex- 
perimental Forest  (WREF)  or  the  sur- 
rounding Gifford  Pinchot  National 
Forest,  Skamania  County,  Washing- 
ton, and  18  stands  were  in  the  H.  J. 
Andrews  Experimental  Forest 
(HJAEF)  or  Willamette  National  For- 
est, Lane  and  Linn  counties,  Oregon. 
Appendix  A  lists  ages,  elevations, 
and  locations  of  all  stands. 


Stand  Selection  and  Classification 

Initial  stand  selections  were  made  by 
OGWHP  investigators  studying  the 
structure  of  old  growth  (Franklin  and 
Spies  1984,  Spies  and  Franklin  in 
press).  Age  was  the  primary  criterion 
for  establishing  a  stand's  position  on 
the  chronosequence.  Topographic 
position  and  understory  vegetation 
provided  a  first  approximation  of  a 
moisture  gradient  (south-  or  west- 
facing  ridges  were  generally  dry, 
whereas  stands  on  north-facing 
slopes  were  usually  moist  to  wet). 
Most  stand  boundaries  were  not 
highly  distinct  (e.g.,  forest  islands 
surrounded  by  clear  cuts)  but  were 
determined  by  several  factors,  in- 
cluding age,  disturbance  history, 
vegetational  composition,  physiogra- 
phy, and  soils.  Stands  were  first  cho- 
sen from  aerial  photographs  and  for- 
est type  maps,  but  an  on-site  inspec- 
tion was  completed  before  any  of  the 
vertebrate  sampling  plots  were  estab- 
lished. Stand  sizes  varied  from  about 
10  to  20  ha. 

Coarse  woody  debris  (CWD), 
vegetation,  and  site  characteristics 
were  sampled  in  five  nested,  circular 
plots  in  each  stand  (Spies  et  al.  in 
press).  Classification  of  downed 
CWD  (=logs)  followed  Franklin  et  al. 
(1981)  and  Maser  and  Trappe  (1984): 
from  class  1  logs  (essentially  unde- 
cayed)  to  class  5  logs  (well  decayed, 
appearing  as  raised  hummocks  in  the 
forest  floor). 

The  chronosequence  consisted  of 
four  categories  beginning  with 
clearcuts  (<  10  years  old),  closed-can- 


opy young  stands  (30-80  years),  ma- 
ture stands  (80-195  years),  and  old 
growth  (195-450  years).  The  latter 
three  categories  were  all  composed 
of  naturally  regenerated  forests. 
Ages  of  young  and  mature  stands 
were  estimated  by  increment  coring 
of  at  least  five  dominant  Douglas-fir 
trees  per  stand  (Spies  et  al.  in  press). 
Ages  of  old-growth  stands  were  esti- 
mated from  increment  cores  and  by 
examining  stumps  in  adjacent 
clearcuts  and  roadsides. 

In  an  ideal  chronosequence  analy- 
sis, age  classes  should  have  similar 
means  and  ranges  of  site  characteris- 
tics. We  were  only  partly  successful 
in  achieving  this  goal,  because  the 
age  classes  were  not  equally  distrib- 
uted over  the  landscajje,  and  other 
criteria  such  as  stand  size,  accessibil- 
ity, and  absence  of  logging  activity 
took  precedence  over  site  uniformity. 
Consequently,  young  and  mature 
stands  spanned  a  wider  range  of  en- 
vironments than  originally  planned 
and  for  some  variables  (such  as  ele- 
vation at  the  HJAEF),  the  younger 
age  classes  differed  from  old  growth. 

We  conducted  analysis  of  mois- 
ture effects  across  the  old-growth 


stands.  Adjustments  were  made  to 
the  preliminary  field  classification  of 
dry  (OGD),  moderate  (OGM),  and 


C3RY  Moer 
DECORANA  AXIS  1 

Figure  2.— Detrended  correspondence 
analysis  (DECORANA)  of  percent  occur- 
rence of  understory  plant  species  in  old- 
growthi  stands  in  Oregon  and  Washington. 
Stands  were  placed  in  nnoisture  categories 
(wet,  moderate,  or  dry)  based  on  ttieir  rela- 
tive positions  on  ttie  two  gradients. 


Figure  3.— A  pitfall  array  in  clearcut  stand  #291  in  Oregon,  Ptioto  by  L.  Hanebury. 


342 


wet  (OGW),  after  conducting  ordina- 
tions of  old-growth  stands  using  de- 
trended  correspondence  analysis 
(DECORANA).  DECORANA  is  a 
weighted  average  technique  that  is 
computationally  related  to  principal 
comp>onents  analysis  (Gauch  1982). 
The  percent  occurrence  of  understory 
plant  species  in  five  1,000-m^  plots  in 
each  stand  was  used  in  separate 
analyses  of  each  study  area  (fig.  2). 
The  first  axis  in  both  areas  separated 
stands  along  a  moisture  gradient  cor- 
related with  indicators  of  topo- 
graphic moisture,  such  as  aspect  and 
slope.  The  second  axis  in  both  analy- 
ses separated  stands  along  a  complex 
gradient  of  temperature  and  mois- 
ture and  was  correlated  with  eleva- 
tion. 


Pitfall  Trapping 

We  installed  a  pitfall  trap  array  (fig. 
3)  in  each  stand.  An  array  included 
two  triads,  25  m  apart,  each  consist- 
ing of  three  5-m  long  aluminum  drift 
fences  with  screen  wire  funnel  traps 
on  each  side  and  pitfall  traps  at  each 
end.  Thus,  each  array  had  six  fences 
and  twelve  pitfall  and  twelve  funnel 
traps.  Bury  and  Com  (1987,  this  vol- 
ume) provide  more  complete  de- 
scriptions and  illustrations. 

The  traps  were  opened  the  last 
week  in  May  1983  and  were  operated 
continuously  for  180  days.  No  water 
was  put  in  traps,  because  this  has  a 
deleterious  effect  on  the  preservation 
of  amphibians,  which  were  a  major 
target  of  the  traps  (Bury  and  Corn, 
this  volume  ).  In  practice,  most  traps 
accumulated  some  water  and  most 
mammals  drowned.  Traps  were 
checked  initially  every  three  days, 
but  as  trap  rate  declined  over  time, 
the  interval  between  checks  in- 
creased to  about  seven  days. 

Mammals  taken  from  traps  were 
identified,  sexed,  measured,  and  pre- 
served as  skulls,  skeletons,  or  skins 
and  skulls.  All  specimens  from  Ore- 
gon and  most  from  Washington  were 
deposited  in  the  National  Museum  of 


Natural  History  (USNM),  where  all 
identifications  were  verified.  Com- 
mon and  scientific  names  used  in  this 
paper  follow  Banks  et  al.  (1987). 

We  encountered  one  problem  that 
significantly  affected  the  data  analy- 
sis. The  high  trap  success  at  the 
WREF  stands  exceeded  the  field 
crew's  ability  to  process  sp>ecimens, 
and  approximately  25%  of  the  mam- 
mals were  discarded  in  the  field. 
When  the  remaining  specimens  were 
examined  later  at  the  USNM,  about 
10%  of  the  field  identifications  of 
Trowbridge's  shrews  (Sorex  trowbr- 
idgii),  montane  shrews  (S.  montico- 
lus),  and  vagrant  shrews  (S.  vagrans) 
were  inaccurate.  Thus,  the  exact 
numbers  of  these  shrews  captured  at 
WR  are  in  doubt  (Bury  and  Com 
1987),  and  analyses  of  overall  species 
richness  and  individual  abundance 
of  these  sp>ecies  were  only  reported 
for  Oregon  data. 

Statistical  Analyses 

We  analyzed  the  mean  abundance 
(total  number  of  captures)  of  each 
species,  mean  total  abundance,  and 
mean  species  richness  across  each 
gradient  with  one-way  analysis  of 
variance  (ANOVA).  No  traps  were 
missing  or  damaged  during  the  180- 
day  trapping  period,  so  it  was  unnec- 
essary to  adjust  raw  abundance  for 
trap  nights.  Scavengers  may  remove 
animals  from  traps  when  there  are 
long  intervals  between  checks  (M.  G. 
Raphael,  personal  communication), 
and  traps  with  water  may  be  more 
effective  than  dry  pitfalls  at  captur- 
ing rodents  with  good  leaping  abil- 
ity. Because  70%  of  all  mammals 
were  captured  in  the  first  60  days  of 
trapping  (Bury  and  Corn  1987),  when 
traps  were  checked  frequently,  we 
feel  these  considerations  are  minor 
and  we  made  no  adjustments  to  the 
data. 

Abundances  were  log  transformed 
before  the  ANOVAs  were  run. 
Clearcuts,  OGW  and  OGD  stands 
were  not  included  in  the  ANOVA  of 


the  chronosequence.  Clearcuts, 
young,  and  mature  stands  were  not 
included  in  the  ANOVA  on  moisture 
(Spies  et  al.  in  press).  A  comparison 
of  species'  abundances  in  clearcuts 
versus  young  stands  is  presented 
separately.  Pearson  correlation  coef- 
ficients were  calculated  between 
abundance  (transformed  as 
ln[abundance  +  1])  and  24  of  the 
habitat  variables  (app>endix  B).  Per- 
centage variables  (e.g.,  %  cover  of 
grasses)  were  arcsin  transformed, 
other  variables  were  log  trans- 
formed. We  also  performed  a  princi- 
pal components  analysis  using  the 
habitat  variables,  but  because  the 
first  three  factors  explained  only  52% 
of  the  variation  among  stands,  we 
report  only  the  significant  (P  <  0.05) 
bivariate  correlations  between  abun- 
dance and  individual  habitat  vari- 
ables. All  analyses  were  performed 
using  the  statistical  program  SYS- 
TAT^  (Wilkinson  1988). 

RESULTS 

The  pitfall  arrays  were  highly  effec- 
tive at  capturing  small  mammals, 
producing  3,877  captures  of  27  spe- 
cies. Insectivores  and  microtine  ro- 
dents were  best  caught  by  pitfalls, 
while  deer  mice  (Peromyscus  manicu- 
latus)  were  under-sampled  (Bury  and 
Com  1987).  Captures  of  each  species 
in  each  stand  are  listed  in  tables  1 
(HJAEF)  and  2  (WREF). 

Mean  species  richness  (number  of 
species)  varied  from  about  nine  in 
mature  stands  to  12  in  OGW  stands 
(fig.  4).  There  was  no  significant  dif- 
ference across  either  the  chronose- 
quence or  the  moisture  gradient.  To- 
tal abundance  was  highest  in  young 
and  mature  stands  and  lowest  in 
OGM  stands,  but  the  difference  was 
not  significant.  There  was  no  appar- 
ent trend  in  small  mammal  abun- 
dance across  the  moisture  gradient. 

^Trade  names  are  provided  for  the 
benefit  of  the  reader:  such  use  does  not 
constitute  an  official  endorsement  by  the 
Fish  and  Wildlife  Service. 


343 


Table  1  —Abundance  of  small  mammals  captured  at  the  H.  J.  Andrews  Experimental  Forest  In  Oregon.  Arrays  of  pit- 
fall traps  with  drift  fences  were  operated  continuously  for  1 80  days  in  1 983, 


Species      Stand  no. 

Trowbridge's  Shrew 
Montane  Shrew 
Vagrant  Shrew 
Pacific  Marsh  Shrev/ 
Northern  Water  Shrew 
Pacific  Shrew 
Unidentified  shrews 
Shrew  Mole 
Coast  Mole 
Western 

Red-backed  Vole 
Creeping  Vole 
Red  Tree  Vole 
Water  Vole 
Heather  Vole 
Townsend's  Vole 
Deer  Mouse 
Pacific  Jumping  Mouse 
Western  Pocket  Gopher 
Others" 


Old  growth 


Wet 

Moderate 

Dry 

Mature 

Young 

Clearcut 

15 

03 

24 

02 

17 

33 

25 

29 

11 

35 

42 

39 

**/ 

*KJ 

75 

291  391 

33 

48 

48 

76 

35 

60 

75 

70 

51 

56 

78 

70 

139 

71 

83 

18 

39 

17 

lO 

1  o 

9R 
zo 

23 

13 

g 

7 

19 

15 

16 

15 

26 

17 

14 

22 

3 

8 

13 

lo 

Z 

1 

9 

4 

0 

7 

2 

17 

3 

5 

6 

1 

1 

74 

7 

14 

o 

1 

1 

1 

7 

9 

2 

2 

4 

13 

1 

4 

2 

1 
1 

1 

1 

1 

3 

5 

3 

1 

1 

4 

■  .  2 

9 

4 

2 

5 

1 
1 

1 

4 

5 

4 

6 

2 

2 

1 

7 

1 

2 

3 

2 

14 

4 

1 

2 

9 

6 

3 

10 

6 

15 

18 

52 

13 

7 

4 

1 

1 

1 

2 

1 

1 

2 

3 

1 

5 

28 

1 

3 

3 

4 

1 

1 

2 

1 

1 

1 

1 

2 

1 

1 

5 

1 

2 

1 

3 

3 

3 

1 

3 

2 

3 

3 

3 

1 

2 

5 

2 

3 

24 

1 

1 

2 

1 

3 

1 

14 

1 

3 

] 

2 

1 

16 

1 

2 

1 

2 

2 

2 

1 

1 

1 

1 

°Townsend's  Chipmunk  (8).  Northern  Rying  Squirrel  (3),  Errriine  (2),  Spotted  Sl(unk  (1),  Sr)owshoe  Hare  (1). 


r 


Table  2.-Abundance  of  small  mammals  captured  at  the  Wind  River  Experimental  Forest  In  Washington,  Arrays  of  pit- 
fall traps  with  drift  fences  were  operated  continuously  for  180  days  in  1983. 


Old  growth 


Species    Stand  No. 


Wet 
14 


Moderate 


12 


21 


20 


Dry 
31 


Mature 


Young 


41 


42 


60 


60 


61 


Clearcut 


70 


71 


Pacific  Marsh  Shrew  10  3 
Other  shrews^  86  73 

Shrew  Mole  6 
Coast  Mole  3 
Southern 

Red-backed  Vole       15  10 
Creeping  Vole  2 
Townsend's  Vole  1 
Other  Microtines*^ 

Deer  Mouse  8  16 

Pacific  Jumping  Mouse  2 
Northern  Pocket  Gopher 
Others^  1 


2 
93 
9 
3 

21 
4 


40 
2 


3 
46 
1 


20 
2 


11 


115 
2 
3 

40 
3 
5 

28 


6 
192 
9 
4 

13 
6 
2 

23 
2 
2 


3 
127 
2 
1 

16 
9 
1 

11 


3 

158 
6 
1 

3 
2 
1 

16 
1 


2 
117 
1 

2 

41 
1 
1 
3 
9 
1 


1 


7 
97 
1 


31 
4 

1 

7 


1 


3 
50 


4 
9 


32 
1 

1 


86 
3 


1 

11 
5 
2 
7 

4 

3 


°unider)tified  (701  Trowbridge's  Shrew  (6967).  Mor)tane  Shrew  (35 1 7),  Vagrant  Shrew  (12071  Masked  Shrew  (7),  and  Northern  Water 
Shrew  (3). 

^unidentified  (6).  Heather  Vole  (1). 

^Ermine  (6).  Townsend's  CNpmunk  (3),  Yellow-pine  Chipmunk  (2).  Snowshoe  Hare  (2),  Northern  Hying  Squirrel  (1),  Pika  (1). 


344 


species- Habitat  Associations 

Trowbridge's  Shirew 

These  shrews  were  the  most  abun- 
dant small  mammal  (about  46%  of  all 
captures).  At  HJAEF,  this  species 
was  most  abundant  in  young  stands 
(fig.  5),  but  the  variation  across  the 
chronosequence  was  not  statistically 
significant.  Most  of  the  high  mean 
abundance  in  young  stands  was  due 
to  one  stand  (#47)  at  HJAEF  (table  1). 
Abundance  on  the  moisture  gradient 
increased  from  OGW  to  (X^D,  but 
the  differences  were  not  significant. 

Habitat  variables  that  were  posi- 
tively correlated  with  abundance  of 
Trowbridge's  shrews  included  the 
total  basal  area  and  mean  diameter  at 
breast  height  (d.b.h.)  of  live  trees,  the 
number  of  decay  class  4  and  5  (most 
decayed)  downed  logs,  and  litter 
depth  (table  3).  Variables  negatively 

MEAN  §  OF  SPECIES 


TOTAL^  /KBUtsO/KNCEI 


MEAN  TOTAL  CAPTURES 


Figure  4.— Mean  species  richness  (HJAEF 
only)  and  total  abundance  (all  stands)  of 
snnall  PDommals  in  closed-canopy  stands. 


correlated  were  percent  cover  by 
herbs  and  grasses  and  the  biomass  of 
least  decayed  logs  (class  1  and  2). 
Montane  Stirew 

This  was  the  second  most  abundant 
species,  occurring  in  similar  numbers 
in  stands  of  different  ages  (fig.  5). 
There  is  a  trend  on  the  moisture  gra- 
dient of  decreasing  abundance  from 
OGW  to  OGD,  but  the  differences 
are  not  significant.  Abundance  of 
montane  shrews  was  positively  cor- 
related with  tree  size  (MDBH)  and 
negatively  correlated  with  percent 
cover  by  grasses  and  number  of  de- 
cay class  1  and  2  logs  (table  3). 

Vagrant  Shirew 

Vagrant  shrews  were  significantly 
less  abundant  in  older  forest  stands 
(fig.  5,  P  =  0.02),  and  variation  across 
the  moisture  gradient  was  not  signifi- 
cant. This  species  reached  its  greatest 
abundance  in  one  clearcut  (see  be- 
low). Abundance  of  vagrant  shrews 
was  negatively  correlated  with  sev- 
eral characters  associated  with  old- 
growth  forests:  number  of  decay 
class  4  and  5  logs,  percent  cover  by 
mosses,  litter  depth,  and  slope  (table 
3). 


Pacific  Marsti  Shrew 

The  Pacific  marsh  shrew  (Sorex 
bendirii)  is  a  large  shrew  generally 
associated  with  small  streams  and 
swamps  (Maser  et  al.  1981,  Whitaker 
and  Maser  1976).  Our  results  agree. 
The  greatest  abundance  was  in  (DGW 
stands  (fig.  5),  and  the  difference 
across  the  moisture  gradient  was  sig- 
nificant (P  <  0.001).  Marsh  shrews 
were  captured  (albeit  in  low  num- 
bers) in  moderate  and  dry  old- 
growth  stands  where  the  pitfall  ar- 
rays were  some  distance  from  flow- 
ing or  standing  water,  but  many  of 
the  younger  stands  in  which  this  spe- 
cies occurred  (e.g.,  stands  11,  35,  and 
75  at  the  HJAEF)  contained  streams 


or  ponds.  Variation  across  the 
chronosequence  was  not  significant, 
but  this  may  be  misleading  given  the 
high  abundance  in  OGYJ  stands.  Our 
study  design  precluded  us  from  de- 
termining whether  Pacific  marsh 
shrews  would  be  abundant  in 
younger  wet  stands. 

Several  habitat  variables  were  as- 
sociated with  abundance  of  Pacific 
marsh  shrews.  Positive  correlations 
reflected  older,  wet  forests  and  in- 
cluded litter  depth,  total  density  of 
live  trees,  mean  d.b.h.,  and  biomass 
of  class  4  and  5  logs.  The  number  of 
decay  class  1  and  2  logs  and  slope 
were  negatively  correlated  with 
abundance  (table  3). 


Stirew  Mole 

Shrew  moles  (Neurotrichus  gibbsii)  are 
small  moles  but  are  more  like  shrews 
in  appearance  and  habits.  Patterns  of 
their  abundance  were  similar  to  the 
Pacific  marsh  shrew  (fig.  5).  Shrew 
moles  were  most  abundant  in  (DGW, 
but  there  were  no  significant  differ- 
ences across  the  moisture  gradient  or 
the  age  gradient.  Unlike  the  marsh 
shrew,  none  of  the  habitat  variables 
were  correlated  with  abundance. 


Coast  Mole 

We  captured  59  coast  moles  (Scapa- 
nus  orarius),  a  form  rarely  taken  by 
conventional  snap-  or  live-trapping 
techniques.  This  species  might  be 
more  active  on  the  surface  than  other 
moles  (Maser  et  al.  1981),  or  our  drift 
fences  (which  were  sunk  20-30  cm 
into  the  ground)  might  have  inter- 
rupted their  burrowing  (Williams 
and  Braun  1983).  There  was  no  sig- 
nificant variation  on  the  chronose- 
quence, but  there  was  on  the  mois- 
ture gradient  (P  =  0.05).  Coast  moles 
were  most  abundant  in  OGM  and 
OGD  stands  and  were  virtually  ab- 
sent from  (DGW  stands  (fig.  5). 

Coast  moles  might  prefer  well- 
drained  soils  (Maser  et  al.  1981).  This 


345 


is  supported  by  their  low  abundance 
in  OGW  stands  where  soils  are  satu- 
rated for  long  periods.  Abundance  of 
coast  moles  was  positively  correlated 
with  percent  cover  by  deciduous 
trees.  Habitat  variables  negatively 
correlated  were  the  number  of  decay 
class  3  logs  and  the  number  of  large- 
diameter  logs. 

Red- Backed  Voles 

We  captured  two  species  of  red- 
backed  voles:  the  southern  red- 
backed  vole  (Clethrionomys  gap-peri)  at 
WREF,  and  the  western  red-backed 
vole  (C.  californicus)  at  HJAEF.  We 
caught  more  southern  than  western 
red-backed  voles  (fig.  6),  but  the  pat- 
terns of  abundance  were  similar. 
Both  species  were  combined  in  the 
ANOVAs  to  maximize  the  sample 
size.  No  differences  were  detected  on 
either  the  age  or  moisture  gradients. 

Habitat  variables  were  tested 
separately  for  each  species,  but  the 
results  were  similar  (table  4).  Abun- 
dance of  western  red-backed  voles 
was  positively  correlated  with  total 
basal  area  of  live  trees,  mean  d.b.h., 
and  percent  cover  by  evergreen 
shrubs  (mainly  Oregon  grape,  Ber- 
beris  spp.,  and  salal,  Gaultheria  shal- 
lon). 

Negative  correlations  were  with 
grass  cover,  biomass  of  decay  class  1 
and  2  logs,  and  aspect  (abundance 
was  greatest  on  southern  exposures). 
Southern  red-backed  voles  were 
positively  correlated  with  density 
and  basal  area  of  live  trees,  and  mean 
d.b.h.,  and  were  negatively  corre- 
lated with  grass  cover. 

Red  Tree  Vole 

This  species  has  been  identified  as  an 
old  growth  associate  (Meslow  et  al. 
1981)  and  is  a  major  food  item  of  the 
spotted  owl  (Forsman  et  al.  1984). 
We  captured  only  17  red  tree  voles 
(Arborimus  longicaudus)  in  the  stan- 
dard arrays,  too  few  to  run  the 


ANOVA.  But,  12  voles  were  cap- 
tured in  the  eight  old-growth  stands 
at  HJAEF,  compared  to  only  five 
voles  in  the  10  younger  stands  (G  = 
4.73,  P  <  0.05).  Corn  and  Bury  (1986) 
provide  a  more  detailed  account  of 
these  results. 


Creeping  Vole 

Creeping  voles  (Microtus  oregoni) 
were  uncommon  in  closed-canopy 
stands  (fig.  6),  and  there  was  no  dif- 
ference in  abundance  on  either  gradi- 
ent. As  with  vagrant  shrews,  this 
species  was  more  abundant  in 


Figure  5.— Mean  abundance  of  insectivores  in  closed-canopy  forest  stands.  Data  for  Trowbr- 
idge's, montane,  and  vagrant  stirews  are  from  HJAEF  only.  Pacific  nrrarsti  stirews,  stirew 
moles,  and  coast  moles  use  data  from  all  stands. 

346 

I 


Table  3.— Significant  (P  <  0.05)  Pearson  correlations  of  insectivore  abun- 
dance and  stand  structure  and  vegetation  variables.  See  appendix  B  for 
descriptions  of  ttie  variables. 


Positive 


Negative 


Species 

Variable 

r 

Variable 

TrnwhriHop'*;  Shrpw 

TOTBA 

HFRR 

-0  7S 

(n=  17-18) 

LNDC45 

0.59 

GRASS 

-0.70 

MDBH 

0,57 

L6DC12 

-0.53 

LfTTER 

0  49 

Montane  Shrew 

MDBH 

0.51 

GRASS 

-0.52 

(n=  17-18) 

LNDC12 

-0.47 

Vagrant  Shrew 

LNDC45 

-0.55 

(n=  17-18) 

MOSS 
SLOPE 
UTTER 

-0.50 
-0.51 
-0.50 

Pacific  Marsh  Shrew 

LITTER 

0.41 

LNDC12 

-0,50 

(n  =  28-30) 

TOTDEN 

0.41 

SLOPE 

-0.37 

MDBH 

LBDC45 

DECTR 

0.44 
0.40 
0,52 

Coast  Mole 

-0.43 

LNDC3 

(n  =  28-30) 


LNDM3 


-0.43 


Figure  6.— Mean  abundance  of  rodents  in  closed-canopy  forest  stands.  Data  from  all  stands 
were  used. 


clearcuts.  Reflecting  this,  creeping 
vole  abundance  was  positively  corre- 
lated with  percent  cover  by  grasses 
and  negatively  correlated  with  sev- 
eral "forest"  variables:  number  and 
biomass  of  decayed  logs,  density,  ba- 
sal area  and  d.b.h.  of  live  trees,  and 
litter  depth. 

Deer  Mouse 

Although  pitfall  traps  are  not  as  ef- 
fective for  catching  deer  mice  as  snap 
traps  (Williams  and  Braun  1983,  Bury 
and  Corn  1987),  we  caught  moderate 
numbers  of  this  species,  particularly 
at  WREF  (table  2).  Deer  mice  were 
most  abundant  in  OGM  stands  and 
least  abundant  in  OGW  and  young 
stands.  Differences  were  not  signifi- 
cant on  either  the  chronosequence  or 
the  moisture  gradient.  Deer  mouse 
abundance  was  negatively  correlated 
with  percent  of  coarse  fragments  in 
the  soil. 


Clearcuts  Versus  Forests 

Pitfall  arrays  were  installed  in  five 
clearcuts,  three  at  HJAEF  and  two  at 
WREF.  We  compared  the  relative 
abundance  of  several  of  the  common 
small  mammals  in  clearcuts  and 
young  stands  (fig.  7).  Trowbridge's, 
montane,  and  vagrant  shrews  were 
compared  only  for  the  three  clearcuts 
and  four  young  stands  at  HJAEF. 

Southern  and  western  red-backed 
voles  were  virtually  absent  from 
clearcuts,  while  creeping  voles  were 
more  than  six  times  more  abundant 
in  clearcuts  than  in  young  stands. 
Most  insectivores  were  two  to  six 
times  more  abundant  in  young 
stands,  but  vagrant  shrews  were 
most  abundant  in  clearcuts.  Much  of 
the  difference  in  the  relative  abun- 
dance of  vagrant  shrews  is  due  to 
their  great  abundance  in  clearcut 
#391  at  HJAEF  (table  1).  Only  one 
vagrant  shrew  was  captured  at  each 
of  the  other  clearcuts  at  HJAEF.  Al- 
though roughly  eshmated,  vagrant 


347 


shrews  were  the  most  common  small 
mammal  at  both  of  the  clearcuts  at 
WREF.  Deer  mice  were  about  three 
times  more  abundant  in  clearcuts 
than  in  young  stands.  A  few  pocket 
gophers  {Thomomys  mazama  at 
HJAEF,  T.  talpoides  at  WREF)  were 
captured  and  are  not  depicted  in  fig- 
ure 7.  Most  pocket  gophers  (20/28) 
were  captured  in  clearcuts;  none 
were  captured  in  old  growth. 

DISCUSSION 

Old-Growth  Species 

Answering  the  question  of  if  a  spe- 
cies is  dependent  on  old-growth  for- 
est for  critical  habitat  is  complex,  in- 
corporating several  aspects  of  ecol- 
ogy and  needs  to  account  for  tempo- 
ral and  random  variation  (Carey 
1984).  Also,  abundance  of  individual 
species  within  a  specific  region  de- 
pend not  only  on  the  multidimen- 
sional niche,  but  on  the  geographic 
distribution  of  each  species  (Brown, 
1984).  The  community  ecology  stud- 
ies of  the  Old-Growth  Program  were 
not  intended  to  provide  definite  an- 
swers on  old-growth  dependencies, 
but  rather  the  results  were  to  be  used 
as  guides  for  designing  species-spe- 
cific research  (Ruggiero  and  Carey 
1984).  Our  results  are  based  on  one 
season's  data  and  must  be  inter- 
preted cautiously,  but  they  are  useful 
for  comparison  with  other  studies 
and  for  suggesting  new  research. 

Only  one  small  mammal,  the  red 
tree  vole,  displayed  a  significant  as- 
sociation with  old-growth  stands, 
and  the  sample  size  for  it  was  small. 
Additional  captures  of  this  species  in 
the  Oregon  Coast  Range  in  1984-1985 
were  almost  exclusively  in  old- 
growth  forests  (Com  and  Bury,  un- 
published data).  Recent  studies  of 
vertebrates  across  a  similar  chronose- 
quence  of  Douglas-fir  forests  in 
northern  California  (Raphael  1984, 
this  volume,  Raphael  and  Barrett 
1984)  found  significant  positive  cor- 
relations between  abundance  of  sev- 


eral species  and  stand  age:  Trowbr- 
idge's shrews.  Pacific  shrews  (Sorex 
pacificus),  coast  moles,  shrew  moles, 
Allen's  chipmunks  (Tamias  senex), 


Townsend's  chipmunks  (T.  town- 
sendii),  Douglas'  squirrels  (Tamias- 
ciurus  douglasii),  dusky-footed 
woodrats  (Neotoma  fuscipes),  deer 


Table  4.— Significant  (P  <  0.05)  Pearson  correlations  of  rodent  abundance 
and  stand  structure  and  vegetation  variables.  See  appendix  B  for  descrip- 
tions of  the  variables. 


Positive 


(n  =  28-30) 


Negative 


Species 

Variable 

r 

Variable 

r 

Western  Red-backed  Vole 

TOTBA 

0.66 

GRASS 

-0.54 

(n=  17-18) 

MDBH 

0.56 

TRASPECT 

-0.51 

EGSHR 

0.48 

LBDC12 

-0.53 

Southern  red-backed  Vole 

TOTDEN 

0.78 

GRASS 

-0.81 

(n=  11-12) 

TOTBA 

0.70 

MDBH 

0.71 

Creeping  Vole 

GRASS 

0.51 

LNDC45 

-0.58 

(n  =  28-30) 

LBDC45 

-0.43 

MDBH 

-0.52 

TOTDEN 

-0.40 

TOTBA 

-0.49 

irrrER 

-0.62 

Deer  Mouse 

TOTOF 

-0.36 

15.4 


YOUNG 


Z  RB5-aACKED   PACFC       SHREW        TROW-     UCNXM^  COAST 

^  VOLES        MARSH        MOLE       BHDQE3       SHREW       MOLE         DBW  VAQRAMT 

LU  .  SHREW  SHCW  MOUSE        ShFEW  VOLE 


Figure  7.— Relative  mean  abundance  of  snnall  mamrrxjls  In  young  stands  and  clearcuts. 
Species  more  abundant  in  young  stands  are  above  \he  horizontal,  species  more  abundant 
In  clearcuts  below.  Values  are  ttie  greater  mean  abundance  divided  by  the  lesser,  so,  for 
example,  red-backed  voles  were  15.4  times  more  abundant  In  young  stands  than  in 
clearcuts. 


348 


mice,  western  red-backed  voles,  and 
fishers  (Martes  penmnti).  Many  of 
these  correlations  were  not  strong, 
however,  with  most  species  repre- 
sented in  the  youngest  forest  stages. 
Mean  species  richness  was  about  10 
in  all  forest  age  classes.  Analysis  of 
the  similarity  of  species  composition 
showed  little  difference  on  the 
chronosequence  (Raphael  1984).  This 
is  very  similar  to  our  results  and  sug- 
gests that  old-growth  forests  do  not 
harbor  unique  communities  of  small 
mammals. 

Anthony  et  al.  (1987)  snaptrapped 
small  mammals  in  riparian  zones  of 
old-growth,  mature,  and  young 
stands  at  HJAEF  in  1983.  They  found 
greater  abundance  of  deer  mice  in 
old-growth  stands,  but  Pacific 
shrews  (S.  pacificus)  were  evenly  dis- 
tributed. They  trapped  14  other  spe- 
cies, though  none  in  sufficient  num- 
bers to  analyze.  Although  both  An- 
thony et  al.  (1987)  and  Raphael  (1984) 
found  more  deer  mice  in  older  for- 
ests, this  species  is  ubiquitous  and 
reaches  its  highest  densities  in  the 
Pacific  Northwest  in  clearcuts  (see 
below). 

Smell  Mammals  in  Managed 
Forests 

Most  studies  of  habitat  relations  of 
small  mammals  in  the  Pacific  North- 
west have  compared  clearcuts  to  for- 
ested stands.  Although  there  is  con- 
siderable variation  among  studies, 
general  trends  are  similar,  likely  re- 
lated to  the  variety  of  factors  exam- 
ined (time  since  logging,  burned,  un- 
bumed,  herbicides  applied,  etc.). 
Populations  of  deer  mice,  creeping 
voles,  and  Townsend's  chipmunks 
increase  after  logging,  while  red- 
backed  voles  and  Trowbridge's 
shrews  decline  (Anthony  and  Morri- 
son 1985,  Gashwiler  1959  1970, 
Hooven  and  Black  1976,  Sullivan  and 
Krebs  1980,  Raphael,  this  volume, 
Tevis  1956).  Red-backed  voles  are 
probably  most  affected  by  clearcut- 
ting.  Western  red-backed  voles  are 


obligate  fungivores,  and  their  food 
supply  disappears  after  clearcutting 
(Maser  et  al.  1978,  Ure  and  Maser 
1982).  Gunther  et  al.  (1983)  found 
southern  red-backed  voles  to  be  the 
most  common  animals  on  the 
clearcuts  they  trapped,  but  they 
trapped  only  three  months  after  log- 
ging and  probably  were  sampling  a 
residual  p)opulation.  Also,  this  spe- 
cies is  less  dependent  on  fungi  (Ure 
and  Maser  1982)  and  might  be  able  to 
persist  for  a  time  after  logging. 

Other  studies  have  not  noted  the 
increase  of  vagrant  shrews  in 
clearcuts  that  we  observed.  Several 
factors  might  be  involved,  including 
random  variation.  Although  mean 
abundance  was  high,  vagrant  shrews 
were  rare  (one  capture  each)  on  two 
of  our  five  clearcuts.  Other  studies 
probably  underestimated  shrew 
abundance,  because  they  used  either 
snap  or  live  traps.  Also,  some  inves- 
tigators might  have  followed  Findley 
(1955)  and  considered  montane  and 
vagrant  shrews  to  be  the  same  spe- 
cies. 

Changes  in  small  mammal  com- 
munities after  logging  can  be  dra- 
matic, but  clearcuts  per  se  might  not 
be  the  main  factor  influencing  species 
diversity  in  managed  forest  land- 
scajjes  in  the  Pacific  Northwest.  In  a 
managed  forest  with  a  90-year  rota- 
tion, about  30%  of  the  area  will  be  in 
clearcuts  and  young  plantations  lack- 
ing canopy  closure.  The  remaining 
70%  of  the  landscape  will  be  in 
stands  30-90  years  old  that  have 
closed  forest  canopies.  The  habitat 
characteristics  of  these  forest  planta- 
tions will  be  a  major  determinant  of 
biological  diversity  in  managed 
lands.  For  example,  the  extensive 
logging  of  low-elevation  old-growth 
forests  in  Oregon  has  probably  elimi- 
nated much  of  the  habitat  of  red  tree 
voles.  The  giant  Douglas-fir  trees, 
which  seem  to  be  preferred  as  nest 
sites,  will  not  occur  in  managed  for- 
ests. Meanwhile,  the  heather  vole 
(Phemcomys  intermedius),  a  species  of 
alpine  meadows,  might  be  benefit- 
ting from  increased  logging  of  high- 


elevation  forests  (Com  and  Bury 
1988). 

Although  we  have  found  few  dif- 
ferences between  old-growth  and 
younger  naturally  regenerated  for- 
ests for  small  mammals  or  the  herpe- 
tofauna  (Bury  and  Com,  this  vol- 
ume), the  same  probably  cannot  be 
said  for  comparisons  of  old-growth 
to  managed  forests.  Our  analysis  of 
habitat  variables  revealed  that  abun- 
dance of  several  species  was  corre- 
lated with  habitat  features  that 
would  be  absent  or  greatly  reduced 
in  managed  forests.  Aside  from  large 
trees,  CWD  is  the  primary  compo- 
nent of  old  growth  that  is  eliminated 
by  current  forestry  practices  (Harris 
et  al.  1982,  Spies  et  al.  in  press).  CWD 
is  correlated  to  abundance  of  shrews 
(this  study),  salamanders  (Bury  and 
Com,  this  volume,  Raphael  1984), 
and  probably  is  required  habitat  for 
red-backed  voles  (Maser  and  Trappe 
1984).  Bury  and  Corn  (this  volume) 
provide  further  discussion  of  the  role 
of  CWD  as  wildlife  habitat. 


Research  Needs 

These  types  of  community  ecology 
studies  provided  baseline  data  on 
nongame  wildlife  in  naturally  regen- 
erated forests  of  the  Pacific  North- 
west. For  example,  we  can  use  the 
data  on  abundance  and  the  correla- 
tions with  habitat  variables  to  begin 
classifying  species  as  to  their  degree 
of  rarity  (Rabinowitz  et  al.  1986). 
Species  with  small  geographic  distri- 
butions, restricted  habitat  specificity, 
and  small  local  populations  (e.g.,  red 
tree  voles.  Pacific  marsh  shrews)  are 
likely  to  be  affected  by  habitat  altera- 
tion. Species  with  large  populations, 
broad  habitat  specificity,  and  either 
large  (deer  mice)  or  small  (Trowbr- 
idge's shrews)  geographic  distribu- 
tions, are  less  likely  to  be  affected  by 
forest  management. 

Our  study  does  not  address 
changes  in  habitats  in  managed  for- 
ests stands  or  the  effects  of  forest 
fragmentation  as  remaining  old 


349 


growth  is  harvested.  Further  studies 
of  small  mammals  should  emphasize 
managed  stands  and  managed  land- 
scapes. 

Even  with  the  creation  of  old- 
growth  habitat  areas  on  National 
Forests,  most  of  the  landscape  will 
probably  be  in  plantations  less  than 
100  years  old.  Research  needs  to  be 
focused  on  the  degree  of  loss  of  di- 
versity in  these  managed  forests  and 
evaluate  silvicultural  options  for 
maintaining  or  enhancing  habitat 
structure. 

Thus  far,  there  is  little  evidence 
that  small  mammal  populations  in 
Douglas-fir  forests  are  strongly  influ- 
enced by  stand  size  or  amount  of  in- 
sularization  (Raphael  1984,  Rosen- 
berg and  Raphael  1986).  As  these  au- 
thors point  out,  however,  forest  frag- 
mentation in  western  coniferous  for- 
ests might  not  have  advanced  far 
enough  or  existed  long  enough  for 
effects  to  be  observed.  Conversely, 
forest  fragmentation  in  the  Pacific 
Northwest  is  not  usually  conversion 
of  forest  to  farmland  or  urban  areas 
as  is  the  case  in  other  temperate  re- 
gions (e.g.,  Wilcove  et  al.  1986, 
Askins  et  al.  1987,  Dickman  1987). 
Rather,  it  results  in  the  replacement 
of  one  forest  habitat  with  another. 
Patches  of  old  growth  in  a  managed 
forest  are  not  strict  analogs  of  oce- 
anic islands  or  isolated  mountain 
tops  (Harris  1984),  so  the  ability  of 
forest-floor  small  mammals  to  main- 
tain populations  in  managed  forests 
is  dependent  on  habitat  availability 
after  logging. 

Our  results  indicate  that  some 
"old-growth  species"  are  found  in 
younger  stands,  but  the  proximity  of 
old  growth  to  younger  forest  might 
partly  explain  their  occurrence.  The 
effect  of  stand  size,  shape,  edge,  and 
juxtaposition  on  small  mammal 
populations  needs  attention.  Where 
old  growth  and  other  habitat  areas 
are  set  aside  to  maintain  biological 
diversity  in  intensively  managed 
landscapes,  the  long-term  viability  of 
these  habitats  and  their  vertebrate 
populations  needs  to  be  monitored. 


ACKNOWLEDGMENTS 

We  thank  S.  Boyle,  L.  Hanebury,  D. 
Hayes,  S.  Martin,  T.  Olson,  and  S. 
Woodis  for  helping  to  install  pitfall 
arrays.  J.  Dragavon,  L.  Jones,  P.  Mor- 
rison, R.  Pastor,  and  D.  Smith 
checked  traps  and  processed  ani- 
mals. R.  Fisher  of  the  USNM  verified 
all  identifications.  A.  McKee  and  J. 
Moreau,  H.  J.  Andrews  Experimental 
Forest,  and  staff  at  the  Wind  River 
Experimental  Forest  and  Carson  Na- 
tional Fish  Hatchery  assisted  with 
housing  and  logistics.  We  appreciate 
the  critical  review  of  this  manuscript 
by  M.  Bogan,  A.  Carey,  M.  Raphael, 
and  F.  Samson.  This  is  contribution 
number  66  of  the  Wildlife  Habitat 
Relationships  in  Western  Washington 
and  Oregon  Project. 

LITERATURE  CITED 

Anthony,  R.  G.,  E.  D.  Forsman,  G.  A. 
Green,  G.  Witmer,  and  S.  K.  Nel- 
son. 1987.  Small-mammal  popula- 
tions in  riparian  zones  of  conifer- 
ous forests  in  western  Oregon. 
Murrelet  68:94-102. 

Anthony,  Robert  G.,  and  Michael  L. 
Morrison.  1985.  Influence  of 
glyphosphate  herbicide  on  small 
mammal  populations  in  western 
Oregon.  Northwest  Science  59:159- 
168. 

Askins,  Robert  A.,  Margarett  J.  Phil- 
brick,  and  David  S.  Sugeno.  1987. 
Relationship  between  the  regional 
abundance  of  forest  and  the  com- 
position of  forest  bird  communi- 
ties. Biological  Conservation 
39:129-152. 

Banks,  Richard  C,  Roy  W.  McDi- 
armid,  and  Alfred  L.  Gardner. 
1987.  Checklist  of  vertebrates  of 
the  United  States,  the  U.S.  Territo- 
ries, and  Canada.  USDI  Fish  and 
Wildlife  Service,  Resource  Publica- 
tion 166.  79  p.  Washington,  D.C. 

Brown,  James  H.  1984.  On  the  rela- 
tionship between  abundance  and 
distribution  of  species.  American 
Naturalist  124:255-279. 


Bury,  R.  Bruce,  and  Paul  Stephen 
Corn.  1987.  Evaluation  of  pitfall 
trapping  in  northwestern  forests: 
trap  arrays  with  drift  fences.  Jour- 
nal of  Wildlife  Management 
51:112-119. 

Carey,  Andrew  B.  1984.  A  critical 
look  at  the  issue  of  species-habitat 
dependency,  p.  346-351.  In  Pro- 
ceedings, Society  of  American  For- 
esters national  convention.  [Port- 
land, Or.,  October  16-20, 1983]  So- 
ciety of  American  Foresters, 
Washington,  D.C. 

Com,  Paul  Stephen,  and  R.  Bruce 
Bury.  1986.  Habitat  use  and  terres- 
trial activity  by  red  tree  voles  (Ar- 
borimus  longicaudus)  in  Oregon. 
Journal  of  Mammalogy  67:404-406. 

Corn,  Paul  Stephen,  and  R.  Bruce 
Bury.  1988.  Distribution  of  the 
voles  Arborimus  longicaudus  and 
Phenacomys  intermedius  in  the  Cen- 
tral Oregon  Cascades.  Journal  of 
Manunalogy  69:427-429. 

Cutler,  M.  Rupert.  1980.  A  wildlife 
policy  for  the  U.S.  Department  of 
Agriculture.  Transactions  of  the 
North  American  Wildlife  and 
Natural  Resources  Conference 
45:56-66. 

Cutler,  M.  Rupert.  1984.  Keynote  ad- 
dress. Old-growth  habitat:  an  en- 
dangered species,  p.  205-211.  In 
William  R.  Meehan,  Theodore  R. 
Merrell,  Jr.,  and  Thomas  A. 
Hanley,  eds..  Fish  and  wildlife  re- 
lationships in  old-growth  forests. 
Proceedings  of  a  symposium.  [Jun- 
eau, Alaska,  April  12-15,  1982]  425 
p.  American  Institute  of  Fishery 
Research  Biologists,  (available 
from  John  W.  Reintjes,  Rt.  4,  Box 
85,  Morehead  City,  NC  28557). 

Dickman,  C.  R.  1987.  Habitat  frag- 
mentation and  vertebrate  species 
richness  in  an  urban  environment. 
Journal  of  Applied  Ecology  24:337- 
351. 

Findley,  James  S.  1955.  Speciation  of 
the  wandering  shrew.  Publication 
of  the  University  of  Kansas  Mu- 
seum of  Natural  History  9:1-68. 

Forsman,  Eric  D.,  E.  Charles  Meslow, 
and  Howard  M.  Wight.  1984.  Dis- 


350 


tribution  and  biology  of  the  spot- 
ted owl  in  Oregon.  Wildlife  Mono- 
graphs 87:1-64. 

Franklin,  J.  F.,  K.  Cromack,  Jr.,  W. 
Denison,  A.  McKee,  C.  Maser,  J. 
Sedell,  F.  Swanson,  and  G.  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-fir  forests. 
USDA  Forest  Service  General 
Technical  Report  PN[W-118, 48  p. 
Pacific  Northwest  Forest  and 
Range  Experiment  Station,  Port- 
land, Or. 

Franklin,  Jerry  F.,  and  Thomas  A. 
Speis.  1984.  Characteristics  of  old- 
growth  Douglas-fir  forests,  p.  328- 
334.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [Portland,  Or.,  October 
16-20, 1983]  Society  of  American 
Foresters,  Washington,  D.C. 

Gashwiler,  Jay  S.  1959.  Small  mam- 
mal study  in  west-central  Oregon. 
Journal  of  Mammalogy  40:128-139. 

Gashwiler,  Jay  S.  1970.  Plant  and 
mammal  changes  on  a  clearcut  in 
west-central  Oregon.  Ecology 
51:1018-1026. 

Gauch,  Hugh  G.,  Jr.  1982.  Multivari- 
ate analysis  in  community  ecol- 
ogy. Cambridge  University  Press, 
Cambridge,  England.  298  p. 

Gunther  Pamela  M.,  Brenda  S.  Horn, 
and  Geoffrey  D.  Babb.  1983.  Small 
mammal  populations  and  food 
selection  in  relation  to  timber  har- 
vest practices  in  the  western  Cas- 
cade Mountains.  Northwest  Sci- 
ence 57:32-44. 

Harris,  Larry  D.  1984.  The  frag- 
mented forest.  Island  biogeogra- 
phy  theory  and  the  preservation  of 
biotic  diversity.  University  of  Chi- 
cago Press,  Chicago,  111.  211  p. 

Harris,  Larry  D.,  and  Chris  Maser. 
1984.  Animal  community  charac- 
teristics, p.  44-68.  In  The  frag- 
mented forest.  Island  biogeogra- 
phy  theory  and  the  preservation  of 
biotic  diversity.  University  of  Chi- 
cago Press,  Chicago,  111.  211  p. 

Harris,  Larry  D.,  Chris  Maser,  and 
Arthur  McKee.  1982.  Patterns  of 
old  growth  harvest  and  implica- 
tions for  Cascades  wildlife.  Trans- 


actions of  the  North  American 
Wildlife  and  Natural  Resources 
Conference  47:374-392. 
Hooven,  Edward  F.,  and  Hugh  C. 
Black.  1976.  Effects  of  some 
clearcutting  practices  on  small- 
mammal  populations  in  western 
Oregon.  Northwest  Science  50:189- 
208. 

Maser,  Chris,  Bruce  R.  Mate,  Jerry  F. 
Franklin,  and  C.  T.  Dymess.  1981. 
Natural  history  of  Oregon  coast 
mammals.  USDA  Forest  Service 
General  Technical  Report  PNW- 
133.  496  p.  Pacific  Northwest  For- 
est and  Range  Experiment  Station. 
Portland,  Or. 

Maser,  Chris  and  James  M.  Trappe. 
1984.  The  seen  and  unseen  world 
of  the  fallen  tree.  USDA  Forest 
Service  General  Technical  Report 
PNW-164.  56  p.  Pacific  Northwest 
Forest  and  Range  Experiment  Sta- 
tion. Portland,  Or. 

Maser,  Chris,  James  M.  Trappe,  and 
Ronald  A.  Nussbaum.  1978.  Fun- 
gal-small mammal  interrelation- 
ships with  emphasis  on  Oregon 
coniferous  forests.  Ecology  59:799- 
809. 

Meehan,  William  R.,  Theodore  R. 
Merrell,  Jr.,  and  Thomas  A. 
Hanley,  eds.  1984.  Fish  and  wild- 
life relationships  in  old-growth 
forests.  Proceedings  of  a  sympo- 
sium. [Juneau,  Alaska,  April  12-15, 
1982]  425  p.  American  Institute  of 
Fishery  Research  Biologists,  (avail- 
able from  John  W.  Reintjes,  Rt.  4, 
Box  85,  Morehead  City,  N.C. 
28557). 

Meslow,  E.  Charles,  Chris  Maser, 
and  Jared  Verner.  1981.  Old- 
growth  forests  as  wildlife  habitat. 
Transactions  of  the  North  Ameri- 
can Wildlife  and  Natural  Re- 
sources Conference  46:329-335. 

Rabinowitz,  Deborah,  Sara  Cairns, 
and  Theresa  Dillon.  1986.  Seven 
forms  of  rarity  and  their  frequency 
in  the  British  Isles,  p.  182-204.  In 
Michael  E.  Soule,  ed..  Conserva- 
tion biology.  The  science  of  scar- 
city and  diversity.  584  p.  Sinauer 
Associates,  Sunderland,  Mass. 


Raphael,  Martin  G.  1984.  Wildlife 
populations  in  relation  to  stand 
age  and  area  in  Douglas-fir  forests 
of  northwestern  California,  p.  259- 
274.  In  William  R.  Meehan,  Theo- 
dore R.  Merrell,  Jr.,  and  Thomas 

A.  Hanley,  eds..  Fish  and  wildlife 
relationships  in  old-growth  for- 
ests. Proceedings  of  a  symposium. 
[Juneau,  Alaska,  April  12-15,  1982] 
425  p.  American  Institute  of  Fish- 
ery Research  Biologists,  (available 
from  John  W.  Reintjes,  Rt.  4,  Box 
85,  Morehead  City,  N.C.  28557). 

Raphael,  Martin  G.,  and  Reginald  H. 
Barrett.  1984.  Diversity  and  abun- 
dance of  wildlife  in  late  succes- 
sional  Douglas-fir  forests,  p.  352- 
360.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [Portland,  Or.,  October 
16-20, 1983]  Society  of  American 
Foresters,  Washington,  D.C. 

Rosenberg,  Kenneth  V.,  and  Martin 
G.  Raphael.  1986.  Effects  of  forest 
fragmentation  on  vertebrates  in 
Douglas-fir  forests,  p.  263-272.  In 
Jared  Verner,  Michael  L.  Morri- 
son, and  C.  John  Ralph,  eds.. 
Wildlife  2000:  modeling  habitat 
relationships  of  terrestrial  verte- 
brates, [symposium  held  October 
7-11, 1984]  470  p.  University  of 
Wisconsin  Press,  Madison. 

Ruggiero,  Leonard  F.,  and  Andrew 

B.  Carey.  1984.  A  programmatic 
approach  to  the  study  of  old- 
growth  forest — wildlife  relation- 
ships, p.  340-345.  In  Proceedings, 
Society  of  American  Foresters  na- 
tional convention.  [Portland,  Or., 
October  16-20, 1983]  Soc.  of  Amer- 
ican Foresters,  Washington,  D.C. 

Schoen,  John  W.,  Olof  C.  Wallmo, 
and  Matthew  D.  Kirchoff.  1981. 
Wildlife-forest  relationships:  is  a 
reevaluation  of  old  growth  neces- 
sary? Transactions  of  the  North 
American  Wildlife  and  Natural 
Resources  Conference  46:531-544. 

Spies,  Thomas  A.  1983.  Study  plan. 
Characterization  of  old-growth 
Douglas-fir  forests  of  western  Ore- 
gon and  Washington.  22  p.  USDA 
Forest  Service,  Corvallis,  Or. 


351 


Spies,  Thomas  A.,  and  Jerry  F.  Fran- 
klin. In  press.  Old  growth  and  for- 
est dynamics  in  the  Douglas-fir 
region  of  western  Oregon  and 
Washington.  Natural  Areas  Jour- 
nal. 

Spies,  Thomas  A.,  Jerry  F.  Franklin, 
and  Ted  B.  Thomas.  In  press. 
Coarse  woody  debris  in  Douglas- 
fir  forests  of  western  Oregon  and 
Washington.  Ecology. 

Sullivan,  Thomas  P.,  and  Charles  J. 
Krebs.  1980.  Comparative  demog- 
raphy of  Peromyscus  maniculatus 
and  Microtus  oregoni  populations 
after  logging  and  burning  of 
coastal  forest  habitats.  Canadian 
Journal  of  Zoology  58:2252-2259. 

Tevis,  Lloyd,  Jr.  1956.  Responses  of 
small  mammal  populations  to  log- 
ging of  Douglas-fir.  Journal  of 
Mammalogy  37:189-196. 

Thomas,  Donald  W.,  and  Stephen  D. 
West.  1984.  On  the  use  of  ultra- 
sonic detectors  for  bat  species 
identification  and  the  calibration 
of  QMC  Mini  Bat  Detectors.  Cana- 
dian Journal  of  Zoology  62:2677- 
2679. 

Ure,  Douglas  C,  and  Chris  Maser. 
1982.  Mycophagy  of  red-backed 
voles  in  Oregon  and  Washington. 
Canadian  Journal  of  Zoology 
60:3307-3315. 


West,  Stephen  D.  1985.  Differential 
capture  between  old  and  new 
models  of  the  Museum  Special 
snap  trap.  Journal  of  Mammalogy 
66:798-800. 

Whitaker,  John  O.,  Jr.,  and  Chris  Ma- 
ser. 1976.  Food  habits  of  five  west- 
ern Oregon  shrews.  Northwest 
Science  50:102-107. 

Wilcove,  David 


system  for  statistics.  Version  4.0. 
Systat  Inc.,  Evanston,  111.  822  p. 
Williams,  Daniel  F.,  and  Suzanne  E. 
Braun.  1983.  Comparison  of  pitfall 
and  conventional  traps  for  sam- 
pling small  mammal  populations. 
Journal  of  Wildlife  Management 
47:841-845. 


S.,  Charles 
H.  McLellan, 
and  Andrew 
P.  Dobson. 
1986.  Habitat 
fragmenta- 
tion in  the 
temperate 
zone.  p.  237- 
256.  In  Mi- 
chael E. 
Soule,  ed.. 
Conservation 
biology.  The 
science  of 
scarcity  and 
diversity. 
584  p.  Sin- 
auer  Associ- 
ates, Sunder- 
land, Mass. 
Wilkinson,  Le- 
land.  1988. 
SYSTAT.  The 


Table  Al  .—Ages,  elevations,  and  locations  of  18  stands  in 
Oregon.  Locations  are  distances  (km)  from  McKenzie  Bridge. 


Stand 

Type 

Age  (yr) 

Elev.  (m) 

Location 

15 

OGW 

450 

795 

Unn  Co.,  9.2  N,  1.2  W 

03 

OGW 

450 

815 

Lane  Co.,  6.4  N, 0.6  E 

24 

OGW 

450 

860 

Lane  Co..  6.0  N,  0,6  E 

02 

OGM 

450 

560 

Lane  Co.,  4.6  N.  6.5  W 

17 

OGM 

450 

790 

Lane  Co.,  6.6  N,  0.3  W 

33 

OGD 

200 

670 

Lane  Co.,  6.0  N,  7.5  E 

25 

OGD 

195 

500 

Lane  Co..  2.4  N,  7.5  W 

29 

OGD 

200 

700 

Lane  Co..  2.6  S 

11 

Mature 

140 

670 

Lane  Co..  5.4  S,  8.8  E 

35 

Mature 

130 

900 

Linn  Co.,  10.5  N,  LOW 

42 

Mature 

150 

1030 

Lane  Co..  3.1  N,3.0W 

39 

Young 

76 

1050 

Lane  Co.,  4.4  S,  14,3  E 

47 

Young 

50 

1110 

Lane  Co.,  3.3  N,2.4W 

48 

Young 

69 

1075 

Unn  Co,,  13.2  N,0,8E 

75 

Young 

30 

560 

Lane  Co.,  1.6  S,  5,2  W 

55 

Clearcut 

9 

830 

Lane  Co.,  2.8  N,6.6W 

291 

Clearcut 

5 

690 

Lane  Co.,  2.6  S,  1.4E 

391 

Clearcut 

5 

1100 

Lane  Co.,  3.8  S,  14.8  E 

r 


Table  A2.— Ages,  elevations,  and  locations  of  12  stands  in 
Wastiington.  Locations  are  distances  (km)  from  Carson,  Ska- 


mania  Co. 

Stand 

Type 

Age  (yr) 

□ev.  (m) 

Location 

14 

OGW 

375 

520 

17.7N,  16.9W 

12 

OGM 

450 

485 

6.4  N,  11,3W 

21 

OGM 

375 

440 

17.2  N,  14.0  W 

20 

OGM 

375 

480 

11.3  N,  11.9  W 

31 

OGD 

375 

670 

18.5  N,  16,5  W 

41 

Mature 

105 

485 

19.3  N,  13,7  W 

42 

Mature 

140 

500 

13.7N,2.4W 

50 

Mature 

130 

610 

16,0  N,  2.1  W 

60 

Young 

65 

475 

13.6  N,  12.1  W 

61 

Young 

65 

640 

8.1  N,6.3W 

70 

Clearcut 

5 

535 

1 1.3  N,  13.4  W 

71 

Clearcut 

5 

730 

16,9  N.  7.2  W 

Table  Bl  .—Stand  structural  and  vegetation  variables. 


Variable  name  Description 


SLOPE 

TRASPECT 

LNDC12 

LNDC3 
LNDC45 

LNDM1 
LNDM2 

LNDM3 
LBDC12 

LBCD3 
LBDC45 

MDBH 
TOTDEN 
TOTBA 
LiTFER 
TOTCF 
MOSS 
FERN 
GRASS 
HERB 
EGSHR 
DESHR 
EVGTR 
DECTR 


Percent  slope 

Transformed  aspect 

Number  of  logs  per  tia,  decay  class  1 

and  2 

Number  of  logs  per  ho,  decay  class  3 
Number  of  logs  per  t^a,  decay  class  4 
and  5 

Number  of  logs  per  ha.  <30cm  diameter 
Number  of  logs  per  ha.  >30cm  and  <60 
cm 

Number  of  logs  per  ha.  >60  cm 
Biomass  (1 ,000  kg  per  ha)  of  logs,  class  1 
and  2 

Biomass  (1 .000  kg  per  ha)  of  logs,  class  3 
Biomass  (1 ,000  kg  per  ha)  of  logs,  class  4 
and  5 

Mean  d.b.h.  (cm)  in  stand 

Density  of  live  trees  (number  per  ha) 

Basal  area  of  live  trees  (m'  per  ha) 

Litter  depth  (01  +  02  horizons;  cm) 

Volume  (%)  of  coarse  fragments  in  soil 

%  cover  by  mosses 

%  cover  by  fems 

%  cover  by  grasses 

%  cover  by  herbaceous  vegetation 

%  cover  by  evergreen  shrubs 

%  cover  by  deciduous  shrubs 

%  cover  by  evergreen  trees 

%  cover  by  deciduous  trees 


352 


Evaluation  of  Small  Mammals 
as  Ecological  Indicators  of 
Old-Growth  Conditions^ 

Kirk  A.  Nordyke^  and  Steven  W.  Buskirk^ 


Abstract.— The  use  of  small  mammals  as 
ecological  indicators  of  old-growth  conditions  was 
evaluated  from  trapping  studies  conducted  in  forest 
stands  reflecting  a  range  of  old-growth  conditions  in 
southeastern  Wyoming.  The  relationship  between 
abundance  of  Clethrionomys  gapperi  an6  old- 
growth  conditions  was  expressed  in  a  quadratic 
function.  Tamias  minimus  and  Peromyscus 
maniculatus  were  negatively  correlated  with  old- 
growth  conditions.  C.  gapperi  '\s  the  most  likely 
candidate  for  a  small  mammal  ecological  indicator 
of  old-growth  conditions  in  spruce-fir  stands. 


Recent  emphasis  in  forest  manage- 
ment has  been  placed  on  an  inte- 
grated multiple-benefit  approach  to 
land  and  resource  planning  and  man- 
agement (Salwasser  et  al.  1982).  The 
National  Forest  Management  Act 
(NFMA)  was  enacted  in  1976  to  es- 
tablish revised  goals  for  the  USDA 
Forest  Service.  NFMA  regulations 
require  that  detailed  plans  be  devel- 
oped and  implemented  in  each  na- 
tional forest.  A  specific  goal  is  to 
manage  wildlife  and  fish  habitats  to 
maintain  viable  populations  of  all 
existing  native  vertebrate  species  in 
the  planning  area  and  to  maintain 
and  improve  habitats  of  management 
indicator  species  (MIS)  (36  CFR 
219.19).  In  addition,  population 
trends  of  MIS  are  to  be  monitored 
and  relationships  of  those  trends  to 
habitat  changes  must  be  determined 
(36  CFR219.19[a][6]). 

Ecological  indicator  species  com- 
prise one  category  of  MIS  and  were 
defined  for  management  purposes  as 
"...plant  or  animal  species  selected 
because  their  population  changes  are 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortt^  America.  (Flag- 
staff. AZ.  July  19-21.  1988.) 

'Kirk  A.  Nordyke  is  a  graduate  student. 
The  University  of  Wyoming.  Department  of 
Zoology  and  Physiology.  University  Station 
Box  3 166.  Laramie.  W/  82071. 

^Steven  W.  Buskirk  is  Assistant  Professor. 
The  University  of  Wyoming.  Department  of 
Zoology  and  Physiology.  University  Station 
Box  3166.  Laramie.  WY  82071. 


believed  to  indicate  the  effects  of 
management  activities  on  other  spe- 
cies of  selected  major  biological  com- 
munities or  on  water  quality"  (36 
CFR  219.19[a][l]).  Ecological  indica- 
tors should  have  a  high  degree  of 
sensitivity  to  perturbation  and  be 
representative  of  habitat  needs  of 
other  species  (Patton  1987).  Thus, 
population  responses  of  an  ecological 
indicator  species  to  habitat  perturba- 
tions should  reflect  similar,  yet  less 
severe,  responses  in  more  tolerant 
species  (Graul  and  Miller  1984).  Eco- 
logical indicators  should  be  easily 
monitored  to  achieve  realistic  goals. 

Compliance  with  the  monitoring 
requirements  of  NFMA  presents  a 
major  challenge  to  national  forest 
management  because  costs  may  be 
high  and  because  methods  are  still 
being  developed  (Verner  1983).  The 
challenge  is  most  pressing  in  old- 
growth  forests:  this  important  habitat 
is  disappearing  at  an  alarming  rate 
and  vertebrate  populations  depend- 
ent on  old-growth  features  are  de- 
clining (Harris  1984).  NFMA  guide- 
lines mandate  that  old -growth  be  a 
significant  element  in  the  diversity  of 
forest  conditions.  To  accomplish  this, 
old-growth  and  associated  fauna 
must  be  characterized  and  monitored 
to  determine  that  management  prac- 
tices will  not  impair  their  productiv- 
ity (Juday  1978). 

Of  22  species  selected  as  ecological 
indicator  species  for  the  Medicine 
Bow  National  Forest  (MBNF),  Gap- 
Iyer's  red-backed  vole  (Clethrionomys 


gapperi)  is  the  only  small  mammal 
ecological  indicator  for  old-growth 
conditions  (USDA  Forest  Service 
1985).  Old-growth  forests  represent 
optimal  habitat  for  C.  gapperi  (Jerry 
1984).  Limiting  factors  to  habitat  use 
by  C.  gapperi  may  include  require- 
ments for  water  (Getz  1968,  Merritt 
1981)  and  log  cover  (Tevis  1956). 
Old-growth  generally  exhibits  more 
mesic  conditions  than  other  forest 
habitats.  Logs  provide  cover  from 
predators  and  weather  (Maser  et  al. 
1979),  pathways  into  new  habitats 
(Franklin  et  al.  1981),  and  mesic  sites 
for  fungal  growth  (Maser  and  Trappe 
1984).  The  importance  of  fungi  as  a 
food  for  C.  gapperi  has  only  recently 
been  recognized  (Martell  1981,  Maser 
et  al.  1978a).  Mesic  conditions  of  old- 
growth  stands  favor  the  occurrence 
of  fungi  (Maser  et  al.  1978b). 

The  indicator  species  concept  was 
adopted  by  the  Forest  Service  in  the 
late  1970s,  but  its  viability  as  a  moni- 
toring approach  has  not  been  investi- 
gated. Certain  parameters  of  C.  gap- 
peri |X)pulations  were  assumed  to 
reflect  changes  in  old-growth  condi- 
tions that  result  from  management 
activity.  This  paper  describes  a  study 
investigating  the  application  of  the 
indicator  species  concept  to  old- 
growth  management.  Oir  objective 
was  to  evaluate  the  responses  of 
small  mammal  populations  to  a 
range  of  old-growth  conditions.  Spe- 
cifically, we  investigated  whether 
abundance  of  C.  gapperi  was  related 
to  old-growth  condition. 


353 


study  Area  and  Methods 

Our  study  area  in  the  Snowy  Range 
included  upper  montane  (2300-2750 
m)  and  subalpine  (2750  m-timberline) 
zones.  Lodgepole  pine  (Pinus  con- 
torta)  was  the  dominant  overstory 
species  in  the  montane  zone;  it  also 
dominated  south  slopes  and  ridge 
tops  at  higher  elevations  (Romme 
and  Knight  1981).  Engelmann  spruce 
(Picea  engelmannii)  and  subalpine  fir 
(Abies  lasiocarpa)  were  generally  co- 
dominant  in  the  subalpine  zone  (Al- 
exander 1974).  Engelmann  spruce 
and  subalpine  fir  are  climax  species 
(Romme  and  Knight  1981)  and  often 
develop  old-growth  conditions.  Old- 
growth  conditions  in  lodgepole- 
dominated  stands  are  less  common 
due  to  a  shorter  fire  interval  and  a 
slower  rate  of  succession  (Romme 
and  Knight  1981).  Understory  vege- 
tation was  sparse  and  generally  con- 
sisted of  common  juniper  (Juniperus 
communis)  and  broom  huckleberry 
(Vaccinium  scoparium). 

Field  studies  were  conducted  in 
the  MBNF  (fig.  1)  from  June  to  Sep- 
tember of  1986  and  1987.  We  estab- 
lished eight  study  plots  in  spruce-fir 
stands  reflecting  a  range  of  old- 
growth  conditions.  Because  plots 
were  located  on  both  the  east  and 
west  slopes,  a  paired  design  was 
used  to  control  for  the  effects  of  ma- 
jor relief.  A  1.42-ha  trap  grid  with  80 
Museum  Special  snap-traps  (8  by  10 
pattern  with  15-m  intervals)  was  lo- 
cated on  each  plot.  In  1987,  four  ad- 


ditional grids  were  located  on  plots 
dominated  by  lodgepole  pine.  Snap- 
traps  were  baited  with  peanut  butter 
and  oatmeal.  Beginning  in  July  1986, 
we  trapped  each  grid  for  three  con- 
secutive nights  and  checked  traps 
daily  in  early  morning.  If  rainfall 
caused  the  release  of  snap-trap 
mechanisms,  trapping  effort  was  ex- 
tended by  as  many  nights  as  it  rained 
(table  1). 

In  1987,  we  rated  the  old-growth 
condition  of  our  study  plots  with  the 
old-growth  scorecard  developed  spe- 
cifically for  the  MBNF  (Marquardt 


1984)  (table  1).  The  scorecard  is  com- 
pleted subjectively  by  Forest  Service 
personnel  and  is  based  on  structural 
characteristics  of  stands.  Structural 
characteristics  that  define  old-growth 
stands  in  the  MBNF  include  trees 
with  large  diameters,  long-lived 
dominant  species  (i.e.,  Engelmann 
spruce  and  subalpine  fir),  a  multi- 
storied  stand  structure,  dense  cano- 
pies, multiple  species,  woody  debris 
on  the  forest  floor,  and  standing 
snags  (Marquardt  1984).  The  score- 
card  incorporates  sub-scores  for  each 
of  these  structural  characteristics  to 


Figure  1  .—Map  of  the  study  area  in  the  Medicine  Bow  National  Forest  of  southeastern  Wyo- 
ming, about  65  km  west  of  Laramie.  Twelve  study  plots  <A-L)  were  establlshied  In  1986  and 
1987  for  Intensive  trapping  and  habitat  characterizations. 


^—  ■ —  —  ~>i 

Table  1.— Information  collected  from  12  study  plots  In  the  l^edicine  Bow  National  Forest  of  southeastern  Wyoming  In 
1986  and  1987. 


Study  plot 

Characteristic       A         B         C         D         E         F         G        H         I  J         K  L 


USFSIocation            201307  201303  201304  205403  205404    201303  201303  208103  201810  201809  204907  205503 
site                         03  07  05  34  20          08  10  23  15  10  07  26 
Dominant  overstory   spruce  spruce  spruce  spruce  spruce    spruce  spruce  spruce  lodge-  lodge-  lodge-  lodge- 
fir  fir  fir  fir  'fir  fir  fir  pole  pole  pole  pole 
Old-growth  rating      51  35  48  44  50         37  40  41  25  22  22  19 
Trapping  effort  1986   240  240  240  320  240  400  240  240  0  0  0  0 
(#  trap  nights)  1987     320  320  320  320  240        240  240  240  240  240  240  240 

 :  :  .   y 


354 


achieve  an  overall  old-growth  rating, 
ranging  from  0  to  60. 

We  quantitatively  determined 
sub-scores  for  characteristics  we  be- 
lieved were  most  important  to  meet- 
ing habitat  needs  of  C.  gapperi.  Log 
density  was  estimated  with  the 
point-quarter  distance  method  (25 
sampling  points)  and  the  diameter  of 
each  log  sampled  (100  logs  were 
sampled)  was  measured  to  deter- 
mine the  mean  log  diameter. 

Data  analyses  were  performed 
with  the  SPSS  computer  package 
(Nie  et  aL  1975).  Analyses  involved 
linear  and  quadratic  correlation  tests 
between  small  mammal  abundance 


40.0 

^  35.0 
o 

2  30.0 

B  25.0 

w 

IT 

S  20.0 

s 

O  5.0 
0.0 


□  1986 

□  1987 


n     /  / 


/ 


mrfl 


7 

^  14 


/ 


D      E      F  G 
Study  PM 


I  J  KL 


Figure  2.— Capture  success  of  Clethriono- 
mys  gapperi  in  12  study  plots  In  \he  Medi- 
cine Bow  National  Forest  of  southieastem 
Wyoming  in  1986  and  1987.  Temporal  vari- 
ation in  abundance  was  extreme  In  five  of 
ttie  eigtit  spruce-fir  plots  sampled  botti 
years.  Plots  l-L  were  dominated  by  lodge- 
pole  pine  and  were  sampled  only  in  1987. 


35.0 

z 

o  30.0 
o 

r  25.0 


ir 

«  20.0 


§  15  0 

10 
e 

5  10.0 

"  5.0 


0.0 


1967 
P  .  0.007 
r«0.81 


0.0  100  200  30.0  40.0 

Stand  Rating 


50.0 


Figure  3.— Capture  success  of  Clethriono- 
mys  gapperi  \n  \he  Medicine  Bow  National 
Forest  of  souttieastem  Wyoming  in  1987  as 
a  function  of  old-growth  ratings.  Ttiis  rela- 
tionshiip  is  best  explained  by  a  quadratic 
correlation.  Dahcened  data  p>oints  repre- 
sent lodgepole-dominated  study  plots; 
open  data  points  represent  spruce-fir- 
dominated  study  plots. 


(as  inferred  from  capture  success) 
and  old-growth  ratings.  While  inter- 
ested primarily  in  the  responses  of  C. 
gapperi  populations,  we  also  evalu- 
ated the  responses  of  other  small 
mammal  species  that  were  captured. 


Results  and  Discussion 

A  total  of  695  small  mammals  were 
captured  in  5,360  trap  nights  (TN).  In 
decreasing  abundance,  these  were  C. 
gapperi,  Tamias  minimus,  Sorex  spp., 
Peromyscus  maniculatus,  Phenacomys 
intermedins,  Sorex  cinereus,  S.  montico- 
lus,  and  Microtus  longicaudus.  Only 
captures  of  C.  gapperi  and  T.  minimus 
were  frequent  enough  to  provide 
data  for  analysis  both  years;  captures 
of  P.  maniculatus  were  adequate  only 
in  1987.  Other  species  were  rarely 
captured. 


Temporal  Fluctuations  in 
Abundance 

Mean  capture  success  increased 
three-fold  from  1986  (5.6/lOOTN)  to 
1987  (18.0/lOOTN).  Capture  success 
of  C.  gapperi  is  representative  of  this 
variation  (fig.  2).  Natural  fluctuations 
in  small  mammal  abundance  are  well 
documented  (Krebs  and  Myers  1974, 
Vaughan  1969).  Such  fluctuations  are 
a  major  source  of  confounding  vari- 
ation and  hinder  the  ability  of  man- 
agers to  monitor  p)opulations  for 
changes  that  result  from  human-in- 
duced disturbance.  Because  of  this 
temporal  variation  in  abundance,  we 
separated  the  data  for  analysis. 


Association  of  C.  gapperi  witti 
Old-Growth  Conditions 

In  1986,  the  abundance  of  C.  gapperi 
was  weakly  correlated  linearly  with 
old-growth  ratings  (r  =  0.62,  P  = 
0.097).  However,  this  result  repre- 
sented only  the  range  of  old-growth 
conditions  found  in  spruce-fir  stands 
(scores  ranged  from  35  to  51).  Four 


lodgepole  pine  study  plots,  which 
rated  lowest  on  the  old-growth 
scorecard  and  provided  a  greater 
range  of  ratings  (19-51),  were  added 
in  1987.  A  more  complete  pattern 
emerged:  C.  gapperi  was  most  abun- 
dant in  the  lowest-scoring  lodgepole 
study  plot,  decreased  in  the  remain- 
ing lodgepole  plots,  further  de- 
creased to  a  minimum  in  the  mid- 
range  spruce-fir  plots,  and  then  in- 
creased in  abundance  with  increasing 
old-growth  condition  in  the  remain- 
ing spruce-fir  plots.  A  quadratic  cor- 
relation model  best  explained  the  re- 
lationship between  abundance  of  C. 
gapperi  and  old-growth  ratings  in 
1987  (r  =  0.81,  P  =  0.007;  fig.  3). 

The  highly  significant  quadratic 
function  that  described  the  relation- 
ship between  abundance  of  C.  gapperi 
and  old-growth  rating  in  1987  should 
be  interpreted  separately  for  the 
lodgepole  pine  and  spruce-fir  seg- 
ments. In  spruce-fir  plots,  the  rela- 
tionship was  positive  (r  =  0.89,  P  = 
0.003),  as  it  was  (suggestively)  in 
1986.  However,  a  comparison  of  C 
gapperi  abundance  in  spruce-fir  plots 
between  1986  and  1987  was  not  sig- 
nificant (r  =  0.43,  P  =  0.290).  This  in- 
dicated that  the  spruce-fir  plots  sup- 
porting high  densities  of  C.  gapperi  in 
1986  were  not  the  same  plots  sup- 
porting high  densities  in  1987.  In 
lodgepole  plots  (1987  only),  abun- 
dance of  C.  gapperi  was  not  signifi- 
cantly correlated  with  old-growth 
raHng  (r  =  -0.88,  P  =  0.116).  There- 
fore, we  are  not  confident  in  the  re- 
sults from  the  lodgepole  plots,  but  an 
interpretation  is  warranted.  The 
abundance  of  C.  gapperi  in  both  serai 
phases  (lodgepole  and  spruce-fir) 
was  strongly  influenced  by  the  abun- 
dance of  woody  debris  (particularly 
logs)  on  the  forest  floor.  However, 
these  two  stand  types  differ  mark- 
edly in  terms  of  the  source,  size  and 
likely  persistence  of  logs. 

In  spruce-fir  plots,  logs  were  large 
(mean  diameter  was  31.0  cm)  and 
were  recruited  through  the  natural 
processes  of  wind  throw  and  snag 
decay.  Log  size  and  biomass  are 


355 


greater  in  older  forests  than  in 
younger  forests  (Franklin  et  al.  1981). 
Thus,  availability  and  size  of  logs  in- 
crease with  time  in  young  spruce-fir 
stands,  and  we  believe  that  this  in- 
crease was  primarily  responsible  for 
the  relationship  we  found  between 
abundance  of  C.  gapperi  and  old- 
growth  rating  of  spruce-fir  plots.  In 
lodgepole  plots,  logs  were  smaller 
than  in  spruce-fir  plots  (mean  diame- 
ter was  22.7  cm;  t  =  7.93,  P  =  0.004) 
and  were  recruited  almost  entirely 
by  thinning.  One  lodgepole  plot  (plot 
L,  in  site  205503-26,  table  1)  had  been 
thinned  13  months  before  we 
sampled  it  and  had  a  high  density  of 
logs  and  the  greatest  abundance  of  C. 
gapperi.  This  single  plot  overwhelm- 
ingly influenced  the  lodgepole  phase 
of  the  quadratic  function. 

Lodgepole  stands  do  not  thin  well 
naturally  (Alexander  1974),  so  log 
recruitment  rates  and  densities  are 
generally  low.  We  predict  that,  be- 
cause they  are  larger  and  are  re- 
cruited at  a  less  variable  rate,  logs  in 
spruce-fir  stands  are  more  persistent 
over  time  than  are  logs  in  lodgepole 
stands.  Kirkland  (1977)  and  Martell 
and  Radvanyi  (1977)  found  high  den- 
sities of  C.  gapperi  in  clearcuts  one 
year  after  logging  spruce  forests. 
Three  years  after  logging,  Martell 
and  Radvanyi  found  that  C.  gapperi 
had  become  rare.  Gunther  et  al. 
(1983)  attributed  the  abundance  of  C. 
gapperi  in  clearcuts  to  high  ground 
cover  created  by  felled  trees  and 
slash  and  to  an  abundant  food  sup)- 
ply  of  lichens. 

Interpretation  of  C.  gapperi  abun- 
dance as  an  indication  of  old-growth 
condition  must  be  undertaken  with 
caution.  C.  gapperi  appears  to  re- 
spond to  natural  processes  of  log  ac- 
cumulation; however,  C.  gapperi 
populations  also  appear  to  respond 
to  accumulation  of  woody  debris  re- 
sulting from  management  actions. 
Stand  thinning  is  more  common  in 
lodgepole  than  in  spruce-fir  stands  in 
the  MBNF  (T.  Cartwright,  pers. 
comm.),  so  use  of  C.  gapperi  as  an  in- 
dicator of  old-growth  conditions  of 


spruce-fir  stands  appears  less  likely 
to  be  confounded  by  this  factor. 


Association  of  T.  minimus  and  P. 
maniculatus  with  Old-Growtti 
Conditions 

The  broad  habitat  affinities  of  these 
two  species  are  well  documented 
(Armstrong  1977).  In  forested  habi- 
tats, they  are  associated  with  early 
successional  stages  (Martell  1984).  In 
our  study,  T.  minimus  abundance  de- 
creased with  increasing  old-growth 
rating  in  1986  (r  =  -0.71,  P  =  0.046; 
fig.  4),  but  the  correlation  was  based 
on  a  narrow  range  of  ratings  so  that 
its  reliability  is  questionable. 
Vaughan  (1974)  noted  this  species' 
dependence  on  stumps  and  rocks  for 
lookout  points.  Certain  structural 
features  that  characterize  old-growth 
conditions  (e.g.,  restricted  average 
sight  distance)  are  inconsistent  with 
the  open  habitat  requirements  of  T. 
minimus.  There  was  no  significant 
correlation  in  1987.  Given  the  high 
population  levels  that  year,  limited 
resources  in  preferred  habitat  may 
have  caused  T.  minimus  to  disperse 
into  less  preferred  habitat. 

Abundance  of  P.  maniculatus  de- 
creased with  increasing  old-growth 
raring  in  1987  (r  =  -0.60,  P  =  0.039; 
fig.  5),  but  the  correlation  was  driven 
by  one  data  point  (study  plot  L,  table 
1).  The  abundance  of  P.  maniculatus 
has  been  shown  to  increase  with 
understory  vegetation  (Tevis  1956).  If 
this  is  due  to  an  affinity  for  cover, 
then  the  conditions  present  in  study 
plot  L  may  explain  the  high  numbers 
of  P.  maniculatus  found  there.  If  the 
data  point  is  excluded  from  the 
analysis,  the  result  supp>orts  the 
broad  habitat  distributions  P.  manicu- 
latus is  known  to  exhibit. 


Conclusions 

We  found  that  abundance  of  C.  gap- 
peri was  correlated  with  old-growth 
ratings  in  spruce-fir  stands,  and  at- 


tribute that  correlation  primarily  to 
the  log  component  of  the  old-growth 
rating.  C.  gapperi  was  strongly  corre- 
lated with  old-growth  conditions  in 
spruce-fir  and  may  be  predictive  of 
old-growth  condition  in  that  stand 
type.  However,  C.  gapperi  appears  to 
respond  to  logs  recruited  from  man- 
agement activities,  and  caution 
should  be  used  in  interpreting  abun- 
dance data. 

Our  results  neither  support  nor 
refute  the  assumption  that  C.  gapperi 
represents  the  habitat  needs  of  other 
species.  Alternative  monitoring  ap- 
proaches may  have  utility  in  forest 
management.  These  include  guild- 
indicator  sf)ecies,  whole-guild,  and 
community-based  monitoring 
schemes. 


4.0 

I 

°  3.0 


2.0 


1.0 


0.0 


1966 
P  =  0.046 
r  =  -0.71 


0.0  10.0  20.0  30.0  40.0 

Stand  FUting 


50.0 


Figure  4.— Capture  success  of  Tamias  mini- 
mus in  the  Medicirie  Bow  National  Forest  of 
southeastern  Wyoming  in  1 986  as  a  function 
of  old -growth  ratings  in  spruce-fir-domi- 
nated  study  plots. 


6.0 


K  5.0 

s 

o 


3.0 


%  2.0 

3 

O  1.0 


0.0 


1967 
P  .  0.039 
r.-0.60 


0.0  10.0  20.0  30.0  40.0 

Stand  Rating 


50.0 


Figure  5.— Capture  success  of  Peromyscus 
maniculatus  in  the  Medicine  Bow  National 
Forest  of  soutt>eastem  Wyoming  in  1987  os 
a  function  of  old-growth  ratings.  Darkened 
data  points  represent  lodgepole-domi- 
nated  study  plots;  open  data  points  repre- 
sent spruce -fir-dominated  study  plots. 


356 


Tamias  minimus  and  P.  maniculatus 
populations  responded  in  a  manner 
consistent  with  their  habitat  affini- 
ties. Thus,  C.  gapperi  may  be  the  only 
choice  for  consideration  as  a  small 
mammal  ecological  indicator  of  old- 
growth  conditions  in  the  MBNF. 

Acknowledgments 

We  wish  to  thank  the  USDA  Forest 
Service,  Medicine  Bow  National  For- 
est, and  the  Wyoming  Game  and  Fish 
Department  for  funding  this  project. 
We  appreciate  very  constructive  re- 
views by  M.  Raphael  and  W.  Block. 

Literature  Cited 

Alexander,  Robert  R.  1974.  Silvicul- 
ture of  subalpine  forests  in  the 
central  and  southern  Rocky  Moun- 
tains: the  status  of  our  knowledge. 
USDA  Forest  Service  Research 
Paper  RM-121, 88  p.  Rocky  Moun- 
tain Forest  and  Range  Experiment 
Station,  Fort  Collins,  CO. 

Armstrong,  David  M.  1977.  Ecologi- 
cal distribution  of  small  mammals 
in  the  Upper  Williams  Fork  Basin, 
Grand  County,  Colorado.  The 
Southwestern  Naturalist  22:289- 
304. 

Franklin,  Jerry  F.,  Kermit  Cromack, 
Jr.,  William  Denison,  Arthur 
McKee,  Chris  Maser,  James  Sedall, 
Fred  Swanson,  and  Glen  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-fir  forests. 
USDA  Forest  Service  General 
Technical  Report  PNW-118, 48  p. 
Pacific  Northwest  Forest  and 
Range  Experiment  Station,  Port- 
land, OR. 

Getz,  Lowell  L.  1968.  Influence  of 
water  balance  and  microclimate 
on  the  local  distribution  of  the  red- 
back  vole  and  white-footed 
mouse.  Ecology  49:276-286. 

Graul,  Walter  D.,  and  Gary  C.  Miller. 
1984.  Strengthening  ecosystem 
management  approaches.  Wildlife 
Society  Bulletin  12:282-289. 


Gunther,  Pamela  M.,  Brenda  S.  Horn, 
and  Geoffrey  D.  Babb.  1983.  Small 
mammal  populations  and  food 
selection  in  relation  to  timber  har- 
vest practices  in  the  western  Cas- 
cade Mountains.  Northwest  Sci- 
ence 57:32-44. 

Harris,  Larry  D.  1984.  The  frag- 
mented forest:  island  biogeogra- 
phy  theory  and  the  preservation  of 
biotic  diversity.  211  p.  University 
of  Chicago  Press,  Chicago,  IL. 

Jerry,  Danielle  G.  1984.  Old-growth 
management  in  the  Idaho  Pan- 
handle National  Forests,  p.  373- 
379.  In  Fish  and  wildlife  relation- 
ships in  old-growth  forests:  Pro- 
ceedings of  the  symposium.  Wil- 
liam R.  Meehan,  Theodore  R.  Mer- 
rell,  Jr.,  and  Thomas  A.  Hanley, 
editors.  [Juneau,  AK,  April  12-15, 
1982]  425  p.  American  Institute  of 
Fishery  Research  Biologists. 

Juday,  Glenn  Patrick.  1978.  Old- 
growth  forests:  a  necessary  ele- 
ment of  multiple  use  and  sus- 
tained yield  national  forest  man- 
agement. Environ.  Law  8:497-522. 

Kirkland,  Gordon  L.,  Jr.  1977.  Re- 
sponses of  small  mammals  to  the 
clearcutting  of  northern  Appala- 
chian forests.  Journal  of  Mam- 
malogy 58:600-609. 

Krebs,  Charles  J.,  and  Judith  H. 
Myers.  1974.  Population  cycles  in 
small  mammals.  Advances  in  Eco- 
logical Research  8:267-399. 

Marquardt,  Steve.  1984.  Old-growth 
scorecard  user's  guide.  USDA  For- 
est Service.  Unpublished  report. 

Martell,  Arthur  M.  1981.  Food  habits 
of  southern  red-backed  voles  (Cle- 
thrionomys  gapperi)  in  northern  On- 
tario. Canadian  Field-Naturalist 
95:325-328. 

Martell,  Arthur  M.  1984.  Changes  in 
small  mammal  communities  after 
fire  in  north-central  Ontario.  Ca- 
nadian Field-Naturalist  98:223-226. 

Martell,  Arthur  M.,  and  Andrew 
Radvanyi.  1977.  Changes  in  small 
mammal  populations  after 
clearcutting  of  northern  Ontario 
black  spruce  forest.  Canadian 
Field-Naturalist  91:41-46. 


Maser,  Chris,  Ralph  G.  Anderson, 
Kermit  Cromack,  Jr.,  Jerry  T.  Wil- 
liams, and  Robert  E.  Martin.  1979. 
Dead  and  down  woody  material, 
p.  78-95.  In  Wildlife  habitats  in 
managed  forests:  the  Blue  Moun- 
tains of  Oregon  and  Washington. 
Jack  Ward  Thomas,  editor.  Agri- 
cultural Handbook  553, 512  p. 
USDA  Forest  Service. 

Maser,  Chris  and  James  M.  Trappe. 
1984.  The  fallen  tree — a  source  of 
diversity,  p.  335-339.  In  New  for- 
ests for  a  changing  world:  Pro- 
ceedings of  the  Society  of  Ameri- 
can Foresters  National  Confer- 
ence. [Portland,  OR,  October  16- 
20, 1983]. 

Maser,  Chris,  James  M.  Trappe,  and 
Ronald  A.  Nussbaum.  1978a.  Fun- 
gal-small mammal  interrelation- 
ships with  emphasis  on  Oregon 
coniferous  forests.  Ecology  59:799- 
809. 

Maser,  Chris,  James  M.  Trappe,  and 
Douglas  C.  Ure.  1978b.  Implica- 
tions of  small  mammal  mycoph- 
agy  to  the  management  of  western 
coniferous  forests.  Transactions  of 
the  North  American  Wildlife  and 
Natural  Resources  Conference 
43:78-88. 

Merritt,  Joseph  F.  1981.  Qethriono- 
mys  gapperi.  Mammalian  Species 
146:1-9.  The  American  Society  of 
Mammalogists. 

Nie,  Norman  H.,  C.  Hadlai  Hull,  Jean 
G.  Jenkens,  Karin  Steinbrenner, 
and  Dale  H.  Bent.  1975.  SPSS:  sta- 
tistical package  for  the  social  sci- 
ences, 2nd  edition.  675  p. 
McGraw-Hill,  Inc. 

Patton,  David  R.  1987.  Is  the  use  of 
"management  indicator  species" 
feasible?  Western  Journal  of  Ap- 
plied Forestry  2:33-34. 

Romme,  William  H.,  and  Dennis  H. 
Knight.  1981.  Fire  frequency  and 
subalpine  forest  succession  along  a 
topographic  gradient  in  Wyoming. 
Ecology  62:319-326. 

Salwasser,  Hal,  Ira  D.  Luman,  and 
Don  Duff.  1982.  Integrating  wild- 
life and  fish  into  public  land  forest 
management.  Western  Association 


357 


of  Fish  and  Wildlife  Agencies:  Pro- 
ceedings 62:293-299. 

Tevis,  Lloyd,  Jr.  1956.  Responses  of 
small  mammal  populations  to  log- 
ging of  Douglas-fir.  Journal  of 
Mammalogy  37:189-196. 

USD  A  Forest  Service.  1985.  Land  and 
resource  management  plan  for  the 
Medicine  Bow  National  Forest  and 
Thunder  Basin  National  Grass- 
land. USD  A  Forest  Service,  Rocky 
Mountain  Region.  Laramie,  Wyo- 
ming. 

Vaughan,  Terry  A.  1969.  Reproduc- 
tion and  population  densities  in  a 
montane  small  mammal  fauna,  p. 
51-74.  In  Contributions  in  Mam- 
malogy. J.  Knox  Jones,  Jr.,  editor. 
Miscellaneous  Publications  of  the 
Museum  of  Natural  History  51:1- 
428.  University  of  Kansas. 

Vaughan,  Terry  A.  1974.  Resource 
allocation  in  some  sympatric,  sub- 
alpine  rodents.  Journal  of  Mam- 
malogy 55:764-795. 

Vemer,  Jared.  1983.  An  integrated 
system  for  monitoring  wildlife  on 
the  Sierra  National  Forest.  Trans- 
actions of  the  North  American 
Wildlife  and  Natural  Resources 
Conference  48:355-366. 


Habitat  Associations  of  Small 
Mammals  in  a  Subalpine 
Forest,  Souttieastern 
Wyoming^ 

Martin  G.  Raphael 


Abstract.— Mammal  capture  rates  were  greatest 
at  sites  with  mature  timber  and  other  old-growth 
attributes.  Shrews  (both  dusky  (Sorex  monticolus) 
and  masked  (S.  cinereus))  and  southern  red-backed 
voles  (Clefhrionomys  gapperi)  were  much  more 
abundant  at  sites  dominated  by  spruce  or  fir 
compared  to  drier  sites  dominated  by  lodgepole 
pine.  Deer  mice  (Peromyscus  maniculofus).  in 
contrast,  were  most  abundant  on  drier,  pine- 
dominated  sites.  The  southern  red-backed  vole, 
because  of  its  high  abundance  and  strong 
association  with  mature  forest,  is  a  good  ecological 
indicator  of  late  serai  conditions  for  forest  planning 
purposes. 


Figure  1.— Map  of  study  area  showing  iocation  of  study  area  and  distribution  of  trapping 
stations. 


Subalpine  forests  of  spruce,  fir,  and 
lodgepole  pine  cover  about  5  million 
ha,  or  38%  of  forested  land  in  the 
central  Rocky  Mountain  region — 
more  than  any  other  forest  type  (Al- 
exander 1974,  USD  A  Forest  Service 
1980).  Subalpine  forest  is  harvested 
heavily,  accounting  for  over  90%  of 
total  sawtimber  volume  in  this  region 
(USDA  Forest  Service  1980).  These 
forests  also  are  managed  to  produce 
water,  and  timber  harvest  practices 
have  been  developed  that  can  sub- 
stantially increase  water  yield  (Tro- 
endle  1983,  Swanson  1987).  The  Coon 
Creek  Water  Yield  Augmentation 
Pilot  Project  (Bevenger  and  Troendle 
1984, 1987)  is  a  large-scale  demon- 
stration of  the  feasibility  and  costs/ 
benefits  of  increasing  water  yield 
through  specially  designed  clearcuts. 
To  evaluate  the  response  of  wildlife 
species  to  such  harvests,  studies 
were  initiated  to  describe  the  pre- 
treatment  structure  and  compxDsition 
of  the  vertebrate  community  (Ra- 
phael 1987b)  and,  ultimately,  to  com- 
pare responses  of  vertebrates  on  the 
treated  watershed  and  on  the  unhar- 
vested  control. 

The  present  study  summarizes  the 
structure  of  the  small  mammal  com- 

' Paper  presented  at  Symposium,  Man- 
agement of  AmphibioDs.  Reptiles,  and 
Small  Mammals  in  Nortti  America  (Ragstaff, 
AZ.July  19-21.  1988). 

'Research!  Ecdogist,  USDA  Forest  Serv- 
ice. Rocky  Mountain  Forest  and  Range  Ex- 
periment Station.  Forestry  Sciences  Labora- 
tory. 222  Southi  22nd  Street.  Laramie.  Wyo- 
ming 82070. 


munity,  describes  habitat  associa- 
tions of  the  dominant  species  during 
the  pretreatment  phase  of  the  longer 
term  project,  evaluates  the  efficacy  of 
an  old-growth  scorecard  to  rate  old- 
growth  characteristics  of  stands,  and 
assesses  designation  of  mammals  as 
ecological  indicators  of  old-growth 
conditions. 


STUDY  AREA 

Studies  were  conducted  within  two 
watersheds,  the  Upper  East  Fork  of 


the  Encampment  River  (911  ha)  and 
Coon  Creek  (1,615  ha).  These  adja- 
cent watersheds  are  part  of  the  Sierra 
Mad  re  range  of  southern  Wyoming, 
located  about  25  km  south  of  the 
town  of  Encampment  (fig.  1).  Eleva- 
tions vary  from  2,600  to  3,300  m. 
Soils  are  50-150  cm  deep  and  are  well 
drained. 

Mean  annual  precipitation  is  about 
100  cm,  70%  falling  as  snow  that  usu- 
ally covers  the  site  from  late  Septem- 
ber through  late  June  at  depths  of  2-4 
m  in  winter.  Forest  cover  is  domi- 
nated by  lodgepole  pine  (~  60%  of 


359 


land  area),  and  a  nuxture  of  Engel- 
mann  spruce  and  subalpine  fir.  Pole 
stands  with  trees  <23  cm  d.b.h.  occur 
on  24%  of  the  two  watersheds,  ma- 
ture stands  occur  on  72%,  and  mead- 
ows or  rock  outcrops  cover  4%. 

METHODS 

Vegetation  Sampling 

In  each  watershed,  90  sampling  sta- 
tions were  established  at  200-m  inter- 
vals along  N-S  lines  that  were  400  m 
apart  (fig.  1).  At  each  of  the  180  sta- 
tions, an  observer  measured  basal 
area  of  each  tree  species  using  a  1- 
factor  metric  reloskop.  Canopy  cover 
was  estimated  from  the  average  of 
four  readings  taken  at  cardinal  direc- 
tions with  a  spherical  densiometer. 
Slope  was  measured  with  a  clinome- 
ter and  aspect  was  measured  with  a 
hand-held  compass.  All  snags  >20 
cm  d.b.h.  and  1.8  m  tall  were 
counted  within  a  0.04-ha  circular  plot 
centered  at  the  station;  cover  per- 
centages of  shrubs,  forbs,  grasses, 
rocks,  litter,  and  bare  ground  were 
visually  estimated  over  the  same 
0.04-ha  plot.  Hard  (class  1,2)  and  soft 
[class  3,4,5  (Maser  et  al.  1979)]  logs 
also  were  counted  on  each  plot. 
Height  and  d.b.h.  of  one  representa- 
tive tree  were  measured  at  each  sta- 
tion with  a  clinometer  and  metric 
d.b.h.  tape. 

All  stands  on  each  watershed 
were  assessed  by  personnel  of  the 
Medicine  Bow  National  Forest  and 
assigned  an  old-growth  rating  based 
on  canopy  structure,  d.b.h.,  tree 
height,  snag  size  and  density,  and  log 
size  and  density  (app)endix).  Possible 
scorecard  values  range  from  0  (no 
old-growth  characteristics)  to  60 
(maximum). 

Stand  maps  were  used  to  associate 
a  sampling  station  with  the  old- 
growth  scorecard  value  for  the  stand 
in  which  the  station  was  located. 
Habitat  types  were  also  assigned  to 
each  station  based  on  classifications 
used  by  Medicine  Bow  National  For- 


est personnel.  Also  recorded  was  the 
presence  or  absence  of  permanent 
streams  within  100  m  of  each  sam- 
pling station. 

Red  Squirrel  Counts 

Three  observers  visited  each  sample 
station  twice  each  year  (totaling  six 
visits/station/yr)  from  13  June  to  25 
July  1985, 18  June  to  23  July  1986, 
and  15  June  to  17  July  1987.  At  each 
visit,  the  observer  recorded  all  red 
squirrels  seen  or  heard  within  a  100- 
m  radius  of  the  station  center.  All 
counts  were  begun  within  30  minutes 
after  sunrise;  each  observer  visited  15 


stations  per  day  and  most  counts 
were  completed  before  noon. 

SnKili  Mammal  Trapping 

To  sample  shrews,  six  pitfall  traps 
were  installed  in  a  2  x  3  grid  (15-m 
spacing)  centered  on  each  station. 
Each  pitfall  trap  was  a  3-gal  plastic 
bucket  buried  flush  with  the  ground 
surface  and  covered  by  logs  or  bark. 
To  capture  other  small  mammals, 
two  50-cm  Sherman  livetraps  were 
placed  within  2  m  of  each  pitfall  sta- 
tion. 

Mammals  were  trapped  during 
late  summer  from  1985  to  1987  (20 


Table  1.— Vegetation  and  stand  attributes  on  small  mammal  trapping  sta- 
tions, estimated  or  measured  on  0.04-ha  circular  plots,  among  tidbltat 
types'  on  a  Sierra  Madre  forest,  Wyoming. 


Characteristic 


Lodgepole  pine  Spruce/fir 

Unclossified  Pole     Mature   Mature  Signifi- 
(n=9)     <n=36)     (n=76)    (n=:59>  cance^ 


Basal  area  (mVha) 

Lodgepole  pine 

Engelmonn  spruce 

Subalpine  fir 
Tree  height  (m) 
D.B.H.  (cm) 
Snags/0.04  ha 
Percent  cover 

Shrubs  and  trees  >2  m  tall 

Forbs 

Grasses 

Rocks  >1 5  cm 

Utter 

Bare  soil 

Hard  logs  >20  cm  diameter 
Soft  logs  >20  cm  diameter 
Overstory  canopy 

Old-growth  scorecard  index 

Stream  presence^ 

Solar  radiation  index^ 

Elevation  (10^  m) 


12.P 

19.2^ 

21.3^ 

10.68 

0.01^ 

5.2^ 

6.3^ 

11.9^ 

0.01^ 

10.7 

8.6 

7.4 

10.2 

0.18 

-19  9AC 

18.2^ 

20.3^ 

21.2^ 

0.00 

29.8AB 

27.0^ 

32.9^ 

36.6^ 

0.00 

2.1 

1.2 

2.0 

2.4 

0.29^ 

45.3^c 

54.r 

50.8^ 

38.4«^ 

0.01 

14.0 

6.4 

7.6 

15.0 

0.12^ 

14.8 

7.4 

7.6 

16.4 

0.18^ 

0.8^ 

4.4^^ 

3.9^ 

3.2^^ 

0.00^ 

82.0^ 

85.6^<^ 

82.7^ 

73.6« 

0.07^ 

0.4 

1.0 

1.2 

1.8 

0.30^ 

0.6^ 

1.9^B 

2.3^ 

2.7^ 

0.00^ 

9.6 

10.7 

11.7 

10.9 

0.63^ 

65.8 

69.2 

68.2 

62.4 

0.32^ 

190ABC 

29.4^ 

34.9^ 

41. 

0.00 

22.0 

25.0 

38.2 

37.3 

0.45 

0.45^ 

0,50^ 

0.50^ 

0.48^ 

0.02 

9.0^ 

9.4^ 

9.5» 

9.48 

0.00^ 

'Letter  superscripts  denote  results  ofmuttiple  compartsais  (Tukey-Kramer  or  Dun- 
nett's  simultaneous  procedures);  means  witt)  same  letter  did  not  differ.  Experiment- 
wise  error  rate  maintairted  ata  =  0,05. 

'Significance  of  analysis  of  variance  P-tests  among  habitats:  W  indicates  fhiat 
Welch's  test  was  performed  when  variances  were  unequal. 

'Percent  of  stations  within  100  m  of  stream. 

"Index  of  yearly  solar  radiation  input  (Frank  and  Lee  1966). 


360 


August  through  26  September  1985, 5 
August  to  11  September  1986,  and  4 
August  to  10  September  1987).  Ob- 
servers checked  traps  once  daily  dur- 
ing each  of  three,  10-day  sampling 
sessions  each  year.  Sampling  sessions 
were  separated  by  four  days,  encom- 
passing six  weeks  each  year.  All  cap- 
tured specimens  were  identified,  toe 
clipped,  sex  determined,  aged, 
weighed,  and  checked  for  reproduc- 
tive status  (currently  breeding  or 
not). 

Dead  animals  were  assigned  a 
permanent  catalog  number.  Shrews 
were  preserved  in  70%  ethanol  and 
all  other  species  were  frozen  for  later 
identification. 


Data  Analysis 

Total  rumbers  of  detections  (red 
squirrels)  or  first  captures  (all  other 
species)  were  calculated  at  each  sta- 
tion over  the  3  years.  Thus,  the  total 
numbers  of  captures  represented  the 
results  of  450  trapnights  of  effort  at 
each  of  the  180  stations  (81,000  total 
trapnights).  Despite  efforts  to  close 


pitfall  traps  between  sessions,  some 
mammals  were  captured  before  the 
start  of  each  10-day  session.  These 
specimens  were  retained,  but  num- 
bers were  not  included  in  analyses. 

To  assess  habitat  associations  of 
the  more  abundant  mammals,  I  per- 
formed a  principal  components 
analyses  (with  varimax  rotation)  us- 
ing the  SPSS/PC+  program  package 
(Norusis  1988).  Principal  components 
analysis  derives  linear  combinations 
of  attributes  (in  this  case  vegetation 
characteristics  as  listed  in  table  1). 
All  components  with  eigenvalues 
>1.0  were  retained  for  subsequent 
analyses.  The  equations  were  then 
''solved"  for  each  station,  resulting  in 
a  set  of  scores  that  were  interpreted 
as  habitat  gradients.  I  identified  these 
gradients  from  those  original  habitat 
variables  most  highly  correlated  with 
the  principal  components  scores.  To 
relate  abundance  of  the  more  abun- 
dant mammals  to  habitat  features  at 
each  station,  I  performed  multiple  re- 
gressions of  capture  rates  at  each  sta- 
tion (dependent  variable)  with  the 
habitat  gradients  or  principal  compo- 
nents scores  (independent  variables). 


r 


Table  2.— Habitat  gradients  derived  from  principal  components  analysis  of 
19  variables  (table  1)  describing  vegetation  structure  and  composition  at 
each  small  mammal  sampling  station,  Sierra  Madre,  Wyoming. 


Percent  of  Cumulative 
Gradient  vcffiance^  percent 


Interpretation  of  tiabltat  gradienF 


1 

26.2 

26.2 

2 

16.1 

42.3 

3 

8.7 

51.0 

4 

7.4 

58.4 

5 

6.6 

65.1 

6 

5.6 

70.7 

Greater  cover  of  shrubs  and  litter;  greater 
basal  area  of  lodgepole  pine;  lower  cover 
of  herbs,  grasses, 

Greater  expression  of  old-growth  attrib- 
utes; greater  basal  area  of  Engelmann 
spruce. 

Upland  sites  with  greater  cover  of  soft  logs; 
greater  basal  area  of  subalpine  fir. 
Lower  cover  of  bare  ground;  greater  can- 
opy cover. 

Greater  cover  of  rocks. 
Higher  elevation  sites  with  greater  solar  ra- 
diation (southerly  slopes). 


'Amount  of  total  variance  (among  all  anginal  variabies)  accounted  for  by  eact) 
principal  component. 

'Interpretation  based  on  magnitude  of  correlations  of  original  variables  with  de- 
rived components.  Descriptions  indicate  positive  extreme  of  each  gradient. 


To  summarize  patterns  of  co-oc- 
currence of  the  more  common  mam- 
mal species,  I  performed  an  average- 
linkage-between-groups  cluster 
analysis  [UPGMA  (Norusis  1988)] 
based  upon  Pearson  correlations  be- 
tween abundances  of  all  pairs  of  spe- 
cies among  the  180  stations.  Results 
of  the  cluster  analysis  were  displayed 
using  a  dendrogram  showing  the 
relative  similarities  of  all  species.  The 
similarity  measure,  for  this  display, 
was  rescaled  to  values  ranging  from 
0  (no  similarity)  to  25  (maximum 
similarity). 


RESULTS  AND  DISCUSSION 

Vegetation 

Structure  and  composition  of  vegeta- 
tion (table  1)  were  typical  of  sub- 
alpine forest  in  the  central  Rocky 
Mountains  (Alexander  1974;  Raphael 
1987a,  1987b).  VegetaUon  characteris- 
tics have  been  shown  to  be  similar 
between  the  two  watersheds  (Ra- 
phael 1987b);  therefore,  no  distinc- 
tion was  made  between  the  two  wa- 
tersheds for  this  study. 

Principal  components  analysis  re- 
sulted in  the  creation  of  six  synthetic 
habitat  gradients  that,  together,  con- 
tained 68%  of  the  total  variance  from 
the  19  original  habitat  variables  (table 
2).  I  used  the  variables  that  were 
most  highly  correlated  with  values  of 
each  gradient  to  interpret  the  biologi- 
cal meanings  of  the  gradients  (table 
2). 


MannnKils 

Over  the  3  years  of  study  and  over 
all  sampling  stations,  observers  cap- 
tured 4,553  individuals  of  17  small 
mammal  species  and  recorded  987 
detections  of  red  squirrels  (table  3). 
The  most  abundant  species  was  the 
southern  red-backed  vole,  account- 
ing for  over  50%  of  all  captures. 
Other  dominant  species  included 
masked  shrew  (15%),  deer  mouse 


361 


(15%),  red  squirrel,  dusky  shrew 
(6%),  and  chipmunks  (2  species,  6%). 

Specific  Habitat  Associations 
Masked  Shrew 

Masked  shrews  were  more  abundant 
than  other  shrews  and  were  captured 
more  frequently  in  mature  lodgepole 
and  spruce/fir  sites  (table  3)  with 
higher  cover  of  herbs  and  grasses; 
they  were  less  abundant  on  dry, 
south-facing  sites  (table  4).  Their 
abundance  at  each  station  was  mod- 
eled (R^  =  0.42)  by  a  regression  that 
included  gradients  2,1,6,  and  4  (in 
order  of  their  statistical  significance) 
(table  4).  Other  studies  (Negus  and 
Find  ley  1959,  Spencer  and  Pettus 
1966,  Brown  1967a,  Armstrong  1977) 
also  report  this  species'  preference 
for  moist  sites.  However,  I  did  not 
find  a  strong  association  with  bogs, 
as  reported  by  Brown  (1967a)  and 
Spencer  and  Pettus  (1966). 

Dusky  Shrew 

Dusky  shrews  were  captured  in 
greater  numbers  in  more  moist,  ma- 
ture spruce/fir  sites  (table  3).  They 
were  most  strongly  associated  with 
dense  herbaceous  cover  and  (to  a 
lesser  degree)  with  old-growth  attrib- 
utes. Unlike  the  masked  shrew,  their 
abundance  was  positively  and  sig- 
nificantly correlated  with  gradient  3 
(moist,  streamside  sites;  tables  2,4). 
Like  masked  shrews,  they  were  less 
abundant  on  southerly,  steeper  sites. 
The  regression  model  explained  41% 
of  variance  in  abundance  (table  4). 
Brown  (1967a)  captured  this  shrew  in 
a  greater  variety  of  habitats  and  in 
drier  sites  than  the  masked  shrew. 
Negus  and  Findley  (1959)  also  re- 
jx)rted  use  of  a  greater  variety  of 
habitats;  Spencer  and  Pettus  (1966) 
found  dusky  shrews  in  association 
with  marshy  habitats. 

The  association  of  this  shrew  with 
old-growth  conditions  has  not,  to  my 


knowledge,  been  previously  re- 
ported. 

Least  Chipmunk 

The  abundance  of  least  chipmunk 
was  significantly  and  negatively  cor- 
related with  gradient  4  (bareground) 
and  positively  correlated  with  gradi- 
ent 6  (southerly  exposure).  Although 


the  regression  was  statistically  sig- 
nificant, it  explained  only  5%  of  vari- 
ance in  abundance  (table  4);  thus,  the 
regression  model  was  not  statistically 
meaningful. 

Nonetheless,  the  associations  sug- 
gested by  the  model,  particularly  the 
preference  for  open,  drier  slopes,  are 
in  accordance  with  results  of  other 
studies  (e.g.,  Telleen  1978,  Clark  and 
Stromberg  1987). 


r 


Table  S.—Smali  mammal  capture  rates  amor>g  generalized  habitat  types^ 
in  the  Sierra  Madre,  Wyoming,  1 985-1 987. 


Species 


Toh3l  no.       Lodgepole  pine  Spruce/fir 
individuals     Pole        Mature  Mahjre 
captured    (r>=36)        (f»=76)  (n=59) 


Masked  shrew  700 
(Sorex  cinereus) 

Dusky  shrew  253 
(S.  monficolus) 

Dwarf  ^rew  2 
(S.  nanus) 

Water  shrew  7 
($.  palustris) 

Pygmy  shrew  11 
(S.hoyO 

Least  chipmunk  101 
(Tamias  minimus) 

Uinta  chipmunk  150 
(Tamias  umbrinus) 

Golden-mantled  ground  sq.  11 

(Spermophiius  lateralis) 

Red  squirrel  3987 

(Tamiasciurus  hudsonicus) 

Northern  pocket  gopher  1 

(Thomomys  falpoides) 

Deer  mouse  696 

(Peromyscus  maniculafus) 

Southern  red-backed  vole  2^75 

(Clefhrionomys  gapperi) 

Heather  vole  17 

(Phenacomys  intermedius) 

Montane  vole  32 

(Microfus  montonus) 

Long-tailed  vole  1 1 

(M.  longicandus) 

House  mouse  :  1 

(Mus  musculus) 

Western  jumping  mouse  80 
(Zapus  princeps) 

Ermine  6 
(Musfela  erminea) 


2.6^ 
0.6^ 
0 
0 

0.03 
0.8* 
1.1* 
0,06 
5.8* 
0 

3.3* 
10.8* 
0.08 
0.03* 
0,03 
0 

0.1* 
0 


3.7*8 

0.8* 

0 

0.08 
0.09 
0.4* 
0.9* 
0.11 
4.7* 
0.01 
4.2* 
11.4^ 
0.09 
0.09* 
0.08 
0 

0,4» 
0.06 


5.08 
2.78 
0,03 
0.02 
0,03 
0,7* 
0,6* 
0.02 
6.0* 
0 

3.5* 
18.08 
0,10 
0.37* 
0.07 
0.02 
0.78 
0.02 


0.01 
0,01 
NT 
NT 
NT 
0,50 
0,51 
NT 
0.23 
NT 
0,31 
0,02 
NT 
0,08 
NT 
NT 
0,00 
NT 


'Values  are  mean  capture  rates  (captures/450  trapnigtits)  or  mean  numbers  of 
detections  (red  squirreO  among  tiabitat  types  for  all  years  combined.  Letter  super- 
scripts indicate  results  of  multiple  comparisons;  means  witti  same  letter  did  not  differ 
significantly. 

^Significance  from  one-way  analy^s  of  varlartce:  NT  =  not  tested  because  of  small 
sample  size. 

'Results  are  expressed  as  numbers  of  detections  during  call  counts. 


362 


Uinta  Chipmunk 

Uinta  chipmunks  were  most  abun- 
dant on  rocky  slopes  (gradient  5),  as 
also  reported  by  Clark  and 
Stromberg  (1987).  They  were  rela- 
tively more  abundant  in  younger 
stands  (gradient  2).  The  regression 
model  explained  17%  of  the  variation 
in  abundance  of  this  species  (table  4). 
Compared  with  the  least  chipmunk, 
this  species  is  reported  to  be  more 
restricted  to  subalpine  forest  habitats 
(Negus  and  Findley  1959).  Telleen 
(1978)  found  an  association  with 
closed  canopy,  open  understory 
habitats. 


Red  Squirrel 

Red  squirrel  abundance  was  some- 
what greater  on  dry,  gently  sloping 
sites  (gradients  3, 5),  but  only  16%  of 
variation  in  abundance  was  ex- 
plained by  the  regression  model. 
These  squirrels  were  abundant 
throughout  the  study  area,  which 
seemed  to  be  comprised  of  excellent 
red  squirrel  habitat.  Therefore,  vari- 
ation in  vegetation  among  sites  was 
probably  minor  in  relation  to  the  po- 


tential variation  that  would  distin- 
guish suitable  from  unsuitable  habi- 
tat. Clark  and  Stromberg  (1987)  de- 
scribe red  squirrels  as  widespread 
throughout  coniferous  forest  habitats 
of  Wyoming. 


i:>eer  Mouse 

Deer  mice  were  associated  with 
streamside  sites  having  lower  basal 
area  of  subalpine  fir  (gradient  3).  Al- 
though widespread  on  the  study 
area,  they  tended  to  be  more  abun- 
dant on  open,  lodgepole-dominated 
sites  and  meadows  than  on  spruce/ 
fir  sites.  The  regression  model  ex- 
plained 15%  of  the  variance  in  deer 
mouse  abundance  (table  4).  Contrary, 
to  these  results,  other  studies  (Brown 
1967b,  Campbell  and  Clark  1980, 
Ramirez  and  Homocker  1981)  re- 
ported associations  of  deer  mice  with 
xeric  sites  away  from  streams.  The 
species  is  known  to  be  abundant  on 
cutover  sites  (Ramirez  and  Hor- 
nocker  1981,  Scrivner  and  Smith 
1984),  tolerant  of  a  wide  range  of  eco- 
logical conditions  (Clark  and 
Stromberg  1987),  and  omnivorous 
(Clark  1975). 


Table  4.— Results  of  stepwise  multiple  regressloru  of  small  mammal  abun- 
dance with  habitat  gradients  (principcd  components  from  table  2)  Sierra 
Mddre,  Wyoming. 


Species 


Habitat  gradient' 
3        4  5 


Explained 
variance^ 


Masked  shrew 

(2) 

1 

(3) 

0.42 

Dusky  shrew 

(1) 

2 

(4) 

(3) 

0.41 

Least  chipmunk 

(1) 

2 

0.05 

Uinta  chipmunk 

(2) 

1 

0.17 

Red  squirrel 

(4) 

1 

(3) 

(2) 

0.16 

Deer  mouse 

(1) 

0,15 

S.  red-backed  vole 

(2) 

1 

4 

5 

(6) 

(3) 

0.46 

W.  jumping  mouse 

(1) 

2 

(3) 

0.20 

'Numbers  below  each  gradient  indicate  /he  order  of  entry  of  that  gradient  into  the 
stepwise  regression  (using  F-to-enter  significance  ofP<  0.05).  Parentheses  indicate 
negative  associations. 

^Adjusted    values  indicating  the  proportion  of  variance  in  capture  (or  detection) 
rates  accounted  for  by  gradients  included  in  regression  model.  All  regressions  were 
significant  atP  <  0.001. 


Southern  Red- Backed  Vole 

This  vole,  the  most  abundant  species 
on  the  study  area,  was  most  abun- 
dant in  mature  spruce/ fir  stands 
(table  3).  Its  abundance  was  also 
greater  in  stands  that  had  more  herb 
and  grass  cover  (gradient  1),  on 
northerly  slopes  (gradient  6),  and  on 
sites  with  greater  basal  area  of  sub- 
alpine fir  and  greater  log  cover  (gra- 
dient 3).  Its  abundance  was  modeled 
well  by  the  regression,  which  ac- 
counted for  46%  of  variation  in  red- 
backed  vole  abundance  among  sites 
(table  4). 

The  association  of  red-backed 
voles  and  mature  spruce/ fir  forest  is 
well  documented  (Ramirez  and  Hor- 
nocker  1981,  Allen  1983,  Scrivner  and 
Smith  1984).  This  association  may  be 
due,  at  least  in  part,  to  the  greater 
cover  of  logs  and  other  woody  debris 
that  provides  protection  during  criti- 
cal periods  of  freezing  and  thawing 
(Merritt  1976, 1985;  Merritt  and  Mer- 
ritt  1978,  Sleeper  1979)  and  supports 
fungi  used  as  food  (Williams  1955, 
Clark  and  Stromberg  1987,  Wywia- 
lowski  and  Smith  1988). 


Western  Jumping  Mouse 

Jumping  mice  were  most  abundant 
in  spruce/ fir  and  mature  lodgepole 
habitats  (table  3).  As  reported  in 
other  studies  (Negus  and  Findley 
1959,  Brown  1967b,  Clark  1971, 
Scrivner  and  Smith  1984),  these  mice 
were  associated  with  dense  herba- 
ceous or  grassy  vegetation  (gradient 
1)  along  moist  streamsides  (gradient 
3)  in  more  mature  stands  (gradient  2) 
(table  4).  The  regression  model  ac- 
counted for  20%  of  variation  in  abun- 
dance across  all  stations.  These  mice 
feed  primarily  on  grass  seeds  and 
fungi  (Jones  et  al.  1978,  Vaughan  and 
Weil  1980),  which  may  account  for 
their  close  association  with  grassy 
streamside  habitats. 


363 


General  Relationships 

Moisture  and  stand  maturity  were 
two  habitat  features  that  separated 
patterns  of  abundance  of  the  various 
species.  This  is  illustrated  most  effec- 
tively through  the  cluster  analysis 
based  on  interspecific  correlations  of 
relative  abundance  (fig.  2).  The  den- 
drogram shows  two  groups:  one 
comprised  of  the  two  shrews,  two 
voles,  and  jumping  mouse;  and  one 
comprised  of  the  red  squirrel,  two 
chipmunks,  and  deer  mouse.  The  for- 
mer group  is  associated  with  more 
moist,  old-growth  conditions  (table 
4).  The  latter  group  is  associated  with 
drier,  less  mature  conditions. 

The  association  of  species  with 
old-growth  conditions  is  of  special 
interest  because  of  concern  over 
identifying  species  that  are  ecological 
indicators  of  old-growth  (USDA  For- 
est Service  1985;  Nordyke  and 
Buskirk,  these  proceedings).  The 
Medicine  Bow  National  Forest,  the 
site  of  this  study,  lists  the  southern 
red-backed  vole  as  an  ecological  indi- 
cator representing  late  successional 
stages  in  conifer  forests.  Because  the 
forest  uses  the  old-growth  scorecard 
to  rate  old-growth  conditions, 
whether  or  not  red-backed  vole 
abundance  is  related  to  old-growth 
index  values  is  of  interest.  Raphael 
(1987b)  confirmed  such  a  trend  based 
on  analyses  of  the  first  2  years  of  the 
present  study. 

The  trend  is  even  more  pro- 
nounced when  all  3  years  are  in- 
cluded in  the  analyses  (fig.  3).  South- 
ern red-backed  voles  are  increasingly 
abundant  as  old-growth  scorecard 
index  values  increase.  Similar  trends 
are  evident  for  masked  and  dusky 
shrews  (fig.  3). 


CONCLUSIONS 

The  small  mammal  community,  as 
sampled  in  this  study,  was  similar  in 
composition  to  that  described  in 
other  studies  in  subalpine  forests  of 
the  Rocky  Mountain  region  (cf.  Ra- 


phael 1987a).  The  southern  red- 
backed  vole  was  the  most  abundant 
species  and  can  be  considered  the 
species  most  representative  of  ma- 
ture spruce/ fir  forest  stands.  Stand 
age  and  moisture  conditions  were 
the  two  most  important  generalized 
gradients  that  were  predictors  of 
summer  abundance  of  the  various 
species.  The  southern  red -backed 
vole  was  confirmed  as  a  suitable  eco- 
logical indicator  of  old-growth  forest; 
but,  two  other  species,  the  masked 
shrew  and  the  dusky  shrew,  are 
good  candidates  as  well. 


ACKNOWLEDGMENTS 

I  am  indebted  to  Christopher  Cana- 
day,  Anita  Kang,  Jeffery  Waters, 
Gary  Rosenberg,  Scott  Stoleson,  San- 
dra Spon,  Sandra  Pletschet,  Steven 
Larson,  Lindsay  Hall,  Thomas 
Batchelor,  Daniel  Maltese,  and  Lisa 
Smith  for  their  help  in  the  field.  I  also 


thank  personnel  of  the  Wyoming 
Game  and  Fish  Department,  the 
Medicine  Bow  National  Forest,  espe- 
cially the  Hayden  Ranger  District,  for 
their  cooperation;  Carron  Meaney  of 
the  Denver  Museum  of  Natural  His- 
tory and  Steven  Buskirk  of  the  Uni- 
versity of  Wyoming  for  technical  as- 
sistance and  use  of  museum  facilities; 
and  Gregory  D.  Hayward,  Graham 
W.  Smith,  Mark  R.  Stromberg,  and 
Richard  H.  Yahner  for  comments  on 
the  manuscript. 


LITERATURE  CITED 

Alexander,  Robert  R.  1974.  Silvicul- 
ture of  subalpine  forests  in  the 
central  and  southern  Rocky  Moun- 
tains: the  status  of  our  knowledge. 
Research  Paper  RM-12.  Fort 
Collins,  CO:  U.S.  Department  of 
Agriculture,  Forest  Service,  Rocky 
Mountain  Forest  and  Range  Ex- 
periment Station;  88  p. 


r 


SOMO 

MIMO 

SOCI 

CLGA 

ZAPR 

TAHU 

TAMI 

PEMA 

TAUM 


0 


5         10  15 

RELATIVE  SIMILARITY 


20 


25 


Figure  2.— Dendrogram  showing  relative  similarity  (Pearson  correlations)  of  abundances  of 
snrxill  nrKimmal  species  across  sampling  stations.  Species  are:  Sorex  monticolus  (SOMO), 
Microtus  montanus  (MIMO),  Sorex  cinereus  (SOCI),  Clethrionomys  gapperi  (CLGA),  Zapus 
princeps  (ZAPR),  Tamiasciurus  hudsonicus  (TAHU),  Tamias  minimus  (TAMI).  Peromyscus 
maniculatus  (PEMA),  and  Tamias  umbrinus  (TAUM). 


364 


Allen,  Arthur  W.  1983.  Habitat  suita- 
bility index  models:  Southern  red- 
backed  vole  (Western  United 
States).  U.S.  Dept.  Int.,  Fish  Wildl. 
Serv.  FWS/OBS-82/10.42. 14  p. 


30 


20 


10 


CLGA 


10 


C3 


-a; 


TAMi 


1 


0 


PEMA 


1 


1 


lOf 


0 


SOCI 


Anonymous.  1985.  Medicine  Bow 
National  Forest  old  growth  habitat 
scorecard.  Southwest  Habitater 
6:2-7. 

Armstrong,  David  M.  1977.  Ecologi- 


SOMO 


10 


TAUM 


2.0 


1.5 


1.0 


0.5 


0 


ZAPR 


1 


0    17     27    37    45     60        0    17     27    37  45 

OLD-GROWTH  SCORECARD  VALUE 


60 


10 


0.5 


0.4 


0.3 


0.2 


0.1 


0 


TAHU 


.It  T 


MIMO 


0    17     27    37     45  60 


if 


Figure  3.— Mean  abundance  of  selected  snrxall  mammal  species  In  relation  to  old-growtti 
scorecard  values.  Larger  scorecard  values  Indicate  greater  expression  of  old-growtfi  condi- 
tions. Vertical  lines  wittiin  bars  indicate  95%  CI  of  means.  See  figure  3  for  species  codes. 

365 


cal  distribution  of  small  mammals 
in  the  Upp)er  Williams  Fork  Basin, 
Grand  County,  Colorado.  The 
Southwestern  Naturalist  22:289- 
304. 

Bevenger,  Greg  S.;  Troendle,  Charles 
A.  1984.  The  Coon  Creek  water 
yield  augmentation  project,  p.  240- 
251.  In  Water  for  the  twenty-first 
century:  will  it  be  there?  proceed- 
ings of  the  symposium,  1984  April 
3-5,  Dallas,  TX.  Dallas,  TX,  South- 
em  Methodist  University. 

Bevenger,  Greg  S.;  Troendle,  Charles 
A.  1987.  The  Coon  Creek  water 
yield  augmentation  pilot  project, 
p.  145-149.  In  Water  for  the 
twenty-first  century:  will  it  be 
there?  proceedings  of  the  sympo- 
sium, 1984  April  3-5,  Dallas,  TX. 
Dallas,  TX,  Southern  Methodist 
University. 

Brown,  Larry  N.  1967a.  Ecological 
distribution  of  six  species  of 
shrews  and  comparison  of  sam- 
pling methods  in  the  central  Rocky 
Mountains.  Journal  of  Mam- 
malogy 48:617-623. 

Brown,  Larry  N.  1967b.  Ecological 
distribution  of  mice  in  the  Medi- 
cine Bow  Mountains  of  Wyoming. 
Ecology  48:677-680. 

Campbell,  Thomas  M.  Ill;  Clark,  Tim 
W.  1980.  Short-term  effects  of  log- 
ging on  red-backed  voles  and  deer 
mice.  Great  Basin  Naturalist 
40:183-189. 

Clark,  Tim  W.  1971.  Ecology  of  the 
western  jumping  mouse  in  Grand 
Teton  National  Park,  Wyoming. 
Northwest  Science  45:229-238. 

Clark,  Tim  W.  1975.  Ecological  notes 
on  deer  mice  in  Grand  Teton  Na- 
tional Park,  Wyoming.  Northwest 
Science  49:14-16. 

Clark,  Tim  W.;  Stromberg,  Mark  R. 
1987.  Mammals  in  Wyoming.  Uni- 
versity of  Kansas,  Museum  of 
Natural  History.  Public  Education 
Series  No.  10.  314  p. 

Frank,  Ernest  C;  Richard  Lee.  1966. 
Potential  solar  beam  irradiation  on 
slopes.  Research  Paper  RM-18. 
Fort  Collins,  CO:  U.S.  Department 
of  Agriculture,  Forest  Service, 


Rocky  Mountain  Forest  and  Range 
Experiment  Station.  166  p. 
Jones,  Gwilym  S.;  John  O.  Whitaker, 
Jr.;  Chris  Maser.  1978.  Food  habits 
of  jumping  mice  (Zapus  trinotatus 
and  Z.  princeps)  in  western  North 
America.  Northwest  Science  52:57- 
60. 

Maser,  Chris;  Ralph  G.  Anderson; 
Kermit  Cromack,  Jr.;  Jerry  T.  Wil- 
liams; Robert  E.  Martin.  1979. 
Dead  and  down  wood  material,  p. 
78-95.  In  Thomas,  J.  W.,  ed.  U.S. 
Department  of  Agriculture,  Forest 
Service  Agriculture  Handbook 
553.  Washington,  DC,  U.S.  Gov- 
ernment Printing  Office. 

Merritt,  Joseph  F.  1976.  Population 
ecology  and  energy  relationships 
of  small  mammals  of  a  Colorado 
subalpine  forest.  Ph.D.  disserta- 
tion, Univ.  Colorado,  Boulder.  146 
p.  From  Diss.  Abstr.  Int.  B  Sci. 
Eng.  37:3756B. 

Merritt,  Joseph  F.  1985.  Influence  of 
snowcover  on  survival  of  Clethri- 
onomys  gctpperi  inhabiting  the  Ap- 
palachian and  Rocky  Mountains  of 
North  America.  Acta  Zoological 
Fennica  173:73-74. 

Merritt,  Joseph  F.;  Joanie  M.  Merritt. 
1978.  Population  ecology  and  en- 
ergy relationships  of  Clethrionotnys 
gapperi  in  a  Colorado  subalpine 
forest.  Journal  of  Mammalogy 
59:576-598. 

Negus,  Norman  C;  James  S.  Findley. 
1959.  Mammals  of  Jackson  Hole, 
Wyoming.  Journal  of  Mammalogy 
40:371-381. 

Norusis,  Marija  J.  1988.  SPSS/PC+: 
Advanced  statistics.  SPSS,  Inc., 
Chicago,  IL. 

Ramirez,  Pedro  Jr.;  Maurice  Hor- 
nocker.  1981.  Small  mammal 
populations  in  different-aged 
clearcuts  in  northwestern  Mon- 
tana. Journal  of  Mammalogy 
62:400-403. 

Raphael,  Martin  G.  1987a.  Nongame 
wildlife  research  in  subalpine  for- 
ests of  the  central  Rocky  Moun- 
tains, p.  113-122.  In  Troendle,  C. 
A.,  M.  R.  Kaufmann,  R.  H.  Hamre, 
and  R.  P.  Winokur,  editors.  Pro- 


ceedings, Management  of  sub- 
alpine forests:  building  on  50  years 
of  research.  USDA  Forest  Service, 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  General  Tech- 
nical Report  RM-149. 

Raphael,  Martin  G.  1987b.  The  Coon 
Creek  wildlife  project:  effects  of 
water  yield  augmentation  on  wild- 
life, p.  173-179.  In  Troendle,  C.  A., 
M.  R,  Kaufmann,  R.  H.  Hamre, 
and  R.  P.  Winokur,  editors.  Pro- 
ceedings, Management  of  sub- 
alpine forests:  building  on  50  years 
of  research.  USDA  Forest  Service, 
Rocky  Mountain  Forest  and  Range 
Experiment  Station  General  Tech- 
nical Report  RM-149. 

Scrivner,  Jerry  H.;  H.  Duane  Smith. 
1984.  Relative  abundance  of  small 
manrmnals  in  four  successional 
stages  of  spruce-fir  forest  in  Idaho. 
Northwest  Science  58:171-176. 

Sleeper,  Roger  A.  1979.  Small  mam- 
mal population  fluctuations  in  re- 
lation to  snowpack.  Ph.D.  Disser- 
tation, Colorado  State  Univ.,  Fort 
Collins.  134  p.  From  Diss.  Abstr. 
Int.  B  Sci.  Eng.  40:3047. 

Spencer,  Albert  W.  and  David  Pettus. 
1966.  Habitat  preferences  of  five 
sympatric  species  of  long-tailed 
shrews.  Ecology  47:677-683. 

Swanson,  Robert  H.  1987.  Applying 
hydrologic  principles  to  the  man- 
agement of  subalpine  forests  for 
water  supply,  p.  79-85.  In  Tro- 
endle, C.  A.;  M.  R.  Kaufmann;  R. 
H.  Hamre;  R.  P.  Winokur,  editors. 
Proceedings,  Management  of  sub- 
alpine forests:  building  on  50  years 
of  research.  USDA  Forest  Service, 
Rocky  Mountain  Forest  and  Range 
Experiment  Station  General  Tech- 
nical Report  RM-149. 

Telleen,  Steven  L.  1978.  Structural 
niches  of  Eutamias  minimus  and  E. 
umbrinus  in  Rocky  Mountain  Na- 
tional Park.  Ph.D.  Dissertation, 
Univ.  Colorado,  Boulder.  152  p. 
From  Diss.  Abstr.  Int.  B  Sci.  Eng. 
39:2115. 

Troendle,  Charles  A.  1983.  The  po- 
tential for  water  yield  augmenta- 
tion from  forest  management  in 


the  Rocky  Mountains.  Water  Re- 
sources Bulletin  19:359-373. 

USDA  Forest  Service.  1980.  An  as- 
sessment of  the  forest  and  range 
situation  in  the  United  States.  FS- 
345,  Washington,  DC,  U.S.  Depart- 
ment of  Agriculture,  Forest  Serv- 
ice, 631  p. 

USDA  Forest  Service.  1985.  Final  En- 
vironmental Impact  Statement, 
Medicine  Bow  National  Forest, 
Laramie,  WY.  U.S.  Department  of 
Agriculture,  Forest  Service. 

Vaughan,  Terry  A.;  Wendy  Pearse 
Weil.  1980.  The  importance  of  ar- 
thropods in  the  diet  of  Zapus  prin- 
ceps in  a  subalpine  habitat.  Journal 
of  Mammalogy  61:122-124. 

Williams,  Olwen.  1955.  Distribution 
of  mice  and  shrews  in  a  Colorado 
montane  forest.  Journal  of  Mam- 
malogy 36:221-231. 

Wy  wialowski,  Alice  P.;  Graham  W. 
Smith.  1988.  Selection  of  mi- 
crohabitats  by  the  red-backed 
vole,  Clethrionomys  gctpperi.  Great 
Basin  Naturalist  48:216-223. 


366 


Appendix.— Old  growth  habitat  scorecard  (Anonymous  1 985)  used  to  rate  stands  in  the  Roclcy  Mountain  Region.  Point  val 
ues  from  1  to  5  are  assigned  to  each  category  A-L  Values  are  summed  over  all  rows  and  the  grand  total  is  used  as  the  in 
dex  value. 


Point  value 

5         4         3         2  1 

A.  Overstory 

3  or  more  species  3  or  more  species  2  species  2  species  1  species 

Spruce  and/or  Fir  >50%  Spruce  and/ or  Fir  <50%  Spruce  and/or  Fir  >50%  Spruce  and/or  Fir  <50%  100% 


B.  Midstory 

3  or  more  species 

Spruce  ond/or  Fir  >50% 


3  or  more  species 
Spruce  ond/or  Fir  <50% 


2  species 

Spruce  and/or  Fir  >50% 


C.  Understory 
3  or  more  species 
Spruce  ond/or  Fir  >50% 


3  or  more  species 
Spruce  ond/or  Fir  <50% 


2  species 

Spruce  ond/or  Fir  >50% 


D.  Total  Canopy  Cover 
70%+   


70-50% 


E.  Overstory.  Canopy  Cover 

50-30%    70-50%  or  30-10% 

F.  Midstory  Canopy  Cover 

40-20%    70-40%  or  20-10% 


G.  Overstory  Ave.  DBH  (Live) 

16'+   

H.  Midstory  Ave.  DBH  (Live) 

9'+   


15--13- 


8"-6" 


50-30% 
100-70%  or  10-1% 
100-70%  or  10-1% 

12--10" 
5"-3- 


I.  Standing  Snogs  Ave.  DBH  (Record  only  those  snags  above  6"  in  height.) 

16"+    15"-13"    12"-10" 


J.  Standing  Snogs  #/Acre  (Record  only  those  snags  above  6'  in  height  and  7"  DBH.) 
6+    6-4    3-1 


K.  Dead.  Down  Logs  Ave.  DBH 
16"+   


15"- 13" 


L.  Dead.  Down  Logs  #/Acre  (Record  only  those  above  7"  DBH.) 
12+    12-6   

Column 

Totals     


12--10- 
6-2 


2  species 

Spruce  and/or  Fir  <5CD% 


2  species 

Spruce  and/or  Fir  <50% 


1  species 
100% 


1  species 
100% 


30-10% 


9'-7" 


<10% 


9'-7"   <7" 

<3-  

9"-7"  


367 


Differences  in  tlie  Ability  of 
Vegetation  IVIodels  to  Predict 
Small  Mammal  Abundance 
in  Different  Aged  Douglas- Fir 
Forests^ 

Cathy  A.  Taylor,^  C.John  Ralph,^  and  Ariene 
T.  Doyle^ 


Abstract.— Three  trapping  techniques  for  small 
mammals  were  used  in  47  study  stands  in  northern 
California  and  southern  Oregon  and  resulted  in 
different  capture  frequencies  by  the  different 
techniques.  In  addition,  the  abundances  of 
mammals  derived  from  the  different  techniques 
produced  vegetation  association  models  which 
were  often  quite  different.  Only  the  California  red- 
backed  vole  (Clefhrionomys  californicus)  showed 
any  association  with  stand  age,  and  no  species  had 
any  marked  associations  with  the  moisture  regime  of 
the  stands  or  the  geographical  region. 


Habitat  association  patterns  have 
been  presented  for  many  small  mam- 
mal species  (e.g.  Rosenzweig  1973, 
M'Closkey  1975,  Dueser  and  Shugart 
1978,  MacGracken,  et  al.  1985).  In 
most  instances,  models  representing 
habitat  use  have  been  derived  for  a 
single  species  using  a  single  trapping 
technique.  Most  community  based 
studies  have  also  used  a  single  trap- 
ping technique.  Individual  species, 
however,  have  different  sensitivities 
to  capture,  making  it  difficult  to  com- 
pare capture  rates  across  species  (Se- 
ber  1981). 

To  better  understand  the  habitat 
associations  across  a  sequence  of  for- 
est ages  in  the  Pacific  Northwest,  we 
studied  the  population  status  in  se- 
lected forest  stands  in  northern  Cali- 
fornia and  southern  Oregon  during 
summer  and  fall  of  1984  and  1985. 
This  study  was  part  of  a  U.S.  Forest 
Service  research  project  extending 
from  northern  California  through 
Oregon  and  north  into  Washington 
(e.g.  Ruggiero  and  Carey  1984, 
Manuwal  and  Huff  1987).  The  im- 
pacts of  the  harvesting  of  old-growth 
forests  on  vertebrate  populations  in 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortt)  America  (Ragstaff. 
AZ.July  19-21.  1988). 

'USDA  Forest  Service,  Pacific  Southwest 
Forest  and  Range  Experiment  Station,  1700 
Bayy^ew  Drive,  Areata,  California  95521. 

^USDA.  Tongass-Chatham  Area  National 
Forest,  Juneau  Ranger  District,  8465  Old 
Dairy  Road.  Juneau,  Alaska  99801. 


this  area  are  uncertain  (Hagar  1960, 
Raphael  and  Barrett  1984,  Raphael  et 
al.  in  press). 

We  trapped  mammals  over  a  gra- 
dient of  different-aged  forest  stands 
using  three  techniques.  Our  primary 
objectives  were:  (1)  to  determine  if 
the  relative  abundance  of  each  spe- 
cies differed  between  the  stands;  (2) 
to  determine  which  habitat  variables 
were  associated  with  the  relative 
abundances  of  each  species;  and  (3) 
to  study  the  efficiency  of  different 
trapping  techniques.  In  this  paper  we 
discuss  differences  in  habitat  models 
derived  from  different  techniques  for 
the  five  most  abundant  species  of 
small  mammals. 


Methods  and  Materials 


Study  Area 

We  selected  47  study  stands  in  three 
regions  of  northwestern  California 
and  southwestern  Oregon.  These 
stands  represented  a  successional 
gradient  typical  of  the  Douglas-fir 
communities  of  the  region.  Stands 
ranged  in  elevation  from  414  m  to 
1,556  m  and  were  generally  domi- 
nated by  Douglas-fir  in  association 
with  tanoak  (Lithocarpus  densiflora) 
and  madrone  (Arbutus  menziesii). 
Three  low  elevation  stands  had  a 
redwood  (Sequoia  sempervirens)  com- 
ponent; four  high  elevation  stands  in 
the  Cave  Junction  region  were  domi- 


nated by  white  fir  (Abies  concolor). 

Fifteen  stands  were  located  at  each 
of  three  regions  (in  the  vicinities  of 
Branscomb  and  Willow  Creek,  Cali- 
fornia, and  Cave  Junction,  Oregon), 
with  an  additional  2  stands  at  Butte 
Creek,  near  Dinsmore,  California. 
These  stands  were  divided  into  three 
age  classes  based  on  core  samples  of 
2  to  10  of  the  dominant  Douglas-firs 
in  each  young  and  mature  stand  (up 
to  approximately  180  years  of  age) 
(B.  Bingham,  USFS  Pacific  Southwest 
Station,  pers.  commun.).  In  old- 
growth  stands,  tree  cores  could  not 
always  be  taken  because  of  large  tree 
size  and  rotten  tree  cores,  thus  some 
stand  ages  were  based  on  rings 
counted  on  stumps  in  adjacent 
clearcuts,  along  roads,  or  on  core 
samples  provided  by  local  Forest 
Service  offices.  Each  forest  stand  was 
assigned  to  one  of  three  age  classes: 
young  forest  <  100  years;  mature  for- 
est 100-180  years;  and  old-growth 
forest,  >  180  years.  Those  that  were 
classified  as  old-growth  forest  were 
further  classified  as  to  moisture  re- 
gime: dry,  mesic,  or  wet,  based  on 
species  composition  and  percent 
cover  of  the  herb  and  shrub  layers  of 
the  stand  (B.  Bingham,  pers.  com- 
mun.). All  young  and  mature  stands 
represented  the  modal,  or  mesic 
moisture  class. 

An  index  to  the  yearly  solar  radia- 
tion was  derived  by  the  method  of 
Frank  and  Lee  (1966),  which  is  based 
on  slope,  aspect,  and  latitude.  Values 
are  largest  on  south-facing,  moderate 


368 


slopes,  and  lowest  on  north-facing, 
steep  slopes. 


Mammal  Trapping 

A  single  trapping  grid  for  snap/liv- 
etrapping  was  established  in  each 
stand  in  12  rows,  with  12  trap  sta- 
tions per  row.  Trap  stations  were 
placed  at  15-meter  intervals  resulting 
in  a  grid  165  m  x  165  m.  Each  stand 
was  relatively  homogeneous,  and 
grids  located  in  each  stand  were,  in 
most  cases,  separated  from  different 
habitat  types  by  at  least  100-m  of  the 
same  habitat. 

In  1984,  two  snaptraps  ("Museum 
Special")  were  placed  in  1984  at  each 
trap  station  within  1.5  meters  of  the 
grid  coordinate  on  all  47  stands.  Six 
stands  were  trapped  simultaneously 
(two  in  each  region)  for  five  days 
(four  nights)  until  all  stands  were 
sampled  (July  3  to  August  31).  In 
1985,  a  single  Sherman  livetrap  (7.6  x 
8.9  x  22.0  cm)  was  used  at  the  same 
stations  on  43  of  the  stands;  six 
stands  again  were  trapped  during 
each  five-day  session  from  July  9  to 
August  30.  We  did  not  livetrap  four 
stands  (two  in  Branscomb  area  and 
two  in  Cave  Junction  area).  In  both 
years,  each  trap  was  placed  along- 


side downed  logs,  brushy  vegetation, 
or  rodent  runways.  Baited  with  pea- 
nut butter  and  oat  groats  in  1984,  and 
oat  groats  and  sunflower  seeds  (3:1 
ratio)  in  1985,  the  traps  were  left  in 
place  for  four  nights.  We  feel  that  the 
four  nights  of  trapping  did  not  sig- 
nificantly alter  populations  between 
years.  In  the  analyses  below,  we  stan- 
dardized captures  to  the  number  per 
1000  trap-nights. 

We  used  pitfall  traps  to  sample 
small  mammal  populations  on  all  47 
stands  during  both  1984  and  1985.  A 
pitfall  grid  consisted  of  six  rows  of 
six  pitfall  traps  per  row  at  15  m  spac- 
ing in  each  stand.  Grids  were  located 
usually  more  than  100  m  from  snap 
and  pitfall  grids.  Traps  were  two  No. 
10  cans  taped  together  and  sunk  until 
the  top  was  flush  with  the  ground.  A 
funnel  collar  to  prevent  animals  from 
escaping  was  made  from  a  margarine 
container  with  the  bottom  cut  out. 
We  propped  a  cedar  shake  3-4  cm 
above  the  opening  to  the  pitfall  trap 
to  act  as  a  cover. 

Traps  were  examined  at  5-day  in- 
tervals for  50  days  in  October  and 
November  1984,  and  for  30  days  in 
October  1985.  In  the  analyses  that 
follows,  we  used  the  number  of 
mammals  captured  unstandardized 
for  effort. 


r 


Table  1.— Vegetation  variables  measured  for  each  cluster  of  trapping  sta- 
tions in  a  study  of  small  mamma!  abundance  in  Douglas-fir  forests  of 
southem  Oregon  and  northem  California,  1984-85. 


Ground  cover 


Vegetation  variables 


Rock 
Soil 

Small  Litter 

Moss 

Lichen 

First  litter  layer 
Second  litter  layer 
Solar  index 

Decoy  class°1  and  2  logs 
Decay  class  3. 4. 5  logs 


Herb 

Grass 

Fern 

Douglas-fir 
Tanoak 

Pacific  madrone 
Live  oaks 
Oregon  grape 
Redwood 
Poison  oak 
True  fir 
Alders 


Dogwood 
California  hazel 
Pines 

White  and  black  oaks 
Said 

Manzanita 
Rosa  spp. 
/?ut»L/s  spp. 
California  laurel 
Huckleberry 
Big-leaf  maples 
False  cedars 


°Jhomas  0979:80). 


The  complication  that  not  all  cap- 
ture methods  were  used  in  both 
years  of  the  study,  resulted  in  an  un- 
known year-effect  that  may  influence 
capture  frequency.  Despite  this  prob- 
lem, we  feel  that  the  data  are  instruc- 
tive as  to  the  variety  of  models  pro- 
duced, and  the  implications  for  in- 
vestigators. 


Vegetation  Sampling 

Vegetation  for  each  snap /livetrap 
grid  was  measured  on  16  plots  over- 
laying the  144  trap  stations.  Nine 
vegetation  plots  were  uniformly  dis- 
tributed among  the  36  pitfall  stations. 
Vegetation  and  structure  were  meas- 
ured in  5.6  m  and  15  m  radii  circular 
plots.  On  each  plot,  we  recorded: 
percent  cover  of  ground  cover  vari- 
ables (i.e.  rocks,  woody  debris);  per- 
cent cover  of  vegetation  at  five  height 
strata;  and  counts  of  trees  and  snags 
in  varying  size  classes. 

We  averaged  the  percent  cover 
values  for  25  vegetation  stations  (16 
in  the  snap/ livetrap  grids  plus  nine 
stations  in  the  pitfall  grids),  to  obtain 
mean  values  of  percent  cover  for  11 
ground  cover  variables  and  24  spe- 
cies of  plants  (or  groups  of  species) 
in  each  of  the  47  stands  (table  1).  We 
combined  some  taxa  into  genera 
prior  to  calculating  means:  the  true 
firs  {Ahies  spp.),  alders  {Alnus  spp.), 
huckleberries  {Vaccinium  spp.),  live 
oaks  iQuercus  spp.),  manzanita  (Arte- 
mesia  spp.),  various  roses  {Rosa  spp.), 
and  Rubus  spp.  The  vegetation  data 
were  vertically  stratified  into  five 
levels:  ground  (0-0.5  m),  shrub  level 
(>0.5-2.0  m),  mid-canopy  (>2.0  m- 
midlevel),  canopy  (those  trees  at  the 
average  height  of  the  stand),  and 
supercanopy  (those  trees  substan- 
tially above  the  canopy).  Mean  val- 
ues for  cover  by  stand  were  com- 
bined into  two  strata:  ''understory" 
included  ground  and  shrub  layers, 
while  ''canop/'  incorporated  mid- 
canopy,  canopy,  and  supercanopy. 

The  small  and  medium  trees  (<50 
cm  dbh)  were  counted  on  a  5.6  m  cir- 


369 


cular  plot,  while  large  trees  (>50  cm 
dbh)  were  counted  on  a  15  m  circular 
plot.  The  counts  of  18  species  of  trees 
were  averaged  over  the  stations  for 
each  grid  and  were  used  in  an  all- 
subsets  regression. 


Analyses 

We  used  one-way  analysis  of  vari- 
ance (ANOVA)  to  evaluate  differ- 
ences in  mammal  abundances  rela- 
tive to  three  stand  age  classes,  three 
moisture  classes  of  the  old-growth 
stands  in  each  of  the  three  regions 
(Branscomb  and  Butte  Creek  area. 
Willow  Creek  area,  and  Cave  Junc- 
tion area). 

These  analyses  were  done  on  the 
three  separate  sets  of  data,  without 
reference  to  the  each  other.  Interac- 
tion among  the  factors  was  ignored 
in  these  analyses.  When  significant 
differences  were  found  among  cap- 
ture frequencies  of  individual  species 
by  classes  of:  age,  moisture,  or  study 
area,  a  multiple  comparison  test  was 
used  to  determine  which  of  the 
groups  were  significantly  different.  A 
Tukey  test  (Zar  1984:186)  was  per- 
formed if  variances  were  found  to  be 
equal,  while  a  Games  and  Howell 
modification  was  used  in  the  case  of 
unequal  variances  (Keselman  and 
Rogan  1978). 

Pearson  product  moment  correla- 
tion coefficients  were  calculated  be- 
tween capture  frequencies  for  each 
combination  of  trapping  techniques 
and  between  capture  frequencies  and 
vegetation  means  over  all  stands. 
Variables  from  ground  cover,  herb 
and  shrub  cover,  and  canopy  trees 
were  included  in  all-possible-subsets 
regression  analyses  for  each  small 
mammal  species  when  a  significant 
correlation  existed  with  capture  fre- 
quency from  any  capture  technique. 
Five-variable  models  were  selected 
for  each  species  when  greater  than 
100  individuals  were  captured  by  a 
particular  technique.  Vegetation  vari- 
ables were  excluded  when  found  on 
less  than  25%  of  the  stands. 


Results  and  Discussion 

Twenty-three  species  of  small  mam- 
mals were  captured  during  the 


study,  though  several  were  repre- 
sented by  only  a  few  individuals 
(table  2).  The  three  techniques  dif- 
fered in  their  effectiveness  in  captur- 


r 


Table  2.— Number  of  captures  by  species  and  trapping  technique,  from  a 
study  of  small  mammals  in  northern  Califomia  and  southern  Oregon,  1984 
and  1985. 


Species 


Pitfalls     Snaptraps  Uvetraps 


Trowbridge's  Shrew  892 
(Sorex  frowbridgii) 
Pacific  Shrew  33 
(Sorex  pocificus) 

Vagrant  Shrew  1 
(Sorex  vagrons) 

Pacific  Water  Shrew  1 
(Sorex  bendidi) 

Shrew-Moie  40 

(Neurofrichus  gibbsii) 

Coast  Mole  1 

(Scapanus  orahus) 

Chipmunks  2 

(Tomias  spp.) 

Golden-mantled  Ground  SquirrelO 
(Sperm  ophilus  lateralis) 
Northern  Flying  Squirrel 
(GlaucorDys  sabrinus) 
Botta's  Pocket  Gopher 
(Thomomys  boffae) 
Deer  Mouse 

(Peromyscus  manicufatus) 
Pinyon  Mouse 
(Peromyscus  fruei) 
Dusky-footed  Woodrat 
(Neofoma  fuscipes) 
Bushy-tailed  Woodrat 
(Neofoma  cinerea) 
California  Red-backed  Vole  572 
(Clethrionomys  califomicus) 
Red  Tree  Vole 
(Arborimus  longicaudus) 
California  Vole 
(Microfus  califomicus) 
Long-tailed  Vole 
(Microfus  longicaudus) 
Creeping  Vole 
(t^icrofus  oregoni) 
Black  Rat 
(Ratfus  raffus) 
Pacific  Jumping  Mouse 
(Zapus  frinofafus) 
Short- tailed  weasel 
(Musfela  erminea) 
Number  of  Trapnights^ 


6 
5 
115 
16 
2 
0 


1 
14 
2 
6 
1 

3 
0 

135,360 


357 
70 
1 

0 
27 
0 
33 
0 
1 
2 

524 
205 
4 
0 
161 
0 
14 
0 
5 
0 

n 

0 

55,284 


101 


n 


0 


0 


0 


282 


Total 


1350 


114 


72 


317 


8 

15 

0 

7 

404 

1043 

213 

434 

28 

34 

5 

5 

101 

834 

0 

1 

5 

33 

0 

2 

10 

21 

0 

1 

0 

14 

6 

6 

23,367 

214.011 

°Totals  were  adjusted  for  traps  damaged  by  bears,  etc. 


370 


ing  different  species  of  mammals. 
Five  species  had  sufficient  captures 
(>  100  individuals  or  more,  by  one  or 
more  of  the  trapping  techniques)  to 
undergo  intensive  analyses.  These 
were  the  California  red-backed  vole, 
deer  mouse,  pinyon  mouse,  Trowbr- 
idge's shrew,  and  the  combined  chip- 
munk species. 

Associations  with  Area,  Age,  and 
Moisture  Class 

Most  mammals  were  found  in  all 
three  areas,  with  the  exception  of 
three  species  with  only  1-2  captured. 
The  California  red-backed  vole  had 


significantly  fewer  captures  in  the 
more  southerly  Branscomb  region 
than  in  the  central  and  northern  re- 
gions (table  3).  The  vole's  abundance 
was  significantly  correlated  with  true 
firs  (r  =  0.46,  P  <  0.05),  which  were 
found  on  11  stands  in  the  north  and 
no  stands  in  the  south.  The  two  mice 
(Peromyscus)  species  exhibited  the 
opposite  trend  with  captures  signifi- 
cantly greater  in  the  south  than  in  the 
north.  The  pinyon  mouse  was  corre- 
lated with  solar  index  which  is  gen- 
erally greater  in  the  southern  area. 
The  shrews  and  chipmunks  were 
found  equally  in  all  areas. 

The  red-backed  vole  was  the  only 
species  to  have  a  significant  associa- 


tion with  age  of  the  forest  stand  (P  < 
0.01).  This  confirms  the  study  of  Ra- 
phael and  Barrett  (1984)  and  Raphael 
(this  volume)  in  the  Willow  Creek 
area.  Our  capture  frequency  was 
fairly  low  in  stands  aged  at  less  than 
150  years,  while  greater  densities 
were  evident  in  many  older  stands 
(fig.  1).  No  such  relationship  was 
found  for  the  deer  mouse,  although 
Raphael  and  Barrett  (1984)  earlier 
showed  a  significant  association  with 
age  in  the  Willow  Creek  area. 

We  tested  the  abundance  of  small 
mammals  in  the  three  moisture 
classes  of  old-growth  forests:  dry, 
mesic,  an  wet.  Among  the  five  mam- 
mal species  with  large  sample  sizes, 
there  were  no  differences  in  capture 
frequency  according  to  the  various 
moisture  classes. 

Therefore,  we  found  that  within 
our  study  areas  in  the  Douglas-fir 
type,  there  were  few  significant  or 
strong  associations  between  five 
small  mammals  and  age  of  the  forest 
stand.  The  stands  chosen  to  represent 
the  different  age  and  moisture 
classes  in  this  study  were  all  natu- 
rally occurring.  The  young  stands 
originated  from  fire  or  other  cata- 
strophic events,  rather  than  by  tim- 
ber  harvest,  and  therefore  often  were 
heterogeneous  in  character  with 
structural  and  floristic  components 
similar  to  old-growth  stands.  Scat- 
tered old  trees  and  abundant  dead 
and  down  material  were  sometimes 
present  in  young  stands,  characteris- 
tics which  are  absent  from  stands 
that  originated  from  clearcuts;  results 
in  even-aged  stands  may  be  very  dif- 
ferent. 


Effectiveness  of  Capture 

Captures  of  small  mammals  varied 
greatly  by  trapping  technique  (table 
1).  The  two  mice  were  most  effec- 
tively captured  by  baited  snap  and 
livetraps.  Very  few  individuals  were 
collected  in  unbaited  pitfalls.  Microt- 
ine  voles,  shrews,  and  moles  were 
trapped  most  efficiently  by  the  pitfall 


t/i    30  - 
25- 


3: 
o 


I 

a. 

2 


o 
o 
o 


UJ 

OL. 

to 

UJ 

=> 
I— 

Q- 
-< 

o 


20- 
15- 
10- 


0- 


o  =  BR 
•  =  WC 

A  =  CJ 


T 

50 


100 


150  200 
FOREST  AGE 


Figure  1.— Captures  of  Cdlfomia  red-backed  voles  per  1000  trapnights  in  a  study  of  snrxill 
mamnrxji  atHjndance  relative  to  stand  age,  1984-1985.  BR  =  Branscomb  stands,  WC  =  Willow 
Creek  stands,  CJ  =  Cave  Junction  stands. 


Table  3.— Significance  of  differences  in  capture  frequency  by  area  for  five 
species  of  small  mammals.  The  areas  are  CJ  =  Cave  Junction,  WC  =  Willow 
Creek,  and  BR  =  Branscomb  and  Butte  Creek.  Methods  with  no  significant 
differences  in  capture  frequencies  at  the  various  areas  are  Indicated  by 
NS;  dashed  lines  indicate  inadequate  sample  size. 


Species 


Snaptrap 


Uvetrap 


Pitfalls 


California  red-backed  vole  NS  NS  BR<WC+CJ 

Deer  mouse  CJ<WC+BR  CJ<WC+BR  NS 

Plnyor^  mouse  CJ<BR  CJ<BR  — 

Trowbridge's  shrew  NS  NS  NS 

Chipmunks  —  NS  — • 


371 


traps  and  somewhat  by  snaptraps. 
Sciurids  and  woodrats  were  cap- 
tured almost  exclusively  by  livetraps. 

We  correlated  the  captures  of  each 
species  by  the  different  techniques. 
We  found  significant  correlation  be- 
tween capture  frequencies  only  in 
those  techniques  effective  at  sam- 
pling large  numbers  of  a  particular 
species  (table  4).  Demonstrating  the 
closest  agreement  between  tech- 
niques were  the  pinyon  mouse  (r  = 
0.88  between  snap  and  livetraps)  and 
the  vole  (r  =  0.73  between  the  two 
years  of  pitfall  traps).  The  Trowbr- 
idge's shrew,  on  the  other  hand, 
showed  no  relationship  between  cap- 
tures by  pitfalls  and  snaptraps  (r  = 
0.14),  or  pitfalls  and  livetraps  (r  = 
0.09).  Biological  interpretation  of 
such  varied  results  may  be  very  diffi- 
cult, as  discussed  in  the  following. 


Vegetation  Models 

Depending  on  which  method  was 
used  to  predict  the  dependent  vari- 
able, we  obtained  very  different 
vegetation  models,  potentially  result- 
ing in  very  different  biological  inter- 
pretations. Models  from  snap  and 
livetrapping  show  that  areas  with 
high  captures  of  pinyon  mice  were 
characterized  by  high  densities  of 
pacific  madrone  and  tanoak,  high 
solar  index,  and  bare  soils  (r^  =  0.64 
and  0.65)  (table  5).  Four  of  the  five 
habitat  variables  were  identical  in 
both  models  suggesting  that  within 
our  study  area,  the  pinyon  mouse 
used  dryer,  southern  exposures  with 
exposed  soils  and  large  amounts  of 
hardwoods. 

Models  developed  for  the  Trowbr- 
idge's shrew  from  snaptrap  and  pit- 
fall methods  were  quite  different 
(table  6).  Only  one  variable  was  in- 
cluded in  both  models,  and  the  asso- 
ciation with  dogwood  trees  switched 
from  negative  to  positive.  Both  mod- 
els included  some  indication  of 
greater  use  of  older  stands,  i.e.,  the 
model  using  snaptrap  data  included 
well  decayed  logs  and  the  livetrap 


model  incorporated  the  decomposed 
litter  layer,  representing  a  well  devel- 
oped layer  of  organic  soil.  The  incon- 
sistency in  these  vegetation  models 
was  predicted  by  the  lack  of  correla- 
tion between  capture  frequencies  by 
the  two  techniques.  It  appears  that  in 
our  Douglas-fir  habitat  type,  the 
shrew  may  be  broadly  distributed, 
independent  of  finer  vegetation  com- 
position. 

Models  for  the  red-backed  vole 
developed  from  capture  frequencies 
associated  with  different  trapping 
techniques  (table  7)  were  more  simi- 
lar than  those  for  the  shrew,  but  less 


similar  than  those  for  the  pinyon 
mouse.  In  models  developed  from 
snap  and  livetrap  captures,  three  of 
the  five  variables  were  selected  by 
both  models.  Models  from  pitfall  and 
snaptrap  data  shared  two  of  the  five 
variables  selected.  Models  from  pit- 
fall and  livetrap  capture  data  also 
shared  two  of  the  five  variables  se- 
lected, but  one  of  these  variables 
switched  from  a  positive  to  a  nega- 
tive association.  Only  the  response  to 
an  abundant  herbaceous  layer  was 
consistent  in  models  from  all  three 
trapping  techniques.  Interpretation 
of  the  snaptrap  model  suggests  that 


r 


Table  4.— Correlation  between  years  and  methods  of  the  capture  fre- 
quency of  four  small  mammal  species  in  snaptraps  (Snap),  livetraps  (Live), 
or  pitfall  traps  (Pits).  (Chipmunks  were  only  caught  in  significant  numbers  In 
livetraps  and  could  not  be  compared). 


Between  years 


Within  yecffs 


Snap84/live85    Pits:84/85     84:pits/snap  85:pits/live 


California  red-backed  vole  0.540"  0.727**  0.459**  0.162 

Deer  mouse  0.392**  0.015  -0.092  0.320* 

Pinyon  mouse  0.884**  0.124  0.250  0.320* 

Trowbridge's  shrew  0.102  0.332*  0.141  0.088 

'=P<Q.Qb. 

"=p<om. 


Tcrf>le  5,— Habitat  association  mod- 
els for  the  pinyon  mouse  deter- 
mined from  capture  frequencies 
by  two  different  trapping  tech- 
niques used.  NS  indicates  the  vari- 
able was  not  selected.  +  or  -  Indi- 
cates a  positive  or  negative  asso- 
ciation with  capture  frequency. 


Selected  predictor 
variables 


Snaptrap  Livetrap 


Exposed  rock  NS  + 

Bar©  soil  +  + 

Solar  index  +  + 

Poison  oak  -  NS 

Tanoak  +  ■»■ 

Pacific  madrone  +  + 

0,64  0.66 

Correlotion  between  capture  fre- 
quencies ofttie  two  tectiniques  =  0.88. 


Table  6.— Habitat  association  mod- 
els for  the  Trowbridge's  shrew  de- 
termined from  ccqDture  frequencies 
by  two  different  trapping  tech- 
niques used.  Symbols  as  In  table  5. 


Selected  predictor 
variobles 


Snaptrap  Livetrap 


Highly  decayed  logs 

+ 

NS 

Fern 

-»■ 

NS 

Dogwood  shrub 

NS 

Dogwood  tree 

+ 

Deciduous  oaks 

+ 

NS 

True  firs 

NS 

+ 

Tanoak 

NS 

California  hazel 

NS 

+ 

Deep  titter  layer 

NS 

+ 

0.59 

0.55 

Correlation  between  capture  fre- 
quencies by  two  techniques  =  0.14, 


372 


the  vole  is  associated  with  a  fairly 
moist  habitat  (abundant  herbs  and 
presence  of  huckleberr}^.  The  pitfall 
nnodel  also  suggests  an  association 
with  a  moist  habitat  (more  herbs  and 
lichens  and  less  solar  index).  The  liv- 
etrap  model  includes  some  indication 
of  moist  habitats  (herbs,  Rosa  spp., 
and  huckleberry)  but  also  a  sugges- 
tion of  a  dryer  habitat  (solar  index). 


The  deer  mouse,  despite  its  abun- 
dance, had  large  differences  between 
variables  selected  in  habitat  models 
(table  8).  Its  relative  abundance  did 
not  appear  to  be  associated  with  the 
same  habitat  variables  in  the  same 
way  for  the  three  different  trapping 
techniques.  Only  two  of  the  12  vari- 
ables selected  in  these  models  were 
included  in  more  than  one  model 


r 


Jab\e  7.— Habitat  association  models  for  the  Calitomia  red-backed  vole 
determined  from  capture  frequencies  by  three  different  trapping  tech- 
niques. Symbols  as  in  table  5. 


Selected  predictor 
variables 

Herbs 
Rose 

Huckleberry 
False  cedar 
Douglas-fir 
solar  index 
Live  oaks 
Lichen 
Grass 

R2 


Correlation  between  capture  frequencies:  snaptrap  and  llvetrap  =  0.64  (P  <  0.0?A' 
\^  snaptrap  and  pitfall  =  0.50(P  <  0.01)j  pitfall  andJivetrap  =  0. 16  (NS). 


Table  8. ~ Habitat  association  models  for  the  deer  mouse  determined  from 
capture  frequencies  by  three  different  trapping  techniques.  Symbols  as  In 
table  5. 


Snaptrap 

Llvetrap 

Pitfall 

+ 

+ 

+ 

+ 

NS 

+ 

+ 

NS 

+ 

NS 

NS 

+ 

NS 

+ 

NS 

+ 

NS 

NS 

NS 

NS 

+ 

NS 

NS 

+ 

0.58 

0.55 

0.63 

Selected  precfictor 
variables 


Snaptrap 


Llvetrap 


Pitfall 


Lichen 

True  firs 

Douglas-fir 

California  laurel 

Pacific  madrone 

Manzanita 

Rose 

Dogwood 
Deciduous  oaks 
Lower  litter  layer 
Herbs 

False  cedars 

R2 


NS 
NS 
NS 
NS 
NS 
NS 
NS 

0.33 


NS 
NS 
NS 
+ 

NS 


+ 
+ 

NS 
NS 
NS 


0.38 


NS 
NS 
+ 

NS 
NS 
NS 

NS 
NS 

+ 

+ 

0.63 


Correlation  between  capture  frequencies;  snaptrap  and  livetrap  =  0.39(9  <  0.01): 
snaptrap  and  pitfall  =  -0.09  (NS);  pitfall  arid  livetrap  =  0.32  (P  <  0.05). 


with  the  same  sign  (avoidance  of 
Rosa  spp.  and  preference  for  areas 
with  California  laurel).  Model  dis- 
parity may,  of  course,  simply  indi- 
cate that  one  or  more  of  the  tech- 
niques estimated  the  dependent  vari- 
able with  considerable  bias,  thus  pro- 
ducing an  erroneous  model. 

The  chipmunks  were  captured  pri- 
marily by  livetrapping.  The  resulting 
5-variable  model  suggests  that  chip>- 
munks  were  more  common  in  the 
true  fir  stands  at  high  elevation  that 
had  an  understory  of  live  oaks  and 
huckleberries  (table  9). 

While  we  are  sure  that  there 
would  be  some  seasonal  differences 
in  the  habitat  association  patterns 
from  autumn  captures  in  pitfalls  and 
summer  captures  in  snap  and  liv- 
etraps,  we  suggest  that  this  seasonal 
effect  would  be  much  less  than  the 
differences  that  we  noted,  because  of 
the  relatively  low  vagility  of  the 
small  mammals  involved. 

All  capture  methods  are  assumed 
to  sample  individuals  of  a  given  spe- 
cies at  some  unknown  proportion  of 
their  true  abundance.  These  propor- 
tions, within  a  species,  likely  differ 
by  capture  method.  If  the  capture  ef- 
ficiency of  all  methods  were  consis- 
tent across  sampled  areas,  then  the 
rank  correlation  of  abundance  be- 
tween methods  should  be  close  to 
1.0.  However,  for  most  species  that 
we  studied,  correlations  of  capture 
frequencies  between  methods  were 
low  and  the  ranking  of  stands  based 


Table  9.— Habitat  association  mod- 
els for  the  chipmunks  determined 
from  capture  frequencies  by  liv- 
etrapping. Symbols  as  in  table  5. 

Selected  predictor 

variables  Llvetrap 

True  fir  + 
Douglas-fir 

Lichen  + 

Vaccinium  + 

Live  Oaks  + 

R2  0.59 


373 


on  capture  frequencies  varied  con- 
siderably depending  on  technique 
used.  This  suggests  that  the  assump- 
tion of  a  constant  proportion  of  cap- 
tures, within  a  given  method,  across 
sampled  areas  was  violated. 

Acknowledgments 

We  thank  the  many  members  of  the 
old-growth  field  crew  for  their  hours 
of  work  under  difficult  conditions. 
Linda  Doerflinger  and  Howard  Sakai 
were  helpful  in  many  ways,  keeping 
the  field  crews  supplied,  organizing 
data,  and  efficiently  expediting  other 
aspects  of  complex  field  work.  We 
are  also  grateful  to  Andrew  B.  Carey, 
Richard  Golightly,  Barry  R.  Noon, 
Nancy  A.  Tilghman,  and  Martin  G. 
Raphael  who  all  made  helpful  com- 
ments on  the  manuscript. 

Literature  Cited 

Dueser,  R.D.  and  H.H.  Shugart.  1978. 
Microhabitats  in  a  forest-floor 
small  mammal  fauna.  Ecology 
59:89-98. 

Frank,  E.  C.  and  R.  Lee.  1966.  Poten- 
tial solar  beam  irradiation  on 
slopes.  U.S.  Dept.  Agric.  For.  Serv. 
Res.  Pap.  RM-18. 

Hagar,  D.C.  1960.  The  interrelation- 
ships of  logging,  birds,  and  timber 
regeneration  in  the  Douglas-fir  re- 
gion of  northwestern  California. 
Ecology  41:116-125. 

Keselman,  H.J.  and  J.C.  Rogan.  1978. 
A  comparison  of  the  Modified- 
Tukey  and  Scheffe  methods  of 
multiple  comparisons  for  pairwise 
contrasts.  Journal  of  the  American 
Statistical  Association  73:47-52. 

MacGracken,  J.G.,  D.W.  Uresk,  and 
R.M.  Hansen.  1985.  Habitat  used 
by  shrews  in  southeastern  Mon- 
tana. Northwest  Science  59:24-27. 

M'Closkey,  R.T.  1975.  Habitat  Di- 
mensions of  white-footed  mice, 
Peromyscus  leucopus.  American 
Midland  Naturalist  93:158-167. 


Manuwal,  D.A.  and  M.H.  Huff.  1987. 
Spring  and  winter  bird  popula- 
tions in  a  Douglas-fir  forest  sere. 
Journal  of  Wildlife  Management 
51:586-595. 

Raphael,  M.G.  and  R.H.  Barrett. 
1984.  Diversity  and  abundance  of 
wildlife  in  late  succession 
Douglas-fir  forests,  p.  352-360.  In 
New  Forests  for  a  Changing 
World.  Proceedings  1983  Conven- 
tion of  the  Society  of  American 
Foresters.  650  pp. 

Raphael,  M.G.,  K.V.  Rosenberg,  and 
B.G.  Marco t.  In  press.  Large-scale 
changes  in  bird  populations  of 
Douglas-fir  forests.  Northwestern 
California.  Bird  Conservation  3. 

Rosenzweig,  M.L.  1973.  Habitat  se- 
lection experiments  with  a  pair  of 
coexisting  heteromyid  rodent  spe- 
cies. Ecology  54:111-117. 

Ruggiero,  L.  F.  and  A.  B.  Carey.  1984. 
A  programmatic  approach  to  the 
study  of  old-growth  forest-wildlife 
relationships,  p.  340-345.  In  Pro- 
ceedings of  the  1983  convention  of 
the  Society  of  American  Foresters: 
New  forests  for  a  changing  world. 
Society  of  American  Foresters,  Be- 
thesda,  MD. 

Seber,  G.A.F.  1981.  The  estimation  of 
animal  abundance  and  related 
parameters.  Second  Edition.  Ox- 
ford University  Press,  New  York. 

Thomas,  J.  W.  (Ed.).  1979.  Wildlife 
habitats  in  managed  forests. 
USDA  Forest  Service  Agriculture 
Handbook  No.  553. 

Zar,  J.H.  1984.  Biostatistical  Analysis. 
Prentice-Hall,  Inc.  Englewood 
Cliffs,  N.J.  718  pp. 


374 


Small  Mammals  in 
Streamside  Management 
Zones  in  Pine  Plantations^ 

James  G.  Dickson^  and  J.  Howard 
Williamson^ 


Abstract.— Small  mammals  were  captured  in  live 
traps  in  6  mature-forested  streamside  management 
zones  of  3  widths,  narrow  (<  25  m),  medium  (30-40 
m),  and  wide  (50-90  m).  which  traversed  young, 
brushy  pine  plantations.  More  small  mammals  were 
captured  in  the  narrow  zones  (165)  than  in  the  me- 
dium (82),  or  wide  zones  (65). 


Many  second-growth  pine-hardwood 
stands  in  southern  forests  are  being 
cut  and  replaced  by  pine  plantations, 
especially  on  industrial  land.  From 
1971  to  1986,  the  amount  of 
Mid  south  timberland  in  pine  planta- 
tions increased  from  6  to  8%  (Birdsey 
and  McWilliams  1986).  White-tailed 
deer  adapt  well  to  young  brushy 
clearcuts  with  ample  forage  and  soft 
mast.  Also,  many  species  of  birds  are 
abundant  in  this  diverse  brushy  habi- 
tat (Dickson  and  Segelquist  1979). 
But  the  effects  of  clearcutting  and 
planting  on  all  vertebrate  species  are 
not  well  assessed  or  defined. 

Various  environmental  conces- 
sions are  being  implemented  along 
with  stand  conversion.  One  practice 
used  to  protect  water  quality  and 
enhance  wildlife  habitat  is  to  retain 
mature  forest  stands  along  intermit- 
tent and  permanent  streams  when 
adjacent  stands  are  cut  and  planted 
to  pines  (Dickson  and  Huntley  1986, 
Seehorn  1986).  These  areas  of  mature 
pine  or  pine-hardwoods  are  called 
riparian  zones,  filter  strips,  stringers, 
streamers,  or  streamside  manage- 
ment zones  (SMZ).  These  areas  en- 

' Paper  presented  at  symposium,  Man- 
ogement  of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff. AZ.  July  19-21,  1988.) 

'James  G.  Dickson.  Supervisory  Research) 
Wildlife  Biologist,  Wildlife  Habitat  Labora- 
tory. Southern  Forest  Experiment  Station, 
USDA  Forest  Service,  Nacogdoches.  Texas. 

^J.  Howard  Williamson.  Forestry  Techni- 
cian. Wildlife  Habitat  Laboratory.  Southern 
Forest  Experiment  Station.  USDA  Forest  Serv- 
ice. Nacogdoches.  Texas. 


hance  habitat  diversity  and  "edge," 
offer  suitable  habitat  for  wildlife  spe- 
cies associated  with  mature  stands, 
serve  as  travel  corridors  for  animals, 
and  may  permit  genetic  interchange 
between  otherwise  isolated  popula- 
tions of  animals.  Retention  of  SMZ 
for  reduction  of  non-point  pollution 
and  for  wildlife  has  been  widely  rec- 
ommended. 

These  mature  hardwood  strips  can 
be  good  squirrel  habitat.  In  Missis- 
sippi (Warren  and  Hurst  1980)  and  in 
eastern  Texas  (McElfresh  et  al.  1980), 
gray  (Sciurus  carolinensis)  and  fox  (S. 
niger)  squirrel  numbers  were  higher 
in  riparian  areas  than  in  adjacent  up- 
land stands.  In  another  facet  of  the 
present  investigation,  gray  and  fox 
squirrels  were  abundant  in  SMZ 
wider  than  50  m  but  virtually  absent 
from  zones  less  than  40  m  wide 
(Dickson  and  Huntley  1986).  A  wide 
variety  of  reptiles  and  amphibians 
were  abundant  in  zones  greater  than 
30  m  wide,  where  a  closed  canopy 
offered  shaded  understory,  but  were 
scarce  in  SMZ  less  than  25  m  wide, 
which  were  dominated  by  low, 
brushy  vegetation  (Rudolph  and 
Dickson  In  Press).  The  relationships 
of  SMZ  and  other  wildlife  species  are 
largely  unknown. 

The  objective  of  this  study  was  to 
determine  the  relationship  of  SMZ 
width  to  small  mammal  communi- 
ties. We  assessed  the  effects  of  nar- 
row (<25  m),  medium  (30-40  m),  and 
wide  (>50  m)  SMZ  widths  on  small 
mammal  captures  in  6  SMZ  in  east- 
ern Texas. 


Study  Areas  and  Mettiods 

Study  areas  consisted  of  6  pine  plan- 
tations on  the  western  edge  of  the 
southern  coastal  plains  in  eastern 
Texas.  Mature  pine  and  hardwood 
trees  on  the  areas  had  previously 
been  harvested.  The  plantations  had 
been  planted  to  loblolly  pine  (Pinus 
taeda)  seedlings  5  to  6  years  before 
this  study  was  begun  and  were  vege- 
tated by  diverse  flora,  dominated  by 
hardwood  and  other  woody  brush. 
Oaks  {Quercus  spp.)  and  sweetgum 
(Liquidambar  styraciflua)  sprouts, 
American  beautyberry  (Callicarpa 
americam),  blackberry  and  dewberry 
(Rubus  spp.),  and  sumac  {Rhus  spp.) 
were  abundant. 

Each  of  the  6  study  areas  was  trav- 
ersed by  a  SMZ  of  mature  vegeta- 
tion. Dominant  trees  (>  13  cm  dbh)  in 
decreasing  order  of  abundance  and 
stem  density  (No. /ha)  were  as  fol- 
lows: sweetgum,  63;  white  oak  (Q. 
alba),  36;  southern  red  oak  (Q.  fal- 
cata),  28;  red  maple  (Acer  rubrum),  19; 
black  gum  (Nyssa  sylvatica),  14; 
shortleaf  pine  (P.  echinata),  14;  and 
eastern  hophornbeam  (Ostrya  virgini- 
ana),  14.  Dominant  understory  vege- 
tation (5-13  cm  dbh)  and  stem  den- 
sity (No./ha)  included  sweetgum, 
140;  eastern  hophornbeam,  71;  black 
gum,  40;  flowering  dogwood  (Cornus 
florida),  40;  loblolly  pine,  21;  and  red 
maple,  19. 

Assigned  treatments  were  3  SMZ 
widths:  narrow  (<25  m),  medium 
(30-40  m),  and  wide  (>50  m).  Two 
replications  of  each  treatment  were 


375 


sampled  at  2  locations.  In  each  of  the 
6  study  areas  two  200-nn  transects 
were  established  along  each  of  the  6 
streamside  zones.  Distance  from 
points  along  the  transects  to  the  SMZ 
edge  was  variable  because  each  zone 
orientation  changed  somewhat  with 
stream  meanders.  Thirteen  Sherman 
live  traps  were  placed  12.5  m  apart 
on  each  of  the  12  transects.  Trapping 
was  conducted  4  consecutive  nights 
in  each  of  2  consecutive  weeks  (8 
nights)  during  February  and  March 
in  1986  and  again  in  1987  (52  traps/ 
treatment  X  8  nights  X  2  years  =  832 
trap  nights).  Traps  were  baited  with 
oatmeal  each  morning  and  checked 
the  following  morning. 

Captures  per  treatment  were  ap- 
proximately normally  distributed  ac- 
cording to  the  Kolmogorov-Smirnov 
Goodness  of  Fit  Test.  Each  of  the  3 
treatments  was  tested  for  differences 
between  years  with  the  T-Test.  There 
were  no  significant  differences  be- 
tween years  (P  >  .10);  therefore,  cap- 
ture data  were  combined  for  both 
years.  Treatment  effects  (captures/ 
treatment)  were  tested  for  differ- 
ences by  ANOVA  and  the  Duncan's 
Multiple  Range  Test  at  the  0.05  level 
of  confidence. 

White-footed  mice  (Peromyscus  leu- 
copus)  and  cotton  mice  (P.  gossypinus) 
were  grouped  together  because  of 
difficulty  in  positive  field  identifica- 
tion. Davis  (1974)  determined  that 
white-footed  mouse  adults  were 
smaller  (15  to  25  g,  as  opposed  to  > 
30  g  for  the  cotton  mouse)  and  had 
brighter  colors.  Also,  adult  hind-foot 
length  was  shorter  (21  mm,  as  op- 
posed to  23  mm  for  cotton  mice). 
However,  numerous  sub-adults  were 
captured  during  the  trapping  period, 
making  identification  extremely  diffi- 
cult. 


Results  and  Discussion 

Significantly  more  small  mammals 
were  captured  in  the  narrow  SMZ 
(165)  than  were  captured  in  the  me- 
dium (82)  or  wide  (65)  SMZ  (table  1). 


The  absence  of  tree  canopy  in  the 
narrow  zones  permitted  dense, 
brushy  vegetation  growth,  abundant 
seeds,  and  dense  logging  slash  cover, 
but  medium  and  wide  zones  were 
characterized  by  shaded  sparse 
understories  under  closed  canopies. 
Other  studies  have  shown  higher 
densities  of  small  mammals  in  young 
brushy  stands  than  in  mature  stands. 
In  an  earlier  study  in  eastern  Texas, 
64  small  mammals  were  captured  in 
a  6-year-old  clearcut,  but  only  24  in  a 
pine-hardwood  stand  more  than  35 
years  old.  Small  mammal  species  di- 
versity was  also  higher  in  the  young 
stand  (Fleet  and  Dickson  1984).  In 
pine  plantations  in  Georgia,  small 
mammal  abundance  was  higher  in  1- 
to  5-year-old  pine  plantations  than  in 
older  stands  with  closed  canopies 
(Atkeson  and  Johnson  1979).  Seed- 
eaters  were  abundant  in  the  1 -year- 
old  plantation,  but  herbivores  were 
abundant  in  older  young  brushy 
stands. 

In  Pennsylvania,  relative  abun- 
dance of  small  mammals  was  greater 
in  recent  clearcuts  of  both  northern 
hardwood  and  oak  forests  than  in 
adjacent  mature  stands  (Kirkiand 
1978).  A  similar  pattern  was  noted  in 


deciduous  and  boreal  forests  in  West 
Virginia  (Kirkiand  1977).  After 
clearcutting,  small  mammal  abun- 
dance and  diversity  increased  and 
remained  relatively  high  until  stands 
returned  to  forest.  In  Arizona,  rodent 
populations  were  higher  in  thinned 
ponderosa  pine  (P.  ponderosa)  stands 
with  slash  than  in  unthinned  stands 
(Goodwin  and  Hungerford  1979). 

The  most  abundant  species,  the 
fulvous  harvest  mouse  (Reithrodonto- 
mys  fulvescens)  and  the  white  footed 
mouse/cotton  mouse  complex,  were 
much  more  abundant  in  the  narrow 
zone.  For  the  fulvous  harvest  mouse, 
there  were  73  captures  in  the  narrow, 
4  in  the  medium,  and  3  in  the  wide 
zones. 

Apparently,  the  dense  brushy 
vegetation  with  ample  down  logging 
slash  provided  ideal  habitat  for  this 
species.  There  was  abundant  vegeta- 
tive forage,  seeds,  and  dense  log  and 
brush  cover.  Schmidly  (1983)  de- 
scribed the  best  habitats  for  fulvous 
harvest  mice  in  the  pineywoods  as 
grassland,  pine-grass  ecotone,  and 
grass-brush.  In  an  earlier  study  in 
eastern  Texas  (Fleet  and  Dickson 
1984),  fulvous  harvest  mice  were 
captured  regularly  in  a  young  pine 


Table  Number  of  small  mammals  captured  In  streamside  management 
zones  in  pine  plantations  (832  trap  nights)  per  treatment. 


SMZ  width 

Narrow 

Medium 

Wide 

Hispid  Cotton  Rat 

(Sigmodon  hispidus) 

9 

3 

Fulvous  Harvest  Mouse 

(Reithrodontomys  fulvescens) 

73 

4 

3 

Eastern  Harvest  Mouse 

(Reithrodontomys  humulis) 

1 

1 

White-footed  and  Cotton  Mouse 

(Peromyscus  leucopus  and  gossypinus) 

76 

67 

50 

Golden  Mouse 

(Perom  yscus  muttalii) 

3 

4 

Florida  Wood  Rat 

(Neotoma  floridana) 

3 

5 

4 

Short-tailed  Shrew 

(Blarina  brevicauda) 

2 

4 

Totals 

165 

82 

65 

376 


plantation,  but  were  not  captured  in 
the  adjacent  mature  pine-hardwood 
stand.  In  a  study  of  small  mammal 
populations  in  5  pine  stands  in  Lou- 
isiana, fulvous  harvest  mice  were 
captured  most  frequently  in  a  pine 
seed-tree  harvest  cut  having  dense 
hardwood  brush  (Hatchell  1964). 

Differences  among  treatments 
were  less  pronounced  for  the  Pero- 
myscus  complex,  with  captures  of  76 
in  the  narrow,  67  in  the  medium,  and 
50  in  the  wide  SMZ.  In  a  1 -year-old 
pine  plantation  in  Georgia,  the  white- 
footed  mouse  was  the  dominant  sp)e- 
cies  (Atkeson  and  Johnson  1979).  It 
also  was  the  most  abundant  species 
in  the  mature  oak-hickory  forest  type 
in  eastern  Tennessee  (Dueser  and 
Shugart  1978).  Cotton  mice  were  cap- 
tured regularly  in  5  mature  pine 
stands  in  Louisiana  (Hatchell  1964) 
and  in  a  pine-hardwood  stand  in 
eastern  Texas  (Fleet  and  Dickson 
1984).  Neither  species  was  captured 
in  a  pine  plantation  in  the  Texas 
study.  Schmidly  (1983)  describes  pre- 
ferred habitat  of  the  cotton  mouse  as 
flatland  hardwood,  flatland  hard- 
wood-pine, and  lower  slope  hard- 
wood-pine. McCarley  (1954)  associ- 
ated the  white-footed  mouse  with 
upland  forest  habitat. 

Six  other  species  were  not  cap- 
tured frequently  enough  for  conclu- 
sions concerning  habitat  preference. 
Habitat  preferences  have  been  docu- 
mented to  some  degree  in  other  stud- 
ies. The  hispid  cotton  rat  is  often 
very  abundant  and  normally  is  asso- 
ciated with  low,  dense  vegetation 
(Atkeson  and  Johnson  1979,  Reet  and 
Dickson  1984,  Goertz  and  Long  1973, 
Schmidly  1983).  It  has  occasionally 
been  found  in  habitats  dominated  by 
early  successional  grasses  and  forbs. 

The  golden  mouse  is  associated 
with  forested  stands  having  low, 
dense  vegetation  (Fleet  and  Dickson 
1984,  Hatchell  1964,  McCarley  1958). 
The  Florida  wood  rat  occupies  for- 
ested upland  and  streamside  habitat 
and  thrives  in  bottomland  hardwood 
stands  with  low  brushy  understories 
(Schmidly  1983).  Short-tailed  shrews 


were  captured  in  the  medium  (2)  and 
wide  zones  (4).  Other  investigations 
have  found  them  inhabiting  a  variety 
of  mature  stands  (Fleet  and  Dickson 
1984,  Hatchell  1964,  Schmidly  1983). 

In  conclusion,  more  small  mam- 
mals, especially  fulvous  harvest 
mice,  were  captured  in  narrow  SMZ 
than  in  medium  and  wide  SMZ.  Ap- 
parently, this  is  related  to  the  abun- 
dance of  low,  dense  vegetation,  with 
ample  forage,  fruits,  and  seeds;  and 
down  logs  and  logging  slash.  But 
medium  and  wide  SMZ  with  closed 
tree  canopies  provide  limited  mature 
habitat  for  some  species  associated 
with  mature  stands,  such  as  the 
short-tailed  shrew,  and  are  positive 
for  a  variety  of  other  wildlife. 

Acknowledgments 

We  thank  Jimmy  C.  Huntley  for  trap- 
ping assistance  and  James  A.  Neal 
and  W.  V.  Robertson  for  reviewing 
an  earlier  draft  of  this  manuscript. 

Literature  Cited 

Atkeson,  Thomas  D.  and  A.  Sydney 
Johnson.  1979.  Succession  of  small 
mamimals  on  pine  plantations  in 
the  Georgia  piedmont.  American 
Midland  Naturalist  101:385-392. 

Birdsey,  R.  A.  and  W.  H.  McWil- 
liams.  1986.  Midsouth  forest  area 
trends.  USDA  Forest  Service  Re- 
source Bulletin  SO-107. 17  p. 
Southern  Forest  Experiment  Sta- 
tion, New  Orleans,  La. 

Davis,  William  B.  1974.  The  mam- 
mals of  Texas.  Texas  Parks  and 
Wildlife  Department  Bulletin  No. 
41,  Ausrin,  TX.  294  p. 

Dickson,  James  G.  and  Jimmy  C. 
Huntley.  1986.  Streamside  man- 
agement zones  and  wildlife  in 
southern  forests:  the  problem  and 
squirrel  relationships,  p.  37-39.  In 
Managing  southern  forests  for 
wildlife  and  fish — a  proceedings. 
J.G.  Dickson  and  O.E.  Maughan, 
eds.  USDA  Forest  Service  General 


Technical  Report  SO-65.  85  p. 
Southern  Forest  Exf)eriment  Sta- 
tion, New  Orleans,  La. 

Dickson,  James  G.  and  C.  A. 
Segelquist.  1979.  Breeding  bird 
populations  in  pine  and  pine- 
hardwood  stands  in  East  Texas. 
Journal  of  Wildlife  Management 
43:549-555. 

Dueser,  Raymond  D.  and  H.  H. 
Shugart,  Jr.  1978.  Microhabitats  in 
a  forest-floor  small  mammal 
fauna.  Ecology  59:89-98. 

Fleet,  Robert  R.  and  James  G. 

Dickson.  1984.  Small  mammals  in 
two  adjacent  forest  stands  in  East 
Texas,  p.  264-269.  In  Proceedings 
of  workshop  on  management  of 
nongame  species  and  ecological 
communities.  [Lexington,  Ky., 
June  11-12, 1984]  W.  C.  McComb, 
editor.  404  p.  University  of  Ken- 
tucky, Lexington. 

Goertz,  John  W.  and  Roland  C.  Long. 
1973.  Habitats  of  five  species  of  rat 
in  Louisiana.  American  Midland 
Naturalist  90:460-465. 

Goodwin,  John  G.,  Jr.,  and  C.  Roger 
Hungerford.  1979.  Rodent  popula- 
tion densities  and  food  habits  in 
Arizona  ponderosa  pine  forest. 
U.S.  Department  of  Agriculture, 
Forest  Service  Research  Paper  RM- 
214. 12  p.  Fort  Collins,  Co. 

Hatchell,  Glyndon  E.  1964.  Small- 
mammal  species  and  populations 
in  the  loblolly-shortleaf  pine  forest 
type  of  Louisiana.  U.S.  Forest 
Service  Research  Paper  SO-10, 12 
p.  Southern  Forest  Exp)eriment 
Station,  New  Orleans,  La. 

Kirkland,  Gordon  L.,  Jr.  1977.  Re- 
sponses of  small  mammals  to  the 
clearcutting  of  northern  Appala- 
chian forests.  Journal  of  Mam- 
malogy 58:600-609. 

Kirkland,  Gordon  L.,  Jr.  1978.  Initial 
responses  of  small  mammals  to 
clearcutting  of  Pennsylvania  hard- 
wood forests.  Proceedings  of  the 
Pennsylvania  Academy  of  Science 
52:21-23. 

McCarley,  Howard.  1958.  Ecology, 
behavior  and  population  dynam- 
ics of  Peromyscus  nuttalli  in  eastern 


377 


Texas.  Texas  Journal  of  Science 
10:147-171. 

McCarley,  W.  H.  1954.  The  ecological 
distriburion  of  Peromyscus  leucopus 
species  group  in  eastern  Texas. 
Ecology  35:375-379. 

McElfresh,  Robert  W.,  Jack  M.  Inglis, 
and  Bennett  A.  Brown.  1980.  Gray 
squirrel  usage  of  hardwood  ra- 
vines within  pine  plantations. 
Louisiana  State  University  Annual 
Forestry  SymfX)sium  19:79-89. 

Rudolph,  D.  Craig  and  James  G. 
Dickson.  Reptiles  and  amphibians 
in  streamside  management  zones 
of  clearcuts.  (In  press). 

Schmidly,  David  J.  1983.  Texas  mam- 
mals east  of  the  Balcones  Fault 
zone.  399  p.  Texas  A  &  M  Univer- 
sity Press.  College  Station. 

Seehorn,  Monte  E.  1986.  The  influ- 
ence of  silvicultural  practices  on 
fisheries  management:  effects  and 
mitigation  measures,  p.  54-63.  In 
Managing  southern  forests  for 
wildlife  and  fish — a  proceedings. 
J.  G.  Dickson  and  O.  E.  Maughan, 
eds.  USDA  Forest  Service  General 
Technical  Report  SO-65.  85  p. 
Southern  Forest  Experiment  Sta- 
tion, New  Orleans,  La. 

Warren,  R.  C.  and  G.  A.  Hurst.  1980. 
Squirrel  densities  in  pine-hard- 
wood forests  and  streamside  man- 
agement zones.  Proceedings  of  the 
Annual  Conference  of  the  South- 
east Association,  Fish  and  Wildlife 
Agencies  34:492-498. 


378 


Patterns  of  Relative  Diversity 
Within  Riparian  Snnall 
Mammal  Communities,  Platte 
River  Watershed,  Colorado^ 

Thomas  E.  Olson^  and  Fritz  L.  KnopP 


Abstract.— Relative  diversity  within  and  between 
small  mammal  assemblages  of  riparian  and  upland 
vegetation  was  evaluated  at  6  study  areas  across 
an  elevational  gradient.  In  contrast  to  avian  diversity 
analyses  conducted  at  the  same  sites,  species  rich- 
ness, relative  diversity,  and  faunal  similarity  of  small 
mammals  were  greater  among  upland  rather  than 
riparian  communities  across  the  dine.  Beta  diversity 
between  riparian  and  upland  small  mammal  com- 
munities is  greater  at  higher  elevations  within  the  wa- 
tershed. These  higher  elevation  portions  of  water- 
sheds must  be  emphasized  in  management  strate- 
gies to  conserve  regional  integrity  of  native  small 
mammal  faunas. 


Figure  1  .—Location  of  study  areas  wittiin  the  Platte  River  drainage,  norttiern  Colorado,  1981 . 


Riparian  communities  in  the  western 
states  are  mesic  vegetative  associa- 
tions occurring  along  ephemeral,  in- 
termittent, and  perennial  streams. 
Although  relatively  limited  in  area, 
these  communities  contribute  more 
biotic  diversity  within  a  region  than 
upland  vegetation  communities 
(Thomas  etal.  1979). 

Riparian  communities  have  been 
substantially  affected  by  land-use 
changes  such  as  conversion  to  agri- 
culture, grazing,  and  water  manage- 
ment (Knopf  et  al.  1988).  Further  al- 
terations in  the  western  United  States 
have  been  caused  by  the  widespread 
naturalization  of  salt  cedar  {Tamarix 
spp.)  (Horton  1977)  and  Russian- 
olive  (Elaeagnus  angustifolia)  (Olson 
and  Knopf  1986).  Because  of  the  bio- 
logical significance  and  potential  for 
perturbations  caused  by  conflicting 
land  uses,  riparian  communities  have 
been  the  focus  of  numerous  technical 
conferences  during  the  past  10  years 
(Knopf  et  al.  1988). 

An  earlier  study  of  the  pattern  of 
avian  species  diversity  in  riparian 
and  upland  study  areas  within  a  wa- 
tershed (Knopf  1985)  showed  that 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortii  America.  (Flag- 
staff .  AZ.  July  19-21.  1988.) 

'Wildlife  Biologist.  Dames  &  Moore.  175 
Cremona  Drive.  Suite  A-E.  Goleta.  Califor- 
nia 931 17. 

^Leader,  Riparian  Studies.  U.S.  Fist)  and 
Wildlife  Service.  1300  Blue  Spruce  Drive.  Fort 
Collins,  Colorado  80524-2098. 


although  more  species  of  birds  occur 
in  riparian  vegetation,  upland  sites 
contribute  more  to  avifaunal  diver- 
sity between  habitats  (beta  diversity) 
and  within  a  region  (gamma  diver- 
sity). Those  findings  were  attributed 
to  greater  similarity  among  riparian 
avifaunas  across  the  altitudinal  cline 
due  to  the  riparian  vegetation  pro- 
viding a  corridor  for  movement  of 
birds  within  a  region.  Beta  diversity 
between  upland  and  riparian  avian 
assemblages  was  greatest  at  the  up- 
per and  lower  ends  of  a  watershed, 
and  the  study  concluded  that  avifau- 


nal conservation  efforts  should  be 
concentrated  at  those  sites. 

Implications  of  the  earlier  avian 
study  to  conservation  of  small  mam- 
mal assemblages  are  unclear.  Numer- 
ous studies  (Anderson  et  al.  1980; 
Honeycutt  et  al.  1981;  Kirkland  1981) 
have  examined  small  mammal  distri- 
bution along  environmental  gradi- 
ents, but  with  a  focus  on  upland 
rather  than  riparian  species  assem- 
blages. The  objectives  of  this  study 
were  to  evaluate  diversity  within, 
and  between,  small  mammal  assem- 
blages of  riparian  and  upland  vegeta- 


379 


Figure  2.— Study  areas:  South  Platte  River  (SPR),  1200  m;  Lone  Pir>e  Creek  (LPC),  1909m;  Sheep 
Creek  (SC),  2341  m;  Illinois  River  (IR),  2S00m;  Laramie  River  (LR).  2631  m;  South  Fork  of  the  Cache 
la  Poudre  River  (SFPR),  2747  m. 


380 


tion  across  an  elevational  gradient. 
We  believed  that  the  results  would 
indicate  relative  elevations  within 
watersheds  at  which  small  mammal 
conservation  efforts  should  be  fo- 
cused. Such  efforts  could  include 
policies  of  state  and  federal  agencies 
concerning  type  of  land  use  within 
portions  of  watersheds. 

Study  Areas 

Six  study  areas  ranged  in  elevation 
from  1200  to  2747  m  within  the  Platte 
River  drainage  of  northern  Colorado 
(fig.  1).  With  the  exception  of  an  al- 
pine area,  riparian  communities 
within  each  major  life  zone  of  upland 
vegetation  along  the  Front  Range 
were  represented  (fig.  2).  Within  each 
upland,  we  located  a  riparian  site 
that  contained  a  permanent  stream. 
Cattle  grazing  had  not  occurred  on 
any  of  the  study  areas  for  at  least 
three  years  prior  to  1981. 

The  South  Platte  River  (SPR)  study 
area  was  on  the  South  Platte  Wildlife 
Management  Area,  2  km  south  of 
Crook,  Logan  County  (elevation  1200 
m).  This  community  was  dominated 
by  sand  sagebrush  mixed-prairie. 
Several  species  of  grass  and  1  woody 
species,  sand  sagebrush  (Artemisia 
filifolia),  occurred  on  the  upland 
sandhills.  Dominant  riparian  species 
were  plains  cottonwood  (Populus 
sargentii),  western  snowberry  (Sym- 
phoricarpos  ocddentalis),  and  willows 
(Salix  spp.). 

The  Lone  Pine  Creek  (LPC)  study 
area  was  11  km  west  of  Livermore, 
Larimer  County,  at  1909  m  elevation. 
This  area  of  mountain  shrub  transi- 
tion vegetation  was  dominated  by 
true  mountain  mahogany  (Cercocar- 
piis  montanus),  antelope  bitterbrush 
(Purshia  tridentata),  and  gooseberry 
(Ribes  spp.)  in  the  upland  site.  The 
riparian  site  was  dominated  by 
plains  cottonwood,  willows,  and 
common  chokecherry  (Prunus  virgini- 
am).  Rocky  Mountain  junipers 
(Juniperus  scopulorum)  were  scattered 
throughout  both  sites. 


The  Sheep  Creek  (SO  study  areas 
was  21  km  north  of  Rustic,  Larimer 
County,  at  an  elevation  of  2341  m 
(fig.  2).  Ponderosa  pine  (Pinus  ponder- 
osa)  forest,  along  with  scattered  big 
sagebrush  (Artemisia  tridentata)  domi- 
nated the  upland  site.  Riparian  vege- 
tation was  dominated  by  narrowleaf 
cottonwood  (Populus  angustifolia), 
willows,  and  alders  (Alnus  spp.). 

The  Illinois  River  (IR)  study  area 
contained  sagebrush  steppe  vegeta- 
tion and  was  within  the  Arapaho  Na- 
tional Wildlife  Refuge,  10  km  south 
of  Walden,  Jackson  County  (eleva- 
tion 2500  m).  Upland  vegetation  was 
predominantly  big  sagebrush.  The 
riparian  site  included  eight  species  of 
shrub  willows  dominated  by  S.  gey- 
eriana  (Cannon  and  Knopf  1984). 

The  Laramie  River  (LR)  study  area 
was  6.5  km  north  of  Chambers  Lake, 
Larimer  County  (elevation  2631  m). 
Aspen  (Populus  tremuloides)  domi- 
nated the  upland  site,  along  with 
Douglas  fir  (Pseudotsuga  menziesii) 
and  lodgepole  pine  (Pinus  contorta). 
The  riparian  site  was  comprised  of 
shrub  willows. 

The  highest  study  area  (elevation 
2747  m)  was  along  the  South  Fork  of 
the  Cache  la  Poudre  River  (SFPR),  at 
the  Pingree  Park  Campus  of  Colo- 
rado State  University,  Larimer 
County.  Upland  vegetation  was  com- 
px)sed  of  lodgepole  pine,  limber  pine 
(Pinus  flexilis),  Engelmann  spruce 
(Picea  engelmannii),  Douglas-fir,  sub- 
alpine  fir  (Abies  lasiocarpa),  and  a 
sparse  understory  of  aspen.  The  ri- 
parian site  was  exclusively  shrub  wil- 
lows. 


Methods 

Small  mammal  trapping  was  con- 
ducted in  1981  to  determine  the  rela- 
tive abundances  of  small  mammal 
species  at  the  6  study  areas.  In  each 
study  area,  two  400-m  survey  lines 
were  established,  including  one  ri- 
parian and  one  upland  site.  Riparian 
survey  lines  were  within  riparian 
vegetation  and  generally  paralleled 


the  stream  course.  Upland  survey 
lines  began  500  m  from  the  stream 
and  were  oriented  perpendicular  to 
the  direction  of  the  stream. 

Trap  surveys  were  conducted  be- 
tween 30  July  and  26  August  1981 . 
Survey  lines  included  20  trap  stations 
spaced  20  m  apart.  Each  trap  station 
contained  1  rat  trap  and  2  museum 
special  snap  traps  located  within  a 
1.8-m  radius  of  the  measured  point. 
Three  traps  were  used  at  each  station 
to  minimize  any  bias  in  the  data  to- 
ward more  aggressive  species,  such 
as  Peromyscus  maniculatus.  Traps 
were  baited  with  a  mixture  of 
ground  raisins,  carrots,  and  chipped 
beef,  blended  in  a  peanut  butter  base, 
and  set  for  3  consecutive  nights  in 
the  riparian  and  upland  sites  of  a 
study  area  sim.ultaneously.  Traps 
were  checked  in  the  morning  and 
evening  during  the  72  hours.  Thus, 
trap  effort  p)er  study  area  was  360 
trap-nights,  including  180  trap-nights 
each  in  the  upland  and  riparian  sites. 
Total  number  of  trap-nights  for  all 
study  areas  was  2160. 

Diversity  indices  were  calculated 
to  compare  species  diversity  within 
(alpha)  and  between  (beta)  riparian 
and  upland  sites  across  the  altitudi- 
nal  cline.  Because  preference  for  type 
of  index  varies,  we  selected  two  each 
of  the  most  commonly  used  indices 
to  measure  alpha  (Simpson  Index, 
Shannon-Weiner  Index)  and  beta  (co- 
efficient of  community,  percentage 
similarity)  diversity  (Whittaker  1975: 
95,118). 

The  former  two  differ  in  the  gen- 
eral relationship  between  output 
value  and  species  diversity.  Shan- 
non-Wiener Index  (H')  varies  directly 
with  number  of  species  trapped, 
while  the  Simpson  Index  (C)  varies 
inversely.  Coefficient  of  community 
(CO  values  are  ratios  of  the  number 
of  species  common  to  both  riparian 
and  upland  sites  to  the  total  number 
of  species  occurring  in  the  two  sites 
combined.  Those  values  are  based 
only  on  presence  or  absence  and  vary 
directly  with  diversity.  Although 
percentage  similarity  values  are 


381 


based  on  the  differences  in  impor- 
tance values  between  the  two  sites, 
they  also  vary  directly  with  diversity. 


Results 

A  total  of  471  small  mammals  of  22 
species  was  trapped  in  all  study  ar- 
eas in  1981  (table  1).  Three  species 
(14%  of  all  species  captured)  were 
trapped  in  riparian  sites  only,  9  spe- 
cies (41%)  were  trapped  in  upland 
sites  only,  and  10  species  (47%)  were 
trapped  in  both.  Nine  species  (41%) 
were  rare,  being  represented  by  2  or 
fewer  captures. 


Within-Habitat  Comparisons 

Species  composition  within  riparian 
sites  differed  among  the  study  areas. 
Deer  mice  (Peromyscus  maniculatus), 
voles  (Microtus  spp.),  and  jumping 
mice  (Zapus  princeps)  accounted  for 
182  of  189  (96%)  total  captures  at  the 
3  lower  study  areas,  although  jump- 
ing mice  did  not  occur  at  SPR.  In 
contrast,  shrews  (Sorex  spp.)  ac- 
counted for  69%  of  all  captures  at  the 
remaining,  higher  areas.  Of  68  small 
mammals  trapped  at  the  higher  sites, 
only  14  (20%)  were  either  voles  or 
jumping  mice.  No  deer  mice  were 
trapped  in  riparian  sites  at  elevations 
higher  than  2293  m. 

Changes  in  species  composition  of 
small  mammals  in  upland  sites  were 
not  distinct.  Deer  mice  were  the  most 
frequently  trapped  of  all  species  at 
the  4  intermediate  study  areas.  Over- 
all, 112  of  214  (52%)  small  mammals 
trapped  in  the  uplands  were  deer 
mice.  The  next  3  species  in  abun- 
dance (least  chipmunk  [Tamias  mini- 
mus], northern  grasshopper  mouse, 
[Onychomys  leucogasterj  and  prairie 
vole  [Microtus  ochrogaster])  accounted 
for  only  67  of  214  (31%)  total  cap- 
tures. Of  these  4  species,  only  the 
deer  mouse  was  trapped  at  all  6  sites. 

Species  richness  varied  among  ri- 
parian and  upland  sites.  The  number 
of  small  mammal  species  trapped  in 


riparian  sites  was  least  at  the  lowest 
elevation  study  area  (SPR)  and  great- 
est at  the  second  highest  study  area 
(LR)  (table  2).  All  other  riparian  sites 
were  intermediate  in  species  richness 
with  no  apparent  altitudinal  trend. 
Values  for  Simpson's  Index  (C)  (a 
measure  of  the  concentration  of 
dominance)  and  Shannon-Wiener 
Index  (HO  (Whittaker  1975:95) 
yielded  similar  results. 

The  highest  diversity  among  ripar- 
ian sites  occurred  at  LR,  which  had 
the  lowest  dominance.  The  SPR 
study  area,  which  had  a  high  C 
value,  also  contained  very  low  spe- 
cies diversity. 

The  number  of  small  mammal  spe- 
cies trapped  in  upland  sites  was 
comparatively  high  at  2  of  3  study 


areas  under  2500  m  (LPC  and  SO 
and  at  the  highest  elevation  study 
area  (SFPR)  (table  2).  Simpson's  In- 
dex values  varied  from  a  high  at  IR 
(2500  m  elevation)  to  a  low  at  SFPR 
(2747  m).  Shannon-Wiener  values  in 
upland  sites  ranged  from  a  low  of 
0.22  at  IR  to  a  high  of  0.74  at  SFPR. 

A  matrix  of  percentage  similarity 
values  (Whittaker  1975:118)  revealed 
a  mean  similarity  of  0.29  +  0.06 
among  upland  sites  and  0.18  ±  0.05 
among  riparian  sites.  These  results 
suggest  that  small  mammal  commu- 
nities in  upland  sites  were  more  simi- 
lar across  the  cline  than  were  those  in 
riparian  sites.  Overall,  beta  diversity 
along  the  altitudinal  gradient  was 
greater  (less  faunal  mixing)  in  ripar- 
ian sites. 


r 


Table  1  .—Species  of  small  mcmimals  trapped  at  6  study  areas  across  an 
altltudirKil  cline,  northern  Colorado,  1981. 

Study  arecP 


SPR 


[PC 


SC 


IR 


LR  SFPR 


Rip"  Upl^  Rip  Up!  Rip  Up!  Rip  Upl  Rip  Up!  Rip  Up! 


Sorex  cinereus 
S.  monficdus 
S.  spp. 

Sylvilagus  nutfallii 
Lepus  americanus 
Tamias  minimus 
T,  quadriviifafus 
T.  umbrinus 

Spermophilus  spilosoma 
S.  lateralis 

Thomomys  talpoides 
Dipodomys  ordii 
Reittvodontomys  megafotis 
Peromyscus  moDiculatus 
Onychomys  leucogaster 
Neotoma  mexicana 
Clefhrionomys  gapperi 
Microtus  longicaudus 
M.  ochtrogaster 
M.  spp. 

Lagurus  curfatus 
Zapus  princeps 

Totals 


1 

15 
2 


20 
10 
1 


17 


2 
10 


2 
66 


A 
2 
1 

10 


1 

42 
1 


33 
2 


1 

68 


1 

n 


7  16 


18 
2 


23  1 


1 


67 


18    83    84    39    38    17    21    20    21    31  32 


°Study  areas:  SPR  =  Soutti  Platte  River;  LPC  =  Lone  Pine  Creek;  SC  =  Stieep  Creek;  IR 
Illinois  River;  LR  =  Laramie  River;  SFPR  =  Soutti  Fork  ofCactie  la  Poudre  River. 


''Rip  =  Riparian  site. 
'=Uj^  =  Upland  site. 


382 


Between-Habrtot  Comparisons 

Species  richness  was  substantially 
higher  in  upland  sites  than  in  adja- 
cent riparian  sites  at  the  lowest  and 
highest  study  areas  (table  2).  The  val- 
ues were  similar  at  3  study  areas  of 
intermediate  elevation.  Only  at  LR 
(the  second-highest  area)  was  species 
richness  higher  in  the  riparian  site. 
At  that  study  area,  number  of  species 
trapped  in  riparian  was  greater  than 
the  upland  even  when  captures  of 


Lepus  americanus  and  Thomomys 
talpoides  were  excluded. 

Coefficient  of  community  (CC) 
values  (Whittaker  1975:118)  suggest 
that  small  mammal  communities  in 
riparian  and  adjacent  upland  sites 
were  relatively  similar  at  lower  ele- 
vations, and  became  more  dissimilar 
at  2500  m  and  higher  (table  3).  More 
species  (3)  were  common  to  both  ri- 
parian and  upland  sites  at  the  3 
lower  shidy  areas  than  at  the  higher 
areas.  Percentage  similarity  (PS)  val- 


]^^LlnZ^^''^\'^l^''^''  ^^'""P'^"  '"^^^  =  ^       relative  alpha  diversity 


Riparian 


Upland 


Study  Area 

South  Ratte  River  (SPR) 
Lone  Pine  Creek  (LPC) 
Sheep  Creek  (SC) 
Illinois  River  (IR) 
Laramie  River  (LR) 
South  Fork  of  Cache  la 
Poudre  River  (SFPR) 


Number  of 
species  (0« 


2 
6 
5 
3 
7 
3 


0.94 

0.42 

0.40 

0.45 

0.21 

0,52 


0.06 
0.46 
0.50 
0.38 
0.74 
0.33 


Number  of 
species  <0 


5 
7 
7 
3 
3 
7 


0.38 
0.67 
0.34 
0.75 
0.68 
0,21 


m 

0.53 
0.30 
0.58 
0.22 
0.26 
0.74 


s  s 
I  p2  =  X  (n,/N)2 
i=l  i=l 
s 

-  I  p,  log  p 
=1 


In^riJ^ria";;  n^^^  ^.^^^^  ^^^''^  °'  communities 


study  area 

South  Ratte  River  (SPR) 
Lone  Pine  Creek  (LPC) 
Sheep  Creek  (SC) 
Illinois  River  (IR) 
Laramie  River  (LR) 
South  Fork  of  Cache  la 
Poudre  River  (SFPR) 


No.  species  Species  common 
(riparian/upland)    to  both  sites 


2/5 
6/7 
5/7 
3/3 
7/3 
3/7 


2 
4 
3 
0 
1 

2 


0.57 

0.08 

0.62 

0.66 

0.50 

0.26 

0.00 

0.00 

0.20 

0,25 

0.40 

0.12 

°CC  (Coefficient  of  community)  =  2S ^(S^  + 
''PS  (Percentage  simiiarity)  =  min  (p^  or 


ues  indicate  the  same  trend,  with  the 
exception  of  the  lowest  study  area. 
The  low  value  at  that  study  area  is 
due  primarily  to  the  abundance  of 
Peromyscus  maniculatus  dominating 
this  calculation  (table  1). 


Discussion 

To  date,  shidies  of  small  mammal 
distribution  along  environmental 
gradients  (Anderson  et  al.  1980; 
Armstrong  et  al.  1973;  Honeyciitt  et 
al.  1981;  Kirkland  1981)  have  been 
conducted  in  upland  sites.  Knopf 
(1985)  compared  distribution  of 
breeding  birds  in  riparian  and  adja- 
cent upland  sites  within  the  6  areas 
used  in  this  study.  The  focus  of  this 
study  was  to  analyze  patterns  of 
small  mammal  faunal  similarity 
within  and  between  riparian  and  ad- 
jacent upland  sites  in  the  same  water- 
shed. Such  patterns,  although  based 
on  relatively  small  sample  sizes,  may 
indicate  elevations  along  the  gradient 
at  which  management  should  be  em- 
phasized to  conserve  regional  diver- 
sity. 

A  pronounced  change  in  species 
composition  occurred  within  riparian 
sites  at  2500  m  elevation.  The  study 
areas  below  that  elevation,  represent- 
ing foothills  and  plains,  were  domi- 
nated by  deer  mice  and  voles.  At 
2500  m  and  above,  dominance 
shifted  primarily  to  shrews.  The 
means  for  PS  values  comparing  the  3 
lower  study  areas  (0.31  ±  0.10)  and  3 
higher  study  areas  (0.43  ±  0.02)  were 
both  considerably  higher  than  the 
mean  for  all  study  areas  (0.18  +  0.05). 
Faunal  similarity  changed  as  riparian 
sites  shifted  from  cotton  wood- wil- 
low to  willow  shrub  systems.  This 
shift  in  small  mammal  community 
composition  could  have  reflected  a 
shift  from  xeric  site  willows  (S. 
amygdaloides,  S.  exigua)  to  mesic  site 
willows  as  described  in  Cannon  and 
Knopf  (1984).  Other  factors  may  have 
influenced  composition  of  small 
mammal  communities.  Among  those 
suggested  in  previous  research  are 


383 


soil  type,  nutrient  availability,  and 
vegetation  structure  (Huntley  and 
Inouye  1984,  Moulton  et  al.  1981). 
Others  have  found  specific  mi- 
crohabitat  components  to  be  impor- 
tant (cf.  M'Closkey  1981,  Szaro  and 
Belfit  1987). 

Dominance  by  deer  mice  was  par- 
ticularly obvious  at  the  lowest  site, 
SPR,  where  65  of  67  captures  were  of 
this  species.  The  remaining  2  small 
mammals  trapped  were  western  har- 
vest mice  (Reithrodontomys  megalotis). 
These  findings  were  supported  by  an 
earlier  study  of  total  small  mammal 
richness  conducted  in  the  same  study 
area.  During  the  1982  and  1983  field 
seasons  of  that  study,  98.3%  of  all 
small  mammals  captured  in  25,000 
trap-nights  were  deer  mice  and  west- 
ern harvest  mice  (Bennett  1984). 

High  numbers  of  deer  mice 
trapped  could  indicate  behavioral 
differences  (deer  mice  being  more 
aggressive),  rather  than  a  dominance 
in  absolute  numbers.  We  believe, 
however,  that  the  number  trapped 
reflected  higher  relative  abundances 
of  Peromyscus  maniculatus  for  several 
reasons.  First,  although  this  species 
was  the  most  frequently  trapped  spe- 
cies, it  dominated  only  4  of  12  total 
sites,  and  was  infrequent  to  absent  at 
7  sites  (table  1).  Total  captures  in  180 
trap-nights  at  each  of  those  4  sites 
(riparian  at  SPR,  upland  at  IR,  both 
sites  at  LPC),  ranged  from  18  to  68. 
That  is,  deer  mice  captures  ac- 
counted for  no  more  than  38  percent 
of  all  available  traps  at  any  site. 
Moreover,  in  the  riparian  site  at  SPR 
(where  deer  mice  were  most  com- 
monly caught),  the  percentages  of  all 
captures  that  were  deer  mice  were 
similar  for  this  study  (97%)  and  that 
of  Bennett  (1984)  (95%). 

Dominance  by  ecological  general- 
ists  at  the  lowest  site,  SPR,  likely  is 
explained  by  periodic  catastrophic 
events,  specifically  flooding.  In  con- 
trast to  periodic  severe  flooding  ob- 
served in  floodplains  of  the  western 
Great  Plains,  riparian  systems  at 
higher  elevations  are  not  subject  to 
severe  overbank  flooding.  During  a 


study  of  riparian  avifauna  at  SPR, 
annual  spring  flooding  varied  tre- 
mendously (Knopf  and  Sedgwick 
1988).  Maximum  mean  daily  flow  in 
1982  was  44  m^/ sec,  compared  to  405 
m^/  sec  in  1983,  when  all  of  the  ripar- 
ian zone,  as  well  as  portions  of  adja- 
cent upland  habitat  were  flooded.  No 
overbank  flooding  occurred  in  1982. 
Habitats  of  small  mammals  in  lower 
riparian  systems  are  periodically 
subjected  to  total  inundation  for  vari- 
able amounts  of  time.  Those  habitats 
appear  to  be  too  unstable  to  assure 
prolonged  survival  by  species  popu- 
lations, and  are  recolonized  by  indi- 
viduals from  the  uplands  following 
each  perturbation. 

Changes  in  small  mammal  com- 
munities among  upland  sites  were 
less  pronounced.  Faunal  similarity 
was  greatest  at  the  intermediate  sites, 
especially  LPC  (1909  m),  SC  (2293  m) 
and  IR  (2500  m).  The  mean  of  PS  val- 
ues comparing  those  sites  was  0.57  + 
0.12,  compared  to  the  overall  mean 
of  0.29  +  0.06.  Deer  mice  were  a 
dominant  species  at  all  sites  but  SPR 
(sand  sagebrush  mixed-prairie)  and 
LR  (aspen).  The  distribution  of  other 
species  appeared  to  be  influenced  by 
changes  in  upland  vegetation  types 
along  the  altitudinal  gradient.  For 
example,  northern  grasshopper  mice 
were  relatively  abundant  at  the  low- 
est site,  which  contained  grassland 
areas.  Boreal  redback  voles  (Clethri- 
onomys  gapperi)  were  similarly  abun- 
dant at  the  highest  site  in  spruce-fir. 
Neither  species  was  trapped  else- 
where. Honeycutt  et  al.  (1981)  also 
reported  that  the  distribution  of 
some  species  along  an  altitudinal 
gradient  in  Utah  was  strongly  influ- 
enced by  type  of  vegetation.  We 
(Knopf  and  Olson  1984)  have  noticed 
regional  differences  in  small  mam- 
mal communities  in  riparian  zones  of 
similar  woody  communities  but  dif- 
ferent herbaceous  composition  that 
can  be  attributed  to  variations  in  site 
dryness. 

Beta  diversity  was  low  (high  CC 
values)  at  elevations  of  less  than  2500 
m  (SPR,  LPC,  and  SC),  indicating 


that  small  mammal  communities  in 
riparian  and  adjacent  upland  sites 
were  quite  similar.  At  2500  m  (IR), 
the  CC  value  declined  to  0  (no  spe- 
cies common  to  both  sites),  then  re- 
mained low  at  the  higher  study  areas 
that  contained  aspen  and  spruce-fir 
uplands.  With  the  exception  of  an 
extremely  low  value  at  SPR  (caused 
by  the  overwhelming  dominance  of 
deer  mice  in  the  riparian  site),  PS  val- 
ues followed  the  same  pattern.  Thus, 
within  the  Platte  River  watershed, 
beta  diversity  between  riparian  and 
upland  small  mammal  communities 
is  much  greater  at  the  upper  end  of 
the  altitudinal  cline. 

These  results  differ  from  the 
avifaunal  studies  of  Knopf  (1985) 
who  found  beta  diversity  between 
riparian  and  upland  sites  to  be  great- 
est at  the  higher  and  lower  ends  of 
the  watershed,  and  upland/ riparian 
assemblages  to  be  similar  at  interme- 
diate study  areas.  Also  in  contrast  to 
Knopf's  (1985)  findings  were  greater 
relative  diversity  in,  and  faunal  simi- 
larity among,  upland  communities. 
In  support  of  the  avian  study  conclu- 
sions, however,  riparian  sites  at  the 
higher  elevations  contributed  sub- 
stantially to  small  mammal  beta  and 
gamma  (regional)  diversity. 

Implications  to  Conservation 

Historically,  management  of  riparian 
zones  has  occurred  primarily  on  ar- 
eas at  lower  elevations.  Management 
that  is  concentrated  in  a  limited  num- 
ber of  habitats  or  at  selected  eleva- 
tions may  result  in  higher  local  (al- 
pha) diversity  at  the  expense  of  beta 
and  gamma  (regional)  diversity 
(Sannson  and  Knopf  1982).  Despite 
different  beta  diversity  patterns,  our 
findings  support  the  conclusion  by 
Knopf  (1985)  that  greater  emphasis 
needs  to  be  placed  upon  conserva- 
tion of  riparian  communities  at 
higher  elevations  regionally. 

Knopf  et  al.  (1988)  have  recom- 
mended that  agencies  develop  guide- 
lines for  region  wide  rather  than  local 


384 


management  of  riparian  systems.  Re- 
spective agencies  should  realize  that 
small  mammal  communities  at 
higher  elevations  contribute  more  to 
regional  diversity  than  those  at  lovs^er 
elevations.  In  order  to  conserve  re- 
gional integrity  in  native  small  mam- 
mal faunas,  land  uses  allowed  in, 
and  adjacent  to,  high  elevation  ripar- 
ian zones  should  be  critiqued  as  care- 
fully as  those  in  lowland  floodplains. 
For  example,  livestock  grazing  can 
affect  structure  of  small  mannmal  as- 
sociations by  reducing  understory 
vegetation  (Moulton  et  al.  1981). 
Grazing  and  other  activities  that  po- 
tentially reduce  understory  vegeta- 
tion in  higher  elevation  riparian 
zones  can  seriously  affect  abun- 
dances of  certain  species  such  as 
shrews  that  are  not  present  at  lower 
sites.  The  consequences  to  regional 
diversity  of  small  mammals  would 
be  greater  than  livestock  grazing  at 
lower  elevations  because  our  find- 
ings suggest  that:  (1)  higher  elevation 
(above  2500  m)  sites  contribute  more 
to  regional  diversity  of  small  mam- 
mals; and  (2)  small  mammal  commu- 
nities in  some  lower  elevation  ripar- 
ian zones  are  composed  mostly  of 
species  populations  of  ecological  gen- 
eralists  that  are  regulated  by  cata- 
strophic, natural  perturbations. 

Acknowledgments 

We  thank  Richard  W.  Cannon,  John 
F.  Ellis,  and  Elizabeth  A.  Ernst  for 
field  and  analytical  assistance.  Ron 
Desilet  assisted  in  locating  field  sites. 
Eugene  C.  Patten  of  Arapaho  Na- 
tional Wildlife  Refuge  and  Marvin 
Gardner  of  the  South  Platte  Wildlife 
Management  Area  granted  access  to 
the  IR  and  SPR  sites,  respectively. 
The  U.S.  Forest  Service  and  Colorado 
State  University  allowed  us  to  work 
within  their  holdings.  This  research 
is  a  product  of  Cooperative  Agree- 
ment 2463-4  between  the  Colorado 
Division  of  Wildlife  and  U.S.  Fish 
and  Wildlife  Service.  This  manu- 
script was  improved  by  the  com- 


ments of  Steven  J.  Bissell,  Michael  A. 
Bogan,  Lawrence  E.  Hunt,  Mel 
Schamberger,  Robert  C.  Szaro,  Don 
E.  Wilson,  and  an  anonymous 
reviewer. 


Literature  Cited 

Anderson,  Douglas  C,  James  A. 
MacMahon,  and  Michael  L.  Wolfe. 
1980.  Herbivorous  mammals 
along  a  montane  sere:  community 
structure  and  energetics.  Journal 
of  Mammalogy  61:500-519. 

Armstrong,  David  M.,  Benjamin  H. 
Banta,  and  Edward  J.  Pokropus. 
1973.  AlHtudinal  distribution  of 
small  mammals  along  a  cross-sec- 
tional transect  through  the  Arkan- 
sas River  Watershed,  Colorado. 
Southwestern  Naturalist  17:315- 
326. 

Bennett,  Lisa.  1984.  The  initial  effects 
of  cattle  introduction  on  a  riparian 
small  mammal  community,  north- 
eastern Colorado.  M.S.  Thesis, 
Colorado  State  University,  Fort 
Collins. 

Cannon,  Richard  W.  and  Fritz  L. 
Knopf.  1984.  Sp)ecies  composition 
of  a  willow  community  relative  to 
seasonal  grazing  histories  in  Colo- 
rado. Southwestern  Naturalist 
29:234-237. 

Honeycutt,  Rodney  L.,  Michael  P. 
Moulton,  Jerry  R.  Roppe,  and 
Lynn  Fifield.  1981.  The  influence 
of  topography  and  vegetation  on 
the  distribution  of  small  mammals 
in  southwestern  Utah.  Southwest- 
ern Naturalist  26:295-300. 

Horton,  Jerome  S.  1977.  The  develop- 
ment and  perpetuation  of  the  per- 
manent tamarix  type  in  the  phrea- 
tophyte  zone  of  the  Southwest,  p. 
124-127.  In  Importance,  preserva- 
tion and  management  of  riparian 
habitat:  a  symposium.  USDA  For- 
est Service  General  Technical  Re- 
port RM-43.  Rocky  Mountain  For- 
est and  Range  Experiment  Station, 
Fort  Collins,  Colo. 

Huntley,  Nancy,  and  Richard  S. 
Inouye.  1987.  Small  mammal 


populations  of  an  old-field 
chronosequence:  successional  pat- 
terns and  associations  with  vege- 
tation. Journal  of  Mammalogy 
68:739-745. 

Kirkland,  Gordon  L.  1981.  The  zoo- 
geography of  the  mammals  of  the 
Uinta  Mountains  regions.  South- 
western NaturaHst  26:325-339. 

Knopf,  Fritz  L.  1985.  Significance  of 
riparian  vegetation  to  breeding 
birds  across  an  altitudinal  cline.  p. 
105-111.  In  Riparian  ecosystems 
and  their  management:  reconciling 
conflicting  uses.  First  North 
American  Riparian  Conference. 
USDA  Forest  Service  General 
Technical  Report  RM-120, 523  p. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Fort  Collins, 
Colo. 

Knopf,  Fritz  L.,  R.  Roy  Johnson,  Ter- 
rell Rich,  Fred  B.  Samson,  and 
Robert  C.  Szaro.  1988.  Conserva- 
tion of  riparian  ecosystems  in  the 
United  States.  Wilson  Bulletin 
100:278-284. 

Knopf,  Fritz  L.,  and  Thomas  E. 
Olson.  1984.  Naturalization  of 
Russian-olive:  implications  to 
Rocky  Mountain  wildlife.  Wildlife 
Society  Bulletin  12:289-298. 

Knopf,  Fritz  L.,  and  James  A. 

Sedgwick.  1988.  Latent  population 
responses  of  summer  birds  to  a 
catastrophic  climatological  event. 
Condor  89:869-873. 

M'Closkey,  Robert  T.  1981.  Mi- 
crohabitat  use  in  coexisting  desert 
rodents — the  role  of  population 
density.  Oecologia  50:310-315. 

Moulton,  Michael  P.,  Jerry  R.  Choate, 
and  Steven  J.  Bissell.  1987.  Small 
mammals  on  revegetated  agricul- 
tural land  in  eastern  Colorado. 
Prairie  Naturalist  13:99-104. 

Moulton,  Michael  P.,  Jerry  R.  Choate, 
Steven  J.  Bissell,  and  Robert  A. 
Nicholson.  1981.  Associations  of 
small  mammals  on  the  central 
high  plains  of  eastern  Colorado. 
Southwestern  Naturalist  26:53-57. 

Olson,  Thomas  E.,  and  Fritz  L. 
Knopf.  1986.  Naturalization  of 
Russian-olive  in  the  western 


385 


United  States.  Western  Journal  of 
Applied  Forestry  1:65-69. 

Samson,  Fred  B.,  and  Fritz  L.  Knopf. 
1982.  In  search  of  a  diversity  ethic 
for  wildlife  management.  Transac- 
tions of  the  North  American  Wild- 
life and  Natural  Resources  Confer- 
ence 47:421-431. 

Szaro,  Robert  C,  and  Scott  C.  Belfit. 
1987.  Small  mammal  use  of  a  des- 
ert riparian  island  and  its  adjacent 
scrub  habitat.  USD  A  Forest  Serv- 
ice Research  Note  RM-473, 5  p. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Fort  Collins, 
Colo. 

Thomas,  Jack  Ward,  Chris  Maser, 
and  Jon  E.  Rodiek.  1979.  Wildlife 
habitats  in  managed  rangelands — 
the  Great  Basin  in  southeastern 
Oregon.  Riparian  zones.  USDA 
Forest  Service  General  Technical 
Report  PNW-80, 18  p.  Pacific 
Northwest  Forest  and  Range  Ex- 
periment Station,  Portland,  Ore. 

Whittaker,  Robert  H.  1975.  Commu- 
nities and  ecosystems.  MacMillan 
Co.,  New  York,  385  p. 


Estimated  Carrying  Capacity 
for  Cattle  Competing  with 
Prairie  Dogs  and  Forage 
Utilization  In  Western  South 
Dakota^ 

Daniel  W.  Uresk  and  Deborah  D.  Paulson^ 


On  the  Great  Plains,  black-tailed 
prairie  dogs  (Cynomys  ludovidanus) 
compete  with  livestock  for  forage 
and  have  been  a  major  concern 
among  livestock  producers  since  the 
late  1800's  (Merriam  1902).  For  live- 
stock producers,  increased  cattle-car- 
rying capacity  on  range  land  is  the 
primary  objective  of  large-scale  prai- 
rie dog  control  programs  (Collins  et 
al.  1984).  However,  carrying  capaci- 
ties for  cattle  have  not  been  fully 
evaluated  comparing  effects  in  the 
presence  versus  the  absence  of  prai- 
rie dogs.  Carrying  capacities  for 
cattle  competing  with  prairie  dogs 
for  forage  have  historically  been  de- 
termined by  estimating  standing 
crop  of  herbage  and  then  arriving  at 
range  condition  and  estimated  carry- 
ing capacity.  Information  on  diets  of 
cattle  and  prairie  dogs,  consumption 
rates,  production  of  forage,  and  prai- 
rie dog  densities  has  never  been  col- 
lectively evaluated  to  determine  car- 
rying capacities  on  rangelands  sup- 
porting both  cattle  and  prairie  dogs. 


'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff ,  AZ,  July  19-21,  1988.) 

'Daniel  W.  Uresk  and  Debar att  D.  Paul- 
son are  Research)  Biologist  and  Wildlife  Bi- 
ologist, respectively,  at  the  Rocky  Mountain 
Forest  and  Range  Experiment  Station's  Re- 
search Work  Unit  in  Rapid  City,  SD  57701,  in 
cooperation  with  the  South  Dakota  School 
of  Mines.  Station  headquarters  is  in  Fort 
Collins,  in  cooperation  with  Colorado  State 
University. 


Abstract.— Carrying  capacities  for  cattle  compet- 
ir^g  with  black-tailed  prairie  dogs  (Cynomys  ludov'h 
cianus)vjeTe  estimated  by  a  linear  programming 
technique  for  management  of  cool-season  grasses 
in  western  South  Dakota.  Forage  utilization  was  al- 
lowed to  range  from  20%  to  80%.  Under  manage- 
ment for  cool-season  grasses  (western  wheatgrass 
(Agropyron  smifhii):  needlegrasses  iStipa  spp.)), 
stocking  rotes  of  cows  ranged  from  43  to  214  per 
hectare  over  a  6-month  grazing  season,  and  cow- 
calf  stocking  rotes  ranged  from  43  to  214  per  hec- 
tare over  a  6-month  grazing  season,  and  cow-calf 
stocking  rates  ranged  from  23  to  161 .  Needlegrasses 
and  needleleaf  sedge  (Corex  eleocharis)v/eTe  key 
forage  species. 


This  study  utilized  a  linear  pro- 
gramming approach  (GOAL)  to  de- 
termine carrying  capacities  of  cattle 
as  limited  by  prairie  dog  town  sizes 
and  forage  utilization  while  still 
maintaining  pastures  in  a  near  climax 
stage  of  mixed  perennial  cool-season 
grasses.  Cool-season  grasses  in- 
cluded western  wheatgrass  (Agropy- 
ron smithii)  and  needlegrasses  (Stipa 
spp.). 


Study  Area  and  Methods 

The  study  was  conducted  in  Conata 
Basin,  approximately  29  km  south  of 
Wall,  S.  Dak.  Average  annual  pre- 
cipitation at  the  Cedar  Pass  Visitor 
Center,  Badlands  National  Park,  ap- 
proximately 21  km  east  of  the  study 
area,  is  39.7  cm,  of  which  79%  falls 
from  April  through  September.  Av- 
erage annual  temperature  is  10°C. 
Effective  forage-year  (October  1  to 
September  30)  precipitation  for  plant 
growth  was  46.3  cm. 

Major  graminoids  of  the  study 
area  included  blue  grama  (Bouteloua 
gracilis),  buffalograss  (Buchloe  dacty- 
loides),  western  wheatgrass, 
needleleaf  sedge  (Carex  eleocharis), 
and  red  threeawn  (Aristida  longiseta). 
Common  forbs  were  scarlet 
globemallow  (Sphaeralcea  coccinea), 
Patagonia  Indianwheat  (Plantago  pat- 
agonica),  and  prairie  dogweed  (Dysso- 
dia  papposa).  Shrubs  were  snakeweed 
(Xanthocephalum  sarothrae)  and  silver 
sagebrush  (Artemisia  carta). 


The  area  is  grazed  by  cattle  and 
black-tailed  prairie  dogs.  Prairie  dogs 
graze  within  towns  and  were  active 
throughout  most  of  the  year.  Cattle 
grazed  the  entire  area  from  approxi- 
mately mid-May  to  the  last  of  Octo- 
ber. Stocking  levels  of  cattle  varied 
from  year  to  year  depending  upon 
moisture  levels  and  available  forage. 

We  applied  the  GOAL  computer 
program  to  a  resource  decision  prob- 
lem using  data  from  a  2,100-ha  pas- 
ture following  similar  procedures  by 
Bartlett  et  al.  (1976),  Bottoms  and 
Bartlett  (1975),  and  Connolly  (1974). 
Basic  data  collected  on  or  near  the 
pasture  included  cattle  diet  composi- 
tion (Uresk  1986),  black-tailed  prairie 
dog  diet  composition  (Uresk  1984), 
prairie  dog  densities  (Cincotta  1985), 
and  forage  production  (Uresk  1985). 
Forage  consumption  of  a  cow  and 
cow-calf  unit  was  estimated  as  355 
kg/month  [1  AUM  (Animal  Unit 
Month)],  and  485  kg/month  (1.32 
AUM),  respecrively  (USDA  1968). 
Forage  consumption  of  a  black-tailed 
prairie  dog  over  a  12-month  period 
was  estimated  at  10.95  kg  (Hansen 
and  Ca vender  1973).  Prairie  dog  den- 
sities were  estimated  as  44  animals/ 
ha  (Cincotta  1985). 

Serai  stages  (table  1)  were  esti- 
mated for  the  entire  pasture,  based 
on  discriminant  functions  developed 
for  canopy  cover  and  frequency  of 
occurrence  of  major  plants.  Climax 
or  near-climax  (serai  stage  A)  was 
dominated  by  western  wheatgrass; 
serai  stage  B  was  high  in  blue  grama; 


387 


while  serai  stage  C  was  high  in  buffa- 
lograss.  Range  serai  stage  D  con- 
sisted of  approximately  equal  but 
smaller  amounts  of  all  three  plant 
species.  Estimates  of  forage  produc- 
tion and  area  occupied  by  prairie  dog 
towns  were  specified  separately  for 
each  range  serai  stage  in  the  analysis. 

In  the  analysis,  forage  utilization 
was  varied  for  the  entire  pasture  at 
four  levels  (20%,  40%,  60%  and  80%) 
when  both  cattle  and  prairie  dogs 
were  grazing.  Prairie  dog  towns 
were  allocated  to  serai  stages  B,  C, 
and  D;  but  not  to  range  condition 
class  A  because  prairie  dogs  do  not 
occur  in  or  near  climax  vegetation. 
Prairie  dogs  were  confined  to  areas 
that  totalled  from  20  to  40  ha  for  the 
entire  pasture.  Forage  utilization  on 
these  areas  was  adjusted  to  100%. 

Major  forage  plants  of  both  herbi- 
vores included  western  wheatgrass 
blue  grama,  buffalograss,  needleleaf 
sedge,  sand  dropseed  ( Sporobolus 
cryptandrus),  needlegrasses,  scarlet 
globemallow  and  categories  of  other 
graminoids,  and  other  forbs  (Uresk 
1984,  Uresk  1986).  Shrubs  were  ex- 
cluded because  they  were  minor 
components  of  the  diets  and  range- 
land.  Average  herbivore  diets  for  the 
season  were  used  in  this  linear  pro- 
gramming analysis. 

With  linear  programming,  man- 
agement options  for  amounts  of  for- 
age utilization  and  area  occupied  by 
prairie  dog  towns  were  analyzed  un- 
der management  for  cool-season 
grasses.  Under  management  for  cool- 
season  grasses,  no  forage  species  was 
utilized  over  the  selected  percent- 
ages. 

For  the  GOAL  programming 
analysis,  the  following  assumptions 
were  made: 

1.  Adequate  forage  of  major 
plant  species  were  available 
within  limits  of  prescribed 
utilization  so  that  herbivores 
did  not  adjust  their  normal 
diets  and  consumption  in  re- 
sponse to  a  decrease  in  for- 
age. 


2.  Common  use  of  the  range  by 
the  two  herbivores  did  not 
alter  the  preference  for  for- 
age within  established  utili- 
zation limits. 

3.  Forage  consumption  was 
proportional  to  population 
densities  of  the  herbivore 
species. 

Cattle  stocking  numbers  were  esti- 
mated as  follows.  Diet  composition 
and  forage  consumption  rates  of  both 
herbivores  were  specified  and  held 
constant.  Forage  availability  was 
specified  for  each  species  by  serai 
stage  and  held  constant.  Prairie  dog 
density  per  hectare  of  town  was 
specified  and  held  constant.  The 
management  variables — percent  for- 
age utilization  and  hectares  in  dog 
towns — were  varied  within  specified 
limits. 

Finally,  the  GOAL  program  solved 
cattle-stocking  numbers  that  could  be 
supported  by  the  available  forage  for 
a  given  forage  utilization  {percentage 
and  hectares  in  prairie  dog  towns. 
When  present,  prairie  dogs  were 
given  first  priority  for  forage. 


Results 


Plant  Production 

Forage  production  for  individual 
species  was  greatest  for  western 
wheatgrass,  followed  by  buffalograss 
and  blue  grama  (table  1).  The  pasture 
at  or  near  climax  serai  stage  (A)  had 
the  lowest  plant  production  (1970 
kg/ha);  serai  stage  C  had  the  greatest 
overall  production  (2267  kg/ha). 
Most  of  the  pasture  was  at  or  near 
cHmax  serai  stage  A  (58%)  and  did 
not  have  prairie  dogs,  a  factor  that 
results  in  a  relatively  low  impact  by 
prairie  dogs.  Serai  stages  B,  C,  and  D 
made  up  3%,  7%,  and  32%,  respec- 
tively, of  the  pasture.  All  had  prairie 
dogs  residing. 


Carrying  Capacity 

Carrying  capacity  for  mature  cows 
without  calves  (6-month  grazing  pe- 
riod) on  range  with  no  prairie  dogs 
competing  ranged  from  55  to  221 
cows/2100  ha  when  forage  utiliza- 
tion levels  were  from  20%  to  80% 


Table  1.— Estimated  peak  plant  production  (kg/ha)  (Uresk  1985)  by  range 
class  on  a  2, 1 00-ha  pasture. 


Range  serai  stages'  (ha) 


A 

B 

c 

D 

Plant  taxa 

(1226) 

(55) 

(144) 

(675) 

Western  wheatgrass  (Agropyron  smifhii) 

1354 

514 

72 

301 

Blue  grama  (Boufeloua  gracilis) 

204 

441 

396 

88 

Buffalograss  (BucNoe  dacfyloides) 

47 

601 

1172 

192 

Needleleaf  sedge  (Carex  eleocharis) 

9 

12 

43 

38 

Needlegrasses  (Sf/pa  spp.) 

32 

44 

0 

55 

Sand  dropseed  (Sporobolus  cryptandrus) 

0 

1 

5 

48 

Other  graminoids 

138 

180 

253 

372 

Total  graminoids 

1784 

1793 

1941 

1094 

Scarlet  globemallow  (Sphaera/ceo  cocclnea) 

36 

36 

47 

96 

Total  forbs 

150 

388 

279 

1046 

Total  production^ 

1970 

2217 

2267 

2236 

'A  =  climax;  D  =  low  serai  stage.  Uresk,  D.  W.  submitted.  A  quantitative  methiod  for 
estimating  ecological  stages  in  a  mixed-grass  prairie  witt)  multivariate  tectiniques.  J. 
Range  t^anage. 

'Shrubs  ore  not  included  in  production  estimates. 


388 


(table  2).  In  estimating  stocking  rates, 
no  single  forage  species  was  allowed 
to  be  utilized  at  levels  greater  than 
the  set  levels  from  20%  to  80%.  Thus, 
a  range  of  1.6  to  6.4  ha/AUM  was 
required.  Numbers  of  cows  de- 
creased as  hectares  of  prairie  dog 
towns  increased;  stocking  rates  de- 
creased by  approximately  3  for  every 
additional  20  ha  of  prairie  dogs  (880 
prairie  dogs  or  293  prairie  dogs/ 
cow)  up  to  40  ha  on  the  pasture. 

Cow-calf  stocking  rates  ranged 
from  40/2,100  ha  to  161  (1  cow-calf 
unit  =  7.92  AUMs  for  6-months) 
when  utilization  levels  varied  from 
20%  to  80%  without  prairie  dogs 
(table  2).  At  these  stocking  rates,  ap- 
proximately 2.1  to  8.7  ha  were  re- 
quired for  each  AUM.  Stocking  rates 
decreased  by  approximately  2  cow- 
calf  units  for  every  additional  20  ha 
of  prairie  dogs. 


Discussion 

Needlegrasses  and  needleleaf  sedge 
limited  carrying  capacity  for  cattle  on 
pastures  managed  for  cool-season 
grasses.  Western  wheatgrass  was 
never  a  limiting  species;  that  is,  con- 
sumption of  western  wheatgrass  by 
both  herbivores  never  exceeded  the 
amount  available.  The  80%  level  of 


utilization  of  some  cool-season 
grasses  is  too  high  to  maintain  the 
viability  of  these  plants,  and  lower 
utilization  levels  (30-45%)  are  recom- 
mended (Lewis  et  al.  1956).  With 
fewer  cattle  grazing  under  manage- 
ment for  cool-season  grasses,  cattle 
gain  more  weight  per  day,  but  fewer 
kilograms  per  hectare  (Black  et  al. 
1937,  Lewis  et  al.  1956,  Bement  1969). 

Prairie  dog  expansion  can  be  re- 
duced under  management  for  cool- 
season  grasses  because  vertical  cover 
and  grass  heights  increase  (Cincotta 
1985).  Prairie  dogs  did  not  signifi- 
cantly expand  over  a  4-year  period 
on  areas  where  cattle  were  excluded 
(Uresk  et  al.  1982).  Furthermore,  a 
lower  stocking  rate  (management  for 
cool-season  grasses)  would  increase 
vertical  grass  cover  on  the  range  and 
would  thereby  further  reduce  prairie 
dog  expansion.  Snell  and  Hlavachick 
(1980)  and  Snell  (1985)  reported  re- 
duced expansion  rates  and  elinuna- 
tion  of  prairie  dog  colonies  by  using 
a  summer-deferred  grazing  system. 
Prairie  dogs  prefer  habitat  managed 
for  warm-season  grasses  [blue  grama 
(Bouteloua  gracilis),  buffalograss 
(Buchloe  dactyloides)].  Increased  stock- 
ing rates  of  cattle  and  shortgrass  stat- 
ure with  low  vertical  cover  allows  for 
prairie  dog  expansion  (Uresk  et  al. 
1982,  Cincotta  1985). 


r 


Table  2. —Estimated  6-month  carrying  capacity  for  mature  cows  with  and 
without  calves  with  management  for  cool-season  grasses.  Stocking  rates 
are  related  to  tiectares  of  prairie  dogs  and  allowable  forage  utilization  on  a 
2,100-ha  pasture  In  westem  South  Dakota.  Consumption  of  needlegrasses 
and  needleleaf  sedge  Is  1 00%  on  prairie  dog  occupied  areas. 


Forage 


Prairie  dogs  occupied  areas  (ha) 


utilization  % 

0 

20 

40 

0 

20  40 

Cow  numbers^ 

Cow-calf  numbers' 

20 

55 

53 

50 

40 

39  37 

40 

110 

108 

105 

81 

79  77 

60 

166 

163 

160 

121 

119  117 

80 

221 

218 

214 

161 

159  157 

'355  kg  of  forage  consumed/cow/monfh  (1  AUM). 

'485  kg  of  forage  cor\sumed/cow-calf/monfh  (1.32  AUM). 


Cattle  stocking  rates  estimated  in 
this  study  were  conservative,  be- 
cause upper  limits  of  forage  con- 
sumption and  prairie  dog  densities 
(44  animals/ ha)  were  used  in  the 
analyses.  The  guidelines  reported 
here  for  cow  or  cow-calf  stocking 
rates  for  cool-season  grasses  repre- 
sent viable  options  for  management. 
Key  forage  species  used  to  estimate 
cattle  numbers  and  monitor  utiliza- 
tion for  management  of  cool-season 
grasses  included  needlegrasses  and 
needleleaf  sedge.  Generally,  stocking 
rates  were  limited  by  production  and 
use  of  needlegrasses,  although 
needleleaf  sedge  and  sand  dropseed 
also  influenced  cow  numbers.  When 
hectares  of  prairie  dogs  are  high, 
needleleaf  sedge  can  become  the  ma- 
jor limiting  factor  in  determining 
cow  numbers.  Needlegrasses  were 
generally  the  limiting  plant  compo- 
nent in  determining  cow-calf  units. 
Sand  dropseed  can  be  limiting  when 
the  area  with  prairie  dogs  is  greater 
than  or  equal  to  200  ha. 

This  study  only  presents  estimates 
for  up  to  40  ha  of  prairie  dog  colonies 
(approximate  current  levels  of  prairie 
dogs)  on  a  2,100-ha  pasture,  and  lim- 
ited extrapolation  is  suggested  be- 
yond data  in  table  2.  An  additional 
constraint  is  availability  of  needle- 
grasses and  needleleaf  sedge.  Ex- 
trapolation of  results  to  pastures 
with  lower  availability  of  these  spe- 
cies should  be  done  cautiously.  In 
fact,  where  forage  availability  and 
composition  are  much  different  from 
the  pasture  studies,  extreme  care 
should  be  used  in  extrapolating  re- 
sults to  other  areas.  The  assumptions 
and  required  constraints  for  GOAL 
linear  program  analysis  imposes 
some  limitations  on  biological  sensi- 
tivity. 


Acknowledgments 

Appreciation  is  extended  to  Ne- 
braska National  Forest  for  providing 
study  areas.  Partial  funding  of  this 
study  was  provided  by  National  Ag- 


389 


ricultural  Pesticide  Impact  Assess- 
ment Program  (NAPIAP)  and  Ne- 
braska National  Forest. 


Literature  Cited 

Bartlett,  E.  Thomas,  Kenneth  E.  Bot- 
toms, and  Randal  P.  Pope.  1976. 
GOAL-multiple  objective  pro- 
gramming. Colorado  State  Univer- 
sity, Ft.  Collins,  CO.  157  p. 

Bement,  R.  E.  1969.  A  stocking-rate 
guide  for  beef  production  on  blue 
grama  range.  J.  Range  Manage. 
22:83-86. 

Black,  W.  H.,  A.  L.  Baker,  V.  I.  Clark, 
and  O.  R.  Mathews.  1937.  Effect  of 
different  methods  of  grazing  on 
native  vegetation  and  gains  of 
steers  in  northern  Great  Plains. 
U.S.  Dept.  Agr.  Forest  Service, 
Rapid  City,  SD  Tech.  Bull.  547. 18 

P- 

Bottoms,  Kenneth  E.,  and  E.  T. 
Bartlett.  1975.  Resource  allocation 
through  GOAL  programming.  J. 
Range  Manage.  28:442-447. 

Cincotta,  Richard  P.  1985.  Habitat 
and  dispersal  of  black-tailed  prai- 
rie dogs  in  Badlands  National 
Park.  Ph.D.  Diss.  Colo.  State 
Univ.,  Ft.  Collins,  CO. 

Collins,  Alan  R.,  John  P.  Workman, 
and  Daniel  W.  Uresk.  1984.  An 
economic  analysis  of  black-tailed 
prairie  dog  (Cynomys  ludovicianus) 
control.  J.  Range  Manage.  35:358- 
361. 

Connolly,  J.  1974.  Linear  program- 
ming and  the  optimum  carrying 
capacity  of  range  under  connmon 
use.  J.  Agric.  Sci.  83:259-265. 

Hansen,  Richard  M.,  and  Barbara  R. 
Ca vender.  1973.  Food  intake  and 
digestion  by  black-tailed  prairie 
dogs  under  laboratory  conditions. 
Acta  Theologica  18:191-200. 

Lewis,  James  K.,  George  M.  Van 
Dyne,  Leslie  R.  Albee,  and  Frank 
W.  Whetzal.  1956.  Intensity  of 
grazing:  its  effect  on  livestock  and 
forage  production.  South  Dakota 
Agr.  Expt.  Sta.  Brookings  Bull. 
459. 44  p. 


Merriam,  C.  Hart.  1902.  The  prairie 
dog  of  the  Great  Plains,  p.  257-270. 
In  Yearbook  of  the  United  States 
Department  of  Agriculture,  Wash- 
ington, DC.  U.S.  Gov.  Print.  Off. 

Snell,  Glen  P.  1985.  Results  of  control 
of  prairie  dogs.  Rangelands  7:30. 

Snell,  Glen  P.,  and  Bill  D.  Hlavachick. 
1980.  Control  of  prairie  dogs — the 
easy  way.  Rangelands  2:239-240. 

Uresk,  Daniel  W.  1984.  Black-tailed 
prairie  dog  food  habits  and  forage 
relationships  in  western  South 
Dakota.  J.  Range  Manage.  37:325- 
329. 

Uresk,  Daniel  W.  1985.  Effects  of  con- 
trolling black-tailed  prairie  dogs 
on  plant  production.  J.  Range 
Manage.  38:466-468. 

Uresk,  Daniel  W.  1986.  Food  habits 
of  cattle  on  the  South  Dakota  High 
Plains.  Prairie  Nat.  18:211-218. 

Uresk,  Daniel  W.,  James  G.  Mac- 
Cracken,  and  Ardell  J.  Bjugstad. 
1982.  Prairie  dog  density  and 
cattle  grazing  relationships,  p.  199- 
201.  In  R.  M.  Timm  and  R.  J. 
Johnson,  eds.  Fifth  Great  Plains 
Wildlife  Damage  Control  Work- 
shop Proc.  October  13-15, 1981. 
Univ.  Neb.,  Lincoln. 

USDA  Forest  Service.  1968.  Range 
Environmental  Analysis  Hand- 
book. U.S.  Dept.  of  Agr.  Forest 
Service  Handbook  2209.21.  244  p. 


390 


Cattle  Grazing  and  Small 
Mammals  on  the  Sheldon 
National  Wildlife  Refuge, 
Nevada^ 

John  L.  Oldemeyer^  and  Lydia  R.  Allen- 
Johnson^ 


Abstract.— We  studied  effects  of  cattle  grazing  on 
small  mamnnal  microhabitat  and  abundance  in 
northwestern  Nevada.  Abundance,  diversity,  and 
microhabitat  were  compared  between  a  375-ha 
cattle  exclosure  and  a  deferred-rotation  grazing  al- 
lotment which  hod  a  three-year  history  of  light  to 
moderate  use.  No  consistent  differences  were  found 
in  abundance,  diversity,  or  microhabitat  between 
the  two  areas. 


Grazing  by  livestock  is  a  common 
and  economically  important  practice 
throughout  much  of  the  western 
United  States.  Because  grazing  alters 
wildlife  habitat,  much  attention  has 
centered  on  its  impact  on  wildlife 
abundance,  diversity,  and  habitat 
use.  However,  relatively  little  infor- 
mation exists  on  effects  of  grazing  on 
small  mammal  communities.  Such 
information  would  aid  development 
of  effective  grazing  progran\s  where 
small  mammals  are  a  management 
concern. 

Several  authors  have  demon- 
strated that  removal  or  alteration  of 
cover  can  cause  changes  in  small 
mammal  communities  (Bimey  et  al. 
1976,  Geier  and  Best  1980,  Grant  et 
al.  1982,  LoBue  and  Darnell  1959). 
More  specifically,  grazing  altered  ro- 
dent species  diversity  through 
changes  in  plant  species  diversity  on 
several  habitats  in  northeastern  Cali- 
fornia (Hanley  and  Page  1982).  Simi- 
larly, Grant  et  al.  (1982)  found  differ- 
ential changes  in  several  small  mam- 
mal community  parameters  between 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortt)  America.  (Flag- 
staff, AZ,  July  19-21.  1988.) 

'Leader,  Ecology  and  Sysfematics  Sec- 
tion, National  Ecology  Research)  Center. 
U.S.  Rsh  and  Wildlife  Sen/ice,  1300  Blue 
Spruce  Drive,  Fort  Collins,  CO  80524. 

^Biological  technician.  Ecology  and  Sys- 
fematics Section,  National  Ecology  Re- 
search Center.  U.S.  Hsh  and  Wildlife  Sen/ice, 
1300  Blue  Spruce  Drive.  Fort  Collins,  CO 
80524. 


grazed  and  ungrazed  sites  in  four 
western  grassland  communities;  tall- 
grass  and  montane  grasslands  ap- 
peared to  be  most  affected  by  graz- 
ing. 

In  assessing  grazing  impacts  on 
small  mammal  communities,  Hanley 
and  Page  (1982)  stressed  the  impor- 
tance of  evaluating  effects  on  a  habi- 
tat-type basis.  Grant  et  al.  (1982)  con- 
cluded that  the  response  of  a  small 
mammal  community  to  grazing  de- 
pended on  the  site  and  the  original 
mammal  species  composition. 

In  1980,  the  Sheldon  National 
Wildlife  Refuge  (SNWR)  initiated  a 
deferred-rotation  grazing  system  on 
the  6,954-ha  Badger  Mountain  graz- 
ing allotment  to  improve  soil  and 
range  conditions.  The  management 
plan  was  designed  to  graze  1,444  ani- 
mal-unit-months (AUMs)  with  the 
grazing  period  alternating  between 
mid-June  through  early  August  dur- 
ing one  year  and  early  August 
through  late  October  the  next  (five- 
year  average,  David  Franzen,  Range 
Conservationist,  SNWR,  pers. 
comm.).  Prior  to  1979,  the  allotment 
had  been  on  a  season-long  grazing 
system  from  early  April  through  Sep- 
tember with  an  estimated  1,700 
AUMs  being  removed  from  the  unit 
(U.S.  Fish  and  Wildlife  Service  1980). 

In  Spring  1981,  we  constructed  a 
375-ha  cattle  exclosure  on  the  Badger 
Mountain  allotment  to  evaluate  the 
effects  of  cattle  grazing  on  wildlife 
and  their  habitat  (Oldemeyer  et  al. 
1983).  The  purpose  of  this  element  of 


the  study  was  to  evaluate  the  effect 
of  the  grazing  system  on  small  mam- 
mals. Specifically,  we  wanted  to  de- 
termine the  following:  (1)  is  there  a 
difference  in  small  mammal  abun- 
dance and  diversity  between  the  ar- 
eas over  time,  (2)  is  there  a  difference 
in  the  available  small  mammal  habi- 
tat between  areas,  and  (3)  what  mi- 
crohabitat characteristics  are  indica- 
tive of  capture  sites  by  individual 
small  mammal  species  for  the  two 
ecosites?  We  tested  the  null  hypothe- 
sis of  no  significant  difference  be- 
tween the  exclosure  and  the  allot- 
ment. 


Study  Area  and  Mettiods 

The  Badger  Mountain  allotment 
ranges  from  1,890-2,152  m  elevation 
and  is  composed  of  two  dominant 
range  ecosites  (Anderson  1978).  The 
shrubby  rolling  hills  (SRH)  ecosite 
occurs  on  moderate  to  deep  soils  and 
is  dominated  by  big  sagebrush  (Ar- 
temisia tridentata)  and  antelope 
bitterbrush  (Purshia  tridentata)  with 
grass  understory  dominated  by 
Idaho  fescue  (Festuca  idahoensis).  The 
mahogany  rockland  (MR)  ecosite  oc- 
curs on  rocky  ridges  and  slopes  with 
bedrock  outcrops.  Curlleaf  mountain 
mahogany  (Cercocarpus  ledifolius)  is 
predominate  in  this  ecosite  with  a 
grass  understory  dominated  by  west- 
ern needlegrass  (Stipa  occidentalis) 
(fig.  1).  Precipitation  on  Badger 
Mountain  ranges  from  27-33  cm  an- 


391 


nually  with  most  coming  as  snow 
and  as  spring  and  autumn  rains  (U.S. 
Fish  and  Wildlife  Service  1980). 

We  conducted  the  study  during 
the  summers  of  1983  and  1984,  four 
and  five  years,  respectively,  after  ini- 
tiation of  the  deferred-rotation  graz- 
ing system.  Grazing  intensities  were 
1,650  AUMs  from  10  July  to  10  Au- 
gust, 1980, 1,770  AUMs  from  7  Au- 
gust to  30  September,  1981,  and  1,036 
AUMs  from  24  June  to  22  August, 
1982.  In  1983,  cattle  were  grazed  on 
the  allotment  from  1  August  through 
15  October  at  a  rate  of  980  AUMs. 
The  following  year,  the  unit  sup- 
ported 1,337  AUMs  during  a  28  June 
to  18  August  grazing  period  (David 
Franzen,  Range  Conservationist, 
SNWR,  pers.  comm.). 

In  1983,  eight  live  trap  grids  were 
established  with  trap  stations  15  m 
apart.  Four  grids  were  located  inside 
the  exclosure  and  four  were  located 
in  the  allotment.  We  arranged  each  7 
X  7  grid  so  that  approximately  half  of 
the  traps  were  in  the  SRH  ecosite  and 
half  were  in  the  MR  ecosite.  We 
sampled  only  four  grids  (two  in  the 
exclosure  and  two  in  the  allotment) 
in  1984,  but  we  increased  the  size  of 
the  grids  to  64  (8  X  8)  trap  stations. 

We  trapped  from  1  July  through 
11  August  in  1983,  and  19  June 
through  1  July  in  1984.  Only  one  pair 
of  grids  were  trapp>ed  at  a  time  (one 
grid  in  the  exclosure,  one  in  the  allot- 
ment), for  a  total  of  four  trap  sessions 
in  1983,  and  two  sessions  in  1984.  A 
Sherman  live  trap  containing  a  hand- 
ful of  cotton  wool  and  baited  with 
rolled  oats  was  placed  at  each  sta- 
tion. Trapping  began  in  the  afternoon 
and  continued  for  five  consecutive 
days.  Traps  were  opened  each  day 
between  1600-1730  hrs  and  closed  the 
following  morning  between  0730- 
11(X)  hrs  to  prevent  daytime  trap 
mortality.  Species,  trap  number,  age 
(adult  or  juvenile),  sex,  weight  and 
tag  number  were  recorded.  We  used 
toe  clips  or  aluminum  ear  tags  to 
identify  individuals. 

We  estimated  relative  abundance 
of  small  mammals  as  the  total  num- 


ber of  individuals  captured  per  trap 
night  (catch /effort)  for  each  ecosite 
type,  area  and  grid.  Abundance  was 
calculated  for  all  small  mammals  as 
well  as  for  each  individual  sp>ecies. 

Small  mammal  diversity  was  de- 
rived for  each  area  using  Patil  and 
Taillie's  (1979)  diversity  profiles.  This 
is  a  graphic  ordering  of  the  diversity 
of  two  or  more  communities.  The  y- 
axis  represents  the  percent  of  small 
mammals  remaining  in  the  sampled 
population  when  a  species  is  re- 
moved. This  is  plotted  against  the 
number  of  species  that  have  been 
removed  from  the  sampled  popula- 
tion, with  species  removal  being  cu- 
mulative. 

The  profile  of  an  intrinsically  more 
diverse  community  will  plot  above 
that  of  a  less  diverse  community.  If 
profile  lines  intersect,  then  the  com- 
munities do  not  differ  in  diversity. 

Vegetation  measurements  describ- 
ing microhabitat  structure  were 
taken  at  each  station  prior  to  trap- 
ping. The  characteristics  we  meas- 
ured are  similar  to  those  reported  in 
other  small  mammal  studies  (e.g. 
Geier  and  Best  1980,  Hallett  1982). 
These  included: 


1.  Percent  canopy  cover  of 
grass,  forbs,  and  litter  (all 
downed  dead  material;  e.g. 
twigs,  dead  grass,  leaves)  in 
a  1.0  X  0.5  m  quadrat  having 
the  trap  station  stake  as  its 
center; 

2.  Height  (cm)  of  the  nearest 
shrub  (crown  foliage  >2  dm 
in  diameter)  in  each  quarter 
around  the  trap  station  stake; 

3.  Line  intercept  distance  (cm) 
of  living  and  dead  shrubs  (in 
the  25  to  50  cm  layer  above 
the  ground)  occurring  within 
two  perpendicularly  oriented 
2-m  transects  centered  at  the 
trap  station  stake. 

Five  microhabitat  variables  were 
derived  from  these  measurements 
for  analysis.  These  included:  (1)  % 
forb  cover,  (2)  %  grass  cover,  (3)  % 
litter  cover,  (4)  total  shrub  intercep- 
tion distance  (cm),  and  (5)  mean 
height  (cm)  of  the  live  shrubs  around 
each  stake. 

Small  mammal  abundance  data 
were  analyzed  using  a  three-way 
analysis  of  variance  to  determine  if 


Figure  1  .—View  from  the  study  site  on  Badger  Mountain,  Sheldon  National  Wildlife  Refuge. 
Nevada. 


392 


small  mammal  abundance  differed 
between  areas,  years,  and  ecosites. 
We  used  a  one-way  analysis  of  vari- 
ance test  to  detect  differences  be- 
tween areas  for  individual  years  and 
ecosites.  To  determine  the  microhabi- 
tat  preferences  of  individual  species 
we  coded  trap  locations  as  being  ei- 
ther capture  or  non-capture  stations. 
We  employed  a  nested  two-way 
analysis  of  variance  to  test  these  pref- 
erences among  areas  and  codes,  the 
interaction  of  areas  by  codes,  and  the 
nested  interaction  of  grids  within  ar- 
eas. We  considered  P<=0.1  to  be  sig- 
nificant. Subsequent  discussion  of 
small  mammal  microhabitat  selection 
concerns  only  the  two  most  abun- 
dant species,  the  deer  mouse  (Pero- 
myscus  maniculatus)  and  the  least 
chipmunk  (Tamias  minimus). 


Results  and  Discussion 

Species  Composition 

Species  of  small  mammals  occurring 
in  the  two  ecosites  of  our  study  area 
are  widely  distributed  throughout 


the  Great  Basin  (Hall  1946).  These 
species  and  their  percentage  of  the 
total  catch  were:  deer  mouse  (46.7%), 
least  chipmunk  (29.8%),  Great  Basin 
pocket  mouse  (Perognathus  parvus) 
(12.3%),  sagebrush  vole  (Lagurus  cur- 
tatus)  (7.8%),  Townsend's  ground 
squirrel  (Spermophilus  townsendii) 
(1.2%),  golden-mantled  ground 
squirrel  (Spermophilus  lateralis) 
(1.2%),  and  long-tailed  vole  (Microtus 
longicaudis)  (0.6%). 


Abundance 

Total  relative  abundance  of  small 
mammals  did  not  differ  between 
year  or  area  (table  1).  However,  more 
animals  were  captured  in  the  SRH 
ecosite  than  in  the  MR  ecosite 
(P=0.05). 

There  was  a  general  decline  in 
deer  mouse  (P=0.08)  and  least  chip- 
munk (P=0.06)  abundance  from  1983 
to  1984,  although  this  probably  re- 
flects the  difference  in  season  and 
length  of  trapping  between  the  two 
years.  We  found  no  significant  differ- 
ence in  abundance  for  these  two  spe- 


cies between  areas  or  ecosites.  This  is 
not  surprising  given  the  opportunis- 
tic, adaptable,  nature  of  these  small 
mammals.  Others  have  found  that 
heavy  grazing  in  big  sagebrush  habi- 
tat appears  to  promote  an  increase  in 
deer  mice  (Black  and  Frischknecht 
1971,  Larrison  and  Johnson  1973), 
and  least  chipmunk  numbers  (Larri- 
son and  Johnson  1973).  Hanley  and 
Page  (1982)  observed  a  different  re- 
sponse for  the  two  species  on  their 
big  sagebrush-Idaho  fescue  site  60-80 
km  west  of  Badger  Mountain.  In  that 
study,  deer  mice  were  captured  in 
the  same  numbers  in  both  grazed 
and  ungrazed  sites,  while  least  chip- 
munks were  four  times  more  abun- 
dant in  the  grazed  site  than  in  the 
ungrazed  site. 

Great  Basin  pocket  mice  were 
more  abundant  (P<0.01)  in  1983  than 
1984,  and  they  were  more  commonly 
captured  in  the  SRH  ecosite  than  in 
the  MR  ecosite  (P=0.08).  However, 
there  was  no  significant  difference  in 
abundance  between  the  areas.  Others 
have  found  Great  Basin  pocket  mice 
to  be  more  abundant  on  ungrazed 
big  sagebrush  sites  (Black  and  Fris- 


Table  1  .—Abundance  of  small  mammals  (number  caught  per  trap  night)  by  year,  area  and  ecosite  on  the  Sheldon 
National  Wildlife  Refuge,  1 983-84. 


Species 

Area 

Shrubby-Rolling  Hills 

Mohogany  Rocklands 

1983 
#/trapnite(S.E.) 

1984 
#Arapnlte(S.E.) 

1983 
#/trapnite(S.E.) 

1984 
#/trapnlte(S.E) 

Deer  mouse 

Excl. 

0.081(0.012) 

0.047(0.014) 

0.061(0.019) 

0.050(0.011) 

Allot. 

0.063(0.005) 

0.052(0.017) 

0.064(0.027) 

0.016(0.002) 

Least  chipmunk 

Excl. 

0.029(0.007) 

0.020(0.005) 

0.067(0.017) 

0.037(0.006) 

Allot. 

0.046(0.025) 

0.031(0.010) 

0.049(0.011) 

0.005(0.005) 

Great  Basin 

Excl. 

0.013(0.005) 

0.028(0.003) 

0.003(0.003) 

0.029(0.014) 

pocket  mouse 

Allot. 

0.011(0.007) 

0.031(0.004) 

0.005(0.003) 

0.014(0,006) 

Sagebrush  vole 

Excl. 

0.019(0.010) 

0.038(0.020) 

0.002(0.002) 

0.004(0.004) 

Allot. 

0.004(0.003) 

0.019(0.004) 

0 

0 

Long-tailed  vole 

Excl. 

0 

0.004(0.004) 

0 

0 

Allot. 

0 

0.009(0.002) 

0 

0 

Townsend's  ground 

Excl. 

0 

0 

0 

0 

squirrel 

Allot. 

0.005(0.005) 

0 

0 

0.005(0.005) 

Golden-mantled 

Excl. 

0.006(0.006) 

0 

0.007(0.007) 

0 

ground  squirrel 

Allot. 

0 

0 

0 

0 

Total  Catch 

Excl. 

0.149(0.013) 

0.135(0.027) 

0.141(0.024) 

0.120(0.028) 

Allot. 

0.154(0.030) 

0.143(0.005) 

0.119(0.026) 

0.042(0.018) 

393 


chknecht  1971),  or  more  abundant  on 
grazed  sagebrush  sites  (Hanley  and 
Page  (1982). 

Relative  abundance  of  the  sage- 
brush voles  and  long-tailed  voles 
could  not  be  compared  statistically 
because  of  the  small  number  of  voles 
captured.  There  was,  however,  a 
general  trend  for  microtine  rodents 
to  be  more  abundant  in  the  SRH 
ecosite  even  though  grass  and  forb 
cover  in  the  MR  ecosite  were  higher. 
Birney  et  al.  (1976)  and  Grant  et  al. 
(1982)  have  discussed  the  importance 
of  cover  for  microtine  rodents  in 
grasslands.  Although  grass  cover 
was  lower  in  the  SRH  ecosite,  the 
combination  of  higher  litter  cover 
and  shrub  intercept  in  that  ecosite 
may  provide  better  habitat  for  these 
rodents.  The  sagebrush  vole  was 
more  abundant  in  the  exclosure  than 
in  the  allotment.  Although  we  were 
unable  to  test  this  trend,  it  is  possible 
that  the  sagebrush  vole  found  the  ex- 
closure,  with  its  slightly  greater  grass 
and  shrub  cover,  to  be  more  inhabit- 
able. It  is  apparent  from  other  studies 
that  grass  and  shrub  cover  are  im- 
portant components  of  sagebrush 
vole  habitat  (MacCracken  et  al.  1985, 
Maser  et  al.  1974,  Maser  and  Strickler 
1978,  O'Farrell  1972). 


Diversity 

In  1983,  diversity  of  small  mammals 
in  the  exclosure  was  greater  than  in 
the  grazing  allotment  (fig.  2).  Rela- 
tive abundance  of  deer  mice,  the 
most  common  species  (table  1),  was 
similar  in  both  areas;  however,  we 
caught  one  more  species  in  the  exclo- 
sure. In  1984,  small  mammal  diver- 
sity was  greater  in  the  allotment  than 
in  the  exclosure.  During  that  year, 
deer  mice  made  up  a  somewhat 
smaller  relative  proportion  of  the 
small  mammal  total  in  the  allotment 
(table  1);  thus  the  line  for  the  allot- 
ment starts  higher  on  figure  2  indi- 
cating greater  evenness  in  the  per- 
centage each  species  contributed  to 
the  population.  We  captured  one 


more  species  in  the  allotment  than  in 
the  exclosure  which  extended  the  tail 
of  the  profile  further  to  the  right.  Be- 
cause of  this  change  from  one  year  to 
the  next,  we  were  unable  to  conclude 
what  impact  the  grazing  system  had 
on  small  mammal  diversity.  Hanley 
and  Page  (1982)  observed  a  higher 
diversity  index  on  their  ungrazed 


sagebrush-Idaho  fescue  site  60-80  km 
west  of  Badger  Mountain. 


Vegetation  on  the  Small  MamnrKil 
Study  Area 

Generally,  the  SRH  ecosite  had  lower 
grass  and  litter  cover  and  a  greater 


I 
I 


I 
I 


NUMBER  OFSPEC/ES 

Figure  2.— Small  mammal  diversity  profiles  for  the  cattle  exclosure  and  ttie  allotment,  Stiel- 
don  National  Wildlife  Refuge,  1983-84.  If  profile  lines  intersect,  tt>en  diversity  does  not  differ 
betv/een  areas  (Patil  and  Taillie  1 979). 


394 


I 


shrub  intercept  value  than  did  the 
MR  ecosite  (fig.  3).  In  the  SRH 
ecosite,  microhabitat  characteristics 
did  not  differ  between  the  exclosure 
and  allotment,  except  for  1983  when 
shrub  height  in  the  allotment  was 
lower  (P<0.05)  than  that  in  the  exclo- 


sure. In  the  MR  ecosite,  shrub  inter- 
cept was  lower  (P<0.03)  in  the  allot- 
ment than  in  the  exclosure  both  years 
and  grass  cover  was  higher  (P<0.10) 
in  the  exclosure  in  1983.  In  both 
ecosites,  there  was  a  general  trend 
for  cover  of  both  grasses  and  forbs  to 


be  lower  in  the  allotment  than  in  the 
exclosure. 

This  trend  is  probably  due  to  the 
cattle  grazing.  However,  the  fact  that 
the  means  are  relatively  similar  (es- 
pecially in  the  SRH  ecosite)  and  do 
not  differ  significantly  between  areas 
indicates  that  the  grazing  effect  is 
within  goals  established  by  the  ref- 
uge. 

Microhabitat  Characteristics  of 
Deer  Mice  Catch  Sites 

In  the  SRH  ecosite,  traps  where  deer 
mice  were  caught  had  significantly 
greater  litter  cover  (P=0.07  in  1984), 
shorter  shrubs  (P=0.09  in  1984),  and 
greater  shrub  intercept  (P=0.10  in 
1983)  than  traps  where  deer  mice 
were  not  caught  (fig.  4).  These  pat- 
terns tended  to  hold  for  both  years. 

In  the  MR  ecosite,  litter  cover, 
which  is  greater  than  in  the  SRH 
ecosite,  did  not  appear  to  be  a  signifi- 
cant vegetative  characteristic  (fig.  4). 
Grass  cover  in  1984  was  lower 
(P=0.06)  and  shrub  height  (P=0.02) 
and  shrub  intercept  (P=0.08)  were 
greater  at  traps  where  deer  mice 
were  caught  than  where  they  were 
not  caught. 

In  both  the  SRH  and  MR  ecosites, 
deer  mice  appeared  to  use  mi- 
crohabitat that  had  greater  shrub 
intercept.  This  corresponds  with  the 
findings  of  Feldhamer  (1979)  who 
noted  an  increase  in  deer  mouse  den- 
sity with  increased  foliage  in  the 
shrub  layer.  Other  studies  have 
found  that  deer  mice  were  associated 
with  light  cover  in  heavily  grazed 
sites  (Black  and  Frischknecht  1971), 
with  increasing  forb  cover  (Geier  and 
Best  1980),  or  with  no  measured 
habitat  variable  (Hallett  1982). 


70  -1 


SRHFCOS/TE 


MR  FC OS/ ft 


i 

I 


t 


m 

fW 
100 
90 
80 
70 
60 
50 
40 
JO 
20 
W 
0 


8J  84  8J  84  8J  84 
OmSS       FO/^ff  LIUER 


8J  84  8J  84  8J  84 
cms        FORB  LIFTER 


8J  84 

SHRUB 
HF/GHT 


8J  84 

SHRUB 
/NTFRCFRT 


SJ  84 

SHRUB 
HF/GHT 


85  84 

SHRUB 
/HTFRCFPr 


M/CmAB/TAT  mmBLES 


FXCLOSURF 


ALLOWFHT 


Figure  3.— Microhabitat  characteristics  arourxJ  trap  stations  in  tt>e  shrubby-roliing  hiils  and 
nrxjhogany  rockiands,  Sheidon  National  Wildlife  Refuge.  Variables  with  an  "a"  denote  a  P 
value  of  <0.1  between  the  two  areas. 


Microhabitat  Characteristics  of 
Chipmunic  Catch  Sites 

In  the  SRH  ecosite,  shrub  height  was 
lower  (P<0.08  in  1984)  in  catch  loca- 
tions in  the  exclosure  and  the  allot- 


395 


ment  than  in  non-catch  locations. 
This  pattern  held  in  1983  (fig.  5). 

In  the  MR  ecosite  there  were  no 
consistent  patterns  of  chipmunk  mi- 
crohabitat  use  (fig.  5).  Shrub  inter- 
ception, in  1984,  was  greater  (P<0.05) 
in  chipmunk  catch  locations  than 
non-catch  locations;  however  this 
pattern  was  not  evident  in  1983. 

Microhabitat  selection  by  the  least 
chipmunk  lacked  a  consistent  pattern 
for  either  ecosite  or  year.  However, 
the  fact  that  the  least  chipmunk  is  an 
opportunistic  forager  and  is  the  most 
widespread  of  all  North  American 
chipmunks  (Hall  1981),  suggests  that 
this  rodent  adapts  rapidly  to  a  vari- 
ety of  habitat  types.  Sullivan  (1985) 
found  that  the  least  chipmunk  was 
associated  with  a  wide  variety  of 
ecological  situations  in  the  southwest 
and  suggested  that  this  species  may 
be  predisposed  to  exploiting  mar- 
ginal environments. 

Conclusions 

These  results  indicate  that  the  graz- 
ing regime  initiated  on  the  Badger 
Mountain  allotment  had  no  discern- 
ible impact  on  the  relative  abundance 
and  diversity  of  small  mammals, 
four  and  five  years  after  its  implem- 
entation. The  dominance  of  two  op- 
portunistic species  on  the  study  area 
probably  contributed  to  this  lack  of 
difference.  We  suggest  future  moni- 
toring of  the  study  area  to  determine 
the  long-term  response  of  small 
mammals  to  the  grazing  program. 
Particular  attention  should  be  given 
to  the  two  vole  species  which  are  the 
most  sensitive  to  changes  in  cover. 

Acknowledgments 

We  thank  the  Sheldon  National  Wild- 
life Refuge  personnel  for  their  sup- 
port during  this  project.  We  appreci- 
ate V.  Reid's  assistance  with  the  de- 
sign of  this  study.  J.  Sedgwick  pro- 
vided expertise  in  the  statistical 
analysis  of  the  data.  And  we  thank 


the  following  people  for  their  assis- 
tance in  the  field:  B.  Allen-Johnson,  S. 
Boyle,  C.  Halvorson,  B.  Oldemeyer, 
E.  Rominger,  M.  Woodis,  and  S. 


Woodis.  This  manuscript  benefited 
by  reviews  from  M.  Bogan,  D.  Fran- 
zen,  W.  Grant,  M.  Kaschke,  B.  Keatt, 
J.  Sedgwick,  and  K.  Severson. 


SRHECOS/TE 


MRECOS/TE 


I 


I 


8J  84  8J  84  8J  84 
cms       FO/?B  L/TTER 


8J  84  8J  84  84  84 
GRASS       FORB  L/J7FR 


8J  84 
SHRUB 
HE/GHT 


85  84 

SHRUB 
/NTFRCEPr 


8J  84 
SHRUB 
Hf/GHr 


8J  84 

SHRUB 
INIFRGEPr 


M/CROHAB/TAT  MR/ABLES 


CAUGHT 


HOrCAUGHT 


Figure  4.— Microhabitat  characteristics  around  traps  where  deer  mice  were  captured  and 
not  captured  by  year  and  ecosite,  Sheldon  National  Wildlife  Refuge.  Variables  with  an  "o" 
denote  a  P  value  of  <0.1  between  the  two  areas. 


396 


Literature  Cited 

Anderson,  W.  E.  1978.  Range  site 
handbook  for  the  high  desert  re- 

>,n .  SRHfCOS/Tf 


source  province,  Oregon.  Un- 
paged. Lake  Oswego,  Oregon. 
Birney,  Elmer  C,  William  E.  Grant, 
and  Donna  Day  Baird.  1976.  Im- 

MR  ECO  SITE 


portance  of  vegetative  cover  to 
cycles  of  Microtus  populations. 
Ecology  57:1043-1051. 

Black,  Hal  L.,  and  Neil  C.  Fris- 
chknecht.  1971.  Relative  abun- 
dance of  mice  on  seeded  sage- 
brush-grass range  in  relation  to 
grazing.  U.S.  Department  of  Agri- 
culture Forest  Service,  Resource 
Note  INT-147, 8p. 

Feldhamer,  George  A.  1979.  Vegeta- 
tive and  edaphic  factors  affecting 
abundance  and  distribution  of 
small  mammals  in  southeast  Ore- 
gon. Great  Basin  Naturalist  39:207- 
218. 

Geier,  Anthony  R.,  and  Louis  B.  Best. 
1980.  Habitat  selection  by  small 
mammals  of  riparian  communi- 
ties: evaluating  effects  of  habitat 
alterations.  Journal  of  Wildlife 
Management  44:16-24. 

Grant,  William  E.,  Elmer  C.  Bimey, 
Norman  R.  French,  and  David  M, 
Swift.  1982.  Structure  and  produc- 
tivity of  grassland  small  mammal 
communities  related  to  grazing 
changes  in  vegetative  cover.  Jour- 
nal of  Mammalogy  63:248-260. 

Hall,  E.  Raymond.  1946.  Mammals  of 
Nevada.  University  of  California 
Press,  Berkeley  and  Los  Angeles. 
710  p. 

Hall,  E.  Raymond.  1981.  The  mam- 
mals of  North  America.  Volumes  1 
and  2.  John  Wiley  &  Sons,  New 
York.  1181  p. 

Hallett,  James  G.  1982.  Habitat  selec- 
tion and  the  community  matrix  of 
a  desert  small  mammal  fauna. 
Ecology  63:1400-1410. 

Hanley,  Thomas  A.,  and  Jerry  L. 
Page.  1982.  Differential  effects  of 
livestock  use  on  habitat  structure 
and  rodent  populations  in  Great 
Basin  communities.  California  Fish 
and  Game  68:160-174. 

Larrison,  Earl  J.,  and  Donald  R. 
Johnson.  1973.  Density  changes 
and  habitat  affinities  of  rodents  of 
shadscale  and  sagebrush  associa- 
tions. Great  Basin  Naturalist 
33:255-264. 

LoBue,  Joseph,  and  Rezneat  M.  Dar- 
nell. 1959.  Effect  of  habitat  distur- 


I 


84    8J    84    8J    84  8J    84    8J    84    8J  84 

GRASS        FORB       UTTE/?  GRASS        FORB  L/JTFR 


t 
I 


8J  84 
SHRUB 
HF/GHr 


8J  84 
SHRUB 
MFRCFPr 


8J  84 
SHRUB 
HF/GHF 


85  84 
SHRUB 
MFRGFPr 


M/CmAB/TAT  m/ABLES 


CAUGHf 


HOrCAUGHT 


Figure  5.— Microhabitat  characteristics  around  traps  where  least  chipmunks  were  captured 
and  not  captured  by  year  and  ecosite,  Sheldon  National  Wildlife  Refuge.  Variables  with  an 
"a"  denote  a  P  value  of  <0.1  between  the  two  areas. 


397 


bance  on  a  small  mammal  popula- 
tion. Journal  of  Mammalogy 
40:425-437. 

MacCracken,  James  G.,  Daniel  W. 
Uresk,  Richard  M.  Hansen.  1985. 
Rodent  vegetation  relationships  in 
southeastern  Montana.  Northwest 
Science  59:272-278. 

Maser,  Chris,  E.  Wayne  Hammer, 
Cheri  Bzown,  Robert  E.  Lewis, 
Robert  1.  Rausch,  and  Murray  L. 
Johnson.  1974.  The  sage  vole, 
Lagurus  curtatus  (Cope,  1968),  in 
the  Crooked  River  National  Grass- 
land, Jefferson  County,  Oregon:  a 
contribution  to  its  life  history  and 
ecology.  Saugetierkundliche  Mit- 
teilungen  22:193-222. 

Maser,  Chris,  and  Gerald  S.  Strickler. 
1978.  The  sage  vole,  Lagurus  cur- 
tatus, as  an  inhabitant  of  subalpine 
sheep  fescue,  Festuca  ovina,  com- 
munities on  Steens  Mountain — an 
observation  and  interpretation. 
Northwest  Science  52:276-284. 

O'Farrell,  Thomas  P.  1972.  Ecological 
distribution  of  sagebrush  voles, 
Lagurus  curtatus,  in  south-central 
Washington.  Journal  of  Mam- 
malogy 53:632-636. 

Oldemeyer,  J.  L,  S.  J.  Martin,  and  S. 
G.  Woodis.  1983.  A  preliminary 
report  on  the  effects  of  a  deferred- 
rotation  grazing  system  on  wild- 
life at  the  Sheldon  National  Wild- 
life refuge.  Cal-Neva  Wildlife 
Transactions  1983:26-42. 

Patil,  G.  P.,  and  C.  Taillie.  1979.  A 
study  of  diversity  profiles  and  or- 
derings  for  a  bird  community  in 
the  vicinity  of  Colstrip,  Montana, 
p.  23-48.  In  Contemporary  quanti- 
tative ecology  and  related  ecomet- 
rics.  International  Cooperative 
Publishing  House,  Fairland,  Mary- 
land. 

Sullivan,  Robert  M.  1985.  Phyletic, 
biogeographic,  and  ecologic  rela- 
tionships among  montane  popula- 
tions of  least  chipmunks  (Eutamias 
minimus)  in  the  southwest.  Sys- 
tematic Zoology  34:419-448. 


U.S.  Fish  and  Wildhfe  Service.  1980. 
Sheldon  National  Wildlife  Refuge 
renewable  resources  management 
plan-Final  EIS.  Department  of  the 
Interior,  Fish  and  Wildlife  Service, 
Region  1,  Portland,  Oregon. 


398 


Effect  of  Seed  Size  on 
Removal  by  Rodents^ 

William  G.  Standley^ 


Abstract.— Plots  located  in  southeastern  Arizona 
were  seeded  with  small  and  large  grass  seeds.  After 
3  days,  virtually  all  large  seeds  were  removed  by  ro- 
dents, while  small  seeds  were  still  present  36  days  af- 
ter planting.  Thus,  managers  may  increase  seed  sur- 
vival in  this  area,  without  removing  rodents,  by  seed- 
ing with  small  seeds  rather  than  large  seeds. 


Seeding  is  commonly  used  for  restor- 
ing depleted  vegetation.  Many  seed- 
ing projects  fail  because  rodents  eat 
the  seeds  (Bramble  and  Sharp  1949, 
Spencer  1954,  Nelson  et  al.  1970).  A 
variety  of  techniques  for  reducing 
the  impact  of  rodents  have  been 
tested,  but  few  have  been  successful. 
Most  often  resource  managers  poison 
rodent  populations  before  seeding, 
but  this  method  is  largely  unsuccess- 
ful because  of  rapid  immigration  of 
new  individuals  (Sullivan  1979,  Sulli- 
van and  Sullivan  1984).  New  meth- 
ods of  biological  management  could 
be  developed  that  use  information 
gained  from  diet  and  behavior  stud- 
ies to  reduce  destruction  of  seeds  by 
rodents.  Many  studies  show  that  cer- 
tain rodents  prefer  particular  species 
or  sizes  of  seeds  (Reynolds  and  Has- 
kell 1949,  Reynolds  1950,  Abbott 
1962,  Gashwiler  1967,  Smith  1970, 
Lockard  and  Lockard  1971,  Smigel 
and  Rosenzweig  1974,  Everett  et  al. 
1978,  Price  1983).  Thus,  whenever 
alternative  plant  species  are  available 
that  both  meet  the  resource  man- 
ager's objectives  and  have  seeds  not 
preferentially  foraged  by  local  seed- 

' Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  North  America.  (Flag- 
staff .  Al.  July  19-21.  1988.) 

'William  G.  Standley.  formerly  a  gradu- 
ate student.  University  of  Arizona.  Arizona 
Cooperative  Fish  and  Wildlife  Research 
Unit,  is  currently  Animal  Ecologist.  EG&G 
Energy  Measurements.  Inc..  c/o  NPR-1.  P.O. 
Box  127.  Tupman.  CA.  93276. 


eating  rodents,  seeding  could  be  suc- 
cessful even  with  rodents  present. 

In  southwestern  deserts  of  North 
America,  where  range  managers  are 
attempting  to  restore  rangelands  de- 
pleted by  overgrazing  (Cox  et  al. 
1982),  kangaroo  rats  (Dipodomys  sp.) 
and  pocket  mice  (Perognathus  sp.  and 
Chaetognathus  sp.)  are  some  of  the 
primary  seed  eaters  (Brown  et  al. 
1979b).  As  early  as  1950,  Reynolds 
suggested  that  the  influence  of  Mer- 
riam's  kangaroo  rats  (Dipodomys  mer- 
riami)  on  seeding  success  depends  on 
the  size  of  seeds  used.  Brown  et  al. 
(1979b),  Inouye  et  al.  (1980),  and 
Price  (1983)  all  found  that  hetero- 
myids  preyed  selectively  on  large 
seeds.  In  this  study,  I  investigated 
the  prediction  that  fewer  small  seeds 
than  large  seeds  would  be  removed 
by  rodents  in  a  seeded  area  in  south- 
eastern Arizona. 


Study  Area  and  Mettiods 

The  study  area  was  on  the  USDA 
Forest  Service  Santa  Rita  Experimen- 
tal Range  located  45  km  south  of 
Tucson,  Pima  County  AZ,  which  is 
thoroughly  described  by  Martin  and 
Reynolds  (1973).  The  vegetation  was 
typical  Sonoran  desert-scrub,  domi- 
nated by  mesquite  (Prosopis  juliflora), 
burroweed  (Haplopappus  tenuisectus) 
and  cholla  {Opuntia  spp.).  Annual 
precipitation  averages  36  to  43  cm 
and  is  bimodal,  with  p>eaks  in  winter 
and  summer.  Plots  were  seeded  fol- 


lowing all  recommended  procedures 
(Jordan  1981),  using  large  and  small 
seeds,  both  separately  and  together. 
Seeded  plots  were  located  within  a 
slightly  sloped  1-hectare  area  with  a 
Comoro  soil  type,  at  an  elevation  of 
1300  m.  I  compared  the  number  of 
seeds  surviving  on  4  experimental 
plots  to  the  number  of  seeds  surviv- 
ing on  a  control  plot  which  was  pro- 
tected from  rodents. 

The  study  area  was  prepared  by 
removing  large  shrubs  by  hand  and 
plowing  small  plants  with  a  disk.  The 
control  plot  was  protected  from  ro- 
dents with  a  rodent-proof  fence  simi- 
lar to  that  used  by  Brown  et  al. 
(1979a).  All  rodents  within  the  exclo- 
sure  were  removed  by  trapping  be- 
fore seeding.  Each  of  the  5-15  x  17  m 
plots  was  seeded  with  3  evenly 
placed  pairs  of  15  m  rows,  one  pair 
for  each  of  3  treatments  which  were: 
(1)  small  seeds  planted  at  a  rate  of 
175/m  (0.15  g/m),  (2)  large  seeds 
planted  at  a  rate  of  100/m  (4.7  g/m), 
and  (3)  5  small  and  large  seeds 
planted  together  at  88/m  and  50/ m 
(0.07g/m  and  2.35  g/m),  respec- 
tively. Seeding  rates  were  chosen  ac- 
cording to  recommended  rates  for 
similar  sized  seeds  (Jordan  1981). 
The  treatm.ent  assigned  to  each  pair 
of  rows  was  randomly  selected.  The 
small  seeds  were  blue  panicgrass 
(Panicum  antidotale)  which  weighed 
an  average  of  0.85  mg  each.  The  large 
seeds  were  barley  (Hordeum  vulgar e) 
which  weighed  an  average  of  47.0 
mg  each.  All  seeds  were  planted  with 


399 


a  cone  seeder  at  a  depth  of  1  to  2  cm 
on  21  June  1984,  just  before  expected 
summer  rains.  Because  blue  pan- 
icgrass  seeds  are  very  small  and  dif- 
ficult to  recover  from  the  soil,  they 
were  dyed  with  water  soluble  green 
food  coloring  before  planting.  Barley 
seeds  were  also  dyed  to  avoid  a  pos- 
sible bias. 

The  species  of  rodents  on  the  ex- 
perimental plots  were  monitored  by 
placing  100  live  traps  at  10  m  inter- 
vals on  and  around  the  plots  on  the 
5th  and  6th  nights  after  planting. 
Traps  were  baited  with  a  mixture  of 
both  sizes  of  seeds  and  checked  at 
midnight  and  sunrise. 

The  number  of  seeds  surviving  on 
plots  was  monitored  by  collecting 
soil  samples  from  the  rows  immedi- 
ately after  planting  and  at  3, 9, 18, 
and  36  days  after  planting.  One  ran- 
dom sample  was  taken  from  each 
quarter  of  every  row  each  time. 
Samples  were  not  taken  from  the 
outer  meter  of  any  row  because  the 
cone  seeder  applied  seeds  at  a  more 
variable  rate  at  the  beginning  and 
end  of  each  row. 

Soil  samples,  2.5  to  3.5  cm  deep 
and  15  X  25  cm  in  area,  were  taken 
lengthwise  along  each  row  with  the 
aid  of  a  two-sided,  fixed-area  sam- 
pler and  a  trowel.  The  samples  were 
placed  in  paper  bags,  and  oven-dried 
at  50  C  for  24  hours.  Seeds  were  re- 
covered by  shaking  soil  samples 
through  a  series  of  Tyler  sieves 
(#5,#10,#14,#18,#20,and  #25)  for  3 
minutes.  The  number  of  seeds  re- 
maining were  counted  by  examining 
the  contents  of  each  sieve,  both  dry 
and  immersed  in  a  salt  water  solu- 
tion, through  a  lOX  viewing  scope. 

The  average  number  of  seeds  re- 
covered in  the  soil  samples  taken 
from  the  4  experimental  plots  di- 
vided by  the  total  number  found  in 
the  control  plot  times  100  was  used 
as  a  seed  survival  index  (SSI).  This 
dimensionless  index  permits  com- 
parison of  the  removal  of  different 
sized  seeds  by  rodents  even  though 
they  were  planted  at  different  rates. 
It  also  standardizes  for  the  experi- 


mental error  contributed  by  the  diffi- 
culty of  recovering  small  seeds.  The 
granivorous  arthropods  and  birds 
present  on  the  study  area  had  equal 
access  to  control  and  experimental 
plots,  so  should  not  have  biased  SSIs. 


Results 

Eleven  of  17  individual  rodents  cap- 
tured on  or  around  the  plots  were 
heteromyids:  9  were  Merriam's  kan- 
garoo rats,  and  2  were  bannertail 


Seeds  sown  separately 


0  3  9  18  36 

DAYS  AFTER  PLANTING 


Figure  1  .—Seed  survival  index  (average  number  of  seeds  recovered  in  4  experinr»ental  plots 
divided  by  total  nunnber  of  seeds  recovered  in  ttie  control  plot  tinnes  1 00)  for  snrxali  and  large 
seeds.  (A)  Seeds  sown  separately  and  (B)  Seeds  sown  togetlier. 


400 


kangaroo  rats  (D.  spectabilis).  Two 
white-throated  woodrats  (Neotoma 
albigula),  2  southern  grasshopper 
mice  (Onychomys  torridus),  1  deer 
mouse  (Peromyscus  maniculatus)  and 
1  cotton  rat  (Sigmodon  hispidus)  were 
also  captured. 

Whether  large  and  small  seeds 
were  planted  separately  or  together, 
the  SSIs  were  higher  for  small  seeds 
than  for  large  seeds  starting  with  3 
days  after  planting  (fig.  1).  After  36 
days,  the  large  seeds  planted  either 
separately  or  with  small  seeds  were 
virtually  gone  from  experimental 
plots  (SSI  =  0.4  and  2.1,  res|:>ectively). 
The  SSI  for  small  seeds  planted  sepa- 
rately was  76.5  after  36  days,  while 
the  small  seeds  planted  with  large 
seeds  had  an  SSI  of  43.6.  The  SSIs  of 
large  seeds  planted  separately  de- 
creased at  a  faster  rate  than  the  SSIs 
of  large  seeds  sown  with  small  seeds. 
The  SSIs  of  small  seeds  planted  sepa- 
rately decreased  at  a  slower  rate  than 
the  SSIs  of  small  seeds  sown  with 
large  seeds,  however.  Complete  data 
are  presented  in  Standley  (1985). 

Discussion 

I  do  not  present  inferential  statistics 
to  test  for  significant  differences  be- 
tween large  and  small  seed  survival 
because  the  experimental  plots  were 
actually  sub-plots  rather  than  true 
replicates  (Hurlbert  1984).  For  this 
study  site,  however,  striking  differ- 
ences between  the  SSIs  of  large  and 
small  seeds  whether  planted  sepa- 
rately or  together  are  certainly  evi- 
dence that  smaller  seeds  have  a 
much  higher  survival  rate  than  large 
seeds  due  to  differential  predation  by 
rodents. 

The  higher  rate  of  removal  of  large 
seeds  planted  separately  compared 
to  large  seeds  planted  with  small 
seeds  may  have  occurred  because  the 
lower  density  of  large  seeds  in  the 
mixed  rows  made  them  less  attrac- 
tive to  rodents.  The  relatively  higher 
rate  of  removal  of  small  seeds 
planted  with  large  seeds,  compared 


to  small  seeds  planted  separately, 
likely  occurred  because  large  seeds 
attracted  rodents  to  the  rows,  where 
the  rodents  then  ate  both  sizes  of 
seeds.  Sullivan  and  Sullivan  (1982) 
observed  the  opposite  effect  when 
seeding  lodgepole  pine  (Pinus  con- 
torta).  Lodgepole  seed  consumption 
by  rodents  was  reduced  by  planting 
the  relatively  small  lodgepole  seeds 
with  sunflower  seeds,  which  were 
larger  and  more  preferred  by  gra- 
nivorous  rodents  present.  The  oppos- 
ing results  may  be  due  to  differences 
in  method  of  seeding  (Sullivan  and 
Sullivan  broadcast  their  seeds)  or  the 
size  of  plots  (Sullivan  and  Sullivan's 
plots  were  larger).  Another  possibil- 
ity is  that  the  main  granivorous  ro- 
dents in  their  study  area,  deer  mice, 
are  more  selective  than  the  hetero- 
myids  present  in  this  study.  Nine 
days  after  planting  there  was  a  lower 
SSI  for  small  seeds  planted  sepa- 
rately (fig.  la)  than  on  18  or  36  days, 
which  can  only  be  attributed  to  vari- 
ability in  seeding  rate  and  sampling 
error. 

It  is  possible  that  not  all  seeds  re- 
moved by  rodents,  small  or  large, 
were  destroyed.  Reynolds  and  Glen- 
dening  (1949)  found  that  the  seed 
caching  behavior  of  Merriam's  kan- 
garoo rats  actually  increased  spread 
of  some  plant  species. 

Factors  other  than  seed  size  affect 
selection  by  rodents  for  particular 
seed  species,  such  as  percent  soluble 
carbohydrates  (Kelrick  and  MacMa- 
hon  1985,  Kelrick  et  al.  1986;  but  also 
see  Jenkins  1988),  moisture  content 
(Frank  1988a),  and  moldiness  (Frank 
1988b).  For  most  seeds,  however,  re- 
source managers  have  only  the  infor- 
mation on  size  available.  This  study 
only  compared  the  effect  of  size  by 
using  grass  seeds  of  similar  composi- 
tion that  differ  most  in  their  linear 
dimensions.  The  results  of  this  study 
support  other  studies  which  showed 
that  heteromyid  rodents  selected 
large  seeds  and  reduced  standing 
stocks  of  large  seeds  in  the  soil  to  a 
greater  extent  than  small  seeds 
(Brown  et  al.  1979a,  Inouye  et  al. 


1980,  Price  1983).  Therefore,  when 
site  conditions  and  management 
needs  allow  a  choice  of  which  species 
to  seed,  resource  managers  should 
consider  the  size  of  seeds  when  plan- 
ning seeding  in  areas  inhabited  by 
heteromyids. 

Acknowledgments 

Appreciation  is  extended  to  Drs.  J.  H. 
Brown,  H.  L.  Morton,  and  N.  S. 
Smith  for  guidance,  and  to  S.  Collins, 
Dr.  J.  Cox,  S.  Horton,  M.  Podborny, 
B.  Travis,  and  D.  Youkey  for  field 
assistance. 

The  research  was  funded  by  the 
Arid  Land  Ecosystems  Improvement 
Unit  of  the  USD  A,  Agricultural  Re- 
search Service. 


Literature  Cited 

Abbott,  Herschel  G.  1962.  Tree  seed 
preferences  of  mice  and  voles  in 
the  northeast.  Journal  of  Forestry 
60:97-99. 

Bramble,  William  C.  and  Ward  M. 
Sharp.  1949.  Rodents  as  a  factor  in 
direct  seeding  on  s]X)il  banks  in 
central  Pennsylvania.  Journal  of 
Forestry  47:477-478. 

Brown,  James,  H.,  Diane  W. 
Davidson,  and  O.  J.  Reichman. 
1979a.  An  experimental  study  of 
competition  between  seed-eating 
desert  rodents  and  ants.  American 
Zoologist  19:1129-1143. 

Brown,  James,  H.,  O.  J.  Reichman, 
and  Diane  W.  Davidson.  1979b. 
Granivory  in  desert  ecosystems. 
Annual  Review  of  Ecology  and 
Systemarics  10:201-227. 

Cox,  Jerry,  R.,  Howard  L.  Morton, 
Thomas.  N.  Johnson,  Jr.,  Gilbert  L. 
Jordan,  S.  Clark  Martin,  and  Louis 
C.  Fierro.  1982.  Vegetation  restora- 
tion in  the  Chihuahuan  and  Sono- 
ran  deserts  of  North  America. 
United  States  Department  of  Agri- 
culture, Agricultural  Research 
Service,  Western  Series  Number 
28,  August  1982. 


401 


Everett,  Richard  L.,  Richard  O. 
Meeuwig,  and  Richard  Stevens. 
1978.  Deer  mouse  preference  for 
seed  of  commonly  planted  species, 
indigenous  weed  seed,  and  sacri- 
fice foods.  Journal  of  Range  Man- 
agement 31:70-73. 

Frank,  Craig  L.  1988a.  The  influence 
of  moisture  content  on  seed  selec- 
tion by  kangaroo  rats.  Journal  of 
Mammalogy  69:353-357. 

Frank,  Craig  L.  1988b.  The  effects  of 
moldiness  level  on  seed  selection 
by  Dipodomys  spectabilis.  Journal  of 
Mammalogy  69:358-362. 

Gashv^iler,  Jay  S.  1967.  Conifer  seed 
survival  in  a  western  Oregon 
clearcut.  Ecology  48:431-438. 

Hurlbert,  Stuart  H.  1984.  Pseu- 
doreplication  and  the  design  of 
ecological  field  experiments.  Eco- 
logical Monographs  54:187-211. 

Inouye,  Richard  S.,  Gregory  S.  Byers, 
and  James  H.  Brown.  1980.  Effects 
of  predation  and  competition  on 
survivorship,  fecundity,  and  com- 
munity structure  of  desert  annu- 
als. Ecology  61:1344-1351. 

Jenkins,  Stephen  H.  1988.  Comments 
on  relationships  between  native 
seed  preferences  of  shrub-steppe 
granivores  and  seed  nutritional 
characteristics.  Oecologia  75:481- 
482. 

Jordan,  Gilbert  L.  1981.  Range  seed- 
ing and  brush  management  on 
Arizona  rangelands.  Cooperative 
Extension  Service,  Agricultural 
Experiment  Station,  University  of 
Arizona,  College  of  Agriculture, 
T81121. 

Kelrick,  Michael  Ira,  and  James  A. 
MacMahon.  1985.  Nutritional  and 
physical  attributes  of  seeds  of 
some  common  sagebrush-steppe 
plants:  Some  implications  for  eco- 
logical theory  and  management.  J. 
Range  Manage.  38:65-69. 

Kelrick,  M.  I.,  J.  A.  MacMahon,  R.  R. 
Parmenter,  and  D.  V.  Sisson.  1986. 
Native  seed  preferences  of  shrub- 
steppe  rodents,  birds  and  ants:  the 
relationships  of  seed  attributes 
and  seed  use.  Oecologia  68:327- 
337. 


Lockard,  Robert,  B.,  and  Joan  S. 
Lockard.  1971.  Seed  preference 
and  buried  seed  retrieval  of  Dipod- 
omys deserti.  Journal  of  Mam- 
malogy 52:219-221. 

Martin,  S.  Clark  and  Hudson  G.  Rey- 
nolds. 1973.  The  Santa  Rita  Experi- 
mental Range:  your  facility  for  re- 
search on  semidesert  ecosystems. 
Journal  of  the  Arizona  Academy 
of  Science  8:56-67. 

Nelson,  Jack  R.,  A.  M.  Wilson,  and 
Carl  J.  Goebel.  1970.  Factors  influ- 
encing broadcast  seeding  in 
bunchgrass  range.  Journal  of 
Range  Management  23:163-170. 

Price,  Mary  V.  1983.  Laboratory 
studies  of  seed  size  and  seed  spe- 
cies selection  by  heteromyid  ro- 
dents. Oecologia  60:259-263. 

Reynolds,  Hudson  G.  1950.  Relation 
of  Merriam  kangaroo  rats  to  range 
vegetation  in  southern  Arizona. 
Ecology  31:456-463. 

Reynolds,  H.  G.  and  G.  E.  Glenden- 
ing.  1949.  Merriam  kangaroo  rat  as 
a  factor  in  mesquite  propagation 
on  southern  Arizona  range  lands. 
Journal  of  Range  Management 
2:193-197. 

Reynolds,  Hudson,  G.  and  Horace  S. 
Haskell.  1949.  Life  history  notes 
on  Price  and  Bailey  pocket  mice  of 
southern  Arizona.  Journal  of 
Mammalogy  30:150-156. 

Smigel,  Barbara  W.,  and  Michael  L. 
Rosenzweig.  1974.  Seed  selection 
in  Dipodomys  merriami  and  Per- 
ogmthus  penicillatus.  Ecology 
55:329-339. 

Smith,  Christopher  C.  1970.  The  co- 
evolution  of  pine  squirrels  (Tami- 
asciurus)  and  conifers.  Ecological 
Monographs  40:349-371. 

Spencer,  Donald  A.  1954.  Rodents 
and  direct  seeding.  Journal  of  For- 
estry 52:824-826. 

Standley,  William  G.  1985.  The  effect 
of  seed  size  on  reseeding  in  the 
presence  of  heteromyids.  M.S.  the- 
sis. University  of  Arizona,  Tucson. 
15  p. 

Sullivan,  Thomas  P.  1979.  Repopula- 
tion  of  clear-cut  habitat  and  coni- 
fer seed  predation  by  deer  mice. 


Journal  of  Wildlife  Management 
43:861-871. 

Sullivan,  Thomas,  P.  and  Druscilla  S. 
Sullivan.  1982.  The  use  of  alterna- 
tive foods  to  reduce  lodgepole 
pine  seed  predation  by  small 
mammals.  Journal  of  Applied 
Ecology  19:33-45. 

Sullivan,  Thomas  P.  and  Druscilla  S. 
Sullivan.  1984.  Influence  of  range 
seeding  on  rodent  populations  in 
the  interior  of  British  Columbia. 
Journal  of  Range  Management 
37:163-165. 


402 


Habitat  Use  by  Gunnison's 
Prairie  Dogs^ 

C.  N.  Slobodchikoff,^  Anthony  Robinson,^ 
and  Clark  Schaack* 


Abstract.— Gunnison's  prairie  dogs  (Cynomys  gun- 
nisoni)  ore  social,  colonial  mammals  found  in  Colo- 
rado, New  Mexico,  and  Arizona.  Colony  location 
depends  to  a  great  extent  on  the  distribution  and 
abundance  of  plants  used  as  food,  Colonies  with 
the  highest  densities  of  prairie  dogs  occur  in  habitats 
where  there  is  a  high  abundance  of  native  species 
of  plants.  From  a  management  standpoint,  prairie 
dog  populations  can  be  conserved  by  maintaining 
habitats  that  offer  such  resources. 


Prairie  dogs  often  have  been  consid- 
ered "weedy"  species  that  thrive  in 
disturbed  habitats.  However,  uncer- 
tainty remains  about  the  impact  of 
prairie  dogs  on  their  habitat,  and 
about  their  economic  impact  as  com- 
petitors of  domesticated  herbivores. 
Some  studies  of  primarily  black- 
tailed  prairie  dogs  (Cynomys  ludovi- 
cianus)  show  that  they  have  a  nega- 
tive effect  on  their  habitat,  while 
other  studies  show  a  positive  effect. 
Negative  effects  include  decreased 
forb  and  grass  cover  in  prairie  dog 
towns  (Knowles  1982,  Archer  et  al. 
1984),  higher  silicon  concentrations 
in  grasses  found  in  areas  grazed  by 
prairie  dogs  (Brizuela  et  al.  1984), 
and  removal  of  plant  biomass  that 
could  be  utilized  by  cattle  (Crocker- 
Bedford  1976,  Hansen  and  Gold 
1977,  Crocker-Bedford  and  Spillett 
1981).  Positive  effects  include  in- 
creased plant  species  diversity  in 
prairie  dog  towns  (Lerwick  1974, 
Boddicker  and  Lerwick  1976,  Gold 
1976,  Severe  1977,  Beckstead  and 


'Paper  presented  at  symposium,  Man- 
ogemer^t  of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff. Al,  July  19-21.  1988.) 

^C.  N.  Slobodchikoff  is  Professor  of  Biol- 
ogy. Norttiern  Arizona  University.  Flagstaff. 
AZ86011. 

^Anttiony  Robinson  is  a  graduate  stu- 
dent in  ttie  Department  of  Biology.  Norttiern 
Arizona  University.  Flagstaff.  AZ  8601 1. 

^Clark  Sctiaack  is  Assistant  Scientist,  De- 
partment of  Botany,  University  of  Wisconsin, 
Madison.  Wl  53706. 


Schitoskey  1980,  Fagerstone  1981, 
Archer  et  al.  1984);  greater  produc- 
tion of  forbs  and  grasses  (Uresk  and 
Bjugstad  1980,  Agnew  1983);  and  bet- 
ter quality  food  and  growing  condi- 
tions inside  prairie  dog  towns 
(Hassien  1976,  Beckstead  and  Schi- 
toskey 1980,  Fagerstone  1981,  Cop- 
pock  et  al.  1980, 1983a,  1983b,  Det- 
ling  and  Painter  1983).  Prairie  dog 
colonies  have  also  been  shown  to 
provide  habitat  for  many  different 
species  of  vertebrates  other  than 
prairie  dogs  (Campbell  and  Clark 
1981,  O'Meilia  et  al.  1982,  Agnew 

1983,  Clark  et  al.  1982). 

The  economic  effects  of  prairie 
dogs  are  also  currently  unclear.  Al- 
though they  are  considered  pests 
(Uresk  1985),  a  series  of  studies  has 
shown  that  controlling  or  eradicating 
prairie  dogs  has  little  effect  on  in- 
creasing the  amount  of  food  available 
for  cattle  (Crocker-Bedford  1976, 
Klatt  and  Hein  1978,  Collins  et  al. 

1984,  Uresk  1985),  and  experimental 
studies  of  competition  between  prai- 
rie dogs  and  steers  failed  to  show 
that  the  prairie  dogs  had  any  signifi- 
cant negative  impact  on  the  weight  of 
the  steers  (O'Meilia  et  al.  1982). 

Prairie  dogs  have  been  character- 
ized as  being  oriented  to  disturbed 
sites  that  are  overgrazed  by  cattle  or 
buffalo  (Osbom  and  Allan  1949).  The 
relationship  between  prairie  dog  oc- 
currence and  overgrazing,  however, 
is  a  correlational  one:  prairie  dogs 
can  be  found  at  sites  that  are  over- 
grazed by  large  herbivores,  but  this 


does  not  necessarily  imply  that  the 
prairie  dogs  specialize  in  colonizing 
sites  that  are  overgrazed.  Over- 
grazing might  be  occurring  subse- 
quent to  colonization.  For  example, 
bison  are  attracted  to  prairie  dog 
towns  as  grazing  sites,  because  the 
vegetation  associated  with  such 
towns  may  be  more  digestible,and 
have  a  higher  nitrogen  content  than 
the  vegetation  at  sites  not  colonized 
by  prairie  dogs  (Coppock  et  al. 
1983a,  1983b). 

Disturbance  of  a  habitat  can  be 
provided  by  the  activities  of  the  prai- 
rie dogs  themselves.  By  digging  ex- 
tensive burrow  systems  (King  1984), 
prairie  dogs  disturb  soil,  promoting 
the  growth  of  disturbance-oriented 
vegetation  and  increasing  plant  di- 
versity (Gold  1976;  Hansen  and  Gold 
1977).  Because  prairie  dogs  have  a 
system  of  vigilance  that  depends  on 
being  able  to  see  terrestrial  predators 
from  some  distance  away  (Slobod- 
chikoff and  Coast  1980),  they  clip 
shrubs  and  other  tall  vegetation  that 
impede  visual  detection.  This  in  turn 
alters  the  habitat  into  one  that  has 
predominantly  short  grasses  and  an- 
nual forbs,  rather  than  the  taller 
grasses  and  shrubs  that  are  more 
characteristic  of  climax  communities 
(Koford  1958). 

The  goal  of  this  paper  is  to  evalu- 
ate habitat  use  by  Gunnison's  prairie 
dogs  (Cynomys  gunnisoni),  and  to 
consider  this  habitat  use  in  the  con- 
text of  managing  existing  popula- 
tions of  this  species.  Many  previous 


403 


ecological  studies  of  prairie  dogs 
have  focused  on  the  blacktailed  prai- 
rie dogs  (Cynomys  ludovicianus) 
found  in  the  midwestem  states.  Gun- 
nison's prairie  dogs  offer  a  better 
opportunity  to  evaluate  habitat  re- 
quirements, because  this  species  is 
associated  with  habitats  that  have 
been  modified  less  by  man  than  habi- 
tats where  blacktailed  prairie  dogs 
are  currently  found. 

In  an  attempt  to  establish  some 
common  habitat  conditions  that  are 
preferred  by  Gunnison's  prairie 
dogs,  we  have  examined  the  follow- 
ing factors  at  several  prairie  dog 
sites:  (1)  burrow  density  as  an  indica- 
tor of  prairie  dog  population  density; 
and  (2)  plant  diversity,  evenness, 
cover,  and  proportions  of  native  and 
introduced  species. 

Study  Areas 

Seven  colonies  in  the  vicinity  of  Flag- 
staff, Arizona,  were  investigated. 
These  were:  (1)  Humane  Society 
(HS),  within  the  city  limits  at  an  ele- 
vation of  2250  m,  in  a  meadow  sur- 
rounded by  Ponderosa  pine  (Finns 
ponderosa)  trees  on  three  sides  and  a 
heavily-utilized  dirt  road  on  the  re- 
maining side;  (2)  Denny's  (D),  also 
within  the  city  limits  at  an  elevation 
of  2250  m,  in  a  small  meadow  en- 
circled by  a  traffic  loop  that  serves  as 
an  approach  to  the  1-17  freeway;  (3) 
Snow  Bowl  (SB),  10  km  north  of  Flag- 
staff in  an  old-field  pasture  at  an  ele- 
vation of  2400  m;  (4)  Upper  Michel- 
bach  (UM),  on  a  privately  owned 
ranch  20  km  north  of  Flagstaff  at  an 
elevation  of  2650  m;  (5)  Lower  Mich- 
elbach  (LM),  also  at  2650  m  and  lo- 
cated within  1  km  east  of  UM;  (6)  Po- 
tato Lake  (PL),  in  an  alpine  meadow 
surrounded  by  forested  slopes,  25 
km  northeast  of  Flagstaff  at  an  eleva- 
tion of  2850  m;  and  (7)  Bismark  Lake 
(BL),  another  alpine  meadow  20  km 
northeast  of  Ragstaff  at  2900  m. 

Grazing  pressure  on  these  sites 
varied.  The  most  heavily  grazed  site 
was  Upper  Michelbach,  with  grazing 


levels  of  0.8  ha  per  AUM.  The  Hu- 
mane Society  site  was  heavily  grazed 
(1.2  ha  per  AUM)  until  1978,  after 
which  there  was  no  grazing.  Both 
Lower  Michelbach  and  Snow  Bowl 
had  the  same  level  of  grazing  (6  ha 
per  AUM).  The  Potato  Lake  and  Bis- 
mark Lake  sites  had  relatively  light 
levels  of  grazing  (12  ha  per  AUM  at 
PL;  14  ha  per  AUM  at  BL).  The 
Denny's  site  was  not  grazed  at  all  in 
the  last  20  years  (all  grazing  informa- 
tion from  J,  Mundell,  pers.  comm.). 

Methods 

To  estimate  relative  densities  of  prai- 
rie dog  populations,  we  sampled 
burrow  densities  at  six  of  the  sites 
(HS,  SB,  UM,  LM,  PL,  and  BL).  Bur- 
rows were  estimated  by  laying  out 
twelve  50  m  transects,  and  counting 
all  the  burrows  that  were  within  0.5 
m  of  each  side  of  the  transect  line. 
Based  on  the  counts  of  burrows  per 
transect,  mean  numbers  of  burrows  / 
0.005  ha  (mean  number  of  burrows 
per  50  m-sq)  were  calculated  for  each 
colony.  Because  of  the  small  size  of 
the  colony  at  BL,  only  six  transects 
were  used  there.  Although  this 
method  did  not  provide  a  total  num- 
ber of  burrows  per  site  (a  number 
constantly  changing  depending  on 
prairie  dog  construction  activity),  it 
did  provide  a  measure  that  allowed 
comparison  of  the  six  sites. 

As  an  estimate  of  habitat  composi- 
tion, vegetation  at  five  sites  (HS,  SB, 
D,  PL,  and  BL)  was  sampled  from 
May-October,  1986-87.  All  plant  spe- 
cies found  at  each  site  were  identi- 
fied to  species  and  classified  as  na- 
tive non-weedy,  native-weedy,  or 
introduced-weedy.  Reference  speci- 
mens for  each  species  from  each  site 
have  been  deposited  in  the  Herbar- 
ium at  Northern  Arizona  University. 

For  estimates  of  plant  diversity 
and  percent  cover,  we  sampled 
plants  every  month  along  transects  at 
two  sites  (HS  and  SB)  from  May-Oc- 
tober, 1986  and  1987.  Each  site  had 
six  100  m  parallel  transects  spaced  20 


m  apart.  Presence  or  absence  of 
plants  by  species  were  recorded  ev- 
ery 2  m  along  each  transect. 

Similarity  indices  (SI)  were  calcu- 
lated for  plant  species  composition 
between  sites,  as  follows: 

Number  of  Species  Common 
to  Both  Site  A  and  B 
S|=   

Total  Number  of  Species  in 
Site  A  +  Site  B 

This  is  an  index  that  allows  compari- 
sons of  sites  based  on  the  percentage 
of  species  common  to  the  two. 

Prairie  dog  densities  were  deter- 
mined at  two  sites,  HS  and  SB,  by 
actual  counts  of  all  the  animals  at 
each  site.  The  prairie  dogs  were 
trapped  weekly  in  squirrel-sized 
Tomahawk  live  traps  and  marked 
with  hair  dye.  Movements  of  marked 
prairie  dogs  were  observed  and  plot- 
ted with  respect  to  a  100  x  120  m  grid 
of  stakes  set  up  10  m  apart.  Territo- 
ries were  determined  behaviorally, 
on  the  basis  of  aggressive  behaviors 
such  as  chases  between  interterritory 
members,  and  cooperative  behaviors 
such  as  greet-kisses  between  intrater- 
ritory  members.  At  these  two  sites, 
HS  and  SB,  the  number  of  burrows  in 
each  territory  was  counted. 

All  statistical  analyses  were  done 
on  a  Honeywell  Sigma  6  mainframe 
computer,  using  SPSS  statistical 
packages  (Nie  et  al.  1975).  Analyses 
included  regression,  correlation, 
analysis  of  variance,  and  least  signifi- 
cant difference.  Additionally,  eco- 
logical indices  were  calculated:  even- 
ness, percent  cover,  Simpson's  domi- 
nance, Shannon-Weaver  diversity, 
and  H  max  (Poole  1974). 


Results 

Plant  Species  Composition 

Similarity  indices  show  that  some 
sites  were  quite  dissimilar  from  other 
sites  (table  1).  The  HS  and  D  sites 
were  most  similar  (63.7  percent  simi- 
larity), and  SB  was  fairly  similar  to 


404 


Table  1.— Similarity  Indices  for  five  Gunnison's  prairie  dog  colonies,  based 
on  plant  species  composition  at  each  site.  A  similarity  of  100  Implies  that 
all  the  plant  species  at  both  sites  are  the  same.  A  similarity  of  0  implies  ttiat 
no  plant  species  are  common  to  the  two  sites.  Sites  are:  BL  =  Bismark  Lake; 
D  =  Denny's;  PL  =  Potato  Lake;  SB  =  Snow  Bowl;  HS  =  Humane  Society. 


Site 

BL 

D 

PL 

SB 

Humane  Society  (HS) 

6.7 

63.7 

8.6 

54.1 

Snow  Bowl  (SB) 

10.2 

44.1 

12.9 

Potato  Lake  (PL) 

23.4 

4.8 

Denny's  (D) 

8.2 

100 -I 


N  NW  I         N  NW  I         N  NW  I         N  NW  I        N  NW  I 

HS  SB  BL  PL  D 


SITES 

Figure  1  .—Composition  of  plant  species  at  five  Gunnison's  prairie  dog  colonies  near  Rag- 
staff,  Arizona.  Percentages  shown  are  for  Native-nonweedy  species  (N),  Native-weedy  spe- 
cies (NW),  and  Introduced-weedy  species  (I).  Sites  are:  HS  =  Humane  Society;  SB  =  Snow 
Bowl;  BL  =  Bismark  Lake;  PL  =  Potato  Lake;  D  =  Denny's. 


Table  2— Mean  burrow  densities  and  standard  deviations  at  6  Gunnison's 
prairie  dog  colony  sites.  Means  that  are  not  significantly  different  (LSD  Test) 
are  associated  by  the  same  letter. 


Site  Mean  +  SD  LSD  Test 


Upper  Michelbach  (UM)  5.42  +  2. 15  a 

Humane  Society  (HS)  4.17+1.90  a  b 

Snow  Bowl  (SB)  3.17+1.69  be 

Lower  Michelbach  (LM)  2.92  +  2.47  b  c 

Potato  Lake  (PL)  2.83  +  1 .64  be 

Bismark  Lake  (BL)  2.17+1.72  c 

V  J 


405 


the  HS  site  (54.1  percent  similarity) 
and  to  the  D  site  (44.1  percent  simi- 
larity). The  HS,  D,  and  SB  sites  were 
quite  dissimilar  from  the  other  two 
sites,  PL  and  BL,and  the  two  latter 
sites  had  a  low  level  of  similarity 
(23.4  percent)  to  each  other. 

The  five  sites  differed  in  plant  spe- 
cies composition  based  on  the  pro- 
portion of  native-nonweedy,  native- 
weedy,  and  introduced-weedy  plant 
species  (fig.  1).  The  PL  site  had  the 
greatest  proportion  of  native-non- 
weedy species  (93.1  percent),  and  the 
D  site  had  the  lowest  (27.2  percent). 
Conversely,  the  PL  site  had  no  (0 
p>ercent)  native-weedy  species,  while 
the  D  site  had  the  highest  proportion 
(45.7  percent)  of  native-weedy  spe- 
cies. The  BL  site  has  the  greatest  pro- 
portion (33.3  percent)  of  introduced- 
weedy  species  found  at  any  site. 

Prairie  Dog  Burrow  Density 

The  mean  numbers  of  burrows  per 
0.005  ha  found  at  sites  HS,  SB,  UM, 
LM,  PL,  and  BL  are  shown  in  table  2. 
The  highest  burrow  density  was  at 
UM,  and  the  lowest  density  was  at 
BL.  These  differences  between  sites 
were  significant  (LSD  =  1 .62,  P  = 
0.05).  The  two  sites  from  the  Michel- 
bach colonies  (UM  and  LM)  had  sig- 
nificantly different  burrow  densities, 
even  though  these  two  sites  were 
within  1  km  of  one  another. 

Burrow  density  was  positively 
correlated  with  prairie  dog  density  at 
both  sites  (HS  and  SB)  where  prairie 
dog  densities  were  determined  and 
all  burrows  were  counted.  Burrow 
density  significantly  correlated  with 
prairie  dog  density  at  r  =  0.665,  ac- 
counting for  44.2  percent  of  the  vari- 
ance (F  =  10.32,  df  =  1, 13,  P  <  0.01). 

For  a  pooled  15  territories  at  the 
two  sites,  the  mean  burrow  density 
was  13.73  burrows  per  territory  (s  = 
8.3),  and  the  mean  number  of  prairie 
dogs  per  territory  was  6.4  (s  =  6.7). 
Consequently,  on  the  average,  there 
were  twice  as  many  burrows  as  prai- 
rie dogs  per  territory. 


Burrow  Density,  Evenness,  Plant 
Cover,  and  Plant  Species  Diversity 

Plant  cover  and  plant  species  diver- 
sity were  negatively  correlated  with 
burrow  density.  Multiple  regression 
analysis  with  burrow  density  as  the 
dependent  variable  and  plant  even- 
ness, percent  cover,  Simpson's  domi- 
nance, Shannon-Weaver  diversity, 
and  H  max  as  independent  variables 
was  significant  (F  =  5.25,  df  =  5,  7,  P 
<  0.05),  accounting  for  88.8  percent  of 
the  total  variance  in  burrow  density. 
Of  these,  evenness  (F  =  7.47),  percent 
cover  (F  =  10.37),  and  Shannon- 
Weaver  diversity  (F  =  7.39)  were  sig- 
nificant to  the  regression.  Evenness 
had  an  r  =  -0.416,  percent  cover  had 
an  r  =  -0.349,  and  Shannon- Weaver 
diversity  had  an  r  =  -0.427. 

Burrow  Density,  Native  Species, 
and  Introduced  Species 

Burrow  density  was  negatively  cor- 
related with  the  number  of  intro- 
duced-weedy  plant  species  (F  = 
18.14,  df  =  1, 10,  P  <  0.01).  Regression 
analysis  showed  that  burrow  density 
was  correlated  with  introduced- 
weedy  plant  species  at  r  =  -0.673,  ac- 
counting for  45.3  percent  of  the  vari- 
ance in  burrow  density. 

Burrow  density  was  not  signifi- 
cantly correlated  with  either  native- 
non weedy  species  or  native-weedy 
species  when  each  of  these  was  con- 
sidered as  an  independent  variable. 
However,  when  these  two  were  com- 
bined into  a  single  variable,  native 
species,  this  produced  a  highly  sig- 
nificant positive  correlation  of  r  = 
0.803  (F  =  18.14,  df  =  1, 10,  p  <  0.01), 
accounting  for  64.5  percent  of  the 
variance  in  burrow  density. 


Burrow  Density,  Plant  Species,  and 
Levels  of  Grazing 

Burrow  density  was  significantly  cor- 
related with  the  level  of  grazing  (r  = 


0.903,  F  =  17.8,  df  =  1, 4,  P  <  0.05). 
The  more  a  site  was  grazed,  the 
higher  was  the  burrow  density.  Re- 
gression analysis  showed  that  graz- 
ing levels  were  not  significantly  cor- 
related with  either  the  number  of  in- 
troduced species  or  the  number  of 
native  nonweedy  species  at  a  site. 
Grazing  was  significantly  correlated 
with  the  number  of  native  weedy 
species  (r  =  0.975,  F  =  37.9,  df  =  1,2,  P 
<  0.05),  and  weakly  correlated  with 
the  total  number  of  plant  species  (r  = 
0.947,  F  =  17.4,  df  =  1,2,  P  =  0.06). 
Multiple  regression  with  burrow 
density  as  the  dependent  variable 
and  native  species,  introduced  spe- 
cies, and  grazing  level  as  independ- 
ent variables  showed  that  native  spe- 
cies (number  of  native  weedy  and 
native  nonweedy  species  combined) 
explained  97.9  percent  of  the  vari- 
ance in  burrow  density,  while  graz- 
ing level  explained  an  additional  1 .8 
percent  and  introduced  species  ex- 
plained 0.2  percent. 

Discussion 

Our  results  show  that  Gunnison's 
prairie  dogs  thrive  at  sites  with  na- 
tive-nonweedy  and  native-weedy 
Sf)ecies  of  plants.  Gunnison's  prairie 
dogs  apparently  do  not  prefer  sites 
that  have  a  high  proportion  of  intro- 
duced-weedy  species.  This  is  not  sur- 
prising when  one  considers  the  die- 
tary requirements  of  these  animals. 
Shalaway  and  Slobodchikoff  (1988) 
found  that  the  diet  of  Gunnison's 
prairie  dogs  at  three  sites  in  the  Flag- 
staff area  consisted  primarily  of  na- 
tive plants:  native-weedy  and  native- 
nonweedy  species  made  up  60-80 
p)ercent  of  the  animals'  food.  Intro- 
duced-weedy  species  made  up  a  rela- 
tively low  proportion  of  the  diet  of 
Gunnison's  prairie  dogs  in  that 
study. 

Contrary  to  the  findings  of  studies 
with  blacktailed  prairie  dogs  (Ler- 
wick 1974,  Boddicker  and  Lerwick 

1976,  Gold  1976,  Hansen  and  Gold 

1977,  Beckstead  and  Schitoskey  1980, 


Archer  et  al.  1984),  Gunnison's  prai- 
rie dogs  did  not  increase  plant  spe- 
cies diversity,  but  instead  decreased 
it.  This  effect  can  be  produced  by  the 
clipping  action  of  prairie  dogs  on 
plants  that  tend  to  grow  tall  and  ob- 
scure the  animals'  view  of  terrestrial 
predators.  Such  clipping  action  can 
lower  the  competitive  ability  of 
shrubs  and  other  tall  plants,  eventu- 
ally eliminating  them  from  prairie 
dog  towns.  Many  of  these  species  are 
introduced  weedy  plants.  A  similar 
effect  was  described  by  Clements 
and  Clements  (1940)  with  Gunnison's 
prairie  dogs. 

The  effects  of  Gunnison's  prairie 
dogs  on  plant  cover  were  consistent 
with  those  found  by  other  studies 
(Knowles  1982,  Archer  et  al.  1984).  In 
each  case,  prairie  dogs  decreased 
plant  cover.  This  is  to  be  expected, 
since  all  species  of  prairie  dogs  graze 
on  vegetation  and  can  eat  up  some 
24-90  percent  of  the  primary  produc- 
tion of  a  site  (Osbom  and  Allan  1949, 
Hansen  and  Gold  1977,  Crocker- 
Bedford  and  Spillett  1981).  To  the 
extent  that  blacktailed  prairie  dogs 
and  cattle  have  a  dietary  overlap  of 
76  percent  (Kelso  1939),  prairie  dogs 
have  been  construed  as  competitors 
of  large  herbivores  such  as  cattle. 
However,  because  prairie  dogs  feed 
very  selectively  on  plants,  80  percent 
of  the  biomass  they  ingest  may  come 
from  plant  parts  not  utilized  by  cattle 
(Crocker-Bedford  1976).  Also,  any 
potential  competitive  effect  might  be 
minimized  by  the  relatively  small 
size  of  most  extant  prairie  dog  colo- 
nies (King  1955;  Koford  1958;  Smith 
1955),  and  the  beneficial  effects  that 
large  herbivores  may  obtain  from 
plants  that  grow  in  prairie  dog  colo- 
nies (Coppock  et  al.  1983a). 

The  positive  correlation  between 
grazing  level  and  density  of  prairie 
dog  burrows  suggests  that  prairie 
dogs  are  found  more  in  habitats  that 
are  highly  grazed.  However,  merely 
addressing  prairie  dog  management 
in  terms  of  possible  comp)etition  with 
cattle  misses  a  much  more  funda- 
mental issue:  that  of  the  prairie  dog' s 


406 


place  in  a  natural  ecosystem.  While 
our  study  has  found  a  positive  corre- 
lation between  prairie  dog  densities 
and  grazing,  the  presence  of  these 
animals  at  ungrazed  sites  indicates 
that  they  can  establish  themselves  in 
ungrazed  areas  that  have  the  right 
configuration  of  habitat  characteris- 
tics. 

A  much  more  important  point 
than  grazing  is  the  strong  link  be- 
tween the  presence  of  prairie  dogs 
and  the  success  of  native  species  of 
plants.  Introduced  weeds  are  not  fa- 
vored in  prairie  dog  colonies,  even 
though  the  soil  is  disturbed  through 
the  burrowing  actions  of  these  ani- 
mals. Rather  than  being  "weed/' 
pests  who  come  into  overgrazed 
lands,  prairie  dogs  might  actually 
have  the  function  of  repairing  over- 
grazed land,  and  driving  the  plant 
community  toward  a  more  natural 
one. 

The  mechanism  for  how  prairie 
dogs  might  drive  the  ecosystem  to- 
ward more  native  plant  species  is 
still  unclear.  We  have  found  that 
Gunnison's  prairie  dogs  decrease 
both  species  diversity  and  plant 
cover.  The  decrease  in  species  diver- 
sity apparently  comes  from  a  de- 
crease in  the  component  represented 
by  the  introduced  weedy  plant  spe- 
cies, and  not  from  the  native  plant 
species.  The  decrease  in  plant  cover 
comes  from  herbivory  on  the  plants 
growing  in  the  colonies.  Some  native 
plant  species  produce  more  flower- 
ing stalks  and  more  seeds  when  they 
are  grazed  by  herbivores  (Paige  and 
Whitham  1987).  Experimental  evi- 
dence for  black-tailed  prairie  dogs 
shows  that  both  forbs  and  grasses 
increased  in  plots  that  contained  both 
prairie  dogs  and  cows  (Uresk  and 
Bjugstad  1980).  In  the  arid  conditions 
of  the  Southwest,  native  plants  might 
be  better  adapted  to  climatic  condi- 
tions than  introduced  weedy  species, 
and  might  respond  to  herbivory  by 
increasing  their  numerical  abun- 
dance. The  relationship  that  we 
found  between  levels  of  grazing  and 
prairie  dog  burrow  densities  may  be 


the  result  of  herbivory  stimulating 
the  growth  of  plants  necessary  to  the 
diet  of  Gunnison's  prairie  dogs. 

Our  results  suggest  that  Gunni- 
son's prairie  dogs  must  be  conserved 
by  maintaining  habitats  with  a  large 
component  of  native  vegetation. 
Gunnison's  prairie  dogs  are  a  natural 
part  of  native  ecosystems,  and  have 
evolved  alongside  large  herbivores 
such  as  elk,  deer,  and  buffalo,  all  of 
which  feed  to  some  extent  on  native 
species  of  grasses  and  forbs.  Native 
plant  species  have  evolved  to  com- 
pensate for  these  effects  of  herbivory, 
and  possibly  for  this  reason  prairie 
dogs  might  have  a  beneficial  function 
of  restoring  rangeland  that  has  been 
damaged  by  grazing;  this  is  a  man- 
agement question  that  must  be  ad- 
dressed experimentally  in  the  future. 
In  addition  to  the  positive  association 
between  prairie  dogs  and  native 
plant  species,  prairie  dog  towns  are 
habitat  sites  that  are  integral  to  the 
existence  of  large  numbers  of  other 
vertebrates  and  invertebrates,  and 
eradication  of  prairie  dogs  can  have 
detrimental  consequences  to  natural 
ecosystems.  Experimental  and  eco- 
nomic evidence  currently  indicates 
that  eradication  of  prairie  dogs  is  nei- 
ther economically  feasible  nor  par- 
ticularly beneficial  to  cattle.  We  sug- 
gest that  prairie  dogs  should  be 
looked  at  in  a  more  positive  role  that 
reflects  their  impact  on  the  mainte- 
nance of  natural  ecosystems. 

Acknowledgments 

We  thank  Jim  Benedix,  Ed  Creef,  Ch- 
eryl Fischer,  Kitty  Gehring,  Gene 
Hickman,  and  Steve  Travis  for  assis- 
tance in  data  collection  and  summa- 
rization. We  also  thank  Judith  Ki- 
riazis  for  her  helpful  and  insightful 
comments  on  this  manuscript. 

Literature  Cited 

Agnew,  W.  1983.  Flora  and  fauna  as- 
sociated with  prairie  dog  ecosys- 


tems. M.  S.  Thesis,  Colorado  State 
University,  Fort  Collins,  47  p. 

Archer,  S.  R.,  I.  D.  Lebreton,  M.  G. 
Garrett,  and  J.  K.  Detling.  1984. 
Structural  changes  in  a  mixed- 
grass  prairie  dog  community  as  a 
function  of  prairie  dog  coloniza- 
tion history.  Bulletin  Ecological 
Society  of  America  65:162. 

Beckstead,  M.  and  F.  Schitoskey,  Jr. 

1980.  Assimilation  efficiency  of  the 
black-tailed  prairie  dog.  Proceed- 
ings South  Dakota  Academy  of 
Science  59:184-194. 

Boddicker,  M.  J.  and  A.  Lerwick. 
1976,  Vegetation  changes  induced 
by  prairie  dogs  on  shortgrass 
range.  Journal  of  Range  Manage- 
ment 29:221-225. 

Brizuela,  M.  A.,  J.  K.  Detling,  and  M. 
S.  Cid.  1984.  Seasonal  silicon  con- 
tent of  grasses  from  sites  receiving 
different  grazing  pressure  by 
grassland  herbivores.  Bulletin  Eco- 
logical Soc.  of  Amer.  65:161-162. 

Campbell,  T.  M.,  Ill,  and  T.  W.  Clark. 

1981.  Colony  characteristics  and 
vertebrate  associates  of  white- 
tailed  and  black-tailed  prairie 
dogs  in  Wyoming.  American  Mid- 
land Naturalist  105:269-275. 

Clark,  T.  W.,  T.  M.  Campbell,  III,  D. 
C.  Socha,  and  D.  E.  Casey.  1982. 
Prairie  dog  colony  attributes  and 
associated  vertebrate  species. 
Great  Basin  Naturalist  42:572-582. 

Clements,  F.  E.  and  E.  S.  Clements. 
1940.  The  biotic  significance  of  dis- 
turbance. Climate,  Climax,  and 
Conservation:  Carnegie  Institute 
of  Washington  Yearbook  39:174- 
175. 

Collins,  A.  R.,  J.  P.  Workman,  and  D. 
W.  Uresk.  1984.  An  economic 
analysis  of  black-tailed  prairie  dog 
(Cynomys  ludovicianus)  control. 
Journal  of  Range  Management 
37:358-361. 

Coppock,  D.  L.,  J.  K.  Detling,  and  M. 
I.  Dyer.  1980.  Interactions  among 
bison,  prairie  dogs  and  vegetation 
in  Wind  Cave  National  Park.  Final 
Report  to  National  Park  Service, 
Wind  Cave  National  Park,  Hot 
Springs,  South  Dakota.  177  p. 


407 


Coppock,  D.  L.,  J.  K.  Detling,  J.  E. 
Ellis,  and  M.  I.  Dyer.  1983a.  Plant- 
herbivore  interactions  in  a  North 
American  mixed-grass  prairie.  I. 
Effects  of  black-tailed  prairie  dogs 
on  interseasonal  aboveground 
plant  biomass  and  nutrient  dy- 
namics and  plant  species  diver- 
sity. Oecologia  56:1-9. 

Coppock,  D.  L.,  J.  E.  Ellis,  J.  K.  Det- 
ling, and  M.  I.  Dyer.  1983b.  Plant- 
herbivore  interactions  in  a  North 
American  mixed-grass  prairie.  II. 
Responses  of  bison  to  modification 
of  vegetation  by  prairie  dogs. 
Oecologia  56:10-15. 

Crocker-Bedford,  D.  C.  1976.  Food 
interactions  between  Utah  prairie 
dogs  and  cattle.  M.  S.  Thesis.  Utah 
State  University,  Logan.  131  p. 

Crocker-Bedford,  D.  C.  and  J.  J. 
Spillett.  1977.  Home  ranges  of 
Utah  prairie  dogs.  Journal  of 
Mammalogy  58:672-673. 

Detling,  J.  K.  and  E.  L.  Painter.  1983. 
Defoliation  responses  of  western 
wheatgrass  populations  with  di- 
verse histories  of  prairie  dog  graz- 
ing. Oecologia  57:65-71. 

Fagerstone,  K.  A.  1981.  A  review  of 
prairie  dog  diet  and  its  variability 
among  animals  and  colonies,  p. 
178-184.  In  R.  M.  Timm  and  R.  J. 
Johnson,  eds.  Proceedings  Fifth 
Great  Plains  Wildlife  Damage 
Control  Workshop.  Institute  of 
Agriculture  and  Natural  Re- 
sources, University  of  Nebraska, 
Lincoln. 

Gold,  I.  K.  1976.  Effects  of  blacktail 
prairie  dog  mounds  on  shortgrass 
vegetation.  MS  Thesis,  Colorado 
State  University,  Fort  Collins, 
Colorado. 

Hansen,  R.  M.  and  I.  K.  Gold.  1977. 
Blacktail  prairie  dogs,  desert  cot- 
tontails and  cattle  trophic  relations 
on  shortgrass  range.  Journal  of 
Wildlife  Management  30:210-214. 

Hassien,  F.  1976.  A  search  for  black- 
footed  ferrets  in  the  Oklahoma 
panhandle  and  adjacent  area  and 
an  ecological  study  of  black-tailed 
prairie  dogs  in  Texas  County, 
Oklahoma.  M.  S.  Thesis,  Okla- 


homa State  University,  Stillwater. 
Ill  p. 

Kelso,  L.  H.  1939.  Food  habits  of 
prairie  dogs.  USDA  Circular 
Number  529. 

King,  J.  A.  1955.  Social  behavior,  so- 
cial organization,  and  population 
dynamics  in  a  black-tailed  prairie 
dog  town  in  the  Black  Hills  of 
South  Dakota.  Contributions  of 
the  Laboratory  of  Vertebrate  Biol- 
ogy, University  of  Michigan  67:1- 
123. 

King  J.  A.  1984.  Historical  ventila- 
tions on  a  prairie  dog  town.  p. 
447-456.  In  J.  O.  Murie  and  G.  R. 
Michener,  eds.  The  biology  of 
ground-dwelling  squirrels.  Uni- 
versity of  Nebraska  Press,  Lincoln, 
Nebraska. 

Klatt,  L.  E.  and  D.  Hein.  1978.  Vege- 
tative differences  among  active 
and  abandoned  towns  of  black- 
tailed  prairie  dogs  (Cynomys  ludo- 
vicianus).  Journal  of  Rangeland 
Management  31:315-317. 

Koford,  C.  B.  1958.  Prairie  dogs, 
whitefaces,  and  blue  grama.  Wild- 
life Monographs  3:1-78. 

Knowles,  C.  J.  1982.  Habitat  affinity, 
populations,  and  control  of  black- 
tailed  prairie  dogs  on  the  Charles 
M.  Russell  National  Wildlife  Ref- 
uge. Ph.  D.  Dissertation,  Univer- 
sity of  Montana,  Missoula.  171  p. 

Lerwick,  A.  C.  1974.  The  effects  of 
the  black-tailed  prairie  dog  on 
vegetative  com}X)sition  and  their 
diet  in  relation  to  cattle.  M.S.  The- 
sis, Colorado  State  University, 
Fort  Collins.  106  p. 

Nie,  N.  H.,  C.  H.  Hull,  J.  G.  Jenkins, 
K.  Steinbrenner,  and  D.  H.  Bent. 
1975.  SPSS:  Statistical  package  for 
the  social  sciences,  ed.  2,  McGraw- 
Hill,  New  York. 

O'Meilia,  M.  E.  1980.  Competition 
between  prairie  dogs  and  beef 
cattle  for  range  forage.  M.S.  The- 
sis, Oklahoma  State  University, 
Stillwater.  32  p. 

Osbom,  B.  and  P.  F.  Allan.  1949. 
Vegetation  of  an  abandoned  prai- 
rie dog  town  in  tallgrass  prairie. 
Ecology  30:322-332. 


Paige,  K.  N.  and  T.  G.  Whitham. 
1987.  Overcompensation  in  re- 
sponse to  mammalian  herbivory: 
the  advantage  of  being  eaten. 
American  Naturalist  129:407-416. 

Poole,  R.  W.  1974.  An  introduction  to 
quantitative  ecology.  McGraw- 
Hill,  N.  Y.  532  p. 

Severe,  D.  S.  1977.  Re  vegetation  of 
black-tailed  prairie  dog  mounds 
on  shortgrass  prairie  in  Colorado. 
M.S.  Thesis,  Colorado  State  Uni- 
versity, Fort  Collins.  92  p. 

Shalaway,  S.  and  C.  N.  Slobod- 
chikoff.  1988.  Seasonal  change  in 
the  diet  of  the  Gunnison's  prairie 
dog.  Journal  of  Mammalogy,  in 
press. 

Slobodchikoff,  C.  N.  and  R.  Coast. 
1980.  Dialects  in  the  alarm  calls  of 
prairie  dogs.  Behavioral  Ecology 
and  Sociobiology  7:49-53. 

Smith,  R.  E.  1955.  Natural  history  of 
the  prairie  dog  in  Kansas.  Univer- 
sity of  Kansas  Museum  Natural 
History  and  State  Biological  Sur- 
vey Publication,  Number  16. 

Uresk,  D.  W.  1985.  Effects  of  control- 
ling black-tailed  prairie  dogs  on 
plant  production.  Journal  of 
Rangeland  Management  38:466- 
468. 

Uresk,  D.  W.  and  A.  J.  Bjugstad. 
1980.  Prairie  dogs  as  ecosystem 
regulators  on  the  Northern  High 
Plains,  p.  91-94.  In  C.  L.  Kucera, 
ed.  Proceedings  7th  North  Amer- 
ica Prairie  Conference.  Southwest- 
ern Missouri  State  University, 
Springfield. 


408 


Environmental  Contanninants 
and  the  Management  of  Bat 
Populations  in  the  United 
States^ 

Donald  R.  Clark,  Jr.^ 


Abstract. —Food-chain  Residues  of  orgono- 
chlorine  pesticides  probably  have  been  involved  in 
declines  of  some  U.S.  Bat  populations;  examples  in- 
clude free-toiled  bats  at  Carlsbad  Cavern.  New 
Mexico,  and  the  endangered  gray  bat  at  sites  in 
Missouri  and  Alabama.  If  a  long-lived  contaminant 
has  not  been  dispersed  in  large  amounts  over  large 
areas,  its  impact  may  be  controlled  by  administra- 
tive action  that  stops  its  use  or  other  environmental 
discharge,  or  that  results  in  physical  isolation  of  local- 
ized contamination  so  that  it  no  longer  enters  food 
chains. 


Several  species  of  bats  in  the  U.S. 
Form  large  aggregations  in  caves,  old 
mines,  or  other  shelters,  and  many  of 
these  colonies  are  of  management 
concern  to  biologists  working  for  the 
states  or  federal  government  (e.g. 
Prichard  1987).  Four  taxa,  the  gray 
bat  (Myotis  grisescens),  Indiana  bat 
(M.  sodalis),  Ozark  big-eared  bat  (Ple- 
cotus  townsendii  ingens),  and  Virginia 
big-eared  bat  (P.  t.  virginianus),  are  of 
particular  concern  because  they  are 
endangered  (USDI,  FWS  1987). 

Habitat  destruction  such  as  defor- 
estation, water  pollution,  stream 
channelization,  and  stream  sedimen- 
tation (Tuttle  1979,  Prichard  1987)  or 
direct  human  disturbance  and  de- 
struction of  bats  (Tuttle  1979,  for  a 
recent  example  see  Anon.  1987)  are 
primary  known  threats  to  bat  colo- 
nies. However,  environmental  con- 
taminants, such  as  organochlorine 
pesticide  residues  and  heavy  metals, 
probably  have  been  involved  in  some 
declines  of  bat  populations.  In  this 
paper  I  discuss  the  management  im- 
plications of  these  contaminants. 
(Note:  for  purposes  of  this  discus- 
sion, "management'  refers  broadly 
to  human  activities  undertaken  in  the 
interest  of  a  bat  colony  with  the  goal 

' Paper  presented  at  symposium  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortf^  America.  (Flag- 
staff .  AZ.  July  19-21,  1988.) 

^Donald  R.  Clark,  Jr.,  is  Researcti  Wildlife 
Biologist,  U.S.  fish  and  Wildlife  Service, 
Patuxent  Wildlife  Researcti  Center,  Laurel, 
MD  20708. 


that  colony  size  will  remain  at  a 
steady,  sustainable  level  or  will  in- 
crease to  such  a  level.) 

Examples  of  Possible  Food-Ctioin 
Contaminant  Impacts  on  Bat 
Populations 

Free-Tailed  Bats  at  Carlsbad 
Cavern,  New  Mexico 

The  Carlsbad  population  of  Mexican 
free-tailed  bats  (Tadarida  brasiliensis 
mexicam)  was  estimated  at  8.7  mil- 
Hon  bats  in  1936  (Allison  1937)  but 
only  2(X),(X)0  bats  remained  in  1973 
(Altenbach  et  al.  1979).  Several  die- 
offs  occurred  during  this  interval 
(Altenbach  et  al.  1979),  and  none  was 
linked  directly  to  pesticide  poison- 
ing; however,  routine  testing  of  tis- 
sues was  not  available.  The  question 
of  pesticide  involvement  was  ad- 
dressed by  simulating  migratory 
flight  in  young  bats  taken  from  the 
colony  in  1974  (Geluso  et  al.  1976). 
Some  of  these  bats  died  of  DDE  (1,1'- 
(dichloroethylidene)bis[4-chloroben- 
zene])  poisoning  (DDE  is  the  princi- 
pal metabolite  of  DDT;  1,1 '-2,2,2- 
(tricholoroethylidene)bis[4-chloro- 
benzene])  due  to  mobilization  of 
DDE  received  in  their  mother's  milk 
and  stored  in  their  fat  (Geluso  et  al. 
1976).  This  result  suggests  that  DDT 
has  contributed  to  the  decline  of  this 
population. 

High  DDE  concentrations  in  the 
Carlsbad  colony  probably  resulted 


from  heavy  DDT  use  in  New  Mexico 
before  its  ban  in  1972;  however,  other 
more-recent  inputs  have  been  postu- 
lated to  explain  high  DDE  levels  in 
wildlife  in  parts  of  Texas,  New  Mex- 
ico, and  Arizona  (Clark  and  Krynit- 
sky  1983,  Hunt  et  al.  1986,  White  and 
Krynitsky  1986). 

Gray  Bats  in  Missouri 

Dieldrin  (3,4,5,6,9,9-hexachloro- 
la,2,2a,3,6,6a,7,7a-octahydro-2,7:3,6- 
dimethanonaphth[23-b]oxirene) 
killed  gray  bats  in  1976, 1977,  and 
1978  in  two  maternity  colonies  in 
Franklin  County,  Missouri  (Clark  et 
al.  1978b,  1983a).  Residues  of  hep- 
tachlor-related  chemicals 
(1,4,5,6,7,8,8 -heptachloro-3a,4,7,7a- 
tetrahydro-4,7-methanoindene)  in 
bats  from  both  colonies  increased  to 
potentially  dangerous  concentrations 
in  1977  and  remained  elevated  in 
1978  (Clark  et  al.  1983a).  Populadon 
size  at  one  colony  was  estimated  at 
1,800  bats  in  1976  and  1978,  but  no 
bats  were  present  from  1979-82 
(Clark  et  al.  1983a,b).  Dieldrin,  per- 
haps in  conjunction  with  heptachlor, 
may  have  caused  the  decline  and  dis- 
appearance of  this  colony.  Dieldrin 
also  killed  gray  bats  at  three  Boone 
County,  Missouri,  caves  in  1980, 
1981,  and  1982  (Clark  et  al.  1983b, 
Clawson  and  Clark  in  manuscript). 

Death  of  gray  bats  were  attributed 
to  dieldrin  because  this  chemical  was 
measured  in  the  bats'  brains  at  con- 


409 


centrations  known  to  be  lethal  in 
other  species  (Clark  et  al.  1978b).  Di- 
eldrin  and  heptachlor-related  resi- 
dues came  from  the  use  of  aldrin 
(Dieldrin's  parent  compound)  and, 
subsequently,  heptachlor,  to  control 
cutworms  (moth  larvae.  Family  Noc- 
tuidae)  in  corn. 


Gray  Bats  at  Cave  Springs  Cave, 
Alabama 

DDT  was  manufactured  at  Redstone 
Arsenal  near  Huntsville,  Alabama, 
from  1947  to  1970,  and  massive 
amounts  of  DDT  and  its  metabolites 
(DDD;  l,l'-(2,2-dichloroethylidene)- 
bis[4-chlorobenzene]  and  DDE)  were 
discharged  into  the  Tennessee  River 
via  Huntsville  Spring  Branch-Indian 
Creek  (Fleming  and  Atkeson  1980). 
Local  biota  remains  heavily  contami- 
nated (O'Shea  et  al.  1980,  Heming 
and  Cromartie  1981,  Fleming  et  al. 
1984,  Reich  et  al.  1986.). 

Samples  of  dead  or  dying  bats  and 
bat  guano  collected  between  1976 
and  1986  from  four  gray  bat  colonies 
as  far  as  140  km  downriver  contained 
residues  from  this  former  discharge 
(Clark  et  al.  1988).  Residues  were 
identifiable  by  their  high  DDD  to 
DDE  ratio,  which  resulted  from  their 
breakdown  under  anaerobic  condi- 
tions. Cave  Springs  Cave  at  Wheeler 
National  Wildlife  Refuge  houses  the 
colony  nearest  the  contaminant 
source — about  20  km.  Biologists 
judged  that  bat  mortality  at  Cave 
Springs  Cave  was  far  above  normal 
in  1978, 1985,  and  1986.  Residues  of 
DDT,  DDD,  and  DDE  in  brains  of 
dead  or  dying  bats  from  this  cave, 
although  elevated  in  comparison 
with  residues  from  colonies  up- 
stream from  Redstone  Arsenal,  were 
well  below  concentrations  believed 
to  be  lethal  (Clark  et  al.  1988).  The 
single  exception  was  a  bat  collected 
in  1978  with  sufficient  DDD  in  its 
brain  (29  ppm  wet  weight)  to  have 
been  poisoned  (Clark  et  al.  1988).  The 
measured  residues,  therefore,  did  not 
explain  the  observed  mortalities. 


Although  there  is  no  explanation 
for  this  mortality  yet,  another  con- 
taminant may  by  involved.  A  guano 
sample  collected  from  Cave  Springs 
Cave  in  1987  was  analyzed  for  heavy- 
metals  and  cadmium  measured  8.5 
Ppm  (dry  weight).  This  amount  may 
be  compared  with  2.2  Ppm  cadmium 
in  guano  (mixed  gray  and  southeast- 
ern bats,  M.  austroriparius)  from  a 
Florida  cave  where  the  bats  were  ex- 
posed to  contaminations  from  a  bat- 
tery salvage  plant.  Kidneys  of  south- 
eastern bats  from  this  Florida  cave 
averaged  0.89  Ppm  (wet  weight)  cad- 
mium with  a  maximum  of  2.9  Ppm. 
Concentrations  of  cadmium  as  low  as 
3.4  Ppm  in  kidneys  of  voles  (Microtus 
pennsylvanicus)  were  associated  with 
reduced  survivorship  in  enclosed 
populations.  Also,  six  gray  bats 
found  dead  in  Cave  Springs  Cave  in 
June  1986  were  examined  by  the  U.S. 
Fish  and  Wildlife  Service's  National 
Wildlife  Health  Research  Center, 
Madison,  Wisconsin.  There  was  no 
evidence  of  injury  or  infectious  dis- 
ease, but  all  bats  showed  mild  renal 
tubular  degeneration.  Because  cad- 
mium caused  kidney  damage  (Nomi- 
yama  1981),  this  metal,  perhaps  in 
combination  with  DDD  and  DDE, 
may  have  caused  the  recent  die-off  of 
gray  bats  at  Cave  Springs  Cave.  The 
cadmium  source  is  unknown.  Addi- 
tional samples  for  chemical  analysis 
will  be  collected  in  1988. 


Managerrient  of  Contaminant 
Impacts  on  Bat  Populations 

Screening  for  Possible 
Contaminant  Problems  in 
Apparently  Healttiy  Colonies 

Contaminants  that  biomagnify  or 
bioaccumulate  in  ecosystems  include 
organochlorine  pesticides  such  as 
DDT  (and  its  metabolites  DDE  and 
DDD),  dieldrin,  heptachlor-related 
chemicals,  and  the  industrial  poly- 
chlorinated  biphenyls  (PCBs).  Also 
included  are  heavy  metals  such  as 
lead,  cadmium,  chromium,  zinc,  and 


mercury.  For  chemicals  that  biomag- 
nify or  bioaccumulate,  analyses  of 
guano  samples  collected  from  the 
surface  of  a  guano  deposit  can  indi- 
cate body  burdens  in  bats  during 
their  most  recent  activity  season. 
Samples  from  greater  depths  may 
indicate  contaminant  concentrations 
in  previous  years. 

Relationships  between  concentra- 
tions in  guano  and  carcasses  of  bats 
from  the  same  colony  have  been  de- 
scribed for  dieldrin,  heptachlor  epox- 
ide, and  DDE  (Clark  et  al.  1982). 
Limited  data  are  available  on  concen- 
trations of  lead,  cadmium,  chro- 
mium, zinc,  and  mercury  in  guano 
from  contaminated  colonies  (Petit 
and  Altenbach  1973,  Clark  1979, 
Clark  et  al.  1986,  this  paper).  About 
20  grams  of  guano,  dry  weight,  are 
necessary  for  analyses. 

Sublethal  exposure  of  bats  to  the 
newer  organophosphorus  and  car- 
bamate insecticides  is  demonstrated 
by  depressed  brain  cholinesterase 
(ChE)  activity  in  exposed  individu- 
als. Depression  is  determined  by 
comparison  to  normal  ChE  activity 
for  a  sample  of  control  bats  of  the 
same  species.  Measurement  of  ChE 
activity  (for  methods,  see  Ellman  et 
al.  1961,  Hill  and  Fleming  1982)  in- 
volves removal  of  the  brain,  hence 
death  of  the  bat. 


Recognizing  Organoctilorine 
Pesticide- Induced  Mortality  in  Bat 
Colonies 

Managed  colonies  are  usually  cen- 
sused  annually  so  that  any  significant 
decline  will  be  recognized.  By  also 
estimating  numbers  of  dead  and 
dying  bats  at  these  censuses,  manag- 
ers can  differentiate  between  "nor- 
mal" mortality  and  increased  mortal- 
ity, which  may  be  the  first  sign  of  a 
contaminant  problem. 

May  of  the  colonies  considered 
most  important  are  maternity  colo- 
nies, and  in  maternity  colonies,  or- 
ganochlorine chemicals  kill  mostly 
young  bats.  There  are  two  reasons 


410 


for  this.  First,  organochlorines  be- 
come concentrated  in  the  fat  of 
mother's  milk  and  these  chemicals 
continually  and  rapidly  accumulate 
in  the  young  as  they  nurse. 

For  example,  insects  collected  in 
foraging  areas  of  Missouri  gray  bats 
contained  a  maximum  of  3.1  Ppm 
(wet  weight)  dieldrin,  but  milk  taken 
from  the  stomach  of  a  young  dead 
gray  bat  contained  89  ppm  (wet 
weight)  dieldrin  (Clark  and  Prouty 
1984).  Second,  young  bats  are  1.9 
Times  more  sensitive  than  adults  to 
dieldrin  and  1.5  Times  more  sensitive 
to  DDT  (Clark  et  al.  1978a,  1983a). 
Young  bats  dying  of  organochlorine 
poisoning  may  still  have  milk  in  their 
stomachs  unlike  young  dying  of  star- 
vation. Therefore,  increased  infant 
mortality  in  a  maternity  colony  with 
some  young  having  milk  in  their 
stomachs  may  indicate  poisoning  by 
an  organochlorine  chemical. 

Diagnosing  Chemical  Poisoning  in 
Bats 

Diagnosis  for  organochlorine  chemi- 
cals requires  analyses  of  brains  and 
interpretation  of  the  resulting  meas- 
urements. However,  because  concen- 
trations in  brains  are  closely  corre- 
lated with  concentrations  in  carcass 
fat  (Clark  1981a),  analyses  of  car- 
casses may  serve  if  brains  are  un- 
available. For  example,  analysis  of 
carcasses  may  be  the  only  option 
when  bats  are  partly  decomposed. 
Correlations  between  brain  and  car- 
cass fat  concentrations  only  have 
been  quantified  for  DDE,  DDT,  and 
dieldrin  (Clark  1981a). 

Lethal  brain  concentrations  for 
DDE,  DDT,  dieldrin,  and  PCB  (Aro- 
clor  1260)  have  been  determined  for 
at  least  one  sp>ecies  of  bat  (Clark 
1981b).  Because  lethal  brain  levels  are 
fairly  similar  among  mammals  and 
birds,  comparisons  can  provide  clues 
about  the  effect  on  a  populations, 
even  though  the  lethal  level  for  the 
species  under  investigation  has  not 
been  determined  yet. 


Diagnosis  of  death  in  bats  from 
heavy-metal  poisoning  is  less  certain, 
but  interpretations  often  can  be  made 
based  on  other  species  of  mammals 
(Clark  1979,  this  paper).  Diagnosis 
for  heavy  metals  involves  analyzing 
liver  and  kidneys  along  with  histo- 
logical examination  for  damage. 

Death  in  bats  caused  by  the  anti- 
cholinesterase insecticides  could  be 
diagnosed  by  measurement  of  de- 
pressed brain  ChE  in  combination 
with  detection  of  an  anticholinester- 
ase chemical  in  the  contents  of  the 
gastrointestinal  tracts  or  other  tissues 
of  the  affected  bats.  Lethal  depres- 
sion of  brain  ChE  has  been  measured 
in  little  brown  bats  (M.  lucifugus)  in 
the  laboratory  for  methyl  parathion 
(phosphorothioic  acid  0,0-dimethyl 
(3-(4-nitrophenyl)ester)  and 
Orthene®  (acephate;  acetylphosphor- 
amidothioic  acid  0,S-dimethyl  ester) 
(Clark  1986,  Clark  and  Rattner  1987). 

Even  though  a  firm  diagnosis  of 
contaminant-induced  mortality  re- 
quires tissue  analyses,  analysis  of  a 
guano  sample,  as  a  first  step,  may 
indicate  whether  organochlorines  or 
metals  are  involved. 

Chemical  analyses  of  tissues  or 
guano  are  not  something  that  manag- 
ers usually  can  perform  themselves. 
However,  an  Environmental  Con- 
taminant Field  Specialist  from  the 
U.S.  Fish  and  Wildlife  Service  can  be 
contacted  (there  are  1-3  in  each 
state);  if  he  or  she  determines  that  the 
situation  warrants,  analyses  can  be 
done.  The  Specialist  also  may  send 
specimens  to  the  National  Wildlife 
Flealth  Research  Center  if  disease  is 
suspected. 

Bat  specimens  for  diagnostic  study 
generally  should  be  frozen  immedi- 
ately. However,  examinations  for 
diseases  and  histopathology  require 
that  specimens  be  kept  refrigerated 
but  not  frozen  until  organs  can  be 
removed  and  preserved  in  fluid. 
Control  specimens  of  the  same  spe- 
cies are  necessary  for  diagnosis  of 
depressed  brain  ChE  activity.  Guano 
does  not  require  freezing  or  refrig- 
eration. The  Contaminant  Field  Spe- 


cialist can  provide  detailed  instruc- 
tions for  specimen  collection  and 
handling. 

Possible  Impacts  of  New 
Generation  Pesticides  on  Bat 
Colonies 

Most  organochlorine  p)esticides  have 
been  banned  or  their  use  otherwise 
reduced  in  the  U.S.,  And  some  wild- 
life-related problems  have  improved. 
Organochlorines  largely  have  been 
replaced  by  organophosphorus  (e.g., 
Acephate,  diazinon  [phosphorothioic 
acid  0,0-diethyl  0-[6-methyl-2-(l- 
methylethyl)-4-pyrimidinyl]ester], 
and  methyl  parathion)  and  car- 
bamate (e.g.,  Aldicarb  [2-methyl-2- 
(methylthio)propanal  0-[(methyl- 
amino)carbonyl]oxime],  carbaryl  [1- 
naphthalenol  methylcarbamate],  and 
carbofuran  [2,3-dihydro-2,2-di- 
methyl-7-benzofuranol  methylcar- 
bamate]) insecticides.  These  chemi- 
cals are  relatively  short-lived  and 
generally  do  not  accumulate  in  food 
chains.  Exposure  in  bats  probably 
occurs  when  they  feed  over  fields  or 
orchards  that  are  being,  or  have  just 
been,  sprayed.  In  these  cases,  bats 
might  be  sprayed  directly  and  re- 
ceive the  chemical  through  their  skin 
and  lungs.  Pesticides  are  frequently 
sprayed  in  the  evening,  at  night,  or 
early  in  the  morning  to  avoid  killing 
honey  bees,  to  kill  adult  mosquitoes, 
or  to  take  advantage  of  quiet  wind 
conditions  and  thereby  avoid  drift. 
Bats  also  may  be  exposed  by  eating 
insects  that  have  just  been  sprayed 
but  are  still  alive. 

New-generation  pesticides  have 
not  yet  been  linked  to  bat  die-offs, 
but,  in  1968,  ranchers  and  farmers  in 
a  cotton-growing  area  of  Arizona  re- 
ported "...unusual  Numbers  of  dead 
or  dying  (free-tailed)  bats  in  their 
fields.. .Many  Were  found  convuls- 
ing, incapable  of  flight"  (Reidinger 
and  Cockrum  1978).  This  mortality 
was  attributed  to  DDT;  however, 
chemical  analyses  indicated  that  nei- 
ther lethal  residues  of  DDT  nor  its 


411 


metabolites  had  been  present  in  these 
bats  (Clark  1981b).  Because  methyl 
parathion  also  was  commonly  used 
on  cotton  in  this  region,  mortality 
may  have  been  caused  by  this  or- 
ganophosphorus  pesticide.  The  mor- 
tality pattern  described  by  ranchers 
and  farmers  where  bats  were  scat- 
tered on  the  ground  in  an  incapaci- 
tated condition  suggests  quick  intoxi- 
cation after  direct  contact  with  a 
chemical  of  high  acute  toxicity  such 
as  the  organophosphate  methyl  para- 
thion (see  Clark  1986). 


Reducing  Contaminant  Impacts  in 
Bat  Colonies 

What  can  be  done  once  it  is  deter- 
mined that  bats  have  died  from  a 
food-chain  contaminant?  The  answer 
will  depend  on  the  contaminant,  its 
source,  and  on  the  ability  or  author- 
ity of  the  manager  to  change  local 
practices  or  obtain  cleanup  proce- 
dures. 

When  large  quantities  of  a  long- 
lived  chemical  have  been  incorpo- 
rated into  soils  over  vast  areas,  such 
as  DDE  in  New  Mexico  or  dieldrin  in 
Missouri,  the  chemical  will  continue 
to  enter  food  chains  for  many  years. 
The  manager  of  an  affected  bat  col- 
ony can  only  protect  the  colony  form 
other  sources  of  damage  and  hope 
that  it  survives  until  the  contamina- 
tion dissipates.  If  the  colony  is  extir- 
pated, the  manager  can  protect  the 
site  so  that  it  might  be  recolonized 
from  outside  the  contaminated  area 
in  the  future. 

After  a  colony  is  known  to  be 
heavily  contaminated  with  an  or- 
ganochlorine  or  metal,  annual  analy- 
ses of  guano  can  determine  whether 
contamination  is  decreasing,  increas- 
ing, or  remaining  stable,  and  also  can 
alert  the  manager  to  potential  prob- 
lems. For  example,  in  Missouri,  hep- 
tachlor  epoxide  increased  from  mi- 
nor amounts  in  bats  in  1976  to  near 
lethal  levels  in  1977  (Clark  et  al. 
1983a).  Such  information  promptly 
passed  to  the  state  authorities  might 


persuade  them  to  recommend  a  dif- 
ferent pesticide  to  farmers  before  the 
problem  chemical  becomes  heavily 
dispersed  over  wide  areas. 

The  Alabama  example  given  pre- 
viously shows  that  large  cleanup  ef- 
forts are  possible  if  the  contamina- 
tion is,  in  total  or  in  part,  localized. 
State  and  federal  agencies  represent 
routes  open  to  managers.  In  this  in- 
stance, the  U.S.  Environmental  Pro- 
tection Agency  exercised  its  author- 
ity. Whether  a  large  cleanup  effort 
would  be  undertaken  if  only  bats 
were  affected  is  not  known;  however, 
if  organochlorine  contamination  is 
heavy  enough  to  cause  mortality  in 
bat  colonies,  it  probably  affects  other 
wildlife  as  well.  Bat  colonies  are 
good  places  to  look  for  food-chain 
contaminant  problems  because  bats 
feed  over  wide  areas  but  congregate 
in  only  a  few  roosts.  Thus,  problems 
from  many  potential  areas  are 
brought  to  a  single  site  where  symp- 
toms may  be  seen  as  dead  or  dying 
bats.  The  disadvantage  is  that  it  may 
be  difficult  to  locate  the  source  area, 
or  areas,  unless  the  feeding  locations 
of  the  bats  are  known. 

Heavy  metals  in  the  environment 
often  have  industrial  point  sources 
that  are  subject  to  existing  emission 
regulations.  Therefore,  such  contami- 
nation may  be  easier  to  stop. 

Acknowledgments 

I  thank  R.L.  Clawson,  E.L.  Flickinger, 
K.N.  Geluso,  C.E.  Grue,  and  T.H. 
Kunz  for  critical  reviews  of  the 
manuscript. 

Literature  Cited 

Allison,  Vernon  C.  1937.  Evening  bat 
flight  from  Carlsbad  Caverns. 
Journal  of  Mammalogy  18:80-82. 

Altenbach,  J.Scott,  Kenneth  N. 
Geluso,  and  Don  E.  Wilson.  1979. 
Populations  size  of  Tadarida  brasil- 
iensis  at  Carlsbad  Caverns  in  1973. 
p.  341-348.  In  H.H.  Genoways  and 


R.J.  Baker,  eds.  Biological  investi- 
gations in  the  Guadalupe  Moun- 
tains National  Park,  Texas.  Pro- 
ceedings and  Transactions  Series 
Number  4,  National  Park  Service. 
Anonym.ous.  1987.  Thornhill  cave  bat 
killer  convicted.  Bats,  December 
5:2,8. 

Clark,  Donald  R.,  Jr.  1979.  Lead  con- 
centrations: bats  vs.  Terrestrial 
small  mammals  collected  near  a 
major  highway.  Environmental 
Science  and  Technology  13:338- 
341. 

Clark,  Donald  R.,  Jr.  1981a.  Death  in 
bats  from  DDE,  DDT  or  dieldrin: 
diagnosis  via  residues  in  carcass 
fat.  Bulletin  of  Environmental 
Contamination  and  Toxicology 
26:367-374. 

Clark,  Donald  R.,  Jr.  1981b.  Bats  and 
environmental  contaminants:  a 
review.  United  States  Fish  and 
Wildlife  Service  Special  Scientific 
Report— Wildlife,  Number  235. 

Clark,  Donald  R.,  Jr.  1986.  Toxicity  of 
methyl  parathion  to  bats:  mortal- 
ity and  coordination  loss.  Environ- 
mental Toxicology  and  Chemistry 
5:191-195. 

Clark,  Donald  R.,  Jr.,  Fred  M.  Bagley, 
and  W.  Wynon  Johnson.  1988. 
Northern  Alabama  colonies  of  the 
endangered  gray  bat  Myotis  gris- 
esens:  organochlorine  contamina- 
tion and  mortality.  Biological  Con- 
servation 43:213-225. 

Clark,  Donald  R.,  Jr.,  Christine  M. 
Bunck,  and  Eugene  Cromartie. 
1983a.  Year  and  age  effects  on  resi- 
dues of  dieldrin  and  heptachlor  in 
dead  gray  bats.  Franklin  County, 
Missouri— 1976, 1977,  and  1978. 
Environmental  Toxicology  and 
Chemistry  2:387-393. 

Clark,  Donald  R.,  Jr.,  Richard  L. 
Clawson,  and  Charles  J.  Stafford. 
1983b.  Gray  bats  killed  by  dieldrin 
at  two  additional  Missouri  caves: 
aquatic  macroinvertebrates  found 
dead.  Bulletin  of  Environmental 
Contamination  and  Toxicology 
30:214-218. 

Clark,  Donald  R.,  Jr.,  and  Alexander 
J.  Krynitsky.  1983.  DDT:  recent 


412 


contamination  in  New  Mexico  and 
Arizona?  Environment  25:27-31. 

Clark,  Donald  R.,  Jr.,  Thomas  H. 
Kunz,  and  T.  Earl  Kaiser.  1978a. 
Insecticides  applied  to  a  nursery 
colony  of  little  brown  bats  (Myotis 
lucifugus):  lethal  concentrations  in 
brain  tissues.  Journal  Mammalogy 
59:84-91. 

Clark,  Donald  R.,  Jr.,  Richard  K. 
LaVal,  and  Douglas  M.  Swineford. 
1978b.  Dieldrin-induced  mortality 
in  an  endangered  species,  the  gray 
bat  (Myotis  grisescens).  Science 
199:1357-1359. 

Clark,  Donald  R.,  Jr.,  Richard  K. 
LaVal,  and  Merlin  D.  Tuttle.  1982. 
Estimating  pesticide  burdens  of 
bats  from  guano  analyses.  Bulletin 
of  Environmental  Contamination 
and  Toxicology  29:214-220. 

Clark,  Donald  R.,  Jr.,  and  Richard  M. 
Prouty.  1984.  Disposition  of  die- 
tary dieldrin  in  the  little  brown  bat 
and  correlation  of  skin  levels  with 
body  burden.  Bulletin  of  Environ- 
mental Contamination  and  Toxi- 
cology 33:177-183. 

Clark,  Donald  R.,  Jr.,  and  Barnett  A. 
Rattner.  1987.  Orthene®  toxicity  to 
little  brown  bats  (Myotis  lucifugus): 
acetylcholinesterase  inhibition,  co- 
ordination loss,  and  mortality. 
Environmental  Toxicology  and 
Chemistry  6:705-708. 

Clark,  Donald  R.,  Jr.,  Anne  Shapiro 
Wenner,  and  John  F.  Moore.  1986. 
Metal  residues  in  bat  colonies, 
Jackson  County,  Florida,  1981- 
1983.  Florida  Field  Nat.  14:38-45. 

Ellman,  George  L.,  K.  Diane  Court- 
ney, Valentino  Andres,  Jr.,  and 
Robert  M.  Featherstone.  1961.  A 
new  rapid  colorimetric  determina- 
tion of  acetylcholinesterase  activ- 
ity. Biochemistry  and  Pharmacol- 
ogy 7:88-95. 

Fleming,  W.  James,  and  Thomas  Z. 
Atkeson.  1980.  Situation  report: 
heavy  DDT  contamination  at 
Wheeler  National  Wildlife  Refuge. 
Proceedings  of  the  Annual  Confer- 
ence of  the  Southeast  Association 
of  Fish  and  Wildlife  Agencies 
34:453-461. 


Fleming,  W.  James,  and  Eugene  Cro- 
martie.  1981.  DDE  residues  in 
young  wood  ducks  (Aix  sponsa) 
near  a  former  DDT  manufacturing 
plant.  Pesticides  Monitoring  Jour- 
nal 14:115-118. 

Fleming,  W.  James,  Burline  P.  Pullin, 
and  D.M.  Swineford.  1984.  Popu- 
lation trends  and  environmental 
contaminants  in  herons  in  the  Ten- 
nessee Valley,  1980-81.  Colonial 
Waterbirds  7:63-73. 

Geluso,  Kenneth  N.,  J.  Scott  Alten- 
bach,  and  Don  E.  Wilson.  1976.  Bat 
mortality:  pesticide  poisoning  and 
migratory  stress.  Science  194:184- 
186. 

Hill,  Elwood  F.,  and  W.  James  Flem- 
ing. 1982.  Anticholinesterase  poi- 
soning of  birds:  field  monitoring 
and  diagnosis  of  acute  poisoning. 
Environmental  Toxicology  and 
Chemistry  1:27-38. 

Hunt,  W.  Grainger,  Brenda  S. 

Johnson,  Carl  G.  Thelander,  Brian 
J.  Walton,  Robert  W.  Risebrough, 
Walter  M.  Jarman,  Alan  M.  Sprin- 
ger, J.  Geoffrey  Monk,  and  Way- 
man  Walker  II.  1986.  Environ- 
mental levels  of  p,p'-DDE  indicate 
multiple  sources.  Environmental 
Toxicology  and  Chemistry  5:21-27. 

Maly,  Mark  S.  1984.  Survivorship  of 
meadow  voles,  Microtus  pennsyl- 
vanicus,  from  sewage  sludge- 
treated  fields.  Bulletin  of  Environ- 
mental Contamination  and  Toxi- 
cology 32:724-731. 

Nomiyama,  Kazuo.  1981.  Renal  ef- 
fects of  cadmium,  p.  643-689.  In 
J.O.  Nriagu,  ed.  Cadmium  in  the 
environment;  part  2  health  effects. 
Environmental  Science  and  Tech- 
nology: A  Wiley  Series  of  Texts 
and  Monographs.  John  Wiley  and 
Sons,  Inc.,  New  York. 

O'Shea,  Thomas,  J.,  W.  James  Flem- 
ing III,  and  Eugene  Cromartie. 
1980.  DDT  contamination  at 
Wheeler  National  Wildlife  Refuge. 
Science  209:509-510. 

Petit,  Michael  G.,  and  J.  Scott  Alten- 
bach.  1973.  A  chronological  record 
of  environmental  chemicals  from 
analysis  of  stratified  vertebrate 


excretion  deposited  in  a  sheltered 
environment.  Environmental  Re- 
search 6:339-343. 

Prichard,  Dennis  E.  1987.  Managing  a 
national  wildlife  refuge  for  bats. 
Bats,  December  5:6,8. 

Reich,  Andrew  R.,  Jimmy  L.  Perkins, 
and  Gary  Cutter.  1986.  DDT  con- 
tamination of  a  north  Alabama 
aquatic  ecosystem.  Environmental 
Toxicology  and  Chemistry  5:725- 
736. 

Reidinger,  Russell  F.,  Jr.,  and  E.  Len- 
dell  Cockrum.  1978.  Organo- 
chlorine  residues  in  free-tailed 
bats  (Tadarida  brasiliensis)  at  Eagle 
Creek  Cave,  Greenlee  County, 
Arizona,  p.  85-96.  In  R.J.  Olembo, 
J.B.  Castelino,  and  F.A.  Mutere, 
eds.  Proceedings  of  the  Fourth 
International  Bat  Research  Confer- 
ence, Kenya  Literature  Bureau, 
Nairobi. 

Tuttle,  Merlin  D.  1979.  Status,  causes 
of  decline,  and  management  of 
endangered  gray  bats.  Journal  of 
Wildlife  Management  43:1-17. 

United  States  Department  of  the  Inte- 
rior, Fish  and  Wildlife  Service. 
1987.  Endangered  and  threatened 
wildlife  and  plants,  April  10, 1987, 
50  CFR  17.11  and  17.12.  United 
States  Government  Printing  Of- 
fice. 

White,  Donald  H.,  and  Alexander  J. 
Krynitsky.  1986.  Wildlife  in  some 
areas  of  New  Mexico  and  Texas 
accumulate  elevated  DDE  resi- 
dues, 1983.  Archives  of  Environ- 
mental Contamination  and  Toxi- 
cology 15:149-157. 


413 


Habitat  Structure,  Forest 
Composition  and  Landscape 
Dimensions  as  Components 
of  Habitat  Suitability  for  the 
Delmarva  Fox  Squirrel^ 

Raymond  D.  Dueser,^  James  L.  Dooley,  Jr.,^ 
and  Gary  J.  Taylor^ 


Abstract.— Discriminant  function  analysis  compar- 
ing 36  occupied  and  18  unoccupied  sites  revealed 
ti^ot  structural  variables  discriminated  betv\/een 
sample  groups  better  than  compositional  variables, 
and  the  latter  discriminated  better  than  landscape 
variables.  These  results  ore  encouraging  that  habitat 
structure  will  provide  a  reliable  basis  for  a  predictive 
classification  model  of  habitat  suitability.  Such  a 
model  would  be  useful  both  for  pre-screening  the 
biological  suitability  of  potential  release  sites  and  for 
planning,  implementing  and  monitoring  prescriptive 
habitat  management. 


The  Delmarva  fox  squirrel  ( Sciurus 
niger  cinereus)  was  placed  on  the  fed- 
eral endangered  species  list  in  1967 
(32  FR  4001;  U.S.  Department  of  Inte- 
rior 1970).  Remnant  populations 
were  restricted  to  four  counties  in 
eastern  Maryland  (Taylor  and  Flyger 
1973),  representing  less  than  10%  of 
the  historic  range  of  the  subspecies 
on  the  Delmarva  Peninsula.  Forest 
clearing  and  habitat  fragmentation 
throughout  the  range  undoubtedly 
contributed  significantly  to  the  pres- 
ent endangerment  (Taylor  1973). 

The  U.S.  Fish  and  Wildlife  Service 
Recovery  Plan  for  the  restoration  of 
the  Delmarva  fox  squirrel  to  secure 
status  emphasizes  both  the  reintro- 
duction  of  this  subspecies  to  suitable 
habitats  throughout  the  former  range 
and  prescriptive  habitat  management 
for  established  populations  (Taylor  et 
al.  1983).  A  thorough  understanding 


'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles  and  Small 
Mammals  in  Nortt)  America.  (Ragstaff.  AZ, 
July  19-21,  1988.) 

'Raymond  D.  Dueser  is  an  Associate 
Professor  in  ttie  Department  of  Environ- 
mental Sciences,  University  of  Virginia, 
Chiarlotfesville,  VA  22903. 

^ James  L  Dooley,  Jr.,  is  a  graduate  re- 
search) assistant  in  the  Department  of  Envi- 
ronmental Sciences,  University  of  Virginia, 
Charlottesville.  VA  22903. 

''Gary  J.  Taylor  is  Associate  Director  of 
Wildlife.  Department  of  Natural  Resources. 
Maryland  Forest,  Park  and  Wildlife  Service. 
Tawes  State  Office  Building.  Annapolis,  MD 
21401. 


of  habitat  requirements  will  be  essen- 
tial for  both  initiatives  (Dueser  and 
Terwilliger  1988). 

Habitat  requirements  might  be 
expressed  through  any  of  three  sepa- 
rate but  related  components  of  habi- 
tat suitability:  forest  habitat  struc- 
ture, forest  tree  species  composition, 
and  surrounding  landscape  struc- 
ture. Both  habitat  structure  and  for- 
est composition  have  been  shown  to 
influence  the  distribution  and  abun- 
dance of  fox  squirrels  in  heterogene- 
ous landscapes  (Nixon  and  Hansen 
1987). 

Recent  research  has  demonstrated 
the  potential  influence  of  landscape 
composition  and  structure  on  popu- 
lations of  woodland  mammals  occu- 
pying farmland  mosaics  (Wegner 
and  Merriam  1979,  Middleton  and 
Merriam  1983,  Fahrig  and  Merriam 

1985)  .  Furthermore,  changes  in  the 
landscape  of  the  Delmarva  Peninsula 
almost  certainly  played  a  major  role 
on  the  decline  of  the  fox  squirrel 
(Taylor  1973). 

Given  this  background,  the  objec- 
tive of  this  study  was  to  compare  the 
apparent  effects  of  habitat  structure, 
forest  composition  and  landscape  di- 
mensions on  the  presence  and  ab- 
sence of  the  Delmarva  fox  squirrel  on 
54  study  sites  in  eastern  Maryland. 
This  analysis  is  the  first  step  in  the 
development  of  a  predictive  classifi- 
cation model  of  habitat  suitability  for 
this  subspecies  (cf.  Houston  et  al. 

1986)  . 

414 


Methods 

Data  Base 

During  a  12-mo  search  for  remnant 
populations  of  the  Delmarva  fox 
squirrel  on  the  Maryland  Eastern 
Shore,  Taylor  (1976)  located  36  "fox 
squirrel  present''  (Present)  sites  with 
extant  populations  and  18  "fox  squir- 
rel absent"  (Absent)  sites.  The  gray 
squirrel  (Sciurus  caroUnensis)  was 
present  on  all  54  sites.  Taylor  then 
sampled  the  forest  habitat  of  each 
site,  to  compare  Present  forest  stands 
with  Absent  stands.  He  established  a 
representative  4  m  x  200  m  belt 
transect  on  each  site,  on  which  he  re- 
corded the  number  of  trees  by  spe- 
cies per  diameter-breast-height 
(DBH)  size  class  (5-20  cm,  20.1-30 
cm,  30.1-50  cm,  and  50.1+  cm),  per- 
cent crown  cover,  percent  understory 
cover,  understory  density,  and 
understory  species  composition.  All 
habitat  measurements  were  taken 
from  June  through  September  1972 
and  1973.  These  data  formed  the  ini- 
tial data  base  for  this  study. 

Taylor  (1976)  reported  the  number  j| 
of  trees  measured  in  each  of  two  size 
classes:  "small"  trees  (5-30  cm  DBH) 
and  "large"  trees  (>  30  cm  DBH).  We 
assigned  each  tree  to  one  of  five  taxo- 
nomic  groups:  loblolly  pine  (Pinus 
taeda),  combined  oak  species  (Quercus 
spp.),  American  beech  (Fagus  gran- 
difolia),  combined  hickory  sp>ecies 
{Carya  spp.),  and  combined  mixed  | 


hardwoods.  We  estimated  the  ap- 
proximate total  basal  area  for  each 
size-taxonomic  class  by  assuming  an 
average  DBH  of  17.5  cm  for  small 
trees  and  40.0  cm  for  large  trees.  We 
then  estimated  total  basal  area  for  all 
trees  >  5  cm  DBH  and  the  fraction  of 
that  total  basal  area  represented  by- 
each  taxonomic  group.  These  basal 
area  estimates  provide  a  basis  for 
comparing  forest  "composition"  in- 
dependently of  forest  "structure"  as 
reflected,  for  example,  in  the  raw 
percentage  of  trees  counted  in  each 
taxonomic  group. 

Original  data  were  collected  on 
land  use  and  cover  composition  of 
the  landscape  surrounding  a  random 
subset  of  Taylor's  (1976)  study  areas 
(fig.  1).  Landscape  variables  included 
area  of  open  fields,  percentage  of 
area  forested,  internal  forest  perime- 
ter ("edge")  within  the  sample  unit, 
forest  shape  (Blouin  and  Connor 
1985),  and  distance  to  next  nearest 
woodland.  These  variables  are  re- 
ferred to  below  as  landscape  "dimen- 
sions." They  were  measured  by 


planimetry  of  1:10(X)  black-and-white 
photographs  (dated  1978)  obtained 
from  Eastern  Shore  offices  of  the 
USDA  Agricultural  Stabilization  and 
Conservation  Service.  We  initially 
measured  each  landscap)e  variable 
for  a  2-km^  circular  sample  unit  cen- 
tered on  the  sample  woodland.  This 
unit  was  chosen  as  a  first-approxima- 
tion of  "minimal  population  area"  on 
the  basis  of  home  range  size  and  ac- 
tivity (Flyger  and  Smith  1980).  Based 
on  the  results  of  analyses  for  the  2- 
km^  unit,  both  smaller  (1-km^)  and 
larger  (4-km^)  sample  units  subse- 
quently were  described  in  the  same 
way. 


Statistical  Analyses 

This  comparison  of  habitat  compx)- 
nents  is  based  on  multivariate  statis- 
tical analyses  of  three  separate  but 
related  components  of  forest  habitat 
suitability:  (1)  habitat  structure 
("What  does  the  forest  'look  like'  to 
an  observer  passing  through  on  the 


Figure  1  .—Schematic  diagram  of  1-,  2-  and  4-krTP  sample  units  for  rrjeasuring  landscape 
dimensions  of  "fox  squirrel  present"  and  "fox  squirrel  absent"  study  areas  on  the  Eastern 
Shore  of  Maryland.. Each  sample  unit  was  centered  on  one  of  Taylor's  (1976)  study  areas. 


ground?"),  (2)  tree  species  composi- 
tion ("Which  tree  species  predomi- 
nate in  this  forest  and  give  it  its  char- 
acter?"), and  (3)  landscape  dimen- 
sions ("What  are  the  land  use  and 
cover  dimensions  of  the  landscape 
mosaic  in  which  this  forest  is  embed- 
ded?"). Conceptually,  these  compo- 
nents represent  a  gradient  of  scales 
from  "microscopic"  habitat  structure 
to  "macroscopic"  composition  to 
"megascopic"  context. 

Two-group  discriminant  function 
analysis  was  used  to  compare  the 
Present  and  Absent  forest  stands 
identified  by  Taylor  (1976).  Each 
analysis  (1)  computed  the  univariate 
F-ratio  comparing  Present  and  Ab- 
sent sites  for  each  habitat  variable, 

(2)  tested  the  centroids  of  Present 
and  Absent  sites  for  equality  on  the 
basis  of  a  linear  combination  of  the 
habitat  variables  (i.e.,  a  linear  dis- 
criminant function),  using  multivari- 
ate analysis  of  variance  (MANOVA), 

(3)  indicated  the  relative  contribution 
of  each  habitat  variable  to  any  ob- 
served difference  between  centroids, 
based  on  the  correlation  between  the 
variable  and  the  discriminant  func- 
tion, (4)  tested  the  sample  variance- 
covariance  matrices  of  Present  and 
Absent  sites  for  homogeneity  using  a 
Box's  M  test  statistic,  and  (5)  indi- 
cated the  percentage  of  the  variation 
in  group  membership  (Present  or 
Absent)  explained  by  the  discrimi- 
nant function,  based  on  the  correla- 
tion between  the  membership  vari- 
able and  the  discriminant  function. 
(Dueser  and  Shugart  1978).  All  analy- 
ses were  computed  both  with  and 
without  arcsin-square  root  transfor- 
mations of  percentage  variables.  Re- 
sults of  the  parallel  analyses  were 
qualitatively  similar  in  each  case.  For 
purposes  of  interpretability,  only  the 
results  for  untransformed  variables 
are  presented  here.  All  analyses  used 
the  MANOVA  and  DISCRIMINANT 
routines  of  the  Statistical  Package  for 
the  Social  Sciences  (SPSS,  Nie  et  al. 
1975). 

As  an  unbiased  test  of  the  ability 
of  each  set  of  habitat  variables  to 


415 


classify  the  group  membership  of  the 
study  sites  (i.e..  Present  or  Absent),  a 
jackknife  procedure  (Efron  1979)  was 
used  to  classify  each  of  Taylor's 
study  areas.  Each  site  was  deleted  in 
sequence,  DISCRIMINANT  was  run 
for  data  from  the  remaining  53  sites, 
and  a  classification  function  was 
computed  from  these  data.  The  de- 
leted site  was  then  classified  on  the 
basis  of  this  independent  classifica- 
tion function.  The  probabilistic  ("pre- 
dicted") classification  was  then  com- 
pared with  the  actual  ("observed") 
classification  for  each  site. 

Brennan  et  al.  (1986)  have  pro- 
posed an  alternative  solution  to  the 
problem  of  habitat  analysis.  Logistic 
regression  analysis  is  superior  to 
multivariate  analysis  of  variance 
when  one  or  more  of  the  predictor 
(e.g.,  habitat)  variables  is  categorical 
(i.e.,  non-continuous),  when  the  vari- 
ance-covariance  matrices  are  non- 
homogeneous  and /or  when  the  data 
violate  the  assumption  of  multivari- 
ate normality  (Press  and  Wilson 
1978).  Parallel  analyses  demonstrate 
that  logistic  regression  analysis  offers 
no  inherent  advantage  over  discrimi- 
nant function  analysis  in  the  present 
case  (Dooley,  unpublished). 

Results 

Habitat  Structure 

Present  sites  had  a  greater  percent- 
age of  trees  larger  than  30  cm  DBH, 
lower  percentage  shrub-ground 
cover,  and  slightly  lower  understory 
vegetation  density  than  Absent  sites 
(table  l,p<  0.05).  Present  and  Absent 
sites  differ  structurally  on  the  aver- 
age (MANOVA  Chi-square  (5)  = 
14.825,  p  <  0.011).  The  linear  combi- 
nation of  structure  variables  ac- 
counted for  26%  of  the  variation  in 
group  membership.  The  variance- 
covariance  matrices  were  marginally 
homogeneous  (Box's  M  =  20.056,  p  > 
0.06).  Percentage  of  trees  greater  than 
30  cm  DBH  (r  =  -0.735),  understory 
vegetation  density  (0.564),  and  per- 


centage shrub-ground  cover  (0.564) 
are  particularly  important  in  dis- 
criminating between  sites.  Conceptu- 
ally, Present  sites  have  larger  trees, 
less  shrub-ground  cover  vegetation, 
and  less  understory  than  Absent  sites 
(fig.  2).  Present  sites  were  correctly 
classified  79%  of  the  time  in  the  jack- 
knifing  procedure,  and  Absent  sites 
were  correctly  classified  48%  of  the 
time. 


Forest  Tree  Species  Composition 

All  54  study  areas  supported  a  mix 
of  hard-  and  soft-mast  tree  species. 
Although  Present  sites  had  some- 
what greater  basal  areas  for  Ameri- 
can beech  {p  >  0.07)  and  mixed  hard- 
woods (p  >  0.05),  there  were  no  clear- 
cut  univariate  differences  between 
sites  in  forest  composition  (table  2). 
There  also  was  no  difference  in  total 


Table  1.— Comparison  of  average  forest  habitat  structure  for  "fox  squirrel 
present"  and  "fox  squirrel  absent"  study  areas  on  the  Eastern  Shore  of 
Maryland,  based  on  data  of  Taylor  (1976).  Tabled  values  are  means  and 
(standard  deviations). 


Habitat  variables 


Present 


Absent 


(N 

=  36) 

(N  = 

18) 

%  Trees  >  30  cm  DBH 

32.3 

(12.14) 

22.1 

(9.26) 

<0.01 

%  Crown  cover 

75.6 

(17.72) 

70.6 

(16.08) 

>0.30 

%  Shrub-ground  cover 

51.1 

(26.60) 

67.5 

(21.85) 

<0.05 

Understory  "der^ity" 

2.6 

(1.38) 

3.4 

(1.04) 

<0.05 

%  Pine  composition 

10.5 

(10.63) 

17.1 

(22.23) 

>0.10 

Present 
ceritroid 


Absent 
centroid 


,  ."Absent". . 

1       1  1 

1    1    1  1 

1  1 

-2 


1  0  +1 

Discriminant  Score 


+2 


Larger  trees 
Sparse  understory 
Sparse  grourxicover 


SmaJler  trees 
Dense  understory 
Dense  groundcover 


Figure  2.— Interpretation  of  discrimirxition  between  average  "fox  squirrel  present"  and  "fox 
squirrel  absent"  study  areas  on  ttie  Eastern  Stiore  of  Maryland,  based  on  analysis  of  forest 
habitat  structure.  Ttie  tiorizontal  dastied  lines  indicate  ttie  range  of  observations  for  a 
sample  group  (Present  or  Absent). 


416 


basal  area  (F  (1,52)  =  2.300,  p  >  0.13). 
The  two  types  of  sites  were  similar  in 
composition  for  both  small  and  large 
trees  (fig.  3).  Reflecting  this  similar- 
ity, there  was  only  a  marginally  sig- 
nificant multivariate  difference  in 
forest  composition  (MANOVA  Chi- 
square  (5)  =  10.584,  p  >  0.06). 

The  linear  combination  of  compo- 
sition variables  accounted  for  19%  of 
the  variation  in  group  membership. 
The  variance-co variance  matrices 
were  conspicuously  non-homogene- 
ous (Box's  M  =  61.549,  p  <  0.001). 
Present  sites  were  correctly  classified 
79%  of  the  time,  and  Absent  sites 
were  correctly  classified  48%  of  the 
time. 

Although  the  correct  classification 
rates  were  the  same  as  for  structural 
variables,  the  two  sets  of  variables 
misclassified  different  sites. 


Landscape  Dimensions 

Five  landscape  variables  were  meas- 
ured for  the  2-km^  circular  sample 
unit  centered  on  the  target  woodland 
of  38  of  the  Taylor's  (1976)  study  ar- 
eas. Present  sites  were  somewhat 
closer  to  the  next  nearest  forest  tract 
than  Absent  sites  (table  3,p<  0.03). 
Despite  this  modest  difference,  there 
was  no  significant  multivariate  dif- 
ference in  landscape  dimensions  be- 
tween sites  (MANOVA  Chi-square 
(5)  =  8.574,  p>  0.127). 

Present  and  Absent  woodlands 
also  were  similar  in  area,  averaging 
9.4  and  10.0  ha,  respectively,  as  pho- 
tographed in  1978.  The  linear  combi- 
nation of  landscape  variables  ac- 
counted for  23%  of  the  variation  in 
group  membership.  The  variance-co- 
variance  matrices  were  homogene- 


r 


Table  2.— Comparison  of  average  tree  species  composition  for  "fox  squirrel 
present"  and  'lox  squirrel  absent"  study  areas  on  the  Eastern  Shore  of 
Maryland,  based  on  estimated  basal  area  (cm^  per  SOO-m^  sample 
transect)  per  taxonomlc  groip.  Data  from  Taylor  (1976).  Tabled  values  are 
means  and  (standard  deviations). 


Taxonomic  group 


Present 
(N  =  36) 


Absent 
(N  =  18) 


Loblolly  pine 

5359 

(1099,84) 

7339 

(2278.43) 

>0.35 

Oak  species 

9547 

(1061.24) 

9628 

(1378.18) 

>0.95 

American  beech 

3293 

(679.62) 

1400 

(546.56) 

>0.07 

Hickory 

1683 

(611,77) 

1050 

(263.26) 

>0.50 

Mixed  hardwoods 

9498 

(1032.96) 

6514 

(690.13) 

>0.05 

Table  3.— Comparison  of  average  landscape  dimensions  for  "fox  squirrel 
present"  and  "fox  squirrel  absent"  study  areas  on  the  Eastern  Shore  of 
Maryland.  Variables  measured  for  2-km^  circular  sample  unit  centered  on 
study  woodlarKl.  Tabled  vcdues  are  means  and  (standard  deviations). 


Landscape  variables 


Present 


Absent 


(N 

=  27) 

(N  = 

11) 

Area  open  fields  (ha) 

99.3 

(6.4) 

96.3 

(11.9) 

>0.81 

%  Forested  area 

56.4 

(3.6) 

50.1 

(6.0) 

>0.35 

Internal  perim.  (km) 

5.3 

(2.0) 

6.2 

(1.9) 

>0.21 

Forest  "shape" 

136.4 

(54.5) 

153.0 

(44.9) 

>0.38 

Dist.  next  woodlot  (km) 

0.4 

(0.1) 

0.8 

(0.2) 

<0.03 

ous  (Box's  M  =  19.926,  p  >  0.39).  As 
with  forest  composition,  there  was 
no  consistent  difference  in  landscape 
dimensions  between  Present  and  Ab- 
sent sites.  Present  sites  were  correctly 
classified  787o  of  the  time,  and  Ab- 
sent sites  were  correctly  classified 
40%  of  the  time. 

To  evaluate  the  possibility  that  the 
negative  result  in  the  test  for  equality 
of  group  centroids  came  about  be- 
cause we  were  measuring  landscape 
variables  on  an  "incorrect"  spatial 
scale,  we  repeated  the  landscape 
analysis  for  both  smaller  (1-km^)  and 
larger  (4-km^)  circular  sample  units, 
still  centered  on  the  woodland  of 
interest.  Again,  there  were  no  consis- 
tent group  differences  on  either  scale 
ip  >  0.40,  table  4). 

Either  the  landscapes  surrounding 
the  sample  Present  and  Absent 
woodlands  do  not  differ  consistently, 
or  they  differ  on  a  scale  of  measure- 
ment or  in  a  way  not  revealed  by  the 
present  analyses. 


< 

lu  10 

IT 
< 


DBH  5-30cm 

^  8PECC8Aae£NT 
■i    8PECCS  CreBEKT 


JLM_ii  ilIl 


DBH>30CM 


Figure  3.— Average  forest  tree  species 
comF>ositlon  of  "fox  squirrel  present"  and 
"fox  squirrel  absent"  study  areas  on  tt»e 
Eastern  Shore  of  Maryland.  "Other"  cate- 
gory includes  a  variety  of  snr»ail  trees  such 
as  cherry  (Prunus  spp.)  and  flowering 
dogwood  (Cornus  florida). 


417 


Discussion 
Present  Habitat 

The  present  habitat  of  the  Delmarva 
fox  squirrel  consists  primarily  of 
relatively  small  stands  of  mature 
mixed  hardwoods  and  pines  having 
relatively  closed  canopies,  relatively 
open  understory,  and  a  high  propor- 
tion of  forest  edge.  Occupied  tracts 
include  both  groves  of  trees  along 
streams  and  bays  and  small  woo- 
dlots  located  near  agricultural  fields. 
In  some  areas,  particularly  in  south- 
ern Dorchester  County,  Maryland, 
occupied  habitat  includes  tracts 
dominated  by  mature  loblolly  pine 
located  adjacent  to  marshes  and  tidal 
streams.  The  woodland  habitats  now 
occupied  by  the  Delmarva  fox  squir- 
rel are  consistent  with  those  occu- 
pied by  other  subspecies  of  fox  squir- 
rel (Bakken  1952;  Brown  and  Yeager 
1945;  Weigl  et  al.,  in  press). 

The  picture  of  the  Delmarva  fox 
squirrel  that  emerges  from  the  litera- 
ture is  one  of  a  species  relatively 
adept  at  utilizing  a  dissected,  hetero- 
geneous landscape  dominated  by  ag- 
riculture and  woodlot  forestry.  Fox 
squirrels  are  more  cursorial  than 
gray  squirrels,  and  often  are  found 
on  the  ground  several  hundred  me- 
ters from  the  nearest  woodlot.  They 
occupy  larger  home  ranges  than  gray 
squirrels  (30  ha  vs.  3  ha),  travel  far- 
ther between  captures  (307  m  vs.  119 
m),  and  thus  are  generally  more  mo- 
bile (Flyger  and  Smith  1980).  Fox 
squirrels  more  readily  exploit  agri- 
cultural crops  such  com,  oats,  soy- 
beans and  fruit.  They  more  fre- 
quently utilize  forest  edges.  Fox 
squirrels  would  thus  appear  to  be 
relatively  well-adapted  to  exploit  the 
landscape  created  by  settlement  of 
the  coastal  plain. 

One  might  conclude  that  man's 
activities  on  the  Delmarva  Peninsula 
should  have  been  to  the  benefit  of  the 
fox  squirrel.  Land  clearing  has  cre- 
ated woodlots.  Grazing  and  burning 
have  opened  up  the  understory.  Ag- 
riculture has  increased  the  availabil- 


ity of  alternative  food  sources  and 
perhaps  stabilized  the  food  supply. 
Indeed,  Allen  (1943)  and  Nixon  and 
Hansen  (1987)  indicate  that  settle- 
ment and  agriculture  have  worked  to 
the  advantage  of  the  fox  squirrel 
throughout  the  midwestern  United 
States,  resulting  in  increased  abun- 
dance and  an  expanded  geographic 
range. 

Why  has  this  not  occurred  with 
the  Delmarva  subspecies?  Why  has 
the  abundance  of  this  fox  squirrel 
continued  to  decline  throughout  the 
period  of  the  recorded  literature 
(since  approximately  1850)? 

Taylor  (1976)  attributes  the  contin- 
ued decline  of  the  Delmarva  fox 
squirrel  to  habitat  destruction.  While 
many  of  the  landscape  changes  re- 
sulting from  settlement  might  have 
benefited  the  fox  squirrel,  others 
have  been  detrimental.  Taylor  be- 
lieves that  extensive  timber  harvest 
has  been  particularly  detrimental. 
The  removal  of  mature  hardwoods 
has  reduced  the  availability  of  suit- 
able den  trees,  removed  reliable 
sources  of  concentrated  hard  mast, 
promoted  the  luxuriant  growth  of 
understory  vegetation,  and  perhaps 
altered  the  competitive  relationship 
between  fox  and  gray  squirrels  to 
favor  grays.  Furthermore,  coastal 
plain  woodland  management  typi- 
cally has  involved  both  short  timber 
rotations  (i.e.,  frequent  harvests)  and 
reforestation  with  pure  stands  of  lob- 
lolly pine.  Finally,  gradual  urbaniza- 
tion has  added  yet  another  detrimen- 
tal land-use  practice. 


Habitat  Suitability 

It  is  assumed  that  Present  (i.e.,  occu- 
pied) sites  are  more  "suitable"  on  the 
average  than  are  Absent  (i.e.,  unoc- 
cupied) sites.  Present  sites  are  re- 
garded here  as  the  "standard  of  ex- 
cellence" by  which  to  judge  the  habi- 
tat requirements  of  the  Delmarva  fox 
squirrel.  Given  that  a  number  of  un- 
known (and  unknowable)  ecological, 
biogeographical,  and /or  historical 
factors  may  actually  be  responsible 
for  the  absence  of  this  subspecies 
from  any  particular  site  within  the 
historic  range,  this  assumption  is  cor- 
rect only  as  a  first  approximation 
(ref.  Hanski  1982).  It  clearly  would  be 
unwarranted  if  the  distribution  of 
squirrels  among  these  54  study  sites 
were  highly  variable  through  time. 
Nevertheless,  the  chance  presence  of 
the  squirrel  on  "unsuitable"  sites  and 
its  absence  from  "suitable"  sites  be- 
cause of  factors  other  than  habitat 
suitability  per  se  can  only  make  it 
more  difficult  to  distinguish  between 
Present  and  Absent  sites.  These 
analyses  based  on  presence-or-ab- 
sence  population  information  thus 
circumvent  many  of  the  potential  pit- 
falls associated  with  the  use  of  popu- 
lation density  as  an  indicator  of  habi- 
tat suitability  (Van  Home  1983). 

Given  its  present  habitat,  it 
seemed  reasonable  to  propose  that 
the  capacity  of  a  woodland  to  sup- 
port a  population  of  the  Delmarva 
fox  squirrel  could  be  determined  by 
habitat  structure,  forest  composition 
and/or  the  land  use  and  cover  com- 


Table  4.— Comparison  of  average  "fox  squirrel  present"  and  "fox  squirrel 
absent"  study  areas  at  the  1-,  2-  and  4-square  kilometer  scales  of  obser- 
vation. Testing  for  similarity  of  landscape  dimensions  listed  In  table  3. 


Statistic 


l-km2 

2-km2 

4-km2 

7 

27 

7 

9 

n 

9 

4.791 

8.574 

2.750 

5 

5 

5 

0.442 

0.127 

0.738 

Number  "Present"  areas 
Number  "Absent"  areas 
Chi-square 
df 
P 

V  


418 


position  of  the  surrounding  land- 
scape. We  anticipated  originally  that 
each  of  these  components  of  habitat 
suitability  would  prove  to  be  impor- 
tant in  its  own  way,  and  that  each 
would  have  a  perceptible  influence 
on  fox  squirrel  presence  or  absence 
in  Maryland  woodlots  today.  Our 
results  indicate,  however,  that  habi- 
tat structure  is  the  component  most 
likely  to  contribute  meaningfully  to 
the  formulation  of  a  predictive  model 
of  habitat  suitability.  Only  the  struc- 
ture variables  discriminate  strongly 
between  Present  and  Absent  sites: 
Present  sites  have  larger  trees,  less 
ground  cover  and  less  understory 
(fig.  2).  These  variables  account  for 
the  greatest  fraction  of  the  explained 
variation,  their  dispersion  matrices 
are  effectively  homogeneous,  and 
they  classify  sites  to  the  correct 
group  (i.e..  Present  or  Absent)  at 
least  as  well  as  any  of  the  variable 
sets  examined. 

Forest  composition  is  highly  vari- 
able among  locations  in  eastern 
Maryland,  but  this  variation  seems  to 
exert  only  a  marginal  influence  on 
the  likelihood  of  occurrence  of  fox 
squirrels  on  any  given  site.  The  com- 
position variables  classify  sites  as  re- 
liably as  the  structural  variables,  and 
they  account  for  only  a  slightly  lower 
fraction  of  the  explained  variation. 
They  do  not,  however,  discriminate 
strongly  between  Present  and  Absent 
sites  and  their  dispersion  matrices 
are  strongly  non-homogeneous.  Of 
course,  this  conclusion  is  based  on  a 
comparison  of  two  groups  of  sites,  all 
of  which  are  known  to  be  "squirrel 
woods."  Had  there  been  a  "tree 
squirrel  absent"  category  of  study 
area,  forest  composition  might  well 
have  appeared  to  be  more  significant 
(cf.  Nixon  et  al.  1978,  Sanderson  et  al. 
1976). 

Landscape  composition  also  varies 
among  locations,  but  this  variation 
seems  not  to  be  important  on  the  av- 
erage in  discriminating  between  oc- 
cupied and  unoccupied  sites  today. 
The  landscape  variables  account  for  a 
comparable  fraction  of  the  explained 


variation,  they  classify  sites  almost  as 
reliably  as  the  structural  variables, 
and  their  disp>ersion  matrices  are 
homogeneous.  They  do  not,  how- 
ever, discriminate  meaningfully  be- 
tween Present  and  Absent  sites. 
Given  the  suggested  importance  of 
landscape  changes  in  bringing  about 
the  decline  of  the  fox  squirrel,  this 
result  was  somewhat  unexpected. 
The  correct  interpretation  probably 
requires  recognition  that  most  of  the 
Eastern  Shore  landscape  has  been 
altered,  fragmented  and  homoge- 
nized. Most  of  the  remaining  wood- 
lands are  mere  remnants  of  forest  in 
a  mosaic  of  agricultural  fields,  wet- 
lands and  suburban  development. 
There  may  simply  be  little  important 
variation  remaining  among  these  for- 
est patches.  At  the  same  time,  it  must 
be  recognized  that  a  number  of  po- 
tentially important  landscape  vari- 
ables— e.g.,  proximity  to  streams  and 
ponds  (McComb  and  Noble  1981) 
and  proximity  to  roadways  (Flyger 
and  Lustig  1976) — were  not  consid- 
ered in  this  analysis. 


Management  Implications 

The  Recovery  Plan  for  the  Delmarva 
fox  squirrel  calls  for  both  the  translo- 
cation of  squirrels  to  suitable  habitats 
throughout  the  historic  range  and  the 
maintenance  of  occupied  habitat 
(Taylor  et  al.  1983).  Will  objective, 
quantitative  habitat  analysis  be  help- 
ful in  evaluating  potential  release 
sites  and  planning  prescriptive  habi- 
tat management?  Results  of  the 
analyses  presented  here  provide 
some  basis  for  optimism.  A  number 
of  management  implications  follow 
from  these  results: 

1.  Of  the  variable  sets  exam- 
ined, habitat  structure  is  the 
best  indicator  of  biological 
habitat  suitability  for  the  Del- 
marva fox  squirrel  at  the 
present  time.  Even  this  mini- 
mal list  of  structure  variables 
(table  1)  has  the  power  to 


discriminate  meaningfully 
between  occupied  and  unoc- 
cupied forest  stands.  Present 
sites  have  larger  trees,  less 
ground  cover,  and  less 
understory  than  Absent  sites. 
Significantly,  these  results 
corroborate  the  general  habi- 
tat descriptions  rep>orted  by 
Flyger  and  Lustig  (1976). 

2.  In  addition  to  this  clear-cut 
discrimination,  the  structure 
variables  exhibit  the  most 
desirable  combination  of  sta- 
tistical properties,  including 
the  highest  variance  explana- 
tion, homogeneity  of  disper- 
sions, and  high  correct  classi- 
fication rates.  These  proper- 
ties will  simplify  the  formu- 
lation of  a  predictive  classifi- 
cation model  of  habitat  suita- 
bility. 

3.  Although  the  absence  of 
meaningful  discriminating 
information  in  forest  compo- 
sition and  landscape  dimen- 
sions was  somewhat  surpris- 
ing, these  results  have  the 
effect  of  simplifying  the  ef- 
fort to  quantify  habitat  suita- 
bility for  the  Delmarva  fox 
squirrel.  It  would  be  impru- 
dent to  disregard  forest  com- 
position and  landscape  at- 
tributes in  the  evaluation  of 
{X)tential  release  sites;  these 
components  of  habitat  suita- 
bility must  be  imjx)rtant  at 
some  level  (Ryger  and  Lus- 
tig 1976).  There  appears  to  be 
little  potential,  however,  for 
the  variables  analyzed  here 
to  contribute  to  a  predictive 
model  of  habitat  suitability. 

4.  The  discriminating  structure 
variables  are  easy  and  rela- 
tively inexpensive  to  meas- 
ure. Including  site  reconnais- 
sance, approximately  one- 
half  day  of  field  time  is  re- 
quired for  a  team  of  two  ex- 


419 


perienced  observers  to  col- 
lect a  Taylor-type  data  set. 

5.  It  should  therefore  be  practi- 
cal to  pre-screen  potential 
release  sites  for  habitat  suita- 
bility relative  to  Present  sites. 
Objective  pre-screening  has 
not  always  been  possible  be- 
cause no  "standard  of  excel- 
lence" has  been  available. 

6.  It  also  should  be  practical  to 
plan,  implement  and  evalu- 
ate prescriptive  habitat  man- 
agement for  the  benefit  of  the 
Delmarva  fox  squirrel  on  oc- 
cupied sites  or  potential  re- 
lease sites.  The  important 
measures  of  habitat  structure 
(e.g.,  understory  vegetation 
density)  tend  to  be  variables 
which  are  amenable  to 
silvicultural  manipulation 
(Nixon  etal.  1980). 

Conclusions 

We  anticipated  at  the  outset  that  each 
of  three  potentially  important  com- 
ponents of  habitat  suitability — forest 
habitat  structure,  forest  tree  species 
composition,  and  surrounding  land- 
scape dimensions — would  influence 
the  present  occurrence  of  the  Del- 
marva fox  squirrel  in  forest  stands  on 
the  Eastern  Shore  of  Maryland.  The 
analyses  reported  here  produced  a 
number  of  surprises. 

Habitat  structure  is  the  only  com- 
ponent that  both  discriminates  be- 
tween occupied  and  unoccupied  sites 
in  a  meaningful  way  and  exhibits  a 
combination  of  statistical  properties 
favorable  for  the  formulation  of  a 
predictive  classification  model  of 
habitat  suitability. 

The  analysis  of  habitat  structure 
provides  a  basis  for  optimism  that 
such  a  model  would  prove  useful 
both  for  pre-screening  potential  re- 
lease sites  and  for  planning,  imple- 
menting and  monitoring  prescriptive 
habitat  management. 


Acknowledgments 

This  study  would  not  have  been  pos- 
sible without  the  cooperation  of 
those  who  assisted  G.  J.  Taylor  in  the 
identification  of  remnant  populations 
of  the  Delmarva  fox  squirrel  on  the 
Maryland  Eastern  Shore.  G.  D.  Ther- 
res  and  G.  W.  Willey,  Sr.,  of  the 
Maryland  Department  of  Natural 
Resources  assisted  with  locating  sites 
for  the  landscape  analysis.  J.  H.  Por- 
ter assisted  with  air  photo  interpreta- 
tion and  with  the  statistical  analyses. 
J.  Peatross  prepared  the  figures,  and 
L.  M.  McCain  assisted  with  prepar- 
ing the  manuscript.  V.  Flyger  and  K. 
E.  Severson  provided  thoughtful  re- 
views of  the  manuscript.  Finally,  the 
owners  of  the  Maryland  study  sites 
generously  provided  access  to  their 
property  for  purposes  of  habitat 
characterization.  This  study  was  sup- 
ported by  funding  from  the 
Nongame  and  Endangered  Species 
Program  of  the  Virginia  Department 
of  Game  and  Inland  Fisheries  and  by 
the  Virginia  Coast  Reserve  Long- 
Term  Ecological  Research  Program 
(NSF  Grant  BSR-8702333). 

Literature  Cited 

Allen,  D.  L.  1943.  Michigan  Fox 
Squirrel  Management.  Game  Divi- 
sion Publication  100.  Lansing, 
Michigan.  404  p. 

Bakken,  Arnold.  1952.  Interrelation- 
ships of  Sciurus  caroUnensis  (Gme- 
lin)  and  Sciurus  niger  (Linnaeus)  in 
nnixed  populations.  Ph.D  Thesis, 
University  of  Wisconsin.  212  p. 

Blouin,  Michael  S.,  and  Edward  F. 
Connor.  1985.  Is  there  a  best  shape 
for  nature  reserves?  Biological 
Conservation  32:  277-288. 

Brennan,  Leonard  A.,  William  M. 
Block,  and  R.  J.  Gutierrez.  1986. 
The  use  of  multivariate  statistics 
for  developing  habitat  suitability 
index  models,  p.  177-182.  In  J. 
Verner,  M.  I.  Morrison  and  C.  J. 
Ralph,  eds.  Wildlife  2000:  Model- 
ing Habitat  Relationships  of  Ter- 


restrial Vertebrates.  U.  Wisconsin 
Press,  Madison.  470  p. 

Brown,  Louis  G.,  and  Lee  E.  Yeager. 
1945.  Fox  squirrels  and  gray  squir- 
rels in  Illinois.  Illinois  Natural  His- 
tory Survey  Bulletin  23:449-536. 

Dueser,  Raymond  D.,  and  H.  H. 
Shugart,  Jr.  1978.  Microhabitats  in 
a  forest-floor  small  mammal 
fauna.  Ecology  59:89-98. 

Dueser,  Raymond  D.,  and  Karen  Ter- 
williger.  1988.  Status  of  the  Del- 
marva fox  squirrel  (Sciurus  niger 
cinereus)  in  Virginia.  Virginia  Jour- 
nal of  Science  38(4):380-388. 

Efron,  Bradley.  1979.  Computers  and 
the  theory  of  statistics:  thinking 
the  unthinkable.  SI  AM  Review 
21:460-480. 

Fahrig,  Lenore,  and  G.  Merriam. 
1985.  Habitat  patch  connectivity 
and  population  survival.  Ecology 
66:1762-1768. 

Hyger,  Vagn,  and  Loren  W.  Lustig. 
1976.  The  potential  for  re-estab- 
lishing fox  squirrels  in  portions  of 
their  former  range  in  the  north- 
eastern states.  Transactions  of  the 
Northeast  Section  of  the  Wildlife 
Society  33:13-17. 

Flyger,  Vagn,  and  Deborah  A.  Smith. 
1980.  A  comparison  of  Delmarva 
fox  squirrel  and  gray  squirrel 
habitats  and  home  range.  Transac- 
tions of  the  Northeast  Section  of 
the  Wildlife  Society  37:19-22. 

Hanski,  1. 1982.  Dynamics  of  regional 
distribution:  the  core  and  satellite 
species  hypothesis.  Oikos  38:210- 
221. 

Houston,  B.  R.,  Tim  W.  Clark,  and  S. 
C.  Minta.  1986.  Habitat  suitability 
index  model  for  the  black-footed 
ferret:  a  method  to  locate  trans- 
plant sites.  Great  Basin  Naturalist 
Memoirs  8:99-114. 

McComb,  William  C,  and  Robert  E. 
Noble.  1981.  Nest-box  and  natu- 
ral-cavity use  in  three  mid-South 
forest  habitats.  Journal  of  Wildlife 
Management  45:93-101. 

Middleton,  J.,  and  G.  Merriam.  1981. 
Woodland  mice  in  a  farmland  mo- 
saic. Journal  of  Applied  Ecology 
18:703-710. 


420 


Nie,  Norman  H.,  C.  Hadlai  Hull,  Jean 
G.  Jenkins,  Karen  Steinbrenner, 
and  Dale  H.  Brent.  1975.  SPSS:  Sta- 
tistical Package  for  the  Social  Sci- 
ences, 2nd  ed.,  McGraw-Hill,  N.Y. 
675  p. 

Nixon,  C.  M.,  and  L.  P.  Hansen.  1987. 
Managing  forests  to  maintain 
populations  of  gray  and  fox  squir- 
rels. Illinois  Dept.  of  Conservation 
Technical  Bulletin  5.  35  p. 

Nixon,  C.  M.,  S.  P.  Havera,  and  R.  E. 
Greenberg.  1978.  Distribution  and 
abundance  of  the  grey  squirrel  in 
Illinois.  Illinois  Natural  History 
Survey  Biological  Notes  105.  55  p. 

Nixon,  C.  M.,  S.  P.  Havera,  and  L.  P. 
Hansen.  1980.  Initial  response  of 
squirrels  to  forest  changes  associ- 
ated with  selection  cutting.  Wild- 
life Society  Bulletin  8:298-306. 

Press,  S.  James,  and  Sandra  Wilson. 
1978.  Choosing  between  logistic 
regression  and  discriminant  analy- 
sis. Journal  of  the  American  Statis- 
tical Association  73:699-705. 

Sanderson,  H.  Reed,  William  M. 
Healy,  James  C.  Pack,  John  D.  Gill, 
and  Jack  Ward  Thomas.  1976. 
Gray  squirrel  habitat  and  nest-tree 
preference.  Proceedings  of  the 
Annual  Conference  of  the  South- 
eastern Association  of  Game  and 
Fisheries  Commissions  29:609-616. 

Taylor,  Gary  J.  1973.  Present  status 
and  habitat  survey  of  the  Del- 
marva  fox  squirrel  (Sciurus  niger 
cinereus)  with  a  discussion  of  rea- 
sons for  its  decline.  Proceedings  of 
the  Annual  Conference  of  the 
Southeastern  Association  of  Game 
and  Fisheries  Commissions  27:278- 
289. 

Taylor,  Gary  J.  1976.  Range  determi- 
nation and  habitat  description  of 
the  Delmarva  fox  squirrel  in 
Maryland.  M.S.  Thesis,  University 
of  Maryland,  College  Park.  76  p. 

Taylor,  Gary  J.,  and  Vagn  Flyger. 
1973,  Distribution  of  the  Delmarva 
fox  squirrel  (Sciurus  niger  cinereus) 
in  Maryland.  Chesapeake  Science 
14:59-60. 

Taylor,  Gary  J.,  Vagn  Flyger,  N. 
Swink,  and  G.  Willey.  1983.  Del- 


marva fox  squirrel  recovery  plan. 
First  revision.  U.S.  Fish  and  Wild- 
life Service.  Washington,  D.C. 
49  p. 

U.S.  Department  of  Interior,  Bureau 
of  Sport  Fisheries  and  Wildlife. 
1970.  Endangered  Species  of  the 
United  States.  Washington,  D.C. 
4  p. 

Van  Horne,  B.  1983.  Density  as  a  mis- 
leading indicator  of  habitat  qual- 
ity. Journal  of  Wildlife  Manage- 
ment 47:893-901. 

Wegner,  J.F.,  and  G.  Merriam.  1979. 
Movements  by  birds  and  small 
mammals  between  a  wood  and 
adjoining  farmland  habitats.  Jour- 
nal of  Applied  Ecology  16:349-357. 

Weigl,  P.  D.,  M.  A.  Steele,  L.  J.  Sher- 
man, J.  C.  Ha,  and  T.  S.  Sharpe.  In 
press.  The  ecology  of  the  fox 
squirrel  (Sciurus  niger)  in  North 
Carolina:  Implications  for  survival 
in  the  southeast.  Tall  Timbers  Re- 
search Series. 


421 


Effects  of  Treating 
Creosotebush  With 
Tebuthiuron  on  Rodents^ 

William  G.  Standley^  and  Norman  S.  Smith^ 


Abstract.— Three  years  after  creosotebush  (Lauea 
tridentata)  v/os  treated  with  tebuthiuron,  rodent 
abundance  was  71%  higher  on  treated  plots  than 
on  controi  plots  in  southeastern  Arizona.  Arizona  cot- 
ton rats  (Sigwodon  arizonae)  and  Western  harvest 
mice  (Reifhrodontomys  megalotis)v^ere  more  abun- 
dant while  the  abundance  of  Merriam's  kangaroo 
rots  (Dipodomys  merriami)       similar.  We  conclude 
that  tebuthiuron  may  be  safely  used  to  control  creo- 
sotebush in  semidesert  grasslands  unless  the  pres- 
ence of  rare  or  endangered  species  precludes  any 
alterations  to  the  community. 


Herbicides  are  often  used  to  control 
shrubs  such  as  mesquite  (Prosopis  ju- 
Uflora)  and  creosotebush  (Larrea 
tridentata),  which  have  invaded  mil- 
lions of  hectares  of  semidesert  grass- 
lands (Cox  et  al.  1982).  The  reduction 
of  shrub  cover  usually  results  in  an 
increase  in  forage  production  (Box 
1964). 

The  herbicide  2,4-D  has  been  used 
for  more  than  two  decades  and  its 
effects  on  rodent  communities  have 
been  extensively  studied  (Keith  et  al. 
1959,  Johnson  and  Hansen  1969, 
Spencer  and  Barrett  1980).  2,4-D  has 
varied  effects  on  rodent  communi- 
ties, increasing  the  abundance  of 
some  species,  while  decreasing  the 
abundance  of  others  (Johnson  and 
Hansen  1969,  Spencer  and  Barrett 
1980). 

Tebuthiuron  is  a  thiadiazolyl-urea 
herbicide  (Walker  et  al.  1973)  used  to 
control  shrubs  in  the  southwest  (Her- 
bel  et  al.  1985).  No  studies  have  been 
conducted  to  determine  effects  of 
tebuthiuron  treatments  on  rodent 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Northi  America.  (Flag- 
staff. AZ.  July  19-21.  1988.) 

'William  G.  Standley.  formerly  a  gradu- 
ate student.  University  of  Arizona.  Arizona 
Cooperative  Rshi  and  Wildlife  Research» 
Unit,  is  currently  Animal  Ecologist.  EG&G 
Energy  Measurements.  Inc..  c/o  NPR-1.  P.O. 
Box  127.  Tupman.  CA.  93276. 

^Norman  S.  Smitti  is  Assistant  Leader.  Ari- 
zona Cooperative  Fist^  and  Wildlife  Re- 
search! Unit.  University  of  Arizona.  Tucson.  AZ 
85721. 


communities.  We  studied  grasslands 
invaded  by  creosotebush  in  south- 
eastern Arizona  in  order  to  deter- 
mine changes  that  take  place  in  a  ro- 
dent community  due  to  treatment 
with  tebuthiuron.  We  compared 
vegetation  and  nocturnal  rodents 
present  on  control  and  treated  plots. 
Because  tebuthiuron  is  nontoxic  to 
laboratory  mice,  rats,  and  rabbits 
(Morton  and  Hoffman  1976)  we  as- 
sumed that  any  changes  in  the  rodent 
community  would  be  in  response  to 


changes  in  food  supply,  ground 
cover,  or  both. 


Methods 

Two  adjacent  150  x  600  m  plots  were 
fenced  from  cattle,  and  one  was  aeri- 
ally treated  with  tebuthiuron  (1.0  kg/ 
ha)  in  May  1981  as  part  of  an  ongo- 
ing experiment  on  the  USD  A  Forest 
Service  Santa  Rita  Exp)erimental 
Range,  45  km  south  of  Tucson  Ari- 


r 


Table  L— Mean  (%)  vegetative  cover  on  tebuthiuron  treated  and  control 
plots  (N=6). 


Treated 


Control 


Species 


SE 


SE 


Grasses 
Threeawn 

(Aristida  sp.) 
Bush  muhly 

(MuNenbergia  porter!) 
Fluff  grass 

(Tridens  pulchellus) 
Other 

Total 

Shrubs 

Creosotebush 

(Larrea  iridentafa) 
Mesquite 

(Prosopis  juliflora) 
Desert  zinnia 

(Zinnia  pumila) 
Desertbroom 

(Baccharis  sarofhroides) 

Total 


18.5 

3.2 

0.1 

0.1 

10.0 

2.5 

11.0 

3.4 

3.0 

1.3 

0.7 

0.7 

0.3 

0.3 

0.0 

0.0 

31.8 

1.9 

11.8 

3.2 

0.2 

0.1 

33.9 

4.6 

0.0 

0.0 

1.4 

0.9 

0.0 

0.0 

3.3 

1.4 

0.5 

0.5 

0.0 

0.0 

0.7 

0.5 

38.6 

4.5 

422 


zona.  Vegetation  on  the  plots  is 
dominated  by  creosotebush,  with 
sparse  grasses  such  as  threeawn 
{Aristida  sp.)  and  bush  muhly 
(Muhlenbergia  porteri)  (Martin  and 
Reynolds  1973). 

We  sampled  vegetation  and  ro- 
dent communities  in  June  1984,  three 
years  after  herbicide  treatment. 
Vegetation  was  sampled  using  the 
line  intercept  method  (Canfield 
1941).  Six  30  m  parallel  lines  were 
systematically  located  on  each  plot. 
Total  intercepts  of  each  species  were 
averaged  and  transformed  into  per- 
cent ground  and  canopy  cover.  Ro- 
dent communities  were  surveyed  us- 
ing the  removal  method.  Sherman 
live-traps  (7.5  x  7.5  x  25  cm)  were 
used  so  that  rodents  could  be  used 
for  other  studies.  Three  8  by  8  grids 
with  traps  spaced  at  10  m  intervals, 
were  placed  on  each  plot.  Grids  were 
placed  as  far  from  each  other  and 
from  plot  boundaries  as  possible,  re- 
sulting in  a  uniform  distribution. 
Traps,  opened  at  sunset  and  closed  at 


sunrise,  were  baited  with  p>eanut 
butter  and  oats.  We  prebaited  traps 
for  one  night  then  removed  all  ro- 
dents captured  during  the  following 
four  nights.  The  total  number  of  each 
species  captured  on  the  three  grids 
on  each  plot  were  averaged. 


Results 

Average  grass  cover  on  the  tebuth- 
iuron-treated  plot  was  almost  three 
times  that  on  the  control  plot  (table 
1),  with  threeawn  contributing  most 
of  the  difference.  Average  shrub 
cover  on  the  treated  plot  was  98% 
lower  than  on  the  control  plot,  with 
creosotebush  accounting  for  the  big- 
gest difference. 

On  tebuthiuron-treated  grids  we 
captured  162  rodents  of  eight  species, 
and  on  control  grids  95  rodents  of 
eight  sf>ecies  (table  2).  Higher  num- 
bers of  Arizona  cotton  rats  (Sigmodon 
arizonae)  and  western  harvest  mice 
(Reithrodontomys  megalotis)  on  the 


r 


Table  2.— Mean  number  of  rodents  captured  on  tebuthiuron  treated  and 
control  plots  (N=3). 


Treated 


Control 


Species 


SE 


SE 


Merriam's  kangaroo  rat 

(DIpodomys  m  erriami) 
Arizona  pocket  mouse 

(Perognafhus  amplus) 
White-throated  wood  rat 

(Neotoma  olbiguta) 
Western  harvest  mouse 

(Reifhrodonfomys  megalotis) 
Arizona  cotton  rat 

(Sigmodon  arizonae) 
Desert  pocket  mouse 

(Perognafhus  penicillatus) 
Southern  grasshopper  mouse 

(Onychomys  torridus) 
Bailey's  pocket  mouse 

(Perognatfius  baileyi) 
Deer  mouse 

(Peromyscus  maniculatus) 
House  mouse 

(Mus  musculus) 

Total 


15.7 
5.3 
8.0 

10.0 
8.0 
3.0 
3.7 
0.0 
0.0 
0.3 

54.0 


4.3 
2.6 
1.5 
1.5 
3.2 
0.0 
0.3 
0.0 
0.0 
0.3 
5.5 


13.0 
7.0 
4.3 
0.3 
0.0 
4,0 
1.7 
1.0 
0.3 
0.0 

31.6 


2.1 
0.6 
1.5 
0.3 
0.0 
1.0 
0.7 
1.0 
0.3 
0.0 
2.7 


treated  grids  accounted  for  most  of 
the  difference  in  abundance.  Cotton 
rats  and  house  mice  (Mus  musculus) 
were  captured  only  on  the  treated 
grids,  while  Bailey's  pocket  mice 
(Perogmthus  baileyi)  and  deer  mice 
(Peromyscus  maniculatus)  were  caught 
only  on  control  grids. 


Discussion 

The  dramatically  greater  grass  cover 
and  lesser  shrub  cover  on  the  treated 
plot  are  consistent  with  results  of 
other  experiments  with  tebuthiuron 
(Herbel  et  al.  1985),  as  well  as  with 
2,4-D  (Spencer  and  Barrett  1980). 
This  difference  in  vegetative  struc- 
ture app>ears  to  account  for  most  of 
the  differences  in  the  rodent  commu- 
nity. Studies  of  cotton  rats  and  har- 
vest mice  have  shown  that  both  spe- 
cies are  strongly  associated  with 
dense  stands  of  grass  (Goertz  1964, 
Ford  1977).  The  similarity  in  abun- 
dance of  Merriam's  kangaroo  rats  on 
control  and  treated  plots  was  unex- 
f)ected  since  heteromyids  are  gener- 
ally more  abundant  in  areas  with 
sparse  ground  cover  (Stamp  and 
Ohmart  1978). 

We  do  not  present  inferential  sta- 
tistics to  test  differences  in  ground 
cover  or  rodent  numbers  because 
both  the  line  intercept  transects  and 
trap  grids  were  actually  subsamples 
rather  than  true  replicates  (Hurlbert 
1984).  We  are  convinced,  however, 
that  differences  between  plots  in 
numbers  of  cotton  rats  and  harvest 
mice,  are  the  result  of  habitat 
changes  following  treatment  with 
tebuthiuron. 

Because  of  the  low  numbers  of 
deer  mice  and  Bailey's  pocket  mice 
captured  on  the  control  plots,  we  do 
not  feel  their  absence  on  the  treated 
plots  is  significant.  Because  the  re- 
sponses were  either  neutral  or  posi- 
tive, we  feel  that  tebuthiuron  can  be 
safely  used  by  managers  to  control 
shrubs  in  semidesert  grasslands 
without  fear  of  endangering  rodents 
directly.  However,  the  impact  of 


423 


habitat  changes  on  rare  or  endan- 
gered species  should  not  be  ignored. 

Acknowledgments 

This  study  was  funded  by  the  USDA 
Arid  Land  Ecosystems  Improvement 
Unit.  We  thank  J.  Ard,  J.  Brown,  B. 
Kotler,  and  B.  Zoellick  for  field  assis- 
tance. 


Literature  Cited 

Box,  Thadis  W.  1964.  Changes  in 
wildlife  habitat  composition  fol- 
lowing brush  control  practices  in 
south  Texas.  Transactions  of  the 
North  American  Wildlife  Confer- 
ence 29:432-438. 

Canfield,  R.  H.  1941.  Application  of 
the  line  interception  method  in 
sampling  range  vegetation.  Jour- 
nal of  Forestry  39:388-394. 

Cox,  Jerry  R.,  Howard  L.  Morton, 
Thomas  N.  Johnsen  Jr.,  Gilbert  L. 
Jordan,  S.  Clark  Martin,  and  Louis 
C.  Fierro.  1982.  Vegetation  restora- 
tion in  the  Chihuahuan  and  Sono- 
ran  deserts  of  North  America. 
United  States  Department  of  Agri- 
culture, Agricultural  Research 
Service,  Agricultural  Reviews  and 
Manuals,  Western  Series  Number 
28,  August  1982. 

Ford,  Steven  D.  1977.  Range,  distri- 
bution and  habitat  of  the  western 
harvest  mouse,  Reithrodontomys 
megalotis,  in  Indiana.  American 
Midland  Naturalist  98:422-432. 

Goertz,  John  W.  1964.  The  influence 
of  habitat  quality  upon  density  of 
cotton  rat  populations.  Ecological 
Monographs  34:359-381. 

Herbel,  Carlton  H.,  Howard  L.  Mor- 
ton, and  Robert  P.  Gibbens.  1985. 
Controlling  shrubs  in  the  arid 
southwest  with  tebuthiuron.  Jour- 
nal of  Range  Management  38:391- 
394. 

Hurlbert,  Stuart  H.  1984.  Pseu- 
doreplication  and  the  design  of 
ecological  field  experiments.  Eco- 
logical Monographs  54:187-211. 


Johnson,  Donald  R.  and  Richard  M. 
Hansen.  1969.  Effects  of  range 
treatment  with  2,4-D  on  rodent 
populations.  Journal  of  Wildlife 
Management  33:125-132. 

Keith,  James  O.,  Richard  M.  Hansen, 
and  A.  Lorin  Ward.  1959.  Effect  of 
2,4-D  on  abundance  and  food  of 
pocket  gophers.  Journal  of  Wild- 
life Management  23:137-145. 

Martin,  S.  Clark,  and  Hudson  G. 
Reynolds.  1973.  The  Santa  Rita 
Experimental  Range:  Your  facility 
for  research  on  semidesert  ecosys- 
tems. Journal  of  the  Arizona  Acad- 
emy of  Science  8:56-67. 

Morton,  D.  M.,  and  D.  G.  Hoffman. 
1976.  Metabolism  of  new  herbicide 
tebuthiuron  (l(5-(l,l-dimethyl- 
l,3,4-thiadiazol-2-yl)-l,3-dimethyl- 
urea)  in  mouse,  rat,  rabbit,  dog, 
duck,  and  fish.  Journal  of  Toxicol- 
ogy and  Environmental  Health 
1:757-768. 

Spencer,  Stephen  R.  and  Gary  W. 
Barrett.  1980.  Meadow  vole  popu- 
lation response  to  vegetational 
changes  resulting  from  2,4-D  ap- 
plication. American  Midland 
Naturalist  103:32-46. 

Stamp,  Nancy  E.,  and  Robert  D. 
Ohmart.  1978.  Resource  utilization 
by  desert  rodents  in  the  lower 
Sonoran  desert.  Ecology  59:700- 
707. 

Walker,  J.  C,  M.  L.  Jones,  and  J.  E. 
Shaw.  1973.  Total  vegetation  con- 
trol with  tebuthiuron — a  new 
broad  spectrum  herbicide.  Pro- 
ceedings of  the  North  Central 
Weed  Control  Conference  28:39. 


424 


Foraging  Patterns  of  Tassel- 
Eared  Squirrels  In  Selected 
Ponderosa  Pine  Stands^ 

Jack  S.  States,^  William  S.  Gaud,^  W. 
Sylvester  Allred/  and  William  J.  Austin^ 


Abstract.— Pine  seed,  primarily  available  in  the  fall, 
and  hypogeous  fungi,  potentially  available  in  all 
seasons,  were  major  food  items  whose  consumption 
was  associated  with  an  increase  in  biweekly  body 
weights  of  marked  squirrels.  Use  of  alternative  foods 
such  OS  twigs  (inner  bark)  and  apical  buds  occurred 
when  these  food  items  were  unavailable.  Consump- 
tion of  inner  bark  and  buds  was  highest  in  winter 
(93%)  and  spring  (86%).  Although  feed  tree  prefer- 
ence was  noted,  widespread  feeding  occurred  in 
more  than  half  of  the  trees  in  both  study  sites.  The 
resulting  variability  in  physical  evidence  of  foraging 
suggests  caution  in  its  use  for  indexing  squirrel  popu- 
lation densities. 

bark  in  the  absence  of  supplemental 
foods  could  potentially  threaten 
squirrel  survival  during  adverse 
weather  conditions  (Patton  1974). 
The  obvious  selection  of  certain  trees 
by  squirrels  for  inner  bark  consump- 
tion sup]X)rts  the  assumption  that 
there  are  differences  in  quality  of 
trees  in  the  same  stand.  In  a  food 
preference  study  using  captive  Abert 
squirrels,  Farentinos  et  al.  (1981) 
were  able  to  show  a  significant  rela- 
tionship between  selective  consump- 
tion of  inner  bark  and  low  oleoresin 
content.  However,  Pederson  and 
Welch  (1987)  noted  a  strong  feeding 
preference  for  trees  with  inner  bark 
that  was  easily  peeled  with  no  appar- 
ent relationship  between  inner  bark 
oleoresin  content  and  "feed  tree"  se- 
lection. 

Studies  on  the  impacts  of  squirrel 
herbivory  on  ponderosa  pine  have 
lead  to  mixed  conclusions.  Hall 
(1981)  and  Ffolliott  and  Patton  (1978) 
found  that  heavy  utilization  of  pine 
twigs  had  negligible  effect  on  stand 
productivity,  although  Hall  demon- 
strated significant  growth  decreases 
of  individual  feed  trees.  Soderquist 
(1987)  reported  twig  clipping  to  de- 
crease tree  growth  in  ecotonal  stands 
of  ponderosa  pine.  Pearson  (1950) 
and  Larson  and  Schubert  (1970) 
noted  extensive,  but  seasonally  vari- 
able, damage  to  cone  crops.  They 
were  unable  to  determine  the  causes 
of  the  highly  variable  pattern  of  her- 
bivory during  several  years  of  obser- 
vation. 


The  tassel-eared  tree  squirrel  (Sciurus 
aberti)  and  its  several  subspecies,  has 
a  unique  and  apparent  obligatory  as- 
sociation with  Southwestern  ponder- 
osa pine  (Pinus  ponderosa).  The  die- 
tary dependence  of  these  squirrels  on 
pine,  including  inner  bark  and  buds 
of  terminal  twigs  and  both  staminate 
and  ovulate  cones,  identifies  the 
squirrel  as  an  herbivore  having  a  po- 
tentially major  influence  on  the 
growth  and  reproduction  of  ponder- 
osa pine.  Conversely,  extensive  har- 
vest of  pine  for  wood  products  has 
resulted  in  deterioration  of  the  squir- 
rel's habitat  since  the  turn  of  the  cen- 
tury (Keith  1965). 

A  number  of  studies  have  at- 
tempted to  explain  the  '"boom  and 
bust"  p>opulation  fluctuations  that 
seem  to  be  characteristic  of  tassel- 
eared  squirrels.  In  his  observations 

'Paper  presented  at  symposium,  f\/1an- 
agement  of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ,  July  19-21,  1988.) 

^Jack  S.  States  is  Professor  of  Biology, 
Department  of  Biological  Sciences,  Noriti- 
em  Arizona  University,  Flagstaff,  AZ  8601 1. 

^William  S.  Gaud  is  Associate  Professor  of 
Biology,  Department  of  Biological  Sciences, 
Northern  Arizona  University.  Ragstaff,  AZ 
86011. 

^W.  Sylvester  AJIred  is  a  doctoral  candi- 
date in  the  Department  of  Biological  Sci- 
ences, Northern  Arizona  University,  Flagstaff, 
AZ86011. 

^William  J.  Austin  is  a  doctoral  candi- 
date in  the  Department  of  Biological  Sci- 
ences, Northern  Arizona  University.  Flagstaff, 
AZ86011. 


on  the  ecology  of  Abert  squirrels  (S. 
a.  aberti),  Keith  (1965)  attributed  short 
term  fluctuations  to  changes  in  quan- 
tity and  quality  of  major  food  items 
assumed  to  be  provided  by  pine. 
However,  high  mortality  in  some 
years  apparently  resulted  from  some 
factors  other  than  food.  Hall  (1981), 
in  a  study  of  Kaibab  squirrels  (S.  a. 
kaibabensis)  also  observed  population 
fluctuations.  He  suggested  that  sea- 
sonal differences  in  food  resources 
and  snowfall  were  potential  causes  of 
declines  and  recovery.  Availability 
and  use  of  various  food  items  have 
not  been  adequately  studied. 

Stephenson  (1974)  in  a  study  of 
Abert  diets  discovered  that  fungi 
were  a  major  part  of  the  total  food 
consumed.  The  fungi  in  Stephenson's 
samples  were  identified  as  belonging 
to  a  subterranean  group  of  mush- 
rooms popularly  called  truffles  (]. 
States,  unpublished  data),  which  are 
known  to  form  mycorrhizal  associa- 
tions with  pine  roots.  Some  of  these 
fungi  were  found  to  be  new  records 
for  the  Southwest  (States  1983, 1984) 
and  they  were  found  to  be  primarily 
associated  with  blackjack  age-class 
ponderosa  pine  stands  with  high  can- 
opy densities  (States  1985). 

A  telltale  sign  of  the  activity  of  the 
tassel-eared  squirrel  is  the  presence 
of  clipped  twigs  on  the  ground  under 
a  tree  after  the  squirrel  has  removed 
the  terminal  shoot  from  a  branch. 
The  nutritional  value  of  the  inner 
bark  consumed  by  the  squirrel  is 
low.  A  diet  comprised  solely  of  inner 


425 


The  tassel-eared  squirrel's  variety 
of  diet  and  use  of  the  forest  has  lead 
to  differences  of  opinion  regarding 
the  best  management  plan  for  both 
squirrel  and  forest.  Patton  et  al. 
(1985)  considered  tree  density,  size, 
and  patchy  distribution  to  be  major 
factors  constituting  habitat  quality 
since  squirrels  use  pine  for  cover  and 
nesting  as  well  as  for  food. 

The  purpose  of  this  study  was  to 
determine  the  seasonal  patterns  of 
food  resource  utilization  by  Abert 
squirrels  in  selected  ponderosa  pine 
stands  and  to  relate  the  results  to 
squirrel  population  levels  within  the 
stands.  The  use  of  fungi  and  inner 
bark  as  major  food  items  is  discussed 
as  it  pertains  to  stand  characteristics 
and  the  potential  impact  of  herbivory 
on  ponderosa  pine. 

Study  Areas  and  Methods 

Two  sites  in  clumped,  uneven-aged 
ponderosa  pine  stands  were  studied 
in  areas  that  had  not  been  disturbed 
by  fire  and  timber  harvest  for  the 
past  35  years.  A  9.3  ha  site  was  lo- 
cated on  the  property  of  the  Lowell 
Observatory  and  adjacent  to  the  Co- 
conino National  Forest.  The  other  site 
of  2.5  ha  was  located  in  the  Mount 
Elden  Environmental  Study  area  of 
the  Coconino  National  Forest.  The 
elevation  of  both  sites  was  2150  m, 
and  they  were  within  10  km  of  the 
Flagstaff  airport  where  weather  data 
used  in  the  study  was  collected 
(NOAA  1987). 

Squirrels  were  captured,  marked, 
and  released  at  the  observatory  site. 
At  this  location  there  were  90  plots, 
625  m^  each,  in  a  nested  trapping 
grid  similar  to  the  one  described  by 
Patton  et  al.(1985).  The  grid  con- 
tained 42  systematically  spaced 
Tomahawk  live-traps  baited  and  set 
for  eight  daylight  hours  once  each 
week.  Trapping  was  conducted  from 
September  1985  through  June  1987 
and  squirrel  body  weight  was  re- 
corded. Fecal  pellets  deposited  in 
traps  were  collected  and  analyzed  to 


{  \ 

Table  1  .—Estimated  food  availability  for  Abert  squirrels. 


Truffles    Seed  cones/tree  Acorns/tree  Mushroom 
Year  kg/ha  xn=25  xn=20  abundarice 


1983  2.88  144  552  high 

1984  0.86  10  83  low 

1985  0.39  10  123  very  low 

1986  0.72  137^  20^-  very  high 

1987  0,65  10  10»^  moderate 


°Cone  crop  2}%  aborted  due  to  insect  damage;  31%  of  total  cone  crop  harvested 
by  September. 


^Acorn  production  low  due  to  early  frost. 


summer  {a\\ 


BA -branch  inn«rbark    Ml-mistletoe  SC  -staminate  cones 

BU -terminal  buds       MU-mushrooms  TR -truffles 

TW-twig  innerbaric 

co-pine  cones  OP -open  pine  cones 

1 


3     winter  spring 


Figure  1  .—Percentage  of  feeding  time  by  Abert  squirrels  for  each  diet  item  in  each  season 
(from  eighiteen  two-hour  periods  per  secjson,  Coconino  National  Forest,  AZ). 


426 


determine  the  ratio  of  dietary  fungus 
to  plant  matter  and  to  identify  the 
fungi  through  spore  characteristics. 

Observational  data  on  foraging 
behavior  was  collected  using  focal 
animal  sampling  (Altmann  1974).  We 
observed  four  individually  marked 
males  from  July  1986  through  No- 
vember 1987  as  they  foraged  in  the 
study  site.  Data  were  collected  in  18 
two-hour  observation  periods  in  each 
of  four  "ecological"  seasons.  These 
seasons  were  established  by  combin- 
ing months  with  similar  temperature 
and  precipitation  means.  The  seasons 
correspond  to  periods  of  truffle  pro- 
duction: Season  I,  December-March 
(winter);  Season  II,  April-June 


(spring);  Season  III,  July- August 
(summer);  Season  IV,  September- 
November  (fall). 

Resource  availability  and  physical 
signs  of  foraging  acrivity  were  re- 
corded over  a  20-month  period  (June 
1986  through  January  1988).  Cone 
consumption,  twig  clipping,  and  dig- 
ging activity  were  documented  on  26 
contiguous  625  m^  plots  (1.6  ha)  on 
the  observatory  site.  During  each  of 
the  biweekly  censuses,  all  twig 
remnants,  cone  cores,  and  digs 
(truffle  excavations)  were  recorded 
with  a  notation  of  the  nearest  tree. 
Trees  were  characterized  by  age-class 
(blackjack,  or  yellow  pine)  and  di- 
ameter at  breast  height  (DBH).  The 


entire  2.5  ha  Mount  Elden  site  was 
sampled  for  seasonal  foraging  pat- 
terns. Permanently  marked  pines  and 
oaks  in  20  plots  were  censused  yearly 
in  September  for  acorn  and  cone  pro- 
duction. Cones  and  acorns  were 
counted  on  a  quarter  of  the  tree  and 
multiplied  by  four  to  obtain  a  pro- 
duction estimate.  Truffle  production 
estimates  were  made  according  to 
States  (1985).  Relative  mushroom 
abundance  was  determined  by 
counting  numbers  of  mushrooms 
present  within  10  randomly  placed  50 
m^  quadrats  sampled  in  the  fall. 

Results 

Resources  available  as  food  for  tas- 
sel-eared squirrels  showed  consider- 
able annual  variability  (table  1).  The 
four  food  items  were  all  relatively 
abundant  in  1983,  but  availability 
subsequently  dropped.  Production  of 
cones  and  mushrooms  was  relatively 
high  again  in  1986,  but  there  were 
considerably  fewer  truffles  and 
acorns  present  than  in  1983.  In  gen- 
eral, the  quantity  of  truffle  produc- 
tion was  more  consistent  that  it  was 
for  the  other  foods. 

Seasonal  foraging  behavior  of  the 
squirrels  (fig.  1)  reflected  changes  in 
availability  of  food  items.  The  ani- 
mals heavily  utilized  a  large  cone 
crop  in  1986  before  the  seeds  were 
mature,  and  continued  to  utilize  it 
into  November  when  the  remaining 
seeds  were  released.  Cone  and  truffle 
use  dropped  abruptly  from  a  fall 
high  of  80%  feeding  time  when  the 
squirrels  switched  to  intensive  feed- 
ing on  buds  and  inner  bark  of  twigs. 
This  behavior  comprised  85%  of  the 
spring  feeding  time.  Collectively, 
pine  products  constituted  the  largest 
portion  of  the  diet  through  winter 
and  spring.  Seasonal  patterns  were 
apparent  in  the  use  of  different  parts 
of  the  tree. 

The  physical  evidence  left  in  the 
forest  by  the  squirrels  verified  a  sea- 
sonal progression  of  food  item 
availability  (fig.  2).  Numbers  of  digs 


2400  n 
2200- 
2000- 
3  1600 
^  1600- 


o 

i 


TWIGS  CUPPED    CONES  HARVESTED     KUICBEK  OF  DIGS 


BO 

70 
00 
50 
40 
30 


g  20 


10 
0 


m 


mm 


JJASONDJFMAUJJASONDi 
1066  1067  1066 

MONTHS 


Figure  2.— Monthly  resource  usage  by  Abert  squirrels  compared  to  total  precipitation  (solid 
bar)  and  snow  deptti  (steaded  bar),  Coconino  Natiorral  Forest,  AZ. 


427 


peaked  in  late  fall  and  dropped  in 
the  winter,  a  pattern  which  corre- 
sponds to  foraging  time  percentages 
for  truffles  (fig.  1).  Subsequent  digs 
during  winter  and  early  spring  repre- 
sented retrieval  of  cones  buried  the 
previous  fall.  Numbers  of  cone  cores 
left  after  seed  removal  increased  as 
the  cones  matured.  Numbers  of  cone 
cores  susequently  decreased  in  the 
winter  months.  As  the  use  of  cones 
and  truffles  declined,  numbers  of 
clipped  twigs  increased.  Twig  clips 
decreased  to  a  moderate  level  in 
summer  but  again  reached  a  high 
level  the  following  winter.  This  peak 
was  coincident  with  increased  snow 
depth,  a  decrease  in  availability  of 
truffles,  and  the  absence  of  seed 
cones.  By  the  end  of  January  1988,  a 
majority  of  the  1114  trees  (67.9%)  in 
the  observatory  site  had  been  clipped 
at  least  once  with  an  average  of  al- 
most 10  clips  per  tree. 

In  spite  of  relatively  low  food 
availability  in  the  year  from  Septem- 
ber 1985  through  August  1986  (table 
1),  resident  squirrels  maintained  a 
fairly  constant  weight  throughout  the 
winter  (table  2).  The  average  weight 
of  four  male  squirrels  dropped  6% 
during  spring  and  early  summer 
from  the  previous  fall's  high.  Subse- 
quent weight  gains,  5%,  occurred 
concurrently  with  maturation  of  the 
1986  fall  cone  crop.  Winter  weight 
loss  paralleled  the  decrease  in  availa- 
bility of  fungi,  as  evidenced  by  their 
presence  in  fecal  contents  (fig.  3). 

The  number  of  terminal  shoots 
removed  by  squirrels  in  the  Mt. 
Elden  site  from  1984-1987  was  appar- 
ently related  to  snowfall  (table  3). 
Squirrel  densities  remained  relatively 
constant,  but  the  number  of  trees 
clipped  increased  by  30%  and  the  av- 
erage number  of  clips  per  tree  de- 
creased by  81%.  Total  snowfall  was 
greater  in  1985  than  in  either  1986 
and  1987.  The  map  of  clipping  behav- 
ior shows  marked  shifts  in  areas  of 
heaviest  clipping  in  the  2.5  ha  site 
(fig.  4).  The  smallest  area  of  clipping 
intensity  (46%  of  the  site)  occurred  in 
1985,  and  it  also  had  the  highest 


Table  2.— Mean  body  weight  (grams)  and  standard  deviation  of  Individual 
male  squirrels  In  each  season  from  September  1985  through  November 
1986.  The  numbers  of  captures  per  season  are  in  parentheses. 


Squirrel 


IV 
Fall 


Winter 


!l 

Spring 


III 

Summer 


IV 
Fall 


1 

2 
3 
4 


614+  11.7 

(11) 
748  +  37.0 

(5) 
655  +  20.4 

(13) 
737  +  24.7 

(6) 

689 


632  +  30.1 
(12) 

680  +  34.3 
(7) 

681  +  15.2 
(12) 

688  +  31.5 
(5) 
670 


610+  12.7 
(3) 
723  + 

(1) 
671  +  18.0 
(7) 
669 

(1) 
668 


587  +  27.9 

(7) 
703+  18.6 

(5) 
648  +  29.8 

(5) 
662  +  33.6 

(7) 

650 


610  +  27.6 

(9) 
769  +  20.9 

(5) 
674  +  16.9 

(8) 
674  +  33.8 

(11) 
682 


number  of  clips  per  tree.  The  area  of 
clipping  expanded  in  1986  and  1987 
to  67%  and  64%,  respectively.  Of  the 
604  trees  clipped  in  the  three  years, 
45%  were  clipped  once  while  23% 


were  clipped  every  year.  Yellow  pine 
constituted  10%  of  the  stand  and  82% 
of  these  were  clipped.  Most  yellow 
pines  not  clipped  were  isolated  in 
open  areas.  Sixty-one  percent  of  all 


FALL  WINTER  SPRING  SUMMER 


Figure  3.— Weight  loss  (solid  line)  by  Abert  squirrels  as  compared  to  fungal  content  of  feces 
(dotted  line)  during  the  period  September  1985  to  August  1986.  Coconino  National  Forest, 
AZ. 


Table  3.— Twig  clipping  data  for  Yellowpine  (VP)  and  Blackjack  (BJ)  age- 
class  trees  over  ttiree  successive  winter  seasons  In  the  Mount  FIden  study 
area. 


Snow 

SquirrelNumber  Trees 

Clipped 

Clips/ 

Total 

Dry  wt. 

Year 

cm. 

number  YP 

BJ 

total 

tree 

clips 

kg 

1984-85 

345.4 

6  38 

155 

193 

124.7 

24,061 

295 

1985-86 

266.7 

8  52 

268 

320 

65.4 

20,640 

253 

1986-87 

217.7 

6  58 

533 

591 

24.2 

14,288 

175 

428 


blackjack  pine  with  a  DBH  greater 
than  10  cm  were  cHpped.  Eleven  of 
these  and  one  yellow  pine  died  fol- 
lowing virtually  total  canopy  re- 
moval by  squirrels.  The  average 
number  of  twigs  clipped  in  yellow 
pine  was  greater  than  in  blackjack 
pine  but  their  mean  dry  weight,  11.2 
•  8.7  g,  was  less  than  that  of  blackjack 
pine,  13.2  •  4.8  g.  Five  of  the  yellow 
pines  in  1985  had  more  than  1000 
twigs  removed  from  each.  A  majority 
of  the  trees  clipped  only  once  before 
1987  were  also  clipped  in  1987  and 
were  located  in  an  area  with  little 
previous  squirrel  use  (fig.  4). 

Discussion 

The  tassel-eared  squirrel  is  a  whole 
forest  species  in  the  sense  that  essen- 
tially all  age  classes  of  trees  are  util- 
ized. Although  pine  provides  much 
of  the  squirrel's  food,  the  various 
items  are  taken  by  the  squirrel  from 
different  age  classes  of  trees  (table  3). 
The  largest  number  of  cones  is  pro- 


Figure  4.— Map  of  the  2.4  ha  Mount  Elden 
study  site  illustrating  shifts  in  clipping  inten* 
sity  for  each  of  three  years  (spring  to 
spring),  Coconino  National  Forest,  AZ.  Clip- 
ping data  corresponding  to  these  areas  is 
presented  in  table  3. 


duced  by  mature  yellow  pines  (Lar- 
son and  Schubert  1970),  while 
truffles  tend  to  be  associated  with 
pole  sized  blackjack  pines  (States 
1985).  Thus,  prime  squirrel  habitat 
provides  optimal  food  in  stands  con- 
taining a  combination  of  tree  age 
classes  whose  size,  density,  and 
grouping  provides  cover  and  nesting 
sites  as  well  (Patton  1984). 

Major  shifts  in  foraging  by  the  tas- 
sel-eared squirrel  are  apparently  as- 
sociated with  variations  in  the  availa- 
bility of  food  resources  in  the  forest. 
In  1986  squirrels  relied  heavily  on 
pine  seeds  with  moderate  utilization 
of  truffles  and  the  inner  bark  of  twigs 
(fig.  2).  Cone  and  acorn  failure  in 
1987  resulted  in  a  reversal  of  the  rela- 
tive emphasis  on  seeds  and  twigs. 
Observation  of  squirrel  foraging  re- 
vealed a  corresponding  opportunistic 
shift  from  such  ephemeral  foods  as 
staminate  cones  and  developing  pine 
buds  in  the  spring  to  mushrooms  in 
summer  to  truffles  in  the  fall.  Similar 
opportunism  in  food  utilization  has 
also  been  reported  for  other  tree 
squirrels:  the  European  tassel-eared 
squirrel,  Sciurus  vulgaris  (Wauters 
and  Dhondt  1987),  the  American  red 
squirrel,  Tamiasciurus  hudsonicus  (Per- 
ron et  al.  1986),  and  the  western  gray 
squirrel,  S.  griseus  (Stienecker  1977). 

In  spite  of  a  seasonal  emphasis  on 
temporary  supplies  of  certain  food 
items,  the  squirrel  removed  terminal 
shoots  of  ponderosa  pine  throughout 
the  year.  When  cones  and  truffles  be- 
came scarce,  as  in  winter  and  early 
spring,  squirrels  increased  consump- 
tion of  inner  bark  (figs.  1  &  2).  There 
seemed  to  be  a  clear  preference  for 
the  inner  bark  of  certain  trees  to  the 
extent  that  some  individual  trees 
were  nearly  defoliated.  However,  a 
decreasing  amount  of  snowfall  in 
three  years  (table  3)  was  associated 
with  an  increasing  number  of  trees 
from  which  inner  bark  was  taken  and 
a  decreasing  average  number  of  clips 
per  tree.  Thus,  the  identification  of  a 
particular  tree  as  a  favorite  "feed 
tree"  (Pfolliott  and  Patton  1978) 
seemed  to  depend  to  some  extent  on 


other  relevant  environmental  condi- 
tions influencing  access  to  food  sup- 
plies, e.g.,  mobility  over  snow-cov- 
ered ground. 

The  repeated  use  of  individual 
trees  for  inner  bark  was  surprisingly 
high.  We  found  that  23%  of  all 
clipped  trees  had  shoots  removed  in 
each  of  three  years,  while  Ffolliott 
and  Patton  (1978)  reported  only  2% 
over  four  years  of  potential  use.  This 
difference  between  the  two  studies 
may  have  resulted  from  differences 
in  squirrel  population  densities  and/ 
or  from  differences  in  the  availability 
of  alternative  foods.  Nevertheless,  it 
is  important  to  note  that  a  resident 
squirrel  population  may  not  rotate 
feed  trees  to  the  extent  previously 
reported.  In  addition,  more  than  half 
a  stand's  individuals  may  become 
feed  trees.  We  expect  that  continued 
observation  will  increase  that  per- 
centage. 

The  quantity  of  hypogeous  fungi 
remained  a  fairly  consistent  food 
supply,  aside  from  its  unusual  abun- 
dance in  1983  (table  1).  Truffles  ap- 
pear to  be  a  common  component  of 
the  diet  of  tree  squirrels  (Gronwall 
and  Pehrson  1984,  Moller  1983),  if 
not  of  most  small  herbivorous  mam- 
mals (Maser  et  al.  1978).  Judging 
from  the  analysis  of  gut  contents  in 
this  (Vireday  1984)  and  other  squir- 
rels (Gronwall  and  Pehrson  1984, 
Grachev  and  Fedosenko  1974, 
McKeever  1964),  hypogeous  fungi 
constitute  one  of  the  primary  food 
resources.  The  drop  in  winter  squir- 
rel weight  as  inner  bark  replaced 
truffles  in  the  diet  (table  2  and  fig.  1) 
is  also  suggestive  of  the  importance 
of  this  fungal  diet  component.  Ken- 
ward  (1983)  showed  similar  weight 
losses  for  gray  squirrels,  S.  carolinen- 
sis,  feeding  heavily  on  inner  bark. 

Truffle  production  has  been  re- 
ported to  be  correlated  with  high 
canopy  cover  (States  1985),  which  is 
more  characteristic  of  blackjack 
stands  than  of  stands  with  a  high 
proportion  of  yellow  pines.  This  rela- 
tionship between  truffles  and  canopy 
cover  may  explain  the  preponder- 


429 


ance  of  squirrel  foraging  activity  ob- 
served in  the  blackjack  stands. 

Observations  of  food  supply  (table 
3)  and  squirrel  food  use  (fig.  2) 
showed  considerable  variability, 
much  of  it  related  to  precipitation 
patterns.  Consequently  each  year 
presented  a  different  pattern  of  food 
combinations,  which  nnay  take  5  to  10 
years  to  repeat.  Nevertheless,  it  is 
clear  that  squirrels  clipped  twigs  to 
some  extent  every  year,  but  greatly 
increased  clipping  when  cones  were 
scarce.  Moreover,  hypogeous  fungi 
were  a  regularly  used  resource. 

Management  impiications 

The  number  of  twig  clips  found  has 
been  suggested  as  an  index  of  squir- 
rel population  density  (Brown  1982, 
Keith  1965).  However,  the  complex 
pattern  of  clipping  observed  in  these 
three  years  suggests  some  limita- 
tions. We  advocate  restricting  such 
an  index  to  comparisons  of  relative 
population  densities  between  differ- 
ent sites  within  the  same  year,  when 
one  can  reasonably  presume  that 
weather  conditions,  pine  seed  abun- 
dance, and  availability  of  alternative 
foods  to  be  similar  over  a  large  area. 

Maintenance  of  clustered  stands  is 
essential  to  provide  the  canopy  cover 
needed  for  truffle  production  as  well 
as  cover  and  nesting  sites  for  squir- 
rels. Reduction  of  stand  heterogene- 
ity and  removal  of  trees  in  large  dis- 
junct blocks  will  likely  have  a  nega- 
tive impact  on  Abert  squirrel  habitat 
(see  also  Pederson  et  al.  1987).  Over 
time,  squirrels  utilize  a  majority  of 
blackjack  and  yellow  pine  within  the 
stands.  Forest  management  practices 
that  provide  corridors  for  squirrel 
movement  among  stands  will  poten- 
tially reduce  localized  herbivory  and 
avoid  severe  tree  damage. 

Acknowledgments 

We  thank  the  Lowell  Observatory  for 
providing  a  site  for  this  study.  This 


research  was  supported  by  a  faculty 
grant  from  Northern  Arizona  Uni- 
versity. 

Literature  Cited 

Altmann,  J.  1974.  Observational 
study  of  behavior:  sampling  meth- 
ods. Behavior  49:227-265. 

Brown,  David  E.  1982.  The  use  of 
"clippings"  to  index  tassel-eared 
squirrel  population  levels.  Journal 
of  Wildlife  Management  46:520- 
525. 

Farentinos,  Robert  C,  P.  J.  Capretta, 
R.  E.  Kepner,  and  Victoria  M.  Lit- 
tlefield.  1981.  Selective  herbivory 
in  tassel-eared  squirrels:  role  of 
monoterpenes  in  ponderosa  pine 
chosen  as  feeding  trees.  Science 
213:1273-1275. 

Ffolliott,  Peter  F.  and  David  R.  Pat- 
ton.  1978.  Abert  squirrel  use  of 
ponderosa  pine  as  feed  trees. 
USDA  Forest  Service  Research 
Note  RM-362.  4  p.  Rocky  Moun- 
tain Forest  and  Range  Experiment 
Station,  Fort  Collins,  CO. 

Grachev,  Uri  A.,  and  A.  K.  Fe- 
dosenko.  1974.  On  the  ecology  of 
the  Teleutka  squirrel  {Sciurus  vul- 
garis exalhidus  Pall.)  in  the  Dzun- 
garskii  Alatau.  Proceedings  of  the 
Academy  of  Sciences  of  the 
Kazachstan  SSR,  Series  Biology 
(6):32-39. 

Gronwall,  Olavi,  and  A.  Pehrson. 
1984.  Nutrient  content  in  fungi  as 
a  primary  food  of  the  red  squirrel 
Sciurus  vulgaris  L.  Oecologia  (Ber- 
lin) 64:230-231. 

Hall,  Joseph  G.  1981.  A  field  study  of 
the  Kaibab  squirrel  in  Grand  Can- 
yon National  Park.  Wildlife  Mono- 
graphs No.  75.  54  p. 

Keith,  James  0. 1965.  The  Abert 
squirrel  and  its  dependence  on 
ponderosa  pine.  Ecology  46(1  &2): 
150-163. 

Kenward,  Roger  E.  1983.  The  causes 
of  damage  by  red  and  gray  squir- 
rels. Mammal  Review  13:159-166. 

Larson,  M.M.  and  Gilbert  H. 

Schubert.  1970.  Cone  crops  of  pon- 


derosa pine  in  central  Arizona  in- 
cluding the  influence  of  Abert 
squirrels.  USDA  Forest  Service 
Research  Paper  RM-58. 15  p. 
Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Fort  Collins, 
CO. 

Maser,  Chris,  James  M.  Trappe,  and 
Ronald  A.  Nussbaum.  1978.  Fun- 
gal-small mammal  interrelation- 
ships with  emphasis  on  Oregon 
coniferous  forests.  Ecology  59:799- 
809. 

McKeever,  Sturgis.  1964.  Food  habits 
of  the  pine  squirrel  in  northeastern 
California.  Journal  of  Wildlife 
Management  28:402-404. 

Moller,  H.  1983.  Foods  and  foraging 
behavior  of  Red  (Sciurus  vulgaris) 
and  Grey  (Sciurus  carolinensis) 
squirrels.  Mammal  Review  13:81- 
98. 

National  Oceanic  and  Atmospheric 
Administration.  1987.  Local  clima- 
tological  data.  Annual  summaries. 
Flagstaff,  Arizona.  National  Cli- 
matic Data  Center,  Asheville,  N.C. 

Patton,  David  R.  1974.  Estimating 
food  consumption  from  twigs 
clipped  by  the  Abert  squirrel. 
USDA  Forest  Service  Note  RM- 
272. 4  p.  Rocky  Mountain  Forest 
and  Range  Experiment  Station, 
Fort  Collins,  CO. 

Patton,  David  R.  1984.  A  model  to 
evaluate  Abert  squirrel  habitat  in 
uneven-aged  ponderosa  pine. 
Wildlife  Society  BulleHn  12(4):408- 
414. 

Patton,  David  R.,  Richard  L. 

Wadleigh,  and  Howard  G.  Hudak. 
1985.  The  effects  of  timber  har- 
vesting on  the  Kaibab  squirrel. 
Journal  of  Wildlife  Management 
49:14-19. 

Pearson,  G.A.  1950.  Management  of 
ponderosa  pine  in  the  Southwest. 
USDA  Agriculture  Monograph 
No.  6:1-218. 

Pederson,  Jordan  C,  R.C.  Farentinos, 
and  Victoria  M.  Littlefield.  1987. 
Effects  of  logging  on  habitat  qual- 
ity and  feeding  patterns  of  Abert 
squirrels.  Great  Basin  Naturalist 
47:252-258. 


430 


Pederson,  Jordan  C.  and  Bruce  L. 
Welch.  1987.  Comparison  of  Pon- 
derosa  pines  as  feed  and  nonfeed 
trees  for  Abert  squirrels.  Journal 
of  Chemical  Ecology  11:149-157. 

Soderquist,  Todd  R.  1987.  The  impact 
of  tassel-eared  squirrel  defoliation 
on  ecotonal  ponderosa  pine.  Jour- 
nal of  Mammalogy  68:398-401. 

States,  Jack  S.  1983.  New  records  of 
hypogeous  ascomycetes  in  Ari- 
zona. Mycotaxon  16:396-402. 

States,  Jack  S.  1984.  New  records  of 
false  truffles  in  pine  forests  of  Ari- 
zona. Mycotaxon  19:352-367. 


States,  Jack.  1985.  Hypogeous,  my- 
corrhizal  fungi  associated  with 
p)onderosa  pine:  sporocarp 
phenology.  In  Proceedings  of  the 
6th  North  American  Conference 
on  Mycorrhizae.  271  p.  Bend,  Ore- 
gon. R.  Molina  (ed.) 

Stephenson,  Richard  L.  1974.  Sea- 
sonal food  habits  of  Abert' s 
squirrels,  Sciurus  aberti.  Journal  of 
Arizona  Academy  of  Sciences  9, 
1974.  Proceedings  Supplement. 

Vireday,  Carol  C.  1982.  Mycophagy 
in  tassel-eared  squirrels  (Sciurus 
aberti  aberti  and  S.  a.  kaibabensis)  in 
northern  Arizona.  Masters  thesis. 
Northern  Arizona  University. 
Ragstaff,  AZ. 

Wauters,  L.  A.,  and  A.  A.  Dhondt. 
1987.  Activity  budget  and  foraging 
behavior  of  the  red  squirrel 
{Sciurus  vulgaris  L.)  in  a  coniferous 
habitat.  Zeitschrift  Saugetierkunde 
52:341-353. 


431 


Small  Mammal  Response  to 
the  Introduction  of  Cattle  into 
a  Cottonwood  Floodplaln^ 

Fred  B.  Samson,^  Fritz  L.  Knopf,^  and  Lisa  B. 
Hass^ 


Abstract.— Few  differences  between  pastures  in 
small  mammal  communities  were  evident  prior  to 
grazing,  1  month  following  grazing,  and  no  differ- 
ences in  numbers  or  distribution  of  small  mammals 
were  observed  5  months  following  grazing.  Each 
small  mammal  species  exhibited  different  habitat 
use  compared  to  availability  and  few  habitat  vari- 
ables differed  on  grazed  versus  ungrazed  pastures. 
Grazing  at  SCS  recommendations  in  winter  did  not 
appear  to  have  an  initial  effect  on  small  mammal 
populations  or  their  habitats  in  a  Colorado 
floodploin. 


Grazing  by  cattle  in  upland  areas  can 
affect  vegetation  and  wildlife  popula- 
tions (Geier  and  Best  1980,  Moulton 
et  al.  1981,  Madany  and  West  1983), 
but  there  is  little  understanding  of 
how  grazing  influences  wildlife 
populations  and  habitats  in  western 
riparian  areas  (Kaufman  et  al.  1982). 
Riparian  areas  of  the  western  United 
States  provide  habitats  for  greater 
diversities  and  densities  of  wildlife 
than  adjoining  upland  communities 
(Thomas  et  al.  1979,  Knopf  1985),  and 
livestock  grazing  is  one  of  many  uses 
that  impacts  riparian  ecosystems. 

Grazing  of  riparian  zones  gener- 
ally occurs  in  winter  along  the  South 
Platte  River  and  similar  stream  or 
river  systems  in  northeastern  Colo- 
rado. Overgrazing  is  reported,  and  in 
some  cases  all  ground  cover  includ- 
ing shrubs  is  removed  (Beidleman 
1954).  The  purpose  of  this  study  was 
to  determine  if  small  mammal  com- 
munities and  vegetation  structure 
were  similar  in  grazed  and  ungrazed 

'Paper  presented  at  symposium.  Man- 
agement of  Amphibians,  Reptiles,  and 
Small  Mammals  in  Nortti  America.  (Flag- 
staff, AZ.  July  19-21,  1988.) 

'Fred  B.  Samson  is  Regional  Wildlife  Bi- 
ologist. USDA  Forest  Service,  Alaska  Region, 
Juneau,  Alaska  99802-1628. 

^Fritz  L.  Knopf  is  Leader  of  Avian  Studies, 
U.S.  Fisti  and  Wildlife  Service,  National  Ecol- 
ogy Research)  Center,  1300  Blue  Spruce 
Drive.  Fort  Collins,  CO  80524-2098. 

''Lisa  B.  Mass  was  a  graduate  student. 
Department  of  Fisheries  and  Wildlife  Biol- 
ogy, Colorado  State  University.  Fort  Collins. 


riparian  areas  in  northeastern  Colo- 
rado. The  approach  was  to  alter  a 
riparian  area  experimentally  by  in- 
troducing cattle  into  an  area  that  had 
not  been  grazed  for  30  years.  The 
specific  objective  was  to  contrast 
small  mammal  communities  and 
vegetation  structure  before,  during, 
and  after  grazing  and  between 
grazed  and  ungrazed  communities. 

Study  Area  and  Methods 

The  study  was  conducted  on  the 
Colorado  Division  of  Wildlife's 
Tamarack  Ranch  Unit,  South  Platte 
State  Wildlife  Area,  in  Logan  County 
near  Crook,  Colorado,  from  March 
1982  to  August  1983.  The  climate  is 
semi-arid.  Mean  annual  precipitation 
is  47.4  cm  and  average  monthly  tem- 
perature is  22.1  CO.  Shallow  clay- 
gravel  soils  in  highly  stratified  allu- 
vial deposits  supported  an  overstory 
of  mature  plains  cottonwood  (Popu- 
lus  sargentii)  and  understories  of 
shrubs  (Salix  exigua,  S.  interior,  Sym- 
phoricarpos  occidcntalis,  Toxicodendron 
radicans,  Vitis  vulpina,  and  Rhus  radi- 
cans),  forbs  (Phragmitcs  communis, 
Spartina  pectinatus,  Chenopodium  al- 
bum, Conium  maculatum,  Rumexcris- 
pus,  and  Melilotus  alba),  and  grasses 
(Elymus  canadensis  and  Spartina  pecti- 
natus). 

The  riparian  zone  adjoining  the 
South  Platte  River  was  last  grazed  in 
the  early  1950's  (M.  Gardner,  pers. 
comm.).  Ten  16-ha  pastures  were  es- 


tablished within  the  riparian  zone 
and  spaced  at  least  0.4  km  apart  to 
eliminate  interactive  effects  among 
pastures.  Five  pastures  selected  at 
random  were  grazed  from  mid -No- 
vember 1982  to  mid-March  1983  at 
levels  recommended  by  the  U.S.  Soil 
Conservation  Service,  with  35.5,  30.8, 
9.0,  37.2,  and  36.8  AUMs  allocated. 
Pre-treatment  data  were  collected  on 
all  pastures  in  March,  June,  and  Au- 
gust 1982.  Posttreatment  data  were 
collected  on  all  pastures  in  March 
and  August  1983. 

A  lOO-trap  grid  of  Sherman  live 
traps  with  15-m  spacing  between 
rows  and  columns  (135  x  135  m,  2.25 
ha)  was  established  in  each  pasture 
to  sample  small  mammal  communi- 
ties. Three,  five-night  trap  sessions 
were  scheduled  per  year:  prior 
(middle  March),  during  (late  June), 
and  after  (late  August)  the  peak 
small  mammal  breeding  season.  The 
total  number  of  trap  nights  for  the 
study  was  25,000: 15,000  trap  nights 
pretreatment  and  10,000  trap  nights 
post-treatment.  Individuals  were 
marked  with  a  numbered  aluminum 
ear  tag,  and  species,  sex,  age,  breed- 
ing condition,  trap  number,  and 
weight  were  recorded.  Density  esti- 
mates were  made  using  the  com- 
puter program  CAPTURE  (Otis  et  al. 
1978,  White  et  al.  1982).  CAPTURE 
examines  capture-recapture  data, 
gives  population  and  density  esti- 
mates for  five  different  models,  and 
indicates  the  model  most  appropriate 
for  estimation.  Model  M  (H)  was  de- 


432 


termined  to  be  the  most  robust  of  the 
five  estimators. 

For  each  of  the  five  trap  sessions, 
trap  sites  were  categorized  according 
to  trap  success  (no-capture  vs.  cap- 
ture) and  to  the  species  captured  at 
that  site.  In  March  1982,  five  no-cap- 
ture sites  and  five  sites  for  each  spe- 
cies were  selected  at  random  for 
vegetation  sampHng  within  each  pas- 
ture. Beginning  in  June  1982  and 
thereafter,  the  sample  size  per  pas- 
ture was  increased  to  ten  no-capture 
sites  and  ten  capture  sites  for  each 
species. 

Habitat  variables  were  measured 
using  two  line  intercept  transects  5-m 
in  length  at  each  selected  trap  site. 
Variables  included  percentage  cover 
of  sand,  litter,  grass,  forb,  and  shrub 
along  the  5-m  transect.  Transects 
were  centered  on  the  trap  site  and 
oriented  toward  randomly  chosen 
cardinal  compass  directions  (north, 
south,  east,  or  west).  The  linear  inter- 
cept of  each  variable  with  the 


transect  was  measured  with  an  incre- 
mental tape.  Two  additional  meas- 
urements at  each  trap  site  were  dis- 
tance-to-ncarest-undcrstory  (<10  m) 
and  distance-to-ncarest-overstory 
(>10  m).  Vegetation  sampling  oc- 
curred concurrently  or  immediately 
following  each  trap  session. 

Chi-square  tests  were  used  to  test 
for  pretreatment  differences  in  sp>e- 
cies  composition  among  those  pas- 
tures chosen  for  grazing  and  those 
chosen  for  controls.  Chi-square  tests 
were  also  used  to  evaluate  posttreat- 
ment  data.  A  f-tcst  was  performed  to 
examine  differences  in  mean  body 
weight  between  treatment  groups. 

T-tests  were  used  to  compare 
habitat  variables  between  species  and 
between  species-specific  capture  sites 
from  all  other  trap  locations.  In  each 
season,  the  vegetation  variables  asso- 
ciated with  the  capture  sites  of  a  spe- 
cies were  compared  to  the  pooled 
sample  of  vegetation  variables  con- 
sisting of  no-capture  sites  in  addition 


to  sites  for  all  other  species  (Dueser 
and  Shugart  1978).  The  degree  of 
habitat  specificity  was  indicated  by 
the  number  of  variables  for  which 
the  species  sample  differed  from  the 
pooled  sample.  Following  the  sp>e- 
cies-sp)ecific  and  pooled  sample  two- 
group  comparison,  mean  vegetation 
values  associated  with  each  species 
were  compared  on  grazed  and  con- 
trol pastures  using  f-tests.  These  pro- 
cedures determined  whether  habitat 
used  by  a  specific  species  differed 
from  the  average  habitat  available 
and  compared  a  species  habitat  use 
on  control  and  grazed  areas  regard- 
less of  habitat  availability.  Some 
overlap  in  use  of  trap  sites  was  ob- 
served, thus  the  pooled  sample  is  not 
expected  to  be  completely  distinct 
from  the  species  specific  sample 
(Dueser  and  Shugart  1978). 

All  statistical  tests  and  density  es- 
timates were  performed  using  the 
Statistical  Package  for  the  Social  Sci- 
ences (Nie  et  al.  1975). 


Table  1.— Total  numbers  of  small  mammals  captured  in  grazed  vs.  un- 
grazed  pastures,  March  1982  to  August  1983,  South  Platte  River  Wildlife 
Management  Area,  near  Crook  Colorado. 


Species/ 


Pretreatment 


Posttreatment 


Treatment 

March 

June 

August 

March 

August 

Deer  Mouse 

Control 

297 

^372 

M27 

^268 

104 

Grazed 

498 

609 

575 

344 

155 

Western  Harvest  Mouse 

Control 

19 

24 

27 

45 

9 

Grazed 

39 

27 

22 

40 

3 

Prairie  Vole 

Control 

5 

5 

12 

11 

3 

Grazed 

4 

7 

2 

4 

6 

Kangaroo  Rat 

Control 

3 

12 

9 

10 

0 

Grazed 

3 

3 

0 

0 

0 

Others 

Control 

6 

7 

17 

2 

5 

Grazed 

7 

5 

6 

6 

9 

'Significantly  different  than  other  treatment  (P  <.05). 
^Significantly  different  than  other  treatment  (P  <.00l). 

^Includes  house  mouse,  hispid  pocket  mouse,  northern  grasshopper  mouse, 
masked  shrew,  and  spotted  skurtk. 


Results 

Species  Composition 

Nine  species  of  small  mammals  were 
captured  in  1982  and  1983  (table  1). 
The  deer  mouse  (Peromyscus  manicu- 
latus)  was  the  most  abundant  species, 
with  the  western  harvest  mouse  (Rei- 
throdontomys  tnegalotis),  kangaroo  rat 
(Dipodomys  ordii),  prairie  vole  (Micro- 
tus  ochogaster),  house  mouse  (Mus 
musculus),  hispid  pocket  mouse  (Per- 
ognathus  hisidus),  northern  grasshop- 
per mouse  (Onychomys  leucogaster), 
masked  shrew  (Sorex  cinereus),  and 
spotted  skunk  (Spilogale  putorius) 
comprising  less  than  2  %  of  the  9,304 
captures. 

Pretreatment  Sf)ecies  richness  did 
not  differ  among  grazed  versus  un- 
grazed  in  March  1982  (X^  =  2.47,  P  = 
0.650)  but  significant  were  evident  in 
June  (X2  =  15.39,  P  =  0.017)  and  Au- 
gust (X2  =  33.18,  P  =  0.001)  (table  1). 
The  differences  in  June  and  August 
were  caused  by  the  abundance  of 


433 


kangaroo  rats,  prairie  voles,  and 
house  mice  on  control  pastures. 
While  three  species — the  hispid 
pocket  mouse,  masked  shrew,  and 
spotted  skunk — were  found  only  on 
pastures  to  be  grazed.  Number  of 
captures  of  the  two  common  species, 
the  deer  mouse  and  western  harvest 
mouse,  were  not  different  in  June  (X^ 
=  1.71,  P  =  0.187)  or  August  (X^  = 
2.97,  P  =0.091)  between  pastures  to 
be  grazed  and  control  pastures. 

Following  nearly  4  months  of 
grazing,  the  composition  of  small 
mammal  communities  in  control  ver- 
sus grazed  pastures  differed  in 
March  1983  (X^  =  15.9,  P  =  0.001) 
(table  1)  but  not  in  August  (X^  =  6.05, 
P  =  0.109).  The  kangaroo  rat  was  not 
captured  on  treated  pastures  in 
March  or  August  1983  although  pres- 
ent in  two  of  five  pastures  prior  to 
treatment  in  1982.  The  number  of 
harvest  mice  captured  in  grazed  pas- 
tures increased  markedly  from 
March  1982  to  March  1983  (19  vs.  45) 
in  contrast  to  control  pastures  (39  vs. 
40). 

Inundation  of  all  pastures  in  May- 
July  1983  (see  Knopf  and  Sedgwick 
1987)  appeared  to  influence  species 
distributions  and  abundances  in  Au- 
gust. From  March  to  August  cap- 
tures of  deer  mice  on  all  pastures  de- 
clined from  61 1  to  259,  western  har- 
vest mouse  from  85  to  12,  and  kanga- 
roo rats  and  mask  shrews  were  no 
longer  captured. 

Densities  and  Population 
Structures 

Only  the  deer  mouse  was  captured  in 
sufficient  numbers  to  calculate  densi- 
ties accurately.  Deer  mice  densities 
were  consistently  higher  on  grazed 
pastures  before  and  after  treatment 
(table  2).  However,  the  density  of 
deer  mice  decreased  18.7%  from  pre- 
to  posttreatment  on  the  five  control 
pastures  (x=  33.6/ha  vs.  x=27.3/ha) 
versus  42.9%  on  the  five  treated  pas- 
tures (63.2/ha  vs.  36.1 /ha)  for  the 
same  interval. 


Age  ratios  appear  unaffected  by 
grazing  (table  2).  In  contrast,  sex  ra- 
tios in  deer  mice  shifted  significantly 
following  grazing  (X^  =  4.90,  P  = 
0.049)  with  three  of  five  grazed  pas- 
tures having  substantially  more 
males.  Western  harvest  mice  sex  ra- 
tios also  changed  following  grazing, 
with  a  higher  percentage  of  females 
captured,  but  sample  sizes  were  in- 
sufficient for  separate  tests  on  each  of 
the  10  pastures. 

The  percentage  of  female  deer 
mice  in  breeding  condition  was  simi- 
lar on  all  pastures  prior  to  grazing 
except  in  June  1982,  when  a  higher 
percentage  of  females  (X^  =  3.84,  P  = 
0.049)  were  in  breeding  condition  on 
control  pastures.  Following  grazing, 
the  percentage  of  breeding  females 
was  higher  in  March  (X^  =  5.53,  P  = 
0.019)  on  control  pastures  yet  grazed 
pastures  had  a  higher  percentage  of 
breeding  females  (X^  =  5.44,  P  = 
0.020)  in  August  1983.  No  significant 


differences  in  the  percentage  of 
breeding  males  or  females  between 
treatment  groups  was  observed  for 
the  other  species. 

Deer  mice  body  weights  were 
similar  across  pastures  prior  to  graz- 
ing, except  in  June  (t  =  3.18,  P  = 
0.002).  After  treatment,  mean  body 
weights  for  mature  (subadult  plus 
adult)  deer  mice  were  significantly 
less  {t  =  2.66,  P  =  0.008)  on  grazed 
pastures  (18.56  +  0.18g)  than  on  un- 
grazed  pastures  (19.3  +  0.21  g)  when 
data  from  all  replicates  were  com- 
bined. The  divergence  in  mean  deer 
mouse  body  weight  between  control 
and  grazed  pastures  continued  into 
August  1983  (t  =3.02,  P  =  0.003). 

Species  Habitat  Use 

Only  sample  sizes  for  the  deer 
mouse,  western  harvest  mouse,  prai- 
rie vole,  and  kangaroo  rat  were  suffi- 


Toble  2.— Selected  population  characteristics  including  population  density 
(mean  no.  per  ha),  age  ratio  (%  juveniles),  sex  ratio  <%  females),  and 
breeding  condition  (%  breeding  females),  March  1982  to  August  1983, 
South  Platte  River  Wildlife  Area,  near  Crook  Colorado, 


Characteristic/ 

Species/ 

Treatment 


Pretreatment 
Mar  1982  Jun  1982  Aug  1982 


Posttreatment 
Mar  1983  Aug  1983 


Density 
Deer  Mouse 

Control 
■:  Grazed 
Age  Ratios 
Deer  IVIouse 
Control 
Grazed 
Sex  Ratios 
Deer  Mouse 


33.6 
63.2 


1.3 
0.7 


36.3 
55.3 


5.3 
4.4 


27.3 
36.1 


1.9 
4.3^ 


18.7  24.8 
42.7  24.7 


Control 

46.0 

48.8 

48.7 

44.0 

Grazed 

53.0 

51.3 

46.2 

43.8 

Western  Harvest  Mouse 

Control 

60.9 

52,0 

31.6 

0.0 

Grazed 

37.0 

31.8 

42.5 

0.0 

Breeding  Condition 

Deer  Mouse 

Control 

65.8 

67.8 

16.2 

64.6 

Grazed 

49.0' 

69.5 

6.61 

85.1 

'Significanfly  different  than  other  treatment  (?  <.06). 


434 


cient  for  subsequent  analysis.  Habitat 
use  by  deer  mice  differed  from  that 
available  in  34%  (12/35)  of  the  t  tests 
on  control  pastures  and  12%  (4/35) 
of  the  tests  on  grazed  pastures  over 
all  seasons  (table  3).  Deer  mice  were 
most  frequently  associated  with  a 
lower  percentage  of  grass  cover  and 
litter  as  well  as  presence  of  shrubs. 
Although  habitat  near  deer  mouse 
capture  sites  differed  from  that  avail- 
able, habitat  use  was  similar  on  con- 
trol and  grazed  pastures.  Among 
those  habitat  variables  associated 
with  the  deer  mouse,  66.7%  (2/3)  in 
March  1982, 100%  (2/2)  in  June  1982, 
66.7%  (2/3)  in  August  1982,  80%  (4/ 
5)  in  March  1983,  and  0%  (0/5)  in 
August  1983  were  similar  on  control 
and  grazed  pastures. 


Like  deer  mice,  the  harvest  mouse 
used  habitats  differing  from  those 
available  and  preferred  similar  sites 
on  control  and  grazed  pastures  (table 
4).  Thirty-four  percent  (12/35)  of  the 
tests  on  control  pastures  and  37% 
(13/25)  of  the  tests  on  grazed  pas- 
tures were  significantly  different 
whereas  the  majority  (68%,  13/19) 
had  similar  values  on  control  and 
grazed  pastures.  The  occurrence  of 
harvest  mice  was  most  strongly  asso- 
ciated with  a  high  percentage  of  litter 
and  grass  cover  and  a  low  percent- 
age of  sand  around  the  capture  site. 

Prairie  vole  capture  sites  differed 
from  the  average  available  site  for 
only  11%  (4/35)  of  the  habitat  com- 
parisons on  control  pastures  and  17% 
(6/35)  of  the  habitat  comparisons  on 


Table  3.— Comparison  of  mean  vegetation  values  between  deer  mouse 
capture  sites  and  the  pooled  sample  on  grazed  and  ungrazed  pastures, 
March  1982  to  August  ! 983,  South  Platte  River  Wildlife  Management  Area, 
near  Crook  Colorado. 


Variable/ 

Pretreotment 

Posttreatment 

Treatment 

Mar  1982  Jun  1982  Aug  1982 

Mar  1983 

Aug  1983 

Sand  (%) 

Control 

16.9 

4.4^ 

8.1 

3.3 

20.1 

Grazed 

7.5 

6.1 

A.V 

2,1 

10.72 

Litter  (%) 

Control 

74.41 

71.4 

86.2^ 

89.8 

24.9 

Grazed 
Grass  (%) 

83.5 

79.4 

88.7 

87.3^ 

21.8 

Control 

20.6^ 

32.9 

52.9 

38,  V 

23.51 

Grazed 

37.32 

46.62 

68.1 

53.52 

43.22 

Forb  (7o) 

Control 

16.4 

48.7 

55.2 

30,9 

I8.51 

Grazed 

19.4 

38.22 

41.52 

23.42 

25.72 

Shrub  (%) 

Control 

6.9 

10.3 

17,2 

20.91 

31.41 

Grazed 

12.0 

15,6 

25,4 

23.8 

I6.32 

Disto3 

Control 

12.31 

n.8 

10.6 

13.51 

10.0 

Grazed 

9.4 

10.3 

12.2 

12.0 

21.11^ 

Distu'^ 

Control 

5,5 

3.V 

3.5 

2.8 

1.21 

Grazed 

7.1 

3.0 

1.91^ 

2.04.42 

'Significanf  (P  <  0.05)  difference  between  deer  mouse  capture  sites  and  pooled 
sample. 

^Significant  (P  <  0.05)  difference  between  grazed  and  control  pastures. 
^Distance  to  nearest  overstory  (>  10m). 
''Distance  to  nearest  understory  (<10m). 


grazed  pastures  (table  5).  Prairie  vole 
habitat  was  similar  to  habitat  used  by 
western  harvest  mouse,  as  both  ex- 
hibited a  preference  for  sites  with  a 
high  percentage  of  litter.  For  vegeta- 
tion variables  which  were  signifi- 
cantly different  on  prairie  vole  cap- 
ture sites  compared  to  the  pooled 
sample  of  sites,  88%  (7/9)  had  simi- 
lar values  on  control  and  grazed  pas- 
tures. 

Kangaroo  rats  exhibited  the  high- 
est habitat  specificity  among  the  four 
major  mammal  species  (table  6). 
Habitat  variables  from  kangaroo  rat 
capture  sites  differed  from  the 
p)ooled  sample  of  sites  for  64%  (18/ 
28)  of  the  habitat  comparisons  on 
control  pastures  and  50%  (7/14)  of 
the  comparisons  on  pastures  to  be 
grazed.  The  factors  which  appeared 
most  critical  in  determining  the  dis- 
tribution of  kangaroo  rats  was  the 
high  percentage  of  sand,  moderately 
high  percentage  of  forbs,  and  low 
percentages  of  litter  and  grass. 


Discussion  and  Conclusions 

Kaufman  et  al.  (1982)  in  Oregon 
noted  that  small  mammal  densities 
decreased  just  following  grazing  only 
to  increase  to  pre-grazing  levels 
within  a  year.  Riparian  grazing  in 
Oregon,  as  in  most  western  range- 
lands,  is  often  in  late  spring  to  early 
fall.  A  similar  pattern,  however,  is 
evident  following  winter  grazing  in  a 
riparian  area  in  northeastern  Colo- 
rado with  few  detectable  differences 
observed  in  small  mammal  commu- 
nity 5  months  following  grazing. 

The  elimination  of  kangaroo  rats 
from  grazed  areas  appears  to  be  a 
consequence  of  grazing  although 
they  were  never  really  abundant  on 
pastures  to  be  grazed  (table  1).  In 
sandhill  rangeland  of  eastern  Colo- 
rado, Green  (1969)  found  the  density 
of  kangaroo  rats  approximately  the 
same  on  ungrazed  and  grazed  pas- 
tures. Kangaroo  rats  may  not  have 
colonized  riparian  grazed  pastures 
because  of  a  change  in  microhabitat 


435 


prior  to,  or  unrelated  to,  cattle  intro- 
duction. Regardless,  the  riparian 
zone  appeared  to  be  a  marginal  habi- 
tat for  this  upland  species. 

Differences  in  age  ratios  appear 
unrelated  to  grazing.  Abramsky 
(1976)  found  that  juvenile  deer  mice 
do  not  readily  enter  traps  and,  thus, 
may  be  under  represented  in  age- 
class  ratios.  The  Trivers-Willard  hy- 
pothesis suggests  that  a  population 
under  stress  will  produce  an  in- 
creased proportion  of  females  (Myers 
1978).  The  imbalance  in  deer  mouse 
sex  ratios  observed  in  this  study  on 
grazed,  but  not  control  pastures, 
does  not  appear  to  be  related  to 
change  in  primary  sex  ratio  or  sur- 
vival of  young  as  suggested  by  the 
above  hypotheses.  Rather,  most  ani- 
mals captured  in  March  1983  trap 
session  were  adults,  70%  of  which 
were  tagged  in  1982.  The  mean  body 
weight  of  deer  mice  on  grazed  pas- 
tures following  treatment  was  lower 
than  on  control  pastures.  A  more 
parsimonius  hypothesis  for  the  ob- 
served shift  in  sex  ratio  is  emigration 
of  females.  Bowers  and  Smith  (1979) 
found  that  female  deer  mice  inhabit 
more  mesic  microhabitats  than 
males.  Grazing  by  cattle  may  have 
altered  microhabitats  preferred  by 
females  and  or  other  resources,  par- 
ticularly seeds,  may  have  been  more 
abundant  on  control  areas.  There  is 
substantial  evidence  in  other  studies 
that  deer  mouse  populations  are  lim- 
ited by  seasonal  food  availability 
(Gashwiller  1979),  specifically  in  win- 
ter (Taitt  1981). 

Small  mammal  habitat  use  and 
seasonal  habitat  shifts  were  similar 
on  grazed  and  control  pastures.  Each 
species  illustrated  differential  habitat 
use  compared  to  availability,  and 
patterns  in  habitat  use  were  little  af- 
fected by  grazing.  Deer  mice  habitat, 
largely  areas  with  little  grass  cover, 
was  consistently  distinguishable 
from  that  of  other  species  as  reported 
elsewhere  (Bowers  and  Smith  1979, 
Kantak  1983,  Lovcll  1983).  Habitat 
use  and  number  of  captures  of  the 
western  harvest  mouse,  prairie  vole. 


and  kangaroo  rat  reported  in  this 
study  are  also  consistent  with  that 
previously  documented.  The  western 
harvest  mouse  is  reported  to  be 
closely  associated  with  grassy  sites 
(Hill  and  Hubbard  1943,  Lovell  1983) 
and  use  of  sandy  sites  by  kangaroo 
rats  was  noted  by  Green  (1969).  The 
importance  of  vegetative  cover  to  the 
prairie  vole  has  been  well  docu- 
mented (Birney  et  al.  1976,  Green 
1969). 

In  summary,  research  reported  in 
this  paper  was  conducted  in  an  ex- 
perimental framework,  with  five  rep- 
lications, to  evaluate  the  initial  effects 
of  cattle  grazing  in  winter  on  small 
mammal  community  in  a  riparian 
area.  Winter  grazing  of  riparian  areas 
based  on  Soil  Conservation  Service 


recommended  levels  appears  to  have 
little  initial  effect  on  small  mammal 
populations  and  their  habitats.  The 
study  further  indicates  that  pretreat- 
ment  assessment  of  habitat  and  small 
mammal  populations  in  studies  to 
evaluate  effects  of  grazing  in  riparian 
areas  is  important.  Significant  differ- 
ences in  small  mammal  numbers  and 
species-specific  habitat  use  observed 
following  grazing  could  have  been 
attributed  to  treatment  without 
knowledge  of  pretreatment  popula- 
tion and  habitat  conditions. 


Acknowledgments 

The  study  was  partially  funded  by 
the  U.S.  Fish  and  Wildlife  Service 


Table  4.— Comparison  of  mean  vegetation  values  (%)  between  western 
harvest  mouse  capture  sites  and  the  pooled  sample  on  grazed  and  un- 
grazed  pastures,  March  1982  to  August  1983,  South  Platte  River  Wildlife 
f^anagement  Area,  near  Crook  Colorado. 


Pretreatment 


Posttreatment 


Treatment 

Mar  1982  Jun  1982  Aug  1982 

Mar  1983 

Aug  1983 

Sand  (%) 
Control 
Grazed 
Litter  (%) 
Control 

0.0^ 
5.9 

0,V 
1.81 

5,5 
0.0 

0.31 
0.2 

12.9 
0.02 

95.41 

73.8 

89.91 

93.41 

42.6 

Grazed 

93.1 

58.4' 

89.2 

95.01 

16.0 

Grass  (%) 

Control 

61. r 

40.3 

64.4 

78.2 

43.2 

Grazed 

78.61-2 

53.6 

78.0 

79.71 

55.0 

Forb  (%) 

Control 

19.9 

49.6 

53.0 

20.21 

28.2 

Grazed 

14.8 

35.52 

38.6 

12.21-2 

38.7 

Shrub  (%) 

Control 

11.6 

20.51 

22.8 

8.81 

22.2 

Grazed 

4.81 

14.3 

30.1 

25.2 

26,7 

Disto^ 

Control 

9.1 

12.2 

8,9 

11.7 

6.9^ 

Grazed 

8.6 

11.4 

14.11 

12.91 

65.0 

Distu^ 

Control 

7.2 

5.1 

2.7 

3.61 

1.4 

Grazed 

7.6 

4.71 

1.31 

1.52 

4.8 

'Significant  (P  <  0.06)  difference  between  western  harvest  mouse  capture  sites  and 
pooled  sample. 

'Significant  (P  <  0.05)  difference  between  grazed  and  control  pastures. 

^Distance  to  nearest  overstory  (>  Wm). 
"Distance  to  nearest  understory  (<10m). 


436 


through  the  Colorado  Cooperative 
Wildlife  Research  Unit,  and  is  a 
product  of  Cooperative  Agreement 
No.  2463-4  between  the  Colorado 
Division  of  Wildlife  and  the  U.S.  Fish 
and  Wildlife  Service's  National  Ecol- 
ogy Research  Center. 


Literature  Cited 

Abramsky,  Z.  Z.  1976.  Small  mam- 
mal studies  in  natural  and  ma- 
nipulated shortgrass  prairie.  Ph. 
D.  dissertation,  Colorado  State 
University,  Fort  Collins. 

Beidleman,  Richard  G.  1954.  The  Cot- 
tonwood river-bottom  community 
as  a  vertebrate  habitat.  Ph.D.  dis- 
sertation. University  of  Colorado, 
Boulder,  Colorado. 


Birney,  Elmer  D.,  William  E.  Grant, 
and  Duane  D.  Baird.  1976.  Impor- 
tance of  vegetative  cover  to  cycles 
of  Microtus  populations.  Ecology 
57: 1043-1051. 

Bowers,  Michael  A.,  and  Howard  D. 
Smith.  1979.  Differential  habitat 
utilization  by  sexes  of  the 
deermouse,  Peromyscus  manicula- 
tus.  Ecology  60:  869-875. 

Dueser,  Raymond  D.  and  Herman  H. 
Shugart,  Jr.  1978.  Microhabitats  in 
a  forest  floor  small  mammal 
fauna.  Ecology  59:  89-98. 

Gash  wilier.  Jay  S.  1979.  Deer  mouse 
reproduction  and  its  relationship 
to  the  tree  seed  crop.  American 
Midland  Naturalist  102:  95-102. 

Geier,  Arnold  R.,  and  Louis  B.  Best. 
1980.  Habitat  selection  by  small 
mammals  of  riparian  comnmni- 


Table  6,— Comparison  of  mean  vegetation  values  between  prairie  vole 
capture  sites  and  the  pooled  sample  on  grazed  and  ungrazed  pastures, 
March  1982  to  August  1983,  South  Platte  River  Wildlife  Management  Area, 
near  Crook  Colorado. 


Pretreatment 


Posttreatment 


Treatment 

Mar  1982  Jun  1982 

Aug  1982 

Mar  1983 

Aug  1983 

Sand  (%) 

Control 

0.0 

20.0 

13.9 

0.0 

0.0 

Grazed 

0,0 

0.0 

0.0' 

0,0 

0.0 

Litter  (%) 

Control 

88.0 

80.0 

85.4 

71.0 

100,0 

Grazed 

99.7^^ 

85.7 

84.5 

99.5' 

33.5 

Grass  (%) 

Control 

52.2 

52.6 

36.6 

80.0 

100.0 

Grazed 

78.7^ 

43.4 

56.0 

65.0 

63.7 

Forb  (%) 

Control 

25.0 

42.8 

49.0 

2.0 

100.0 

Grazed 

22.8 

29,92 

21.0 

38,0 

40.72 

Shrub  (%) 

Control 

0.0 

4.8 

48.6' 

22,0 

0.0 

Grazed 

5.3 

17,3 

22.5 

20,6 

23.0 

Disto^ 

Control 

5.8^ 

14.1 

10.9 

2.4 

15.0 

Grazed 

15.6^2 

10.6 

4.0'^ 

12.4 

95.7' 

Distu^ 

Control 

5.8 

2.4 

0.2' 

3,2 

7.5 

Grazed 

14.0 

1.8 

5.62 

2,7 

12.8 

'Significont  (P 

<  0.05)  difference  between  the 

prairie  vole 

and  pooled  sample. 

^Significant  (P  <  0.05)  difference  between  grazed  and  control  pastures. 
^Distance  to  nearest  overstory  (>  10m). 
^Distance  to  nearest  understory  (<10m). 


ties:  Evaluating  effects  of  habitat 
alterations.  Journal  of  Wildlife 
Management  44: 16-24. 

Green,  N.  E.  1969.  Occurrence  of 
small  mammals  on  sandhill  range- 
lands  in  eastern  Colorado.  M.S. 
thesis,  Colorado  State  University, 
Fort  Collins. 

Hill,  John  E.  and  Claude  W.  Hub- 
bard. 1943.  Ecological  differentia- 
tion between  two  harvest  mice 
(Reithrodontomx/s)  in  western  Kan- 
sas. Journal  of  Mammalogy  24:  22- 
25. 

Kantak,  Gail  E.  1983.  Behavioral, 
seed  preference  and  habitat  selec- 
tion experiments  with  two 
sympatric  Peromyscus  species. 
American  Midland  Naturalist  109: 
246-252. 

Kaufman,  J.  Boone,  William  C. 
Krueger,  and  M.  Vavra.  1982.  Im- 
pacts of  a  late  seasonal  grazing 
scheme  on  nongame  wildlife  habi- 
tat in  a  Wallowa  Mountain  ripar- 
ian ecosystem,  p.  208-220.  In  Wild- 
life-livestock Relationships  Sym- 
posium. Coeur  d'Alene,  Id.,  [April 
21-22,  1981]  University  of  Idaho 
Forest,  Wildlife,  and  Range  Ex- 
p)eriment  Station,  Moscow,  Idaho. 

Knopf,  Fritz  L.  1985.  Significance  of 
riparian  vegetation  to  breeding 
birds  across  an  altitudinal  cline.  p. 
105-111.  In  Riparian  ecosystems 
and  their  management:  Reconcil- 
ing conflicting  uses.  [Tucson, 
Ariz.,  April  16-18, 1985]  USDA 
Forest  Service  General  Technical 
Report  RM-120,  Rocky  Mountain 
Forest  and  Range  Experiment  Sta- 
tion, Fort  Collins,  Colo. 

Knopf,  Fritz  L.  and  James.  A. 

Sedgwick.  1987.  Latent  Population 
responses  to  a  catastrophic,  clima- 
tological  event.  Condor  89:  869- 
873. 

Lovell,  David  C.  1983.  Succession  of 
mammals,  and  associations  of 
small  mammals,  in  disturbed  habi- 
tats at  Barr  Lake  State  Park, 
Adams  County,  Colorado.  M.S. 
thesis.  For  Hays  State  University, 
Fort  Hays,  Kan. 


437 


Madany,  Michael  H.,  and  Neil  E. 
West.  1983.  Livestock  grazing-fire 
regime  interactions  within  mon- 
tane forests  of  Zion  National  Park, 
Utah.  Ecology  64:  661-667. 

Moulton,  Michael  P.,  Jerry  R.  Choate, 
Stephen  J.  Bissell,  and  Robert  A. 
Nicholson.  1981.  Associations  of 
small  mammals  on  the  central 
high  plains  of  eastern  Colorado. 
Southwestern  Naturalist  26:  53-57. 

Myers,  Judith  H.  1978.  Sex  ration  ad- 
justment under  food  stress:  Maxi- 
mization of  quality  of  numbers  of 
offspring?  American  Naturalist 
112:  381-388. 

Nie,  Norman  H.,  C.  Hadlei  Hull,  Jean 
G.  Jenkins,  Karin  Steinbrenner, 
and  Dale  H.  Bent.  1975.  Statistical 
Package  for  the  Social  Sciences, 
McGraw-Hill  Book  Company, 
New  York,  NY. 

Otis,  David  L.,  Kenneth  P.  Burnham, 
Gary  C.  White,  and  David  R.  An- 
derson. 1978.  Statistical  inference 
from  capture  data  on  closed  ani- 
mal populations.  Wildlife  Mono- 
graph 62: 1-135. 

Taitt,  M.  J.  1981.  The  effect  of  extra 
food  in  small  mammal  popula- 
tions. I.  Deermice  (Peromyscus 
maniculatus).  Journal  of  animal 
ecology  50: 111-124. 

Thomas,  Jack  W.,  Chris  Maser,  and 
Jon  E.  Rodick.  1979.  Riparian 
zones.  In  Wildlife  habitats  in  man- 
aged forests,  the  Blue  Mountains 
of  Oregon  and  Washington,  USDA 
Handbook  553:  40-47. 


Table  6.— Comparison  of  mean  vegetation  values  between  kangaroo  rat 
capture  sites  and  the  pooled  sample  on  grazed  arxi  ungrazed  pastures, 
March  1982  to  August  1983,  South  Platte  River  Wildl'ife  Management  Area, 
near  Crook  Colorado. 


Treatment 


Pretreatment 
Mar  1982  Jun  1982  Aug  1982 


PosttrGatment 
Mar  1983  Aug  1983 


Sand  (%) 

Control 

Grazed 

Litter  (%) 

Control 

Grazed 

Grass  (%) 

Control 

Grazed 

Forb  (%) 

Control 

Grazed 

Shrub  (%) 

Control 

Grazed 

Disto2 

Control 

Grazed 

Dlstu^ 

Control 

Grazed 


53.01 
53.0^ 

69.0 
38.01 

12.0 
33.7 

18.0 
25.3 

21.5 
10.7 

3.61 
7,5 

9.6 
3,21 


67.41 
51.01 

44.4' 
70.0 

10.71 
11.01 

71.11 
60.5 

i.r 

26.0 

11.2 
6.01 

7.51 
0.41 


38.61 


66.5^ 


29.41 


65.6 


lo.r 


6.71 


5.1 


34.91 
4971 
15.11 
44.81 
4.21 
9.8 
4.3 


'Significant  (P  <  0.05)  difference  between  l<ongaroo  rat  capture  sites  and  pooled 
sample. 

'Distance  to  nearest  overstory  (>  10m). 
^Distance  to  nearest  understory  (<  10m), 


White,  Gary  C,  David  R.  Anderson, 
Kenneth  P.  Burnham,  and  David 
L.  Otis.  1982.  Capture-recapture 
and  removal  methods  for  sam- 
pling closed  populations.  Los 
Alamos  National  Laboratories, 
Los  Alamos,  NM. 


Zimmerman,  Earl  G.  1965.  A  com- 
parison of  habitat  and  food  of  two 
species  of  Microtus.  Journal  Mam- 
malogy 46:  605-612. 


438 


Old  Growth  Forests  and  thie 
Distribution  of  the  Terrestrial 
Herpetofauna^ 

Hartwell  H.  Welsh,  Jr.^  and  Amy  J.  Lind^ 


Abstract.— Terrestrial  herpetofauna  were  sampled 
by  pitfall  traps  and  time-constrained  searches  on  42 
stands  of  Douglas-fir/hardwood  forest  in  southwest- 
ern Oregon  and  northwestern  California.  Stands 
ranged  in  age  from  40  to  450  years.  We  found  25 
species  of  herpetofauna.  Species  diversity  was 
greater  in  older  forest  stands  than  in  young  stands. 
Amphibians  were  significantly  more  abundant  in  old 
than  in  young  stands  and  significantly  less  abundant 
in  dry  than  in  moist  stands.  Our  research  indicates 
that  changes  in  forest  structure  due  to  forest  prac- 
tices results  in  reduced  species  diversity  and  abun- 
dance among  the  herpetofauna. 


The  coniferous  forests  of  the  Pacific 
Northwest  are  currently  the  focus  of 
a  national  conflict  between  compet- 
ing interests.  These  ancient  forests, 
previously  more  species  rich  and 
continuous  across  the  continental 
United  States,  have  undergone  a 
natural  decline  since  the  Mesozoic  in 
conjunction  with  broad  climatic  and 
geologic  changes  (Axelrod  1976). 
This  process  eliminated  most  of  the 
wooded  areas  of  the  Midwest,  but 
left  expansive  tracts  of  forest  in  the 
eastern  and  western  United  States.  In 
the  last  hundred  years,  many  of  these 
remaining  ancient  forests  have  been 
harvested  for  wood  products,  chang- 
ing the  species  composition,  struc- 
ture, and  forest  age  (Harris  1984). 
These  natural  forest  ecosystems  have 
been  altered  so  rapidly  that  we  are 
only  now  recognizing  the  loss  of 
some  plant  and  animal  species  and 
the  threat  to  others  [e.g.,  the  spotted 
owl  (Strix  occidentalis)]  (Simberloff 
1987).  Recent  concern  for  the  health 
and  well-being  of  these  forest  ecosys- 
tems, and  the  need  for  more  knowl- 

' Roper  presented  at  tlie  Symposium  on 
Management  of  Amphibians.  Reptiles,  and 
Small  Mammals  in  Nortii  America  (July  19- 
21.  1988.  Flagstaff  Arizona). 

^Wildlife  Biologist,  Pacific  Soutt)west  For- 
est and  Range  and  Experiment  Station. 
Forest  Service.  U.S.  Department  of  Agricul- 
ture. Areata.  California  95521. 

^Biological  Technician,  Pacific  Southwest 
Forest  and  Range  and  Experiment  Station, 
Forest  Service,  U.S.  Department  of  Agricul- 
ture, Areata.  California  95521. 


edge  to  meet  management  goals  and 
the  requirements  of  the  National  For- 
est Management  Act  1976  and  the 
Endangered  Species  Act  1973  has 
prompted  research  into  the  structure 
and  composition  of  the  vertebrate 
communities  of  these  forests 
(Meslow  et  al.  1981,  Raphael  1984, 
Ruggiero  and  Carey  1984). 

From  1981  through  1983,  Raphael 
(1984, 1987,  this  volume)  used  a  vari- 
ety of  sampling  methods  to  collect 
data  on  the  forest  age,  moisture,  and 
habitat  associations  of  birds,  mam- 
mals, reptiles,  and  amphibians  in  for- 
ests of  northwestern  California.  From 
1984  through  1986,  researchers  from 
the  Forest  Service's  Pacific  Southwest 
Forest  and  Range  Experiment  Station 
extended  these  studies  to  include 
southwestern  Oregon.  By  measuring 
differences  in  the  species  composi- 
tion and  relative  abundance  of  the 
herpetofaunal  community  in  altered 
versus  unaltered  habitats  it  is  pos- 
sible to  indicate  biologically  mean- 
ingful differences  in  habitat  quality 
(e.g..  Bury  et  al.  1977,  Busack  and 
Bury  1974,  Jones  1981,  Luckenbach 
and  Bury  1983,  Ortega  etal.  1982). 
Such  information  on  differences  in 
the  composition  of  the  herpetofauna, 
relative  to  forest  age  and  moisture, 
have  scientific  value  as  well  as  practi- 
cal value,  as  indicators  of  habitat 
change,  useful  to  natural  resource 
managers. 

This  paper  reports  on  a  study  to 
determine  the  occurrence  and  abun- 
dance of  the  forest  herpetofauna  rela- 


Oregon 


California 


#  Coastal  Stand 
▲  Inland  Stand 


Figure  1.— Study  stands  In  Douglas-fir  forests 
were  located  in  norttiwestern  California 
and  southwestern  Oregon.  Triangles  = 
stands  in  the  inland  area,  circles  =  stands  In 
the  coostal  area. 

tive  to  forest  age  and  moisture,  and 
to  compare  two  methods  (time-con- 
strained searches  and  pit-fall  trap- 
ping) used  to  sample  this  herpe- 
tofauna in  northwestern  California 
and  southwestern  Oregon. 


439 


STUDY  AREA 

This  study  was  conducted  in 
Douglas-fir  (Pseudotsuga  menziesii)/ 
hardwood  forests  at  low  to  mid-ele- 
vations in  the  Klamath  Mountains 
and  Coast  Range.  We  sampled  54 
stands,  but  we  use  data  from  only  42 
stands,  omitting  nine  higher  eleva- 
tion, white-fir  dominated,  stands  and 
three  stands  on  serpentine  soils  be- 
cause they  differed  so  greatly  from 
our  remaining  stands.  Even-aged 
stands  in  the  above  forest  type  were 
selected  in  three  areas  within  the  Kla- 
math Mountains  and  Coast  Range 
(fig.  1)  in  accordance  with  proce- 
dures outlined  by  Spies  et  al,  (in 
press).  Using  stand  characteristics 
(Franklin  et  al.  1986)  and  tree  age,  we 
assigned  stands  to  one  of  three  age 
classes:  young,  mature,  and  old- 
growth  forests.  Stands  ranged  in  age 
from  40  to  450  years.  Stands  in  old- 
growth  were  further  categorized  into 
three  moisture  classes:  dry,  mesic, 
and  wet  (fig.  2).  Stands  ranged  in  size 
from  21  to  150  hectares,  and  in  eleva- 
tion from  53  m  to  1205  m.  One-half  of 
the  stands  occurred  within  the  Coast 
Range,  an  area  formed  primarily  of 
Franciscan  parent  materials  and 
dominated  by  the  maritime  climatic 
influences  of  the  Pacific  Ocean.  These 
stands  were  classified  as  coastal  for- 
est stands  (fig.  1).  All  stands  were 
dominated  by  Douglas-fir  and  con- 


Old  Crowlh 

Dry; 

Old  Growth 
Mcsic: 

Old  Growth 
Wet: 

1  coastal 
3  inland 

4  coastal 
6  inland 

3  coastal 
3  inland 

MhIuth: 

5  coastal 

6  inland 

Young; 

8  coastal 
3  inland 

Age 

Class 


Moisture  Class 


tained  a  significant  hardwood  ele- 
ment, primarily  tanoak  (Lithocarpus 
densiflora)  and  madrone  (Arbutus 
menziesii);  about  half  also  contained 
coast  redwood  (Sequoia  semperznrens). 

The  other  sites  were  designated 
inland  stands  (fig.  1),  occurring 
within  the  Klamath  Mountains,  pri- 
marily on  granitic  and  metamorphic 
parent  materials.  This  area  is  subject 
to  colder  winters  and  drier,  hotter 
summers  than  the  Coast  Range.  The 


inlands  stands  were  dominated  by 
Douglas-fir  in  association  with 
tanoak,  madrone,  and  to  a  lesser  ex- 
tent, canyon  live  oak  (Quercus 
chrysolepis),  black  oak  (Quercus  kellog- 
gii),  ponderosa  pine  (Pinus  ponder- 
osa),  sugar  pine  (Pinus  lamhertiana), 
and  incense  cedar  (Calocedrus  decur- 
rens).  For  a  more  complete  descrip- 
tion of  the  vegetations  of  these  two 
provinces  see  Raphael  (in  press)  and 
Sawyer  and  Thomburgh  (1977). 


r~  ^ — 

Table  1  .—Structural  features'  of  Douglas-fir  stands  on  which  herpetofauna 
were  sampled  in  northwestern  California  and  southwestern  Oregon.' 

Forest  age  class 


Figure  2.— Distribution  of  study  stands  by 
forest  age  and  moisture  class,  and  by 
coastal  and  Inland  area. 


Young  (9)^ 

Mature  (11) 

Old  (19) 

<  100  yrs 

100-200  yrs 

>  200  yrs 

Structural  feature 

(mean') 

(mean) 

(mean) 

Live  trees 

Age  of  dominant  size  class 

44.3 

129.0 

264.6 

of  Douglas-fir 

+  18.3 

+38.9 

+74.4 

Diameter  at  breast  ineight 

38.4 

85.1 

111.5 

^r)Rl-l^  of  Hominont  <?i7fs 

+  14  ft 

+99  9 

+23  8 

class  of  Douglas-fir  (cm) 

Ig.  conifers— trees/ha 

1.5 

17.1 

34.7 

(>  80  cm  DBH) 

+2.1 

+12.0 

±14,1 

Ig.  hardwoods— trees/ha 

9.1 

10.7 

13.0 

(>  50  cm  DBH) 

±10.8 

+  12.0 

+  10.1 

sm.  trees— trees/ha 

1430.0 

764.1 

630.0 

(conf.-5-80  cm  DBH 

±594.0 

±331.0 

±240.0 

+  hdwds-5-50  cm  DBH) 

Snags 

(conifers  and  hardwoods) 

large— snags/ha 

3.4 

2.2 

6.1 

(>  50  cm  DBH  and 

±6.3 

±2.2 

±5.9 

>  4.5  m  in  height) 

Logs 

Ig.  conifers— logs/ha 

1.3 

0.7 

4.0 

(>  50  cm  DBH  and 

±4.0 

±1.8 

±4.0 

>  15  m  long) 

Ig.  conifers— mt/ha^ 

1.2 

1.0 

6.9 

+3.5 

+2.5 

+6.9 

sm.  conifers— logs/ha 

334.4 

151.4 

192.2 

±179.5 

+  129.9 

+  107.8 

hardwoods— logs/ha 

95.6 

146.4 

113.7 

±64.8 

+123.1 

+59.0 

'Bruce  Bingham,  unpublished  data  on  file  with  Pacific  Southwest  Forest  and  Range 
Experiment  Station,  1700  Bayview  Drive.  Areata,  CA  95621. 

^Sampling  occurred  from  1984-1986. 

'Number  of  stands. 

^f\/lean  ±  1  standard  deviation. 

^  mt  =  metric  tons. 


440 


METHODS 

Herpetofauna  Sampling 

A  herpetofaunal  sampling  design 
was  developed  for  the  USDA  Forest 
Service's  old-growth  wildlife  habitats 
project  in  Oregon  and  Washington 
by  Corn  and  Bury  (in  prep.).  Their 
design  used  two  methods  to  sample 
species  composition  and  relative 
abundance  of  the  herpetofauna:  pit- 
fall traps  (PF)  and  time-constrained 
searches  (TCS)  (Bury  and  Corn,  this 
volume;  Welsh  1987).  The  TCS 
method  employed  in  our  study  dif- 
fered from  that  described  by  Corn 
and  Bury  (in  prep.)  in  that  headwater 
habitats  (springs,  seeps,  and  first  or- 
der streams)  were  included  in  the 
sampling.  Pitfall  trap  grids  consisted 
of  36  cans  buried  at  ground  level  and 
spaced  15  m  apart.  Traps  were  cov- 
ered with  bark  or  cedar  shakes.  We 
sampled  40  stands  in  the  fall  of  1984 
and  1985,  for  50  and  30  nights,  re- 
spectively. Our  total  pitfall  trapping 
effort  amounted  to  115,200  trap- 
nights.  Time-constrained  searches 
consisted  of  intensively  searching  all 
terrestrial  microhabitats  in  the  forest 
environment  for  a  fixed  amount  of 
time.  Only  actual  search  time  was 
counted,  when  an  animal  was  en- 
countered the  timer  was  stopped 
while  data  were  collected.  A  4-per- 
son-hour  TCS  was  conducted  on 
each  of  the  42  stands  in  1984  and 

1985.  An  additional  4-person-hour 
TCS  was  conducted  on  30  stands  in 

1986.  Our  total  effort  for  TCS 
amounted  to  456  person-hours. 

Forest  Age 

Forest  age  was  determined  for  each 
stand  by  increment  borer,  or  by 
counting  rings  on  stumps  in  adjacent 
logged  areas.  Dominant  or  co-domi- 
nant size  class  Douglas-fir  trees  were 
selected  for  aging  and  trees  were 
cored  at  breast  height.  Two  to  10 
trees  (average  3)  were  cored  on  each 
stand  and  the  sample  mean  was  used 


to  estimate  forest  age  for  the  stand. 
On  the  basis  of  tree  coring,  ring 
counts,  and  structural  characteristics 
(Franklin  et  al.  1986),  we  grouped 
stands  into  three  age  classes:  young 
forest,  <100  years;  mature  forest,  100- 
200  years;  and  old-growth  forest, 
200+  years  (table  1). 

Moisture  Class 

Stands  that  were  classified  as  old- 
growth  were  also  assigned  a  mois- 
ture classification  (dry,  mesic,  or 
wet),  depending  on  plant  species 
composition  and  percent  cover  of  the 
herb  and  shrub  layers  within  the 
stand.  The  data  were  independently 
recorded  from  three  to  five  0.1  ha 
circular  plots  selected  at  random 
within  each  stand.  Moisture  class  as- 
signment was  based  on  mean  percent 
cover  values  and  the  absolute  con- 
stancy of  particular  shrub  and  herb 
species  within  each  stand. 

Faunal  Comparisons 

We  tested  the  null  hypotheses  (H^) 
that  mean  capture  frequencies  for 
herpetofauna  did  not  differ  between 
either  forest  age  or  moisture  classes 
(1)  within  the  coastal  and  inland  ar- 
eas, (2)  between  the  coastal  and  in- 
land areas,  and  (3)  among  all  stands 
(coastal  and  inland  areas  combined). 
Only  the  mesic  old-growth  stands 
were  used  in  the  age  analysis  (fig.  2). 
One  coastal  old-growth  dry  stand 
prevented  testing  for  differences  in 
means  among  moisture  classes 
within  the  coastal  area,  and  between 
the  coastal  and  inland  dry  stands. 

We  emphasize  that  our  inferences 
are  drawn  from  observations  and  not 
experimental  manipulations.  Though 
our  results  are  described  in  the  con- 
text of  hypothesis  testing,  our  study 
is  primarily  exploratory.  In  addition, 
the  power  of  our  tests  was  low  be- 
cause our  sample  sizes  were  rela- 
tively small.  Our  approach  yields 
preliminary  results  about  forest  age 


and  moisture  relationships  among 
the  herpetofauna,  but  we  caution 
against  making  broad  inferences. 

Combining  Data  Across  Years 

Data  from  pitfall  trapping  were  to- 
taled, by  stand,  for  each  species,  di- 
vided by  50  (1984  data)  or  30  (1985 
data)  nights  x  36  traps  and  multi- 
plied by  1000  to  yield  captures  per 
1000  trap-nights.  Data  from  time- 
constrained  searches  were  adjusted 
for  unequal  sampling  effort  by  ex- 
pressing abundance  of  each  species 
in  captures  per  person-hour. 

We  performed  paired  t-tests  be- 
tween years  (total  captures  per 
stand)  for  each  data  set.  TCS  samples 
were  not  significantly  different  be- 
tween years:  1984  vs.  1985,  t  =  1.16,  P 
=  0.25;  1984  vs.  1986,  t  =  1.24,  P  = 
0.22;  1985  vs.  1986,  t  =  1.85,  P  =  0.075. 
PF  samples  were  also  not  signifi- 
cantly different  between  years:  1984 
vs.  1985,  t  =  1.85,  P  =  0.072.  Conse- 
quently, we  combined  years  for  each 
sampling  method  for  all  analyses. 

Statistical  Comparisons 

For  each  method,  we  tested  for  statis- 
tical differences  in  mean  capture  fre- 
quencies among  age  and  moisture 
classes,  across,  within,  and  between 
inland  and  coastal  areas.  These  tests 
were  performed  on  the  total  herpe- 
tofauna, taxa  at  the  level  of  class,  or- 
der, and  sub-order,  and  on  those  spe- 
cies captured  on  at  least  one  third  of 
our  stands  in  either  area. 

Mean  capture  frequencies  of  each 
faunal  grouping  were  tested  for  sta- 
tistical differences  among  three  forest 
age  classes  and  three  moisture 
classes.  In  cases  where  group  vari- 
ances were  equal  among  classes,  we 
used  one-way  analysis  of  variance 
(ANOVA).  We  used  Hartley's  F  max 
test  (Milliken  and  Johnson  1984:18) 
with  P  <  0.01  to  determine  the  equal- 
ity of  variances  for  all  three-group 
tests.  We  used  P  <  0.01  because 


441 


ANOVA  is  robust  under  moderate 
violations  of  the  assumption  of  equal 
variances  (Zar  1984:170).  If  a  signifi- 
cant F-statistic  resulted  from  the 
ANOVA  test,  we  tested  further  for 
significant  differences  between  pairs 
in  order  to  isolate  the  source  of  the 
differences  by  using  the  Tukey  test 
(TU)  for  multiple  comparisons  (Zar 
1984:186).  Where  group  variances 
were  not  equal  or  where  one  of  the 
three  age  or  moisture  classes  had  no 
captures,  we  performed  all  pairwise 
tests  (multiple  comparisons)  using 
the  Games  and  Howell  modification 
of  the  Tukey  test  (GHMC)  (Keselman 
and  Rogan  1978). 

To  test  for  statistical  differences  in 
capture  frequencies  in  age  and  mois- 
ture classes  between  coastal  and  in- 
land areas,  we  used  two  sample  t- 
tests  (Zar  1984:131).  We  followed  the 
more  conservative  approach  of  not 
pooling  variances.  Between-area 
comparisons  consisted  of  two  fami- 
lies of  tests:  (Da  single  paired  com- 
parison based  on  all  stands,  and  (2) 
five  pairwise  comparisons  defined  by 
the  different  forest  age  and  moisture 
classes.  Tests  in  the  first  family  were 
considered  statistically  significant  at 
the  P  <  0.05  level.  A  Bonferroni  ad- 
justment (Miller  1981:67)  was  used 
for  tests  done  within  the  second  fam- 
ily to  maintain  an  overall  significance 
level  of  P  <  0.05. 

For  the  species  richness  analyses, 
stand  records  from  the  TCS  and  PF 
data  were  combined.  The  means  of 
the  total  number  of  species  for  each 
forest  age  and  moisture  class  were 
tested  for  differences  described. 

Also,  the  similarity  of  species' 
composition  among  equal  numbers 
of  stands  (selected  randomly)  in  each 
forest  age  class  were  determined  by 
using  Jaccard's  similarity  coefficient 
(Sneath  and  Sokal  1973:131): 


a 


S  = 


o  +  b  +  c 


only,  and  c  =  number  of  species  in 
the  second  class  only. 


RESULTS  AND  DISCUSSION 

We  sampled  25  species.  Amphibians 
accounted  for  97.8%  (salamanders, 
96.3%)  of  all  captures,  and  reptiles 
2.2%.  The  TCS  method  yielded  more 
than  66%  of  all  captures  (table  2), 
sampling  22  species  (table  3)  and  ac- 
counting for  67%  of  the  amphibians 
and  85%  of  the  reptiles.  The  PF 
method  sampled  19  species  (table  4) 
and  accounted  for  slightly  less  than 
1/3  of  all  captures  (table  2). 

Species  Composition,  Richness 

Similarity  Indices 

Based  on  species  presence-absence 
data,  an  analysis  of  faunal  similari- 
ties between  forest  age  classes 


(coastal  and  inland  areas  combined) 
indicated  that  greatest  similarity  in 
species  composition  occurred  be- 
tween the  mature  and  old-growth 
stands  (table  5).  Jacard's  Similarity 
Index  (JSI)  values,  for  comparisons 
between  young  and  old-growth 
stands  and  young  and  mature 
stands,  indicated  that  young  stands 
were  different  in  species  composition 
from  both  classes  of  older  forest 
stands.  These  differences  were  great- 
est between  young  and  old-growth 
stands  (table  5). 


Species  Rictiness 

The  number  of  species  per  stand  for 
all  42  stands  ranged  from  3  to  13  (fig. 
3).  The  coastal  mature  stands  yielded 
the  highest  mean  number  of  species 
overall,  while  the  lowest  mean  num- 
ber of  species  occurred  on  the  inland 
mature  stands  (table  6,  fig.  3).  The 


in  which,  for  any  two  classes,  a  = 
number  of  species  in  common,  b  = 
number  of  species  in  the  first  class 


Table  2.— Captures  of  herpetofauna  by  time-constrolned  searches  (TCS) 
and  pitfall  traps  (PF)  in  Douglas-fir  forests  of  northwestern  California  and 
southwestern  Oregon  from  1 984  to  1 986. 

Salamanders  Frogs    Lizards     Snakes      All  species 


Method 

(mean^) 

(mean) 

(mean) 

(mean) 

(mean)\o\a\  captures 

PF  19842 

13.72 

0.38 

0.07 

0.00 

14.18 

1021 

(40  stands) 

+  14.18 

±1.06 

+0.18 

+  14.27 

PF  19853 

11.20 

0.32 

0.21 

0.02 

11.76 

508 

(40  stands) 

±10.23 

±0,83 

±0.61 

±0.15 

±10.19 

PF  Totals 

1529 

(32.6%)« 

TCS^  1984 

6.46 

0.04 

0.11 

0. 

04 

6.66 

1118 

(42  stands) 

±3.63 

±0.14 

+0.27 

±0. 

13 

+3.61 

TCS  1985 

5.80 

0.01 

0.15 

0.06 

6.11 

1027 

(42  stands) 

±3.80 

±0.20 

±0.30 

±0.16 

+3.81 

TCS  1986 

8,10 

0.18 

0.15 

0.08 

8.51 

1021 

(30  stands) 

±4,10 

±0.41 

±0.31 

±0.23 

±4.09 

TCS  totals 

3166 

(67.4%) 

Totals,  both  methods 

4695 

'Mean  for  pitfall  trapping  =  per  1000  trap-nigtits;  X  for  tirDe-constrolned  searches  = 
per  person-hour  of  search  time:  both  ore  ±  1  standard  deviation. 

'PF  1984  =  50  trap-nights  per  stand. 

^PF  1985  =  30  trap-nights  per  stand. 

"All  TCS  =  4  person-hours  per  stand  per  year. 

^Percentage  of  total  captures. 


442 


Table  3.— Mean  number  of  captures  per  person-hour'  captured  by  time-constrained  searches  (TCS)  In  different  age 
and  moisture  classes  of  Douglas-fir  forests  of  northwestern  California  and  southwestern  Oregon  In  the  springs  of  1984, 
1985,  and  1986.^ 


Young 

Mature 

Old -wet 

Old-mesic 

Old-dry 

Total  old 

Total 

Species 

(11)^ 

(11) 

(6) 

(10) 

(4) 

(20) 

Captures 

Frogs 

Tailed  frog 

0.000 

0.008 

0.000 

0.013 

0.000 

0.006 

2 

(Ascaphus  fruei) 

±0.025 

±0.040 

±0.028 

Pacific  treefrog 

0.049 

0.166 

0.028 

0.117 

0.000 

0.067 

44 

(Hyla  regilla) 

±0.128 

+0.263 

±0.068 

±0.153 

±0.123 

Total 

0.049 

0.174 

0.028 

0.129 

0.000 

0.073 

46 

±0.128 

±0.259 

±0.068 

±0.148 

±0.122 

Salamanders 

Northwestern 

0.000 

0.008 

0.000 

0.000 

0.000 

0.000 

1 

Salamander 

_. 

±0.025 



, , ,  , 

(Ambysfoma  gracile) 

Clouded  salamander 

0.496 

0.390 

0.361 

0.725 

0.146 

0.500 

227 

(Aneides  ferrous) 

+0.914 

±0.457 

±0.215 

±0.451 

±0.172 

±0.415 

Black  salamander 

0.099 

0.121 

0.000 

0.050 

0.000 

0.025 

35 

(A.  fiavipuncfafus) 

+0.178 

±0.272 

±0.070 

±0.065 

Calif,  slender^ 

2.718 

5.533 

4.470 

5.542 

0.417 

4.500 

972 

salamander 

±1.958 

±1.065 

±1.320 

±1.738 

±2.190 

(Bafrachoseps 

affenuafus) 

Pacific  giant 

0.091 

0.008 

0.000 

0.008 

0.021 

0.008 

12 

salamander 

±0.183 

±0.026 



±0.026 

±0.042 

±0.026 

(Dicampfodon 

ensatus) 

Ensatina 

2.265 

2.595 

2.625 

4.508 

2.938 

3.629 

1447 

(Ensafina 

±1.653 

±1.391 

±2.321 

±2.816 

±1.332 

±2.506 

eschscholfzii) 

Del  Norte* 

0.278 

0.396 

1.722 

2.278 

0.208 

1.622 

258 

salamander 

±0.411 

±0.970 

±2.237 

±3.349 

±0.191 

±2.607 

(Plefhodon 

elongafus) 

Olympic  salamander 

0.000 

0.038 

0.070 

0.192 

0.000 

0.116 

31 

(Rhyacofrifon 



±0.086 

±0.111 

±0.258 



±0.203 

olympicus) 

Rough-skinned  newt 

0.038 

0.140 

0.028 

0.192 

0.021 

0.108 

49 

(Taricha  qranulosa) 

+0.101 

+0.183 

+0.068 

+0.399 

+0.042 

+0.290 

Total 

5.041 

6.030 

6.180 

9.260 

3.385 

7.160 

3032 

±1.917 

+2.991 

±1.490 

±4,900 

±1.485 

+4.234 

Total 

5.090 

6.204 

6.208 

9.390 

3.385 

7.233 

3078 

amphibians 

±1.969 

±3.078 

±1.532 

±4,900 

±1.485 

±4.270 

Lizards 

Western  skink 

0.008 

0,045 

0.000 

0,008 

0.063 

0.017 

10 

(Eumeces 

±0.025 

±0.101 

±0,026 

±0.125 

±0.058 

skilfonianus) 

Northern 

0.095 

0.167 

0.014 

0.042 

0.084 

0.042 

42 

Alligator  lizard 

±0.160 

±0.230 

±0.034 

±0.044 

±0.096 

±0.057 

(Elgaria  coeruleus) 

Southern 

0.000 

0.008 

0.000 

0.000 

0.063 

0.013 

3 

Alligator  lizard 

±0.025 

±0.125 

±0.056 

(E.  mulficarinafus) 

(Continued) 


443 


Table  Z —  (continued) 


Species 


Young 
(11)^ 


Mature 
01) 


Western  fence  lizard 
(Sceloporus 
occldentalis) 
Total 


Snakes 
Rubber  boa 
(Charina  botfae) 
Sharp-tailed  snake 
(Contia  tenuis) 
Ringneck  snake 
(Diadophis  punctatus) 
Western  aquatic 
garter  snake 
(Ttiamnophis  couchii) 
Terrestrial 
garter  snake 
(T.  elegans) 
Nortti  western 
garter  snake 
(T.  ordinoides) 
Common  garter  snake 
(T,  sirtalis) 

Total 


Total  reptiles 
All 

herpetofauna 


0.000 


0.102 
±0.170 

0.008 
±0.025 

0,008 
±0.025 

0.000 

0.000 


0,015 
±0.050 

0.023 
+0.075 


0.023 
±0.054 

0.242 
±0.313 

0.000 

0.008 
±0.025 

0.057 
±0.109 

0.008 
±0.025 

0.000 


0.011 
+0.038 


Old-wet 

Old-mesic 

Old-dry 
(4) 

Total  old 
(20) 

0,000 

0,000 

0.083 
±0.118 

0.017 
±0.058 

0,014 
±0.034 

0.058 
±0.068 

0.292 
±0.323 

0.092 
±0.173 

0,000 

0.000 

0.000 

0.000 

0.014 

+0  n'^4 

0.000 

0.083 

±U.  10/ 

0.021 
±U,u/6 

0.000 
0.000 

0.008 
±0.026 
0.000 

0.073 
±0.086 
0,000 

0.019 

+0  048 
0.000 

0.000 

0.000 

0,000 

0.000 

0,000 

0.000 

0.031 

0.006 

Total 
Captures 


+0,063 


'Mean  ±  1  standard  deviation. 

^Data  are  from  inland  and  coastal  stands  combined. 

^Number  of  stands. 

"Absent  from  inland  stands. 

^Absent  from  coastal  stands. 


+0.028 


0.000 

0.000 

0.014 

0,000 

0.000 

0.000 

±0.034 

0.053 

0.083 

0.028 

0.008 

0.187 

0.050 

±0.086 

±0.138 

±0.043 

±0.026 

±0.239 

±0.122 

0.155 

0.326 

0.042 

0,066 

0.479 

0.142 

±0.210 

±0.418 

±0.069 

±0.086 

±0,473 

±0.265 

5.246 

6.530 

6.250 

9.450 

3.865 

7.371 

±2.004 

±3.205 

±1.559 

±4.900 

±1.543 

+4.203 

61 

1 
7 
11 
1 


1 

27 
88 
3166 


coastal  stands  had  significantly  more 
species  per  stand  than  the  inland 
stands  (fig.  3,  table  Al). 

With  coastal  and  inland  areas 
combined,  our  mean  species  values 
indicated  that  species  richness  was 
greatest  on  mature  stands  (table  6), 
but  was  not  statistically  different. 

In  the  inland  area,  the  old-growth 
dry  stands  had  the  greatest  mean 
number  of  species  (table  6)  but  no 
comparisons  yielded  significant  dif- 
ferences (fig.  3).  Within  the  coastal 


area,  mean  numbers  of  species  were 
significantly  different  between  forest 
age  classes.  Multiple  comparisons 
(TU)  indicated  that  the  greatest  dif- 
ferences occurred  between  young 
and  mature  stands  (fig.  3). 

The  significantly  higher  number  of 
species  in  the  coastal  vs.  the  inland 
area  (fig.  3)  is  attributable  to  the 
salamander  Aneides  lugubris  and  four 
snakes  (Thamnophis  couchii,  T.  sirtalis, 
T.  elegans,  and  Charina  bottae),  which 
were  all  sampled  in  very  low  num- 


bers and  only  in  the  coastal  area 
(tables  3-4).  We  believe  this  is  an  arti- 
fact of  the  difficulty  of  sampling  for 
snakes  in  forested  habitats  (Bury  and 
Corn  1987,  Raphael  and  Marcot  1986, 
Welsh  1987).  Most  snake  species  exist 
in  low  densities,  and  available  sam- 
pling methods  only  establish  pres- 
ence. All  of  these  snake  species  occur 
in  the  ii\land  area.  The  arboreal 
salamander,  Aneides  lugubris,  is  ab- 
sent inland  at  the  northern  latitudes 
we  sampled  (Stebbins  1985). 


444 


Table  4.— Mean  number  of  captures  per  1000  trap-nights'  captured  by  pitfall  traps  (PF)  in  different  age  and  moisture 
classes  of  Douglas-fir  forests  of  northwestern  California  and  southwestern  Oregon.  Sampling  occurred  In  the  falls  of. 
1984  and  1985.2 


Young         Mature        Old-wet      Old-mesic      Old-dry       Total  old  Total 
Species  (10)^  (11)  (6)  (9)  (4)  (19)  Captures 


Prone 
n  uyo 

Tnilpd  froa 

0.000 

V  V           1  ^  III    1  V-/  \-A 

0.000 

Pacific  treefroa 

0.139 

^Hvln  rf^nilln) 

+0.293 

Yellow-l©aaed  froa 

0.000 

fQnnci  Hnv/lii) 

Total 

0  139 

±0.293 

Snlnmnnciers 

North  wpstern 

1  '1  \y  1  II   1  VV       O  1       1  1  1 

0.035 

<;alannand©r 

±0.110 

(Ambysfomo  gracile) 

Clouded  salamar^der 

0.035 

±o.no 

Black  salamander 

0.035 

(A.  flavipuncfofus) 

±0.1 10 

Arhorenl  ^nlnmnnder 

0.035 

('/A,  lugubris) 

±0.1 10 

Calif,  slender^ 

0.298 

salamander 

±0.422 

(Bafrach  oseps 

affenuafus) 

Pacific  giant 

0.104 

salamander 

±0.168 

(Dicompfodon 

ensnfijs) 

Ensatina 

W  1    1  \J         III  1 

8.646 

CEnsofino 

±7.107 

eschscholfzii) 

Del  Norte^ 

0,810 

salamander 

±1.120 

(Pie  f hod  on 

elongafus) 

Olympic  salamander 

0.000 

(Rhyacofrifon 

olympicus) 

Rough-skinned  newt 

0.174 

(Toricha  granulosa) 

+0.245 

Total 

9.620 

±7.340 

Total 

9.760 

Amphibians 

±7.480 

Lizards 

Western  skink 

0.000 

(Eumeces 

skilfonionus) 

V 


n  06'^ 

n  nnn 

0  n'^9 

+0  909 

+0  116 

0  06'^ 

0  oon 

0  000 

+0  140 

0.315 

0,058 

0,077 

±0.105 

±0.142 

±0.159 

0  '^lA 

0  058 

0  579 

±0.667 

±0.142 

±1,493 

0  473 

0  116 

0  694 

±0.948 

±0.179 

±1.483 

0.032 

0,116 

0.000 

+0  105 

+0  284 

0  063 

0  058 

0  039 

+0.140 

±0.142 

±0,116 

0.410 

0.000 

0.193 

+  1  9S0 

+0  352 

0.000 

0,000 

0,000 

2.153 

0.463 

2,517 

+  1  162 

±0,401 

±1.037 

0.126 

0.578 

0.154 

±0.234 

±0.474 

±0.252 

10.164 

6.539 

9,375 

±8.996 

±4.328 

±7,209 

0  120 

1.500 

13.060 

±0.280 

±2.310 

±25,370 

0.000 

0.058 

0.000 

±0.142 

0.442 

0.174 

0.579 

+0.493 

+0.290 

+  1.264 

12.280 

8.510 

18.750 

+8.590 

±4.680 

±21.040 

12.750 

8.620 

19.440 

±8.340 

±4.670 

±20.940 

0.095 

0.058 

0.000 

±0.225 

±0.142 

0  nnn 

n  niR 

+n  nfto 

0  087 

0  018 

3 

+0  174 

+0  080 

0.087 

0.073 

9 

±0.174 

±0.145 

0  087 

0  31 1 

±0.174 

±1.035 

27 

n  96n 

n  49n 

±0.332 

±1.039 

0.000 

0.037 

4 

+0  159 

0  nnn 

0  037 

5 

+0.109 

0.087 

0.110 

20 

+0  174 

+0  260 

0.000 

0.000 

1 

0.347 

1.476 

72 

+  1  322 

0.000 

0.256 

21 

±0.381 

14.757 

9.613 

1097 

±11.260 

±7.648 

0  000 

6.340 

213 

±17.330 

0.000 

0.018 

1 

±0,080 

0.087 

0.347 

38 

+0.174 

+0.889 

15.020 

14.730 

1472 

+  11.380 

±15.710 

15.280 

15.150 

1514 

±11.610 

±15.710 

0.087 

0.037 

5 

±0,174 

±0.109 

(Continued) 
 J 


445 


Table  A.— (continued). 


Species 


Young 


Mature        Old-wet      Old-mesic      Old-dry       Total  old  Total 
(11)  (6)  (9)  (4)  <19)  Captures 


Northern 
alligator  lizard 
(Elgoria  coeruleus) 
Southern 
alligator  lizard 
(E.  mulficarinafus) 
Western  fence  lizard 
(Sceloporus 
occidenfalis) 
Total 

Snakes 
Northwestern 
garter  snake 
(T.  ordinoides) 


0.035 
±0.110 

0.000 


0.000 


0.037 
±0.110 

0.000 


0.000 


0.000 


0.032 
±0.105 

0.126 
±0.321 

0.000 


0.000 


0.000 


0.000 


0.058 
±0.142 

0.000 


0.039 
±0.116 

0.039 
0.116 

0.000 


0.077 
±0.153 

0.000 


0.260 
±0.521 

0.000 


0.174 
±0.347 

0.521 
±0.601 

0.087 
+0.174 


0.073 
±0.248 

0.018 
±0.080 

0.037 
±0.159 

0.164 
±0.335 

0.018 
+0.080 


14 


Total 

0.000 

0.000 

0.000 

0.000 

0.087 

0.018 

1 

±0.174 

±0.080 

Total  reptiles 

0.035 

0.126 

0.058 

0.077 

0.608 

0,183 

15 

±0.110 

±0.321 

±0.142 

±0.153 

±0.716 

±0.390 

All 

9.791 

12.877 

8.680 

19.527 

15.888 

15.333 

1529 

herpetofauna 

±7.572 

±8.408 

±4.742 

±20.888 

±11.555 

±15.694 

'Mean  ±  1  standard  deviation. 

^Data  are  from  inland  and  coastal  stands  combined. 

^Number  of  stands. 

^Absent  from  inland  stands. 

^Absent  from  coastal  stands. 


The  fact  that  we  generally  found 
more  species  on  older  stands  and 
that  we  found  a  greater  similarity 
between  mature  and  old-growth 
stands  than  between  either  of  these 
older  classes  and  young  stands  (see 
also  Raphael,  this  volume)  suggests 
that  both  the  mature  and  old  forest 
age  classes  provide  more  suitable 
habitat  and  a  more  diverse  herpe- 
tofauna than  young  forests. 


Relative  Abundance  Analysis 

Differences  Between  TCS  and  PF 

A  notable  aspect  of  our  data  is  the 
differences  between  the  TCS  and  PF 
methods — both  in  terms  of  kinds  of 


species  and  numbers  of  individuals 
captured.  These  differences  follow 
from  the  different  natures  of  these 
sampling  methods.  TCS  is  an  active 
search  method  that  permits  the  in- 
vestigator to  seek  out  animals  where 
they  hide.  PF  is  a  passive  method 
that  relies  on  animal  surface  move- 
ment or  the  seeking  of  shelter  under 
trap  covers  (Welsh  1987.) 

The  results  of  our  comparisons  of 
salamander  captures  between  coastal 
and  inland  areas  using  TCS  and  PF 
data,  which  appear  contradictory, 
serve  to  illustrate  the  pronounced 
differences  between  the  two  meth- 
ods. With  TCS  data,  in  all  compari- 
sons except  the  old-growth  wet  cate- 
gory, the  coastal  area  had  higher 
mean  captures  than  the  inland  area. 


This  result  was  due  to  high  captures 
(over  900  individuals)  of  a  single  spe- 
cies of  salamander,  Batrachoseps  at- 
tenuatus,  a  species  that  occurred  in  all 
age  and  moisture  classes.  This  spe- 
cies is  absent  inland.  However,  sev- 
eral factors  unique  to  the  inland  area 
acted  to  counter  the  effects  of  the 
high  captures  of  B.  attenuatus.  Those 
factors  were  the  high  captures  of  Ple- 
thodon  elongatus  (more  than  250  cap- 
tures), a  species  found  almost  exclu- 
sively on  the  inland  stands,  and 
higher  relative  captures  of  Ensatim 
eschscholtzii  inland  (865  inland  vs.  580 
coastal). 

In  contrast,  results  from  PF,  indi- 
cated significantly  higher  captures  on 
inland  stands  than  on  coastal  stands, 
for  all  stands  combined  (table  Al). 


446 


PF  captured  few  (n=72)  of  the  highly 
sedentary  Batrachoseps  attenuatus 
relative  to  TCS  (n=972).  Captures  of 
the  relatively  more  vagile  salaman- 
der species,  P.  elongatus  and  E.  es- 
chscholtzii,  were  greater  on  the  inland 
stands  than  the  coastal  stands,  for  the 
PF  data. 

TCS  provided  a  more  complete 
data  set,  sampled  more  species  (par- 
ticularly reptiles)  and  had  twice  as 
many  individuals  as  did  PF  (tables  2- 
4).  The  active  nature  of  TCS  accounts 


for  the  disparities  in  capture  num- 
bers, and  in  the  lack  of  consistency  of 
statistically  significant  differences 
among  forest  age  and  moisture 
classes  between  these  data  sets,  even 
for  the  same  species  (table  Al).  Most 
significant  results  from  our  analyses 
derived  from  the  larger  TCS  data  set. 
Subsequent  discussion  of  results  will 
refer  to  these  data  unless  they  are 
identified  as  PF  data.  Mean  captures 
(+  one  standard  deviation)  for  all 
taxa  analyzed  are  found  in  tables  3 


Table  5.— Jaccard  similarity  index  (JSI)  values  for  species  of  herpetofauna 
in  3  age  classes  of  Douglas-fir  forests  of  rK>rthwestem  California  and  south- 
western Oregon.  Values  were  calculated  using  10  randomly  selected 
stands  from  each  forest  age  class,  including  coastal  and  inland  areas. 
Greater  JSI  values  indicate  greater  similarity  in  species  composition. 


All  stands 

(Areas  combined) 


Young 


Mature 


Old-growth 


Mature 
Old-growth 

Total  number  of  species 


.542 
.467 
16 


.846 
21 


15 


Table  6.— Mean  (±  1  standard  deviation)  numbers  of  species  of  herpe- 
tofauna among  three  age  and  three  moisture  classes  of  Douglas-fir  forests 


of  northwestern  California  and  southwestern  Oregon. 

Inland  stands 

Young  Mature  Old-dry  Old-mesic 

Old-wet 

Number  of 

stands 

3 

6 

3 

6 

3 

Mean  number 

5.67 

4.67 

6.67 

5.17 

6.33 

of  species 

±3.06 

±1,51 

±3.51 

±1.33 

±2.08 

Total  number 

of  species 

10 

13 

14 

12 

10 

Coastal  stands 

Number  of 

stands 

8 

5 

1 

4 

3 

Mean  number 

5.50 

10.00 

6.00 

9,25 

5.00 

of  species 

±2.56 

±2.92 

±2.63 

±2.00 

Total  number 

of  species 

16 

21 

6 

15 

10 

All  stands 

Number  of 

stands 

11 

n 

4 

10 

6 

Mean  number 

5.55 

7.10 

6.50 

6.80 

5.67 

of  species 

±2.54 

±3.51 

±2.89 

±2.78 

±1.97 

Total  number 

of  species 

17 

23 

17 

17 

14 

V 

and  4.  Results  of  all  tests  on  both 
data  sets,  and  test  statistics  for  those 
tests  with  significant  differences,  are 
found  in  table  Al. 


Salannanders 

Almost  all  captures  (96.3%)  were 
salamanders  (table  2),  consequently, 
the  results  of  our  analyses  were  es- 
sentially the  same  for  all  herpe- 
tofauna, amphibia,  and  salamanders 
(species  combined)  (table  Al).  Sala- 
manders were  not  equally  distrib- 
uted among  forest  age  classes.  Test- 
ing the  equality  of  mean  captures 
among  age  classes,  with  coastal  and 
inland  areas  combined,  yielded  sig- 
nificant differences.  Multiple  com- 
parisons (TU)  indicated  these  differ- 
ences were  between  the  young  and 
old  stands,  with  more  captures  on 
the  old  stands  (fig.  4). 

Salamanders  were  not  equally  dis- 
tributed among  forest  moisture 
classes.  Multiple  comparisons 
(GHMC),  with  areas  combined,  indi- 
cated a  significant  difference  in  mean 
captures  between  the  old-growth 
mesic  and  old-growth  dry  stands, 
with  more  captures  in  the  mesic 
stands  (fig.  4).  These  differences  are 
probably  a  result  of  the  fact  that  drier 
sites  offer  less  equable  habitat  for 
amphibians.  We  also  captured  fewer 


at 


I  ♦  % 

urn 

I  -  X 


wvuc     (uati    umc    oavt    ai  sukb 
STAND  TTPt 

Figure  3.— Numbers  of  species  of  herpe- 
tofauna captured  \n  the  coastal  arxl  inland 
areas  in  three  forest  age  and  three  forest 
moisture  classes  of  Douglas-fir  dominated 
forests  from  1984-1986.  Captures  were  by 
time-constrained  search  (TCS)  and  pitfall 
traps  (PF). 


447 


amphibians  on  old-wet  stands  than 
old-mesic  stands,  although  the  differ- 
ence is  not  statistically  significant. 

Within  the  coastal  area,  multiple 
comparisons  (TU)  indicated  that  both 
mature  and  old-growth  mesic  stands 
were  significantly  different  from 
young  stands,  but  not  from  each 
other,  with  the  lowest  mean  captures 
occurring  on  the  young  stands  (fig. 
5a).  Between-area  comparisons  for 
salamanders  indicated  a  significant 
difference  in  means  between  coastal 
and  inland  mature  stands  (fig.  5a). 

The  PF  data  yielded  no  significant 
differences  between  mean  captures 
in  age  or  moisture  classes  with 
coastal  and  inland  areas  combined  or 
within  either  area  (table  Al).  How- 
ever, comparisons  between  these  ar- 
eas indicated  a  significant  difference 
with  all  stands  combined  (fig.  5b). 
The  greatest  differences  occurred  be- 
tween the  old-growth  wet  stands; 
however  the  results  were  not  signifi- 
cant (fig.  5b). 

The  greater  number  of  individuals 
in  older  stands  parallel  our  findings 
of  greater  numbers  of  species  in 
older  forest  age  classes  (table  6).  As 
with  the  species  richness  analysis,  the 
number  of  individuals  was  greater  in 
older  forests  of  the  coastal  area  than 
in  the  inland  area.  These  differences 
suggest  that  older  forests  support 
both  a  richer  and  more  abundant 
salamander  fauna. 

The  lower  capture  rates  on  old- 
wet  stands  compared  to  old-mesic 
was  an  unexpected  result.  We  offer 
two  possible  explanations  for  these 
lower  sample  values.  One  possibility 
is  that  the  habitat  structure  is  more 
complex  on  these  wet  forest  stands, 
with  more  and  larger  downed 
woody  material,  a  thicker  duff  layer, 
and  denser  undcrstory  vegetation 
requiring  more  time  to  search  and 
making  it  more  difficult  to  find  ani- 
mals (TCS  method)  and  making  them 
less  likely  to  be  moving  about  on  the 
surface  and  encountering  our  traps 
(PF  method).  A  second  possibility  is 
that  the  wet  stands  actually  contain 
fewer  salamanders. 


Salamanders  play  an  important 
functional  role  in  forest  ecosystems 
because  of  several  unique  aspects  of 
their  ecology.  Though  they  are  small, 
with  90%  of  species  having  adult 
body  masses  less  than  those  of  small 
birds  and  mammals  (Pough  1980), 
they  are  often  a  major  portion  of  the 
vertebrate  biomass  in  a  forest.  At  the 
Hubbard  Brook  Experimental  Forest 
in  New  Hampshire,  a  single  species 
of  salamander  accounted  for  a 
greater  portion  of  biomass  and  sec- 
ondary productivity  than  any  other 
vertebrate  group  (Burton  and  Likens 
1975a,b).  Their  small  size  enables 
them  to  exploit  prey  too  small  to  be 
used  by  birds  and  mammals  and 
subsequently  to  convert  these  prey 
into  biomass  that  is  available  to 
larger  vertebrates  (Pough  1983). 
Pough  et  al.  (1987)  cites  both  direct 
observations  of  predation  and  the 
ubiquity  of  defensive  mechanisms 
among  salamanders  as  evidence  of 
their  importance  as  a  food  source  for 
both  avian  and  mammalian  preda- 
tors. Because  salamanders  are  ectoth- 
erms  and  have  the  lowest  metabolic 
rates  of  any  terrestrial  vertebrates 
(Feder  1983),  this  biomass  conversion 
process  is  extremely  efficient,  with 
40-80%  of  the  energy  invested  being 
used  to  produce  new  biomass 
(Pough  et  al.  1987).  As  a  consequence 
of  these  characteristics,  salamanders 
are  quantitatively  and  qualitatively 
important  components  of  food  webs 


of  many  forest  ecosystems.  The  fact 
that  their  numbers  appear  to  be  re- 
duced by  certain  forest  practices 
could  potentially  affect  energy  flow 
and  biomass  production  at  all  bio- 
logical levels. 

Frogs 

Testing  the  equality  of  mean  captures 
yielded  significant  differences  in  cap- 
tures of  frogs  in  coastal  age  and 
moisture  classes,  with  significantly 
higher  mean  captures  in  old  vs. 
young  stands  and  mesic  vs.  wet 
stands  (table  Al).  These  results  are 
attributable  to  a  single  species,  the 
Pacific  treefrog.  No  other  significant 
differences  were  found  (table  Al). 

17  1 

osMwc    looav  ~j~ 
4  ■  siK       I  -  xa 

IS '  KM  KM 

a:  'J- 


fOJM         WTME        OU  KT       OU  KSK       «fi  OT 

STAWD  TYPE 


Figure  4.— Captures  of  salarrtanders  per 
person-hour  (TCS)  in  tt^ree  forest  age  and 
three  moisture  classes.  Data  are  from  the 
coastal  and  inland  areas  combined,  and 
sampling  occurred  from  1984-1986. 


f  <  mi 

misM  wrna  >  cnei*.  loac 

« ■  \nt 
emeu.  OU  <tx  >  Cttsm  nuc 
« -  loi 
Km 


0COHSTAI. 
HJttl 


ICM  t  SC 

new 
HUN  -  X 


u  90 
z 
I 


HXJNC        IHROT       010  HC      OU)  1£SC      OLD  DBT     MX  SIAMDS 

STAND  TYPE 


L  sua  ><ii  s 

I  -  2JI 

»  .  »II 


T 


mm  «  z 

ION 


YTKMO      iw\j*i     (uwi    aamx    aeon  usunm 
STAHDTYPE 


Figure  5.— Captures  of  salamanders  per  person-hour  (A:TCS)  and  per  1000  trap-nights  (B:PF) 
in  the  coastal  and  inland  areas.  Data  are  from  1984-1986  aCS)  and  1984-1985  (PF). 


448 


Reptiles 

The  reptile  fauna  in  the  forests  of  the 
Pacific  Northwest  is  depauperate 
(Nussbaum  et  al.  1983,  Stebbins  1985) 
with  most  species  occurring  in  rela- 
tively low  abundance  (tables  3-4). 
Distribution  of  reptile  species,  by  age 
and  moisture  class,  indicated  about 
equal  numbers  of  species  in  the 
young,  mature,  and  old-growth  age 
classes,  with  lower  numbers  of  spe- 
cies in  old-growth  wet  forests. 

Based  on  TCS  and  PF  data,  our 
mean  captures  of  reptiles  (species 
combined)  were  higher  on  both  drier 
and  older  stands,  but  the  differences 
were  not  statistically  significant.  Our 
sample  sizes  were  not  sufficient  to 
analyze  for  differences  among  age 
and  moisture  classes  at  the  species 
level,  except  for  the  northern  alliga- 
tor lizard  for  which  our  data  indi- 
cated no  statistically  significant  asso- 
ciation with  a  particular  forest  age  or 
moisture  class  (table  Al). 

We  did  not  sample  in  any  recently 
harvested  areas,  but  given  their  pref- 
erences for  open  areas  and  their  re- 
lated heliothermic  natures,  reptiles, 
particularly  lizards,  probably  in- 
crease following  logging,  and 
through  the  early  serai  stages  of  re- 
generating forests  (see  Raphael,  this 
volume).  Raphael  and  Marcot  (1986) 
indicated  that  the  sagebrush  lizard 
(Sceloporus  graciosus)  was  four  times 


a»  IOC  >  lOK 


a* icac>  M 1CT 

42} 


I 

I 


jTuoc      CLD  w     <u  neve     cm  I 
STAND  TYPE 


Figure  6.— Captures  per  person-hour  (TCS) 
of  the  Pacific  treefrog  (Hyla  regilla),  in  three 
forest  age  and  three  moisture  classes.  Data 
are  fronn  the  coastal  area  from  1 984- 'i  986. 


as  abundant  in  early  vs.  late  shrub 
stages. 


Relative  Abundance  of  Connmon 
Species 

Common  species  (captured  on  at 
least  one  third  of  our  stands  in  either 
area  by  either  sampling  method) 
were  analvzed  for  differences  in 
mean  captures  in  age  and  moisture 
classes,  across,  within  and  between 
coastal  and  inland  areas  (table  Al). 
Besides  the  northern  alligator  lizard, 
these  species  consisted  of  amphibi- 
ans— 2  frogs  and  7  salamanders. 
Other  amphibians  whose  distribu- 
tions relative  to  forest  age  were  con- 
sidered noteworthy  are  also  dis- 
cussed. 

Yellow-Legged  Frog  (Rana 
boylii). — This  species  was  absent 
from  all  young  stands  (table  4),  but 
they  were  also  captured  at  such  low 
frequencies  on  our  inland  stands  as 
to  preclude  analyses  within  this  area. 
Within  the  coastal  area,  no  significant 
differences  were  found  for  capture 
frequencies  of  this  species  in  forest 
age  or  moisture  classes  (table  Al). 

The  yellow-legged  frog  is  a  highly 
aquatic  species  (Stebbins  1985)  and 
therefore  our  PF  captures  (table  4) 
must  be  considered  incidental.  These 
captures  may  have  been  frogs  seek- 
ing terrestrial  overwintering  cover 
above  high  water  levels  (PF  sampling 
was  done  in  the  fall).  However,  this 
frog  was  absent  from  young  stands. 
Three  facts  need  be  considered:  (1) 
all  but  a  single  capture  occurred  in 
the  coastal  area;  (2)  in  general,  the 
coastal  stands  were  closer  to  peren- 
nial streams  and  creeks  than  were 
the  inland  stands;  (3)  within  the 
coastal  area,  only  two  out  of  eight 
young  stands  had  PF  grids  near  suit- 
able aquatic  habitat,  whereas  all  the 
mature  and  old-growth  stands  had 
PF  grids  near  such  habitat.  Thus  we 
can  not  rule  out  the  possibility  that 
this  frog's  absence  from  young 
stands  in  our  sam.ples  is  an  artifact  of 
our  stand  locations  relative  to  avail- 


able and  suitable  aquatic  habitat 
(Bury  and  Corn,  this  volume). 
Twenty-one  records  from  area-con- 
strained aquatic  surveys  (H.  Welsh, 
unpubl.  data)  were  almost  equally 
divided  between  creeks  in  young  and 
mature  forests.  On  the  other  hand,  it 
is  possible  that  older  forests  provide 
some  particulars  of  microhabitat  re- 
quired by  overwintering  yellow- 
legged  frogs  not  present  in  young 
forests. 

Pacific  Treefrog  (Hyla  regilla). — 
The  Pacific  treefrog  is  the  only  frog 
for  which  our  data  indicated  signifi- 
cant differences  in  captures  between 
both  forest  age  and  moisture  classes 
(fig.  6).  Within  the  coastal  area,  this 
frog  was  captured  at  significantly 
different  frequencies  in  both  forest 
age  and  moisture  classes.  However, 
these  differences  were  not  observed 
within  the  inland  area,  probably  due 
to  lower  captures  and  higher  vari- 
ances on  these  stands  (table  Al). 

Because  the  Pacific  treefrog  is  not 
restricted  to  forested  habitat  (Steb- 
bins 1985),  we  are  suspicious  of  our 
data  indicating  greater  abundance  in 
older  forests  (fig.  6).  Conceivably 
older  forests  provide  more  cover  and 
foraging  areas  for  this  species  than 
do  young  forests  and  thus  support 
higher  relative  abundances.  Most  of 
our  captures  of  treefrogs  occurred  in 
association  with  large  downed 
woody  material.  However,  we  can- 
not rule  out  the  ]X)ssible  influence  of 
proximity  of  breeding  sites  on  these 
results  (Bury  and  Com,  this  volume). 
The  older  forest  stands  were  gener- 
ally closer  to  standing  water  than  the 
young  stands  (as  with  Ram  boylii)  in 
the  coastal  area. 

The  difference  in  captures  of 
treefrogs  between  the  mesic  and  wet 
moisture  classes  (fig.  6)  may  be  an 
artifact  of  unequal  detectability.  Most 
treefrogs  were  captured  by  TCS  and 
they  are  more  easily  exposed  and 
seen  by  investigators  in  the  more 
open  understory  of  the  mesic  stands. 
The  alternate  possibility,  that  there 
are  actually  more  treefrogs  on  mesic 
stands,  is  consistent  with  the  in- 


449 


creased  incident  radiation  in  the  me- 
sic  stands  which  would  promote 
higher  productivity  of  invertebrate 
prey,  and  thus  possibly  support 
more  treefrogs. 

The  Tailed  Frog  (Ascaphus 
truei), — This  frog  was  captured  only 
on  mature  and  old-growth  stands 
(tables  2-3);  however,  the  total  num- 
ber of  captures  (5)  was  too  low  for 
statistical  tests.  This  species  is  of 
interest,  nonetheless,  because  of  its 
absence  from  young  stands.  The 
tailed  frog,  like  the  yellow-legged 
frog,  is  highly  aquatic  (Bury  1968, 
Stebbins  1985).  Therefore  these  rec- 
ords based  on  terrestrial  sampling 
are  considered  incidental.  However, 
results  from  another  study  employ- 
ing an  area-constrained  aquatic  sam- 
pling method  yielded  more  than  400 
captures  of  tailed  frogs  (Welsh,  in 
prep.).  These  data  were  consistent 
with  the  incidental  records  reported 
here;  there  were  significant  increases 
in  tailed  frog  abundance  with  in- 
creased forest  age. 

Olympic  Salamander  (Rhyacotri- 
ton  olympicus). — This  species  was 
absent  from  all  young  stands  (tables 
3-4).  Low  captures  prompted  us  to 
combine  moisture  classes  for  the  age 
analysis.  Multiple  comparisons 
(GHMC),  coastal  and  inland  areas 
combined,  indicated  that  older 
stands  had  significantly  greater  num- 
bers of  Olympic  salamander  than 
young  stands  (fig.  7). 

This  species  is  restricted  to  head- 
water habitats,  such  as  seeps, 
springs,  and  small  creeks  in  forests 
where  it  prefers  cold  water  flowing 
over  rocky  substrates  (Anderson 
1968,  Nussbaum  et  al.  1983).  Because 
of  the  relative  scarcity  of  this  mi- 
crohabitat  in  the  areas  of  our  study, 
Rhyacotriton  occurs  in  a  patchy  distri- 
bution. It  can  be  abundant  where 
conditions  are  suitable,  but  we  found 
appropriate  microhabitat  islands  for 
this  species  to  be  few,  small,  and 
widely  scattered  on  our  stands.  This 
resulted  in  relatively  few  captures 
(tables  3-4).  We  found  Rhyacotriton 
absent  in  younger  forests  (fig.  7), 


which  is  consistent  with  results  from 
other  studies  (Bury  1983;  Bury  and 
Corn  1988;  Welsh,  in  prep.).  This  spe- 
cies appears  to  be  sensitive  to  forest 
harvest  practices  because  of  its  par- 
ticular habitat  requirements  (Bury 
and  Corn  1988;  Welsh,  in  prep.).  Cur- 
rent harvest  practices  do  not  protect 
headwater  habitats.  Such  habitats  are 
often  radically  altered  by  harvest 
practices,  which  can  change  water 
flow  and  temperature,  increases  sedi- 
ment loads,  and  change  the  structure 
and  composition  of  the  riparian 
vegetation  (Bury  and  Corn  1988).  The 
result  of  these  changes  is  often  the 
extirpation  of  local  populations  of 
this  species. 

Clouded  Salamander  (Aneidesfer- 
reus). — Multiple  comparisons 
(GHMC)  indicated  significant  differ- 
ences in  mean  captures  of  clouded 
salamanders  between  young  and  old 
stands  in  the  inland  area  but  not  in 
the  coastal  area  (fig.  8a).  Testing  for 
differences  with  coastal  and  inland 
areas  combined  revealed  significant 
differences  in  mean  captures  among 
moisture  classes;  multiple  compari- 
sons (TU)  indicated  that  the  mesic 
stands  had  significantly  higher  mean 
captures  than  did  dry  stands  (fig. 
8b). 

This  species,  a  habitat  specialist, 
occurs  most  often  under  exfoliating 
bark  on  downed  conifer  logs  (Steb- 
bins 1985,  Nussbaum  et  al  1983).  At 
several  coastal  redwood  localities. 


z 


(3  i-o 

c 

3 


I  -  SO 
P  <  GB 


I 


I 


UIM  *  SE 
UCM 

VlUi  -  St 


rOREST  ACE  CUSS 


Of  ncsc) 


Bury  (1983)  and  Bury  and  Martin 
(1973)  found  it  to  be  more  abundant 
in  young  stands  than  older  stands. 
They  attributed  the  differences  to  an 
increase  in  bark  on  downed  woody 
material  from  logging.  Our  data  from 
the  coastal  area  (fig.  8a)  indicated 
slightly  more  A.  ferreus  in  younger 
than  older  forests,  but  the  differences 
were  not  significant.  However,  in  the 
inland  area  the  clouded  salamander 
was  found  in  significantly  higher 
numbers  on  old  vs.  young  stands 
(fig.  8a).  We  suspect  that  these  differ- 
ences are  due  to  the  differences  in 
moisture  regimes  between  the  two 
areas.  This  idea  is  supported  by  our 
findings  of  significant  differences  in 
capture  means  between  mesic  and 


OJ  >  KM 

Km 


WW  «  a 

mm 
ia«  -  s 


roREsr  AC£  cuss 


Figure  7.— Captures  per  person-hour  (TCS) 
of  the  Olympic  salamander  (Rhyacotriton 
olympicus),  in  three  forest  age  classes. 
Data  are  from  the  inland  and  coastal  areas 
combined,  from  1984-86. 


T 


■UHM 

MM  4  a 

ISM 
ICJM  -  X 


rOtEST  MOtSTURt  CLASS 


Figure  8.— (A)  Captures  per  person-hour  (TCS)  of  the  clouded  salamander  (Aneldes  ferreus) 
in  the  coastal  and  inland  areas,  in  three  forest  age  classes.  (B)  Captures  per  person-hour 
(TCS)  in  three  forest  moisture  classes;  data  are  from  coastal  and  inland  areas  combined 
Sampling  occurred  from  1984-86. 


450 


dry  old-growth  sites  (fig.  8b).  We 
suggest  that  logs  on  inland  young 
stands  are  subjected  to  higher 
evapotranspiration  rates  than  are 
logs  on  old-growth  stands  because  of 
greater  incident  radiation.  Possible 
increases  in  clouded  salamanders  on 
young  stands  from  an  increase  in 
slash  and  logs  after  harvesting  may 
be  outweighed  by  the  loss  of  suitable 
microclimatic  conditions  due  to  in- 
creased exposure. 

Black  Salamander  (Aneides 
flavipunctatus). — We  found  signifi- 
cantly greater  numbers  of  this  spe- 
cies in  the  coastal  area  than  in  the  in- 
land area  (fig.  9).  Lynch  (1981) 
pointed  out  that  inland  populations 
occur  in  a  patchy  distribution  charac- 


0a»sm. 


ICW 


•ntii      mniK     tUKS    as  tea:    cuar    «i  stmos 
STAND  TYPE 

Figure  9.— Captures  per  person-hour  (ICS) 
of  the  black  salamander  (Aneides  flavip- 
unctatus) in  coastal  and  inland  areas,  in 
three  forest  age  and  three  nnoisture  classes. 
Data  are  fronn  1984-86. 


-r  I 


I 


I 


MCNI  -  3 
I— 


ST*KD  TYPt 


teristic  of  a  species  on  the  decline. 
Further,  he  attributed  the  inland 
patchiness  to  climatic  constraints  and 
noted  that  the  black  salamander  is 
restricted  to  low-lying  suitable  areas 
receiving  at  least  75  cm  of  annual 
precipitation.  Its  restriction  to  rocky 
habitats  and  its  low  relative  abun- 
dance in  northwestern  California 
preclude  drawing  any  conclusions 
from  our  forest  age  and  moisture 
class  analysis  (table  Al). 

California  Slender  Salamander 
(Batrachoseps  attenuatus). — The 
slender  salamander,  like  the  black 
salamander,  appears  to  be  restricted 
to  low-lying  suitable  areas  with  rela- 
tively high  annual  precipitation 
(Maiorana  1976a).  This  species  was 
absent  from  our  inland  sites,  but  ac- 
counted for  the  highest  captures  of 
any  species  within  the  coastal  area. 
This  was  one  of  the  few  species  we 
captured  in  sufficient  numbers  with 
both  sampling  methods  to  test  both 
data  sets  for  differences  between  for- 
est age  and  moisture  classes  (see 
table  Al).  Within  the  coastal  area, 
both  TCS  and  PF  data  indicated  sig- 
nificant differences  in  mean  captures 
among  forest  age  classes  (figs.  lOa-b). 
Multiple  comparisons  (TU)  indicated 
that  these  differences  were  between 
both  young  and  mature  and  young 
and  old-growth  stands  (figs.  lOa-b). 
Our  findings  here  were  consistent 
with  trends  found  by  others  (Bury 
1983,  Bury  and  Marrin  1973). 


<l 

1.0 

u 
u 
u 

\s 

1.0 

u 

&0 


I 


h 



1 


now  •  St 

low 
IC»  -  3 

HNIUI 


STAHO  TYPt 


Figure  10.— Captures  per  person-hour  (A:TCS),  and  captures  per  1000  trap-nights  (B:PF).  of 
the  California  slender  salamander  (Batrachoseps  atfenuatus),  in  three  forest  age  and  three 
moisture  classes.  Data  are  from  the  coastal  area  from  1984-86  (TCS)  and  1984-85  (PF). 


The  PF  data  showed  a  significant 
difference  between  captures  in  mois- 
ture classes,  with  a  higher  mean  cap- 
tures on  mesic  than  on  old-growth 
wet  stands  (fig.  10b),  but  the  TCS 
data  did  not  (fig.  10a).  For  a  salaman- 
der species  whose  presence  and  rela- 
tive abundance  is  correlated  with 
relatively  high  and  predictable  mois- 
ture (Maiorana  1974, 1976a),  this  re- 
sult is  unexpected  and  may  be  an  ar- 
tifact of  different  sampling  efficien- 
cies between  forest  moisture  classes. 
The  old-growth  wet  stands  appear  to 
contain  habitat  with  relatively  great 
structural  complexity:  a  thick  and 
complex  layer  of  understory,  decom- 
p>osing  woody  material,  and  mossy 
duff.  Such  habitat  provides  abundant 
microhabitat  for  a  ground  dwelling 
and  semi-fossorial  species  like  the 
slender  salamander. 

Slender  salamanders  may  not  fre- 
quent the  surface  as  much  to  forage 
as  they  would  on  drier  stands.  Forag- 
ing in  more  protected  areas  would 
reduce  exposure  to  predation  and 
thus  incur  a  selective  advantage. 
Maiorana  (1976b)  termed  this  sub- 
mergent  behavior  (our  concept  is  a 
slight  variation  of  her  idea;  she  hy- 
pothesized that  a  species  might  actu- 
ally forage  less  at  times  to  avoid  ex- 
posure to  predation).  As  a  result  of 
less  surface  activity,  fewer  slender 
salamanders  are  captured  in  the  pit- 
fall traps.  The  same  logic  can  also  be 
applied  to  the  TCS  method,  in  which 
lower  captures  would  be  expected  in 
the  structurally  more  complex  habi- 
tat per  unit  of  search  time.  With  TCS, 
we  did  get  slightly  lower  captures  on 
old-growth  wet  stands  for  this  spe- 
cies (table  3),  but  the  active  nature  of 
TCS  allowed  us  to  detect  enough 
slender  salamanders  that  the  capture 
rates  between  moisture  classes  were 
not  significantly  different. 

Ensatina  (Ensatina  es- 
chscholtzii). — Ensatina  has  broad 
ecological  tolerances,  occurring  from 
relatively  dry  woodland  habitats  to 
moister  forests  at  high  elevations 
(Stcbbins  1954).  This  species  has  the 
most  extensive  geographic  distribu- 


451 


tion  of  all  the  western  woodland 
salamanders,  ranging  from  British 
Columbia  to  Baja  California  (Stebbins 
1985).  Ensatina  were  captured  in  the 
highest  numbers  of  any  species  we 
sampled  (table  3-4).  There  were  sig- 
nificant differences  in  mean  captures 
among  forest  age  classes,  with 
coastal  and  inland  areas  combined 
(fig.  11).  Multiple  comparisons  (TU) 
indicated  that  old  stands  had  signifi- 
cantly higher  captures  than  young 
stands  (fig.  11). 

Both  PF  and  TCS  data  indicated 
significant  differences  in  mean  cap- 
ture frequencies  between  the  coastal 
and  inland  areas  (figs.  12a-b).  Greater 
numbers  were  found  on  the  inland 
stands.  These  differences  between 
areas  indicate  that  this  species  may 
be  more  abundant  in  the  drier  inland 
area  than  along  the  coast. 

Del  Norte  Salamander  (Plethodon 
elongatus). — Except  for  three  cap- 
tures from  our  most  northern  coastal 
stand,  this  species  was  sampled  only 
on  our  inland  stands.  These  salaman- 
ders are  found  primarily  on  or  in 
rocky  substrates  (Stebbins  1985, 
Nussbaum  et  al.  1983),  and  reach 
high  densities  in  talus  and  outcrops 
of  fractured  metamorphic  rock.  Such 
habitats  were  not  present  on  some  of 
our  stands.  Also,  our  study  region 
encompassed  the  geographic  range 
of  this  species,  and  all  of  our  south- 
ern and  some  of  our  easternmost 
stands  were  beyond  its  geographic 
limits.  Despite  the  patchy  distribu- 
tion of  this  species  due  to  habitat  re- 
strictions, and  absence  from  sites  be- 
yond its  range,  both  methods  indi- 
cated a  higher  relative  abundance  on 
older  forest  stands  and  a  lower  rela- 
tive abundance  on  drier  stands  (figs. 
13a-b,  tables  3-4).  These  differences 
were  not  statistically  significant; 
something  we  attribute  to  high  vari- 
ances within  forest  age  classes  result- 
ing from  this  lack  of  appropriate  mi- 
crohabitat  and  the  inclusion  of  stands 
beyond  the  range  (table  Al).  A  sepa- 
rate analysis  of  only  stands  from 
within  the  geographic  range  of  the 
Del  Norte  salamander  indicated  that 


the  abundance  of  this  species  is  sig- 
nificantly correlated  with  increased 
forest  age  (Welsh,  in  prep.). 

Rough-Skinned  Newt  (Taricha 
granulosa). — Both  TCS  and  PF 
showed  a  marked  increase  in  cap- 
tures of  this  species  in  older  forests 
(figs.  14a-b,  tables  3-4).  Lack  of  statis- 
tically significant  differences  in  cap- 
tures between  forest  age  classes 
(table  Al)  is  probably  related  to  spe- 
cific habitat  requirements  of  this  spe- 
cies. We  suspect  that  the  critical  habi- 
tat component  was  proximity  to 
creeks  or  ponds,  a  breeding  require- 
ment for  this  species  (Stebbins  1985). 
Many  of  our  stands,  particularly 
within  the  inland  area,  were  a  con- 
siderable distance  from  suitable 
breeding  habitat  for  this  newt.  We 
had  no  TCS  captures  of  this  species 
on  old-growth  stands  in  our  inland 
area,  yet  the  rough-skinned  newt  is 
common  there  (Stebbins  1985,  pers. 
observ.). 


CONCLUSIONS 

Our  research  indicates  that  salaman- 
ders comprise  the  majority  of  both 
sf)ecies  and  individuals  among  the 
herpetofauna  of  the  Douglas-fir/ 
hardwood  forests  of  northwestern 
California  and  southwestern  Oregon. 
We  found  species  diversity  of  the  to- 
tal herpetofauna  to  be  greater  in 
older  forest  age  classes.  Amphibians, 


•UM  cnsui. 
u  am  >  «i  siucs 

1  .  213 


]CtMSM 
J  fUM 


h 


««  <  St 

ISW  -  5E 


rate        IMUt       CU  KT      OH)  UCSK     OLD  Om     AU  SUMS 

STAMO  lYPe 


particularly  salamanders,  were  sig- 
nificantly more  abundant  in  older 
forests  and  significantly  less  abun- 
dant in  drier  forests. 

We  found  the  TCS  method,  ac- 
tively searching  for  animals  in  their 
preferred  microhabitats  (usually  as- 
sociated with  downed  woody  mate- 
rials in  these  forest  habitats),  yielded 
more  useful  data  on  herpetofaunal 
diversity  and  abundance  relative  to 
forest  age  and  moisture  class  than 
did  PF.  The  TCS  method  sampled 
more  individuals  and  species  in  ad- 
dition to  taking  less  time  and  ex- 
pense than  PF  (see  Welsh  1987). 

Recent  research  in  forested  habi- 
tats (Bury  and  Corn  1988,  Pough  et 
al.  1987,  Enge  and  Marion  1986,  Bury 


ICAM  4  9E 
ICM 


K3MS  WIUC  an  0C9C) 

roRCST  Acc  cuss 

Figure  11.— Captures  per  person-hour  (ICS) 
of  Ensatina  (Ensatina  eschschotltzii).  In 
three  forest  age  classes.  Data  are  from 
coastal  and  inlartd  areas  combined,  from 
1984-86. 


T  » 


CO»SI»L 


mi  >  ai  sua 

I  •  IM 

t'  aa 


MM « a 

KAN 
ICM  -  X 


yoMC      MTum.     010  vt    ou  icac    <u  on  ta  sumr 
STWe  TYPE 


Figure  12.— Captures  per  person-hour  (A:TCS),  and  captures  per  1000  trap-nights  (B:PF),  of 
Ensatina  (Ensatina  eschschotltzii)  in  coastal  and  inland  areas,  in  three  forest  age  and  three 
moisture  classes.  Data  are  from  1984-86(TCS)  and  1984-85  (PF). 


452 


1983,  Bennett  et  al.  1980,  Bury  and 
Martin  1973)  has  indicated  a  pattern 
of  fewer  species  and  reduced  abun- 
dance of  herpetofauna  after  logging. 
We  also  found  lower  numbers  of 
both  species  and  individuals  on 
younger  stands. 

Greater  species  diversity  and 
greater  relative  abundance,  for  most 
species,  on  mature  and  old-growth 
stands  may  be  related  to  greater 
structural  complexity  in  older  forests 
(Franklin  and  Spies  1984,  Franklin  et 
al.  1981).  Older  forests  also  have  a 
narrower  and  more  stable  range  of 
moisture  and  temperature  than  pre- 
canopy  and  young  forests  (Bury 
1983,  Harris  1984).  Bury  (1983) 
sampled  amphibians  on  four  paired 


plots  in  coastal  redwood  forest,  each 
pair  consisting  of  a  logged  and  an 
old-growth  forest  stand.  He  attrib- 
uted the  lower  diversity  and  relative 
abundance  of  amphibians  on  the 
logged  sites  to  microclimatic  differ- 
ences. Bury  (1983)  also  found  higher 
numbers  of  amphibians  associated 
with  a  greater  volume  of  downed 
woody  material,  but  he  considered 
these  differences  in  cover  habitat  to 
be  of  secondary  importance.  Re- 
cently, Bury  and  Corn  (this  volume) 
found  that  coarse  woody  debris  is 
related  to  salamander  occurrence 
and  abundance  in  the  Oregon  and 
Washington  Cascades. 

We  believe  that  structural  com- 
plexity or  spatial  heterogeneity  (Pi- 


B 


X 


I 


MCUI  -  SE 


iMno       tuTuK      OLD  «n     as  not     OU)  MY 


Sim  TYP£ 


5  30 
!£! 

in 


HEW  «  X 


WW  -  SE 


TMNC      lunnc     OLD  «rr    out  Mac    ou  orr 
STAND  TYPE 


Figure  13.— Captures  per  person-hour  (A:TCS),  and  captures  per  1000  trap-nights  (B:PF),  of 
the  Del  Norte  salamander  (Plethodon  elongatus),  in  three  forest  age  and  three  moisture 
classes.  Data  are  from  the  inland  area,  from  1984-86  (\CS)  and  1984-85  (PF). 


X 


KM  4  Z 


UNftJLW 


route  lur.JM         OLD  WT       01*  MISIC        OLD  c 

SIAHO  TYPt 


8 

o  1:6 


X 


itMt  -  9 


IUT1JI         0U>  ITT       OLD  IC9C 

SUNO  TYPC 


Figure  14.— Captures  per  person-hour  (A:TCS),  and  captures  per  1000  trap-nights  (B:PF),  of 
the  rough-skinned  newt  (Taricha  granulosa),  in  three  forest  age  and  three  moisture  classes. 
Data  are  from  coastal  and  inland  areas  combined,  from  1984-86  (ICS)  and  1984-85  (PF). 


anka  1966)  plays  an  important  role  in 
promoting  the  addition  of  species 
and  numbers  of  individuals  in  older 
forests.  Downed  woody  material,  be- 
sides affording  cover,  creates  micro- 
climatic pockets  that  can  act  to  buffer 
the  moisture  and  temperature  fluc- 
tuations in  the  forest  at  large,  and  it 
provides  protection  from  predation 
as  well.  Maiorana  (1978)  reported 
that  space  (small  cavities  and  bur- 
rows) was  more  important  in  regu- 
lating relative  abundance  between 
two  sympatric  salamanders  (Aneides 
luguhris  and  Batrachoseps  attenuatus) 
than  competition  for  food  resources. 
Therefore,  more  salamander  species 
and  individuals  should  be  expected 
in  more  structurally  complex  habi- 
tats. In  fact,  both  microclimate  and 
cover  are  probably  interrelated,  ulti- 
mate factors  (Baker  1938)  determin- 
ing habitat  suitability  for  temperate 
forest  herpetofauna.  Both  are  clearly 
affected  by  forest  harvest  practices 
and  probably  jointly  account  for 
most  of  the  differences  in  diversity 
and  abundance  observed  in  the  her- 
petofauna between  young,  mature, 
and  old-growth  forests  in  northwest- 
ern California  and  southwestern  Ore- 
gon. 

ACKNOWLEDGMENTS 

We  thank  the  members  of  the  field 
crews  of  the  Pacific  Southwest  Forest 
and  Range  Experiment  Station's  Tim- 
ber/Wildlife Research  Unit  for  their 
help  in  collecting  data;  James  A. 
Baldwin  and  Barry  R.  Noon  for  ad- 
vising on  statistical  methods;  C.  John 
Ralph,  R.  B.  Bury,  and  M.  G.  Raphael 
for  their  reviews  of  the  manuscript; 
and  Dana  L.  Waters  for  his  help  with 
the  figures  and  tables  in  the  manu- 
script. 

LITERATURE  CITED 

Anderson,  James  D.  1968.  Rhyacotri- 
ton  and  R.  olympicus.  American 
Society  of  Ichthyologists  and  Her- 


453 


petologists.  Catalog  of  American 
Amphibians  and  Reptiles  68.1- 
68.2. 

Axelrod,  Daniel  1. 1976.  History  of 
coniferous  forests,  California  and 
Nevada.  University  of  California 
Publications  in  Botany  70:1-62. 

Baker,  J.R.  1938.  The  evolution  of 
breeding  seasons,  p.  161-177.  In 
Evolution,  essays  presented  to  E.S. 
Goodrich.  G.R.  de  Beer,  editor. 
Oxford  University  Press,  Oxford. 

Bennett,  Stephen  H.,  J.  Whiti^ield  Gib- 
bons, and  Jill  G.  Glansville.  1980. 
Terrestrial  activity,  abundance, 
and  diversity  of  amphibians  in  dif- 
ferent managed  forest  types. 
American  Midland  Naturalist 
103(2):412-416. 

Burton,  Thomas  M.,  and  Gene  E.  Lik- 
ens. 1975a.  Energy  flow  and  nutri- 
ent cycling  in  salamander  popula- 
tions in  the  Hubbard  Brook  Ex- 
perimental Forest,  New  Hamp- 
shire. Ecology  56:1068-1080. 

Burton,  Thomas  M.,  and  Gene  E.  Lik- 
ens. 1975b.  Salamander  popula- 
tions and  biomass  in  the  Hubbard 
Brook  Experimental  Forest,  New 
Hampshire.  Copeia  1975:541-546. 

Bury,  R.  Bruce.  1968.  The  distribution 
of  Ascaphus  truei  in  California. 
Herpetologica  24(l):39-46. 

Bury,  R.  Bruce.  1983.  Differences  in 
amphibian  populations  in  logged 
and  old-growth  redwood  forest. 
Northwest  Science  57(3):167-178. 

Bury,  R.  Bruce  and  Paul  S.  Corn. 

1987.  Evaluation  of  pitfall  trapping 
in  northwestern  forests:  trap  ar- 
rays with  drift  fences.  Journal  of 
Wildlife  Management  51:112-119. 

Bury,  R.  Bruce  and  Paul  S.  Corn. 

1988.  Responses  of  aquatic  and 
streamside  amphibians  to  timber 
harvest,  (in  press).  In  Streamside 
management:  riparian  wildlife  and 
forestry  interactions.  K.J.  Raedeke, 
editor.  College  of  Forestry  Re- 
sources, University  of  Washing- 
ton. 

Bury,  R.  Bruce,  Roger  A.  Lucken- 
bach,  and  Stephen  D.  Busack. 
1977.  Effects  of  off-road  vehicles 
on  vertebrates  in  the  California 


desert.  U.S.  Department  of  the  In- 
terior, Fish  and  Wildlife  Service, 
Report  8, 23  p. 

Bury,  R.  Bruce  and  Michael  Martin. 
1973.  Comparative  studies  on  the 
distribution  and  foods  of  Pletho- 
dontid  salamanders  in  the  red- 
wood region  of  Northern  Califor- 
nia. Journal  of  Herpetology 
7(4):331-335. 

Busack,  Stephen  D.,  and  R.  Bruce 
Bury.  1974.  Some  effects  of  off- 
road  vehicles  and  sheep  grazing 
on  lizard  populations  in  the 
Mojave  Desert.  Biological  Conser- 
vation 6(3):179-183. 

Corn,  Paul  S.,  and  R.  Bruce  Bury.  (In 
prep.).  Sampling  terrestrial  am- 
phibians and  reptiles.  In  Monitor- 
ing techniques  for  wildlife  in  Pa- 
cific Northwest  forests.  Andrew  B. 
Carey  and  Leonard  F.  Ruggiero, 
editors.  U.S.  Forest  Service,  Gen- 
eral Technical  Report. 

Enge,  Kevin  M.,  Wayne  R.  Marion, 
1986.  Effects  of  clearcutting  and 
site  preparation  on  herpetofauna 
of  a  north  Florida  flatwoods.  For- 
est Ecology  and  Management 
14:177-192. 

Feder,  Martin  E.  1983.  Integrating  the 
ecology  and  physiology  of  pletho- 
dontid  salamanders.  Herpetol- 
ogica 39:291-310. 

Franklin,  Jerry  F.,  Kermit  Cromack, 
Jr.,  William  Denison,  Arthur 
McKee,  Chris  Maser,  James  Sedell, 
Fred  Swanson,  and  Glen  Juday. 
1981.  Ecological  characteristics  of 
old-growth  Douglas-tir  forests. 
U.S.  Department  of  Agriculture, 
Forest  Service,  General  Technical 
Report  PNW-118, 48  p.  Pacitic 
Northwest  Forest  and  Range  Ex- 
periment Station,  Portiand,  Oreg. 

Franklin,  Jerry  F.,  F.  Hall,  W. 
Laudenslayer,  Chris  Maser,  J. 
Nunan,  J.  Poppino,  C.  John  Ralph, 
and  Thomas  Spies.  1986.  Interim 
definitions  of  old-growth  Douglas- 
fir  and  mixed  conifer  forests  in  the 
Pacific  Northwest  and  California. 
U.S.  Department  of  Agriculture, 
Forest  Service,  PNW  Research 
Note  PNW-447,  7  p. 


Franklin,  Jerry  F.  and  Thomas  Spies. 
1984.  Characteristics  of  old- 
growth  Douglas-fir  forests,  p.  328- 
334.  In  Proceedings,  Society  of 
American  Foresters  national  con- 
vention. [October  16-20, 1983], 
Portland,  Oregon. 

Harris,  Larry  D.  1984.  The  Frag- 
mented Forest.  211  p.  University 
of  Chicago  Press. 

Jones,  K.  Bruce.  1981.  Effects  of  graz- 
ing on  lizard  abundance  and  di- 
versity in  western  Arizona.  South- 
western Naturalist  26(2):107-115. 

Keselman,  H.J.  and  Joanne  C.  Rogan. 
1978.  A  comparison  of  the  modi- 
fied-Tukey  and  Scheffe  methods  of 
multiple  comparisons  for  pairwise 
contrasts.  Journal  of  the  American 
Statistical  Association  73(361  ):47- 
52. 

Luckenbach,  Roger  A.  and  R.Bruce 
Bury.  1983.  Effects  of  off-road  ve- 
hicles on  the  biota  of  the  Al- 
godones  Dunes,  Imperial  County, 
California.  Journal  of  Applied 
Ecology  20:265-286. 

Lynch,  James  F.  1981.  Patterns  of  on- 
togenetic and  geographic  variation 
in  the  black  salamander.  Amides 
flavipunctatus  (Caudata:  Pletho- 
dontidae).  Smithsonian  Contribu- 
tions to  Zoology  324:1-53. 

Maiorana,  Virginia  C.  1974.  Studies 
in  the  behavioral  ecology  of  the 
plethodontid  salamander  Batra- 
choseps  attenuatus.  Ph.D.  disserta- 
tion. University  of  California, 
Berkeley. 

Maiorana,  Virginia  C.  1976a.  Size  and 
environmental  predictability  for 
salamanders.  Evolution  30(3):599- 
613. 

Maiorana,  Virginia  C.  1976b.  Preda- 
tion,  submergent  behavior,  and 
tropical  diversity.  Evolutionary 
Theory  1:157-177. 

Maiorana,  Virginia  C.  1978.  Differ- 
ences in  diet  as  an  epiphenom- 
enon:  space  regulates  salaman- 
ders. Canadian  Journal  of  Zoology 
56(5):1017-1025 

Meslow,  E.  Charles,  Chris  Maser, 
and  Jared  Verner.  1981.  Old- 
growth  forests  as  wildlife  habitat. 


454 


Transactions  of  the  North  Ameri- 
can Wildlife  and  Natural  Resource 
Conference  46:329-335. 

Milliken,  G.A.  and  D.E.  Johnson. 
1984.  Analysis  of  Messy  Data. 
Wadsworth  Inc.  Belmont,  Califor- 
nia. 473  p. 

Miller,  Rupert  G.,  Jr.  1981.  Simultane- 
ous Statistical  Inference,  2nd.  ed. 
Springer-Verlag,  New  York.  299  p. 


z-       Nussbaum,  Ronald  A.,  Edmund  D. 

Brodie,  Jr.,  and  Robert  M.  Storm, 
h-  1983.  Amphibians  and  Reptiles  of 

the  Pacific  Northwest.  University 
n,  of  Idaho  Press,  Moscow  Idaho.  332 

P- 

of       Ortega,  A.,  M.E.  Maury,  and  R.  Bar- 
se  bault.  1982.  Spatial  organization 

n  and  habitat  partitioning  in  a 


mountain  lizard  community  of 
Mexico.  Ecol.  Gen.  3(3):323-330. 

Pianka,  Eric  R.  1966.  Convexity,  des- 
ert lizards,  and  spatial  heterogene- 
ity. Ecology  47:1055-1059. 

Pough,  Harvey  F.  1980.  The  advan- 
tages of  ectothermy  for  tetrapods. 
American  Naturalist  115:92-112. 

Pough,  Harvey  F.  1983.  Amphibians 
m  and  reptiles  as  low-energy  sys- 

tems, p.  141-188.  In  Behavioral  En- 
ergetics the  Cost  of  Survival  in 
Vertebrates.  W.P.  Aspey  and  S.I. 
Lustick,  editors.,  Ohio  State  Univ. 
Press,  Columbus,  Ohio. 

Pough,  Harvey  F.,  Ellen  M.  Smith, 
Donald  H.  Rhodes,  and  Andres 
Collazo.  1987.  The  abundance  of 
salamanders  in  forest  stands  with 
different  histories  of  disturbance. 
]  Forest  Ecology  and  Management 

20:1-9. 

Raphael,  Martin  G.  1984.  Wildlife 
populations  in  relation  to  stand 
age  and  area  in  Douglas-fir  forests 
of  northwestern  California,  p.  259- 
274.  In  Fish  and  Wildlife  Relation- 
ships in  Old-Growth  Forests.  [Jun- 
eau, Alaska,  April  12-15,  1982] 
W.R.  Meehan,  T.R.  Merrelly  Jr., 
and  T.A.  Hanley,  editors..  Pro- 
ceedings of  the  American  Institute 
of  Fishery  Research  Biologists. 
1      Raphael,  Martin  G.  1987.  Wildlife- 
'  tanoak  associations  in  Douglas-fir 

forests  of  northwestern  California. 


In  Multiple-use  management  of 
California's  hardwood  resources. 
Plumb,  T.R.  and  N.H.  Pillsbury, 
editors..  General  Technical  Report. 
PSW-100.  Pacific  Southwest  Forest 
and  Range  Experiment  Stations, 
U.S.  Department  of  Agriculture, 
Forest  Service.  Berkeley,  Califor- 
nia. 

Raphael,  Martin  G.  In  press. 

Douglas-fir  (DFR).  In  A  Guide  to 
Wildlife  Habitats  of  California.  K. 
Mayer  and  W.F.  Laudenslayer,  Jr., 
editors. 

Raphael,  Martin  G.,  and  Bruce  G. 
Marcot.  1986.  Validation  of  a 
Wildlife-Habitat-Relationships 
model:  vertebrates  in  a  Douglas-fir 
sere.  p.  129-138.  In  Wildlife  2000: 
modeling  habitat  relationships  of 
terrestrial  vertebrates.  Jared 
Vemer,  Michael  L.  Morrison,  and 
C.  John  Ralph,  editors.  University 
of  Wisconsin  Press,  Madison. 

Ruggiero,  Lenard  F.  and  Andrew  B. 
Carey.  1984.  A  programmatic  ap- 
proach to  the  old-growth  forest: 
wildlife  relationships,  p.  340-345. 
In  Proceedings  of  1983  convention 
of  the  Society  of  American  Forest- 
ers, Bethesda,  MD. 

Sawyer,  John  O.  and  Dale  A.  Thorn- 
burgh.  1977.  Mixed  evergreen  for- 
est, p.  59-382.  In  Terrestrial  Vege- 
tation of  California.  M.G.  Barbour 
and  J.  Major,  editors.  J.  Wiley  and 
Sons,  New  York. 

Simberloff,  Daniel  1987.  The  spotted 
owl  fracas:  mixing  academic,  ap- 
plied and  political  ecology.  Ecol- 
ogy 68(4):766-779. 

Sneath,  P.H.A.  and  Robert  R.  Sokal. 
1973.  Numerical  Taxonomy.  W.H. 
Freeman  Co.  San  Francisco,  Calif. 
573  p. 

Spies,  Thomas  A.,  Jerry  F.  Franklin, 
and  Ted  B.  Thomas.  Coarse 
woody  debris  in  Douglas-fir  for- 
ests of  western  Oregon  and  Wash- 
ington. Ecology.  (In  Press.) 

Stebbins,  Robert  C.  1954.  Natural  his- 
tory of  the  salamanders  of  the  ple- 
thodontid  genus  Ensatina.  Univer- 
sity of  California  Publications  in 
Zoolology  54(2):46-123. 


Stebbins,  Robert  C.  1985.  A  Field 
Guide  to  Western  Reptiles  and 
Amphibians.  336  p.  Houghton 
Mifflin  Company,  Boston,  MA. 

Welsh,  Hartwell  H.,  Jr.  1987.  Moni- 
toring herpetofauna  in  woodland 
habitats  of  Northwestern  Califor- 
nia and  southwestern  Oregon:  a 
comprehensive  strategy,  p.  203- 
213.  In  Multiple-use  management 
of  California's  hardwood  re- 
sources. Timothy  R.  Plumb  and 
Norman  H.  Pillsbury,  editors. 
USDA  Forest  Service  General 
Technical  Report  PSW-100.  Pacific 
Southwest  Forest  and  Range  Ex- 
periment Station,  Berkeley,  CA. 

Welsh,  Hartwell  H.,  Jr.  (In  Prep.).  Re- 
lictual  amphibians  and  old-growth 
forests. 

Zar,  Jerrold  H.  1984.  BiostaHstical 
Analysis.  Prentice-Hall,  Inc.  Engle- 
wood  Cliffs,  N.J.  718  p. 


455 


Table  Al  .-Comparisons  of  mean  capture  frequencies  of  herpetofauna,  captured  by  two  sampling  methods  time- 
clo^aZ,  Nand  ^^eT'  '"^^  c^sas":nd^|rin 


Spedes  richness 
TCSandPF 
data  combined 

multiple 
comparisons: 
All  herpetofauna 
TCS  data 


multiple 
comparisons: 

PF  data 


multiple 

comparisons: 
Reptiles 

TCS  data 

multiple 

comparisons: 
All  lizards 

TCS  data 

multiple 

comparisons: 
Elgaria 
coeruleus 

TCS  data 

multiple 

comparisons: 
All  snakes 

TCS  data 

multiple 

comparisons: 
Amphibia 

TCS  data 


multiple 
comparisons: 

PF  data 


multiple 
comparisons: 
All  frogs 
TCS  data 


multiple 
comparisons: 
PF  data 
multiple 
comparisons: 


Inland  stands      Coastal  stands       Comparisons  between  coastal  and  Inland  stands    Stands  combined 


Moisture      Age  Moisture 


All  All  Old 

je       stands       old  wet 


F=5.20  C>l 

.025>P>.01  t=2.13 
P=.04 

mature>young 
q=4.160,  P<.05 

F=10.75         *  + 
.0025>P>.001 


old>young 
q=6.035,  P<.05 
mature>young 
q=4.558,  P<.05 


l>C 
t=2.44 
P=.023 


F=9.74 
.005>P>.0025 

old>young,  q=5.091 
mature>young,  q=4.004 
P<.05 

l>C 
t=2.43 
P=.024 


mesiowet 
t=4.34 
P=.023 


old>young 
t=3.97,  P<.05 


Old 


mesic      Mature     Young     Moisture  Age 


C>l 


C>l 
t=3.70 
P=.01 


+ 

t=5.51 
P<.0001 


C>l 
1=5.061 
P<.0001 


F=3.87 


.05>P>.02 


mesiodry  old>young 
t=3.22  q=3.860 
P<.05  P<.05 


F=4.27 
.025>P>.01 

mesiodry  old>young 
t=3.49  q=3.998 
P<.05  P<.05 


(Continued)  ^ 


456 


Table  A 1 . — (continued). 


Inland  stands      Coastal  stands       Comparisons  between  coastal  and  Inland  stands    Stands  combined 


All  All 

Moisture      Age      Moisture      Age       stands  old 


Old  Old 

wet       mesic      Mature     Young     Moisture  Age 


Hyla  regilla 
TCS  data 

multiple 

comparisons: 
Rana  boylii 

PF  data 

multiple 

comparisons; 
All  Salamanders 

TCS  data 


multiple 
comparisons: 

PF  data 


multiple 

comparisons: 
Rhyacotriton  olympicus 

TCS  data 

multiple 

comparisons: 
Aneides  ferreus 
TCS  data 

multiple 
comparisons: 

Aneides  flavipunctatus 
TCS  data  ^ 

multiple 
comparisons: 
PF  data 
multiple 
comparisons: 
Batrachoseps  attenuatus 
TCS  data  4 

multiple 

comparisons: 

PF  data  * 


multiple 
comparisons: 
Ensatina  eschscholtzii 
TCS  data 

multiple 
comparisons: 

PF  data 


meslowet 
t=4.34,P=.023 

okJ>young 
t=3.97,P<.05 


old>young 
t=3.42,  P<.05 


F-8.67 
.005>P>.0025 

old>young,  q=5.601 
mature>young,  q=3.706 
P<.05 

l>C 
t=2.59 
P=.017 


C>l  + 
t=2.12,P=.04 


F=5.82  * 
.05>P>.025 
old>young,  q=3.836,  P<.05 
mature>young,  q=4.108,  P<.05 
mesiowet   F=10.94  * 
t=3.62  .0025>P>.001 
P=.022 

old>young,q=5.799,  P<.05 
mature>young,  q=5.188,  P<.05 

l>C 
t=2.13.P=.04 


l>C 

t=3.34,P=.003 


C>l 
t=5.091 
P<.0001 


F-4.26 
.025>P>.01 

meslodry  old>young 
t=3.42  q=3.970 
P<.05  P<.05 


old>young 
t=2.57,P<.05 


+  F=4.45 
.05>P>.025 


meslodry 
q=3.903,  P<.05 


4  4 


F=3.72 
.05>P>.025 

old>young 
q=3.60,P<.05 


(Continued)  J 


457 


r 


Table  A 1 .  ~  (continued). 


Inland  stands 

Coastal  stands 

Comparisons  between  coastal  and  Inland  stands 

Stands  connbined 

Moisture  Age 

Moisture  Age 

All 
stands 

All          Old  Old 

old        wet       mesic      Mature  Young 

moisiure  Age 

Plethodon  elongatus 

ICS  and  PF 

4 4 

■    .4  ■ 

4                         4  4 

4  4 

4  4 

multiple 

comparisons: 

•  ♦' 

Taricha  granulosa 

ICS  data 

•  .  ■. 

+  + 

multiple 

comparisons: 

■   *  ■ 

»                 ■    ♦  . 

PFdata 

* 

+                     +  + 

multiple 

comparisons: 

* 

♦      ■    ■  ■ 

'  *  =  not  significont  at  P<.05. 
'+  =  not  significant  at  P  <  .01. 

^Capture  frequency  in  designated  category  was  too  low  for  analysis. 
^Species  absent  from  inland  or  coastal  area. 
^Absent  from  young  stands. 


<rU.S.  GOVERNMENT  PRINTING  OFFICE:  1989-674-391/5024 


458 


1022423217