30.10.2012 Views

Proceedings of the Seventh Mountain Lion Workshop

Proceedings of the Seventh Mountain Lion Workshop

Proceedings of the Seventh Mountain Lion Workshop

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PROCEEDINGS OF THE<br />

SEVENTH MOUNTAIN LION WORKSHOP<br />

15-17 MAY 2003 • THE VIRGINIAN LODGE • JACKSON, WYOMING<br />

Editors:<br />

Scott A. Becker<br />

Daniel D. Bjornlie<br />

Fred G. Lindzey<br />

David S. Moody<br />

Organizing Committee<br />

Scott Becker Ron Grogan<br />

Dan Bjornlie Fred Lindzey<br />

Tom Easterly Dave Moody<br />

Sponsored By:<br />

The Wyoming Chapter <strong>of</strong> <strong>the</strong> Wildlife Society<br />

Wyoming Game and Fish Department<br />

Wyoming Cooperative Fish and Wildlife Research Unit<br />

© 2003<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, Wyoming 82520


Suggested Citation Formats<br />

Entire Volume:<br />

Becker, S.A., D.D. Bjornlie, F.G. Lindzey, and D.S. Moody. eds. 2003. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong><br />

<strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>. Lander, Wyoming.<br />

For individual papers:<br />

Author’s name(s). 2003. Title <strong>of</strong> Paper. Pages 00-00 in S.A. Becker, D.D. Bjornlie, F.G.<br />

Lindzey, and D.S. Moody, eds. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>.<br />

Lander, Wyoming.<br />

Purchasing Additional Copies <strong>of</strong> <strong>the</strong> <strong>Proceedings</strong><br />

Please send a check made out to “Wyoming Chapter, TWS” for <strong>the</strong> amount <strong>of</strong> fifteen (15) US<br />

dollars to Tim Thomas, Wyoming Game and Fish Department, PO Box 6249, Sheridan, WY<br />

82801, USA; phone: (307) 672-7418; email: Tim.Thomas@wgf.state.wy.us. Information on<br />

different purchasing options may also be made through Tim.


TABLE OF CONTENTS<br />

Preface....................................................................................................................................................................vii<br />

In Memory<br />

Ian Ross ............................................................................................................................................................. viii<br />

<strong>Mountain</strong> <strong>Lion</strong> Status Reports<br />

Session Chair: Dave Moody, Wyoming Game and Fish Department<br />

STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA<br />

Brian F. Wakeling ....................................................................................................................................................1<br />

CALIFORNIA MOUNTAIN LION STATUS REPORT<br />

Doug Updike............................................................................................................................................................6<br />

COLORADO MOUNTAIN LION STATUS REPORT<br />

Jerry Apker.............................................................................................................................................................14<br />

FLORIDA FISH AND WILDLIFE CONSERVATION COMMISSION STATUS REPORT<br />

Mark Lotz and E. Darrell Land ..............................................................................................................................18<br />

IDAHO MOUNTAIN LION STATUS REPORT<br />

Steve Nadeau..........................................................................................................................................................25<br />

MONTANA MOUNTAIN LION STATUS REPORT<br />

Rich DeSimone and Rose Jaffe..............................................................................................................................29<br />

NEVADA MOUNTAIN LION STATUS REPORT<br />

Russell Woolstenhulme..........................................................................................................................................31<br />

NEW MEXICO MOUNTAIN LION STATUS REPORT<br />

Rick Winslow.........................................................................................................................................................39<br />

STATE OF SOUTH DAKOTA MOUNTAIN LION STATUS REPORT<br />

Mike Kintigh ..........................................................................................................................................................43<br />

MOUNTAIN LION STATUS REPORT FOR TEXAS<br />

John Young ............................................................................................................................................................49<br />

UTAH MOUNTAIN LION STATUS REPORT<br />

Craig R. McLaughlin .............................................................................................................................................51<br />

WASHINGTON COUGAR STATUS REPORT<br />

Richard A. Beausoleil, Donald A. Martorello, and Rocky D. Spencer..................................................................60<br />

WYOMING MOUNTAIN LION STATUS REPORT<br />

Scott A. Becker, Daniel D. Bjornlie, and David S. Moody....................................................................................64<br />

CRYPTIC COUGARS – PERSPECTIVES ON THE PUMA IN THE EASTERN, MIDWESTERN, AND GREAT PLAINS<br />

REGIONS OF NORTH AMERICA<br />

Jay W. Tischendorf ................................................................................................................................................71<br />

MOUNTAIN LION STATUS REPORT: BRITISH COLUMBIA – Abstract<br />

Matt Austin ............................................................................................................................................................87<br />

IMPROVING OUR UNDERSTANDING OF MOUNTAIN LION MANAGEMENT TRENDS: THE VALUE OF CONSISTENT<br />

MULTI-STATE RECORD KEEPING - Abstract<br />

Christopher M. Papouchis and Lynn Michelle Cullens .........................................................................................88<br />

<strong>Mountain</strong> <strong>Lion</strong> Interactions with Humans and Livestock<br />

Session Chair: Kenneth Logan, Colorado Division <strong>of</strong> Wildlife<br />

iii


LESSENING THE IMPACT OF A PUMA ATTACK ON A HUMAN<br />

E. Lee Fitzhugh, Sabine Schmid-Holmes, Marc W. Kenyon, and Kathy Etling ...................................................89<br />

A CONCEPTUAL MODEL AND APPRAISAL OF EXISTING RESEARCH RELATED TO INTERACTIONS BETWEEN<br />

HUMANS AND PUMAS – Abstract<br />

David J. Mattson, Jan V. Hart, Paul Beier, and Jesse Millen-Johnson ................................................................104<br />

RELATIONSHIPS BETWEEN LAND TENURE SYSTEM, MOUNTAIN LION PROTECTION STATUS, AND LIVESTOCK<br />

DEPREDATION RATE – Abstract<br />

Marcelo Mazzolli .................................................................................................................................................105<br />

MOUNTAIN LION MOVEMENTS AND PERSISTENCE IN A FRAGMENTED, URBAN LANDSCAPE IN SOUTHERN<br />

CALIFORNIA – Abstract<br />

Seth P.D. Riley, Raymond M. Sauvajot, and Eric C. York..................................................................................106<br />

PUMA RESPONSES TO CLOSE ENCOUNTERS WITH RESEARCHERS – Abstract<br />

Linda L. Sweanor, Kenneth A. Logan, and Maurice G. Hornocker ....................................................................107<br />

<strong>Mountain</strong> <strong>Lion</strong> Genetics and Disease<br />

Session Chair: Deedra Hawk, Wyoming Game and Fish Department<br />

PRELIMINARY RESULTS OF FLORIDA PANTHER GENETIC ANALYSES – Abstract<br />

Warren E. Johnson, Darrell Land, Jan Mortenson, Melody Roelke-Parker, and Stephen J. O’Brien..................108<br />

GENETIC STRUCTURE OF COUGAR POPULATIONS ACROSS THE WYOMING BASIN: METAPOPULATION OR<br />

MEGAPOPULATION – Abstract<br />

Chuck R. Anderson, Jr., Fred G. Lindzey, and Dave B. McDonald ....................................................................109<br />

ECOLOGICAL SIGNIFICANCE AND EVOLUTION OF A COMMON COUGAR RETROVIRUS – Abstract<br />

Roman Biek and Mary Poss.................................................................................................................................110<br />

<strong>Mountain</strong> <strong>Lion</strong> Ecology<br />

Session Chair: Fred Lindzey, Wyoming Cooperative Fish and Wildlife Research Unit<br />

CHARACTERISTICS OF MOUNTAIN LION BED, CACHE AND KILL SITES IN NORTHEASTERN OREGON<br />

James J. Akenson, M. Cathy Nowak, Mark G. Henjum, and Gary W. Witmer...................................................111<br />

IMPACT OF EDGE HABITAT ON HOME RANGE SIZE IN PUMAS – Abstract<br />

John W. Laundré and Lucina Hernández.............................................................................................................119<br />

EFFECT OF ROADS ON HABITAT USE BY COUGARS – Abstract<br />

Dorothy M. Fecske, Jonathan A. Jenks, Frederick G. Lindzey, and Steven L. Griffin........................................120<br />

ECOLOGY OF SYMPATRIC PUMAS AND JAGUARS IN NORTHWESTERN MEXICO – Abstract<br />

Carlos A. Lopez Gonzalez and Samia E. Carrillo Percastegui.............................................................................121<br />

COUGAR ECOLOGY AND COUGAR-WOLF INTERACTIONS IN YELLOWSTONE NATIONAL PARK: A GUILD<br />

APPROACH TO LARGE CARNIVORE CONSERVATION – Abstract<br />

Toni K. Ruth, Polly C. Buotte, Howard B. Quigley, and Maurice G. Hornocker................................................122<br />

EVALUATION OF HABITAT FACTORS THAT AFFECT THE ABUNDANCE OF PUMAS IN THE CHIHUAHUAN<br />

DESERT – Abstract<br />

Joel Loredo Salazar, Lucina Hernández, and John W. Laundré ..........................................................................123<br />

<strong>Mountain</strong> <strong>Lion</strong>/Prey Dynamics<br />

Session Chair: Steve Cain, Grand Teton National Park<br />

ARE PUMAS OPPORTUNISTIC SCAVENGERS? – Abstract<br />

Jim W. Bauer, Kenneth A. Logan, Linda L. Sweanor, and Walter M. Boyce .....................................................124<br />

COUGAR-INDUCED INDIRECT EFFECTS: DOES THE RISK OF PREDATION INFLUENCE UNGULATE FORAGING<br />

BEHAVIOR ON THE NATIONAL BISON RANGE? – Abstract<br />

David M. Choate, Gary E. Belovsky, and Michael L. Wolfe ..............................................................................125<br />

iv


COUGAR PREDATION ON PREY IN YELLOWSTONE NATIONAL PARK: A PRELIMINARY COMPARISON PRE- AND<br />

POST-WOLF REESTABLISHMENT – Abstract<br />

Toni K. Ruth, Polly C. Buotte, Kerry M. Murphy, and Maurice G. Hornocker ..................................................126<br />

FOUR DECADES OF COUGAR-UNGULATE DYNAMICS IN THE CENTRAL IDAHO WILDERNESS – Abstract<br />

Holly A. Akenson, James J. Akenson, Howard B. Quigley, and Maurice G. Hornocker....................................127<br />

COUGAR TOTAL PREDATION RESPONSE TO DIFFERING PREY DENSITIES: A PROPOSED EXPERIMENT TO TEST<br />

THE APPARENT COMPETITION HYPOTHESIS – Abstract<br />

Hugh Robinson, Robert Wielgus, Hilary Cruickshank, and Ca<strong>the</strong>rine Lambert .................................................128<br />

<strong>Mountain</strong> <strong>Lion</strong> Population Monitoring and Management<br />

Session Chair: Kerry Murphy, Yellowstone National Park<br />

CHARACTERISTICS OF COUGAR HARVEST WITH AND WITHOUT THE USE OF DOGS<br />

Donald A. Martorello and Richard A. Beausoleil................................................................................................129<br />

RESPONSE BY THREE LARGE CARNIVORES TO RECREATIONAL BIG GAME HUNTING ALONG THE YELLOWSTONE<br />

NATIONAL PARK AND ABSAROKA-BEARTOOTH WILDERNESS BOUNDARY – Presentation Only<br />

Howard B. Quigley, Toni K. Ruth, Douglas W. Smith, Mark A. Haroldson, Polly C. Buotte, Charles C.<br />

Schwartz, Steve Cherry, Kerry M. Murphy, Dan Tyers, and Kevin Frey<br />

DEFINING AND DELINEATING DE FACTO REFUGIA: A PRELIMINARY ANALYSIS OF THE SPATIAL DISTRIBUTION<br />

OF COUGAR HARVEST IN UTAH AND IMPLICATIONS FOR CONSERVATION – Abstract<br />

David C. Stoner and Michael L. Wolfe................................................................................................................136<br />

MONITORING CHANGES IN COUGAR SEX/AGE STRUCTURE WITH CHANGES IN ABUNDANCE AS AN INDEX TO<br />

POPULATION TREND – Abstract<br />

Chuck R. Anderson, Jr. and Fred G. Lindzey ......................................................................................................137<br />

MANAGEMENT OF COUGARS (Puma concolor) IN THE WESTERN UNITED STATES – Abstract<br />

Deanna Dawn, Michael Kutilek, Rich Hopkins, Sulehka Anand, and Steve Torres ...........................................138<br />

DYNAMICS AND VIABILITY OF A COUGAR POPULATION IN THE PACIFIC NORTHWEST – Abstract<br />

Ca<strong>the</strong>rine Lambert, Robert B. Wielgus, Hugh S. Robinson, Donald D. Katnik, Hilary Cruickshank, and<br />

Ross Clarke ..........................................................................................................................................................139<br />

PROJECT CAT (COUGARS AND TEACHING): INTEGRATING SCIENCE, SCHOOLS AND COMMUNITY IN<br />

DEVELOPMENT PLANNING – Abstract<br />

Gary M. Koehler and Evelyn Nelson...................................................................................................................140<br />

MONITORING MOUNTAIN LIONS IN THE TUCSON MOUNTAIN DISTRICT OF SAGUARO NATIONAL PARK,<br />

ARIZONA, USING NONINVASIVE TECHNIQUES – Abstract<br />

Lisa Haynes, Don Swann, and Melanie Culver ...................................................................................................141<br />

ESTIMATING COUGAR ABUNDANCE USING PROBABILITY SAMPLING: AN EVALUATION OF TRANSECT VERSUS<br />

BLOCK DESIGN – Abstract<br />

Chuck R. Anderson, Jr., Fred G. Lindzey, and Nate Nibbelink...........................................................................142<br />

EVALUATING MOUNTAIN LION MONITORING TECHNIQUES IN THE GARNET MOUNTAINS OF WEST CENTRAL<br />

MONTANA – Abstract<br />

Rich DeSimone ....................................................................................................................................................143<br />

PRESENCE AND MOVEMENTS OF LACTATING AND MATERNAL FEMALE COUGARS: IMPLICATIONS FOR STATE<br />

HUNTING REGULATIONS – Abstract<br />

Toni K. Ruth, Kerry M. Murphy, and Polly C. Buotte ........................................................................................144<br />

<strong>Mountain</strong> <strong>Lion</strong> Conservation<br />

Session Chair: Christopher Papouchis, <strong>Mountain</strong> <strong>Lion</strong> Foundation<br />

MYSTERY, MYTH AND LEGEND: THE POLITICS OF COUGAR MANAGEMENT IN THE NEW MILLENNIUM –<br />

Abstract<br />

Rick A. Hopkins...................................................................................................................................................145<br />

v


RECONCILING SCIENCE AND POLITICS IN PUMA MANAGEMENT IN THE WEST: NEW MEXICO AS A TEMPLATE<br />

– Abstract<br />

Kenneth A. Logan, Linda L. Sweanor, and Maurice G. Hornocker ....................................................................146<br />

COMMUNITY-BASED CONSERVATION OF MOUNTAIN LIONS – Abstract<br />

Lynn Michelle Cullens and Christopher Papouchis.............................................................................................147<br />

PUMA MANAGEMENT IN WESTERN NORTH AMERICA: A 100-YEAR RETROSPECTIVE – Abstract<br />

Steven Torres, Hea<strong>the</strong>r Keough, and Deanna Dawn............................................................................................148<br />

USING COUGARS TO DESIGN A WILDERNESS NETWORK IN CALIFORNIA’S SOUTH COAST ECOREGION – Abstract<br />

Paul Beier and Kristeen Penrod ...........................................................................................................................149<br />

MOUNTAIN LIONS AND BIGHORN SHEEP: FACING THE CHALLENGES – Abstract<br />

Christopher M. Papouchis and John D. Wehausen ..............................................................................................150<br />

POSTER PRESENTATIONS<br />

Session Chair: Scott Becker, Wyoming Game and Fish Department<br />

FACTORS AFFECTING DISPERSAL IN YOUNG MALE PUMAS<br />

John W. Laundré and Lucina Hernández.............................................................................................................151<br />

COUGAR EXPLOITATION LEVELS AND LANDSCAPE CONFIGURATIONS: IMPLICATIONS FOR DEMOGRAPHIC<br />

STRUCTURE AND METAPOPULATION DYNAMICS – Abstract<br />

David C. Stoner and Michael L. Wolfe................................................................................................................161<br />

ASSESSING GPS RADIOTELEMETRY RELIABILITY IN COUGAR HABITAT – Abstract<br />

Trish Griswold, James Briggs, Gary Koehler, and Students at Cle Elum-Roslyn School District ......................162<br />

USING GPS COLLARS TO DETERMINE COUGAR KILL RATES, ESTIMATE HOME RANGES, AND EXAMINE<br />

COUGAR-COUGAR INTERACTIONS –Abstract<br />

Polly C. Buotte and Toni K. Ruth ........................................................................................................................163<br />

FUNCTIONAL RESPONSE OF COUGARS AND PREY AVAILABILITY IN NORTHEASTERN WASHINGTON – Abstract<br />

Hilary S. Cruickshank, Hugh S. Robinson, Ca<strong>the</strong>rine Lambert, Robert B. Wielgus ...........................................164<br />

WHAT DOES TEN YEARS (1993-2002) OF MOUNTAIN LION OBSERVATION DATA REVEAL ABOUT MOUNTAIN<br />

LION-HUMAN INTERACTIONS WITHIN REDWOOD NATIONAL AND STATE PARKS – Abstract<br />

Gregory W. Holm ................................................................................................................................................165<br />

DEPREDATION TRENDS IN CALIFORNIA – Abstract<br />

Sarah Reed, Christopher M. Papouchis, and Lynn Michelle Cullens ..................................................................166<br />

THE DISTRIBUTION OF PERCEIVED ENCOUNTERS WITH NON-NATIVE CATS IN SOUTH AND WEST WALES, UK:<br />

RELATIONSHIP TO MODELED HABITAT SUITABILITY – Abstract<br />

A.B. Smith, F.E. Street Perrott, and T. Hooper....................................................................................................167<br />

PUMA ACTIVITY AND MOVEMENTS IN A HUMAN-DOMINATED LANDSCAPE: CUYAMACA RANCHO STATE<br />

PARK AND ADJACENT LANDS IN SOUTHERN CALIFORNIA – Abstract<br />

Linda L. Sweanor, Kenneth A. Logan, Jim W. Bauer, and Walter M. Boyce .....................................................168<br />

MODELING OFFSPRING SEX RATIOS AND GROWTH OF COUGARS – Abstract<br />

Diana M. Ghikas, Martin Jalkotzy, Ian Ross, Ralph Schmidt, and Shane A. Richards .......................................169<br />

MOUNTAIN LION SURVEY TECHNIQUES IN NORTHERN IDAHO: A THREE-FOLD APPROACH – Abstract<br />

Craig G. White, Peter Zager, and Lisette Waits...................................................................................................170<br />

MOUNTAIN LIONS IN SOUTH DAKOTA: RESULTS OF A 2002 PUBLIC OPINION SURVEY – Abstract<br />

Larry M. Gigliotti, Dorothy M. Fecske, and Jonathan A. Jenks ..........................................................................171<br />

CRITICAL COUGAR CROSSING AND BAY AREA REGIONAL PLANNING – Abstract<br />

Michele Korpos....................................................................................................................................................172<br />

List <strong>of</strong> Participants............................................................................................................................................173<br />

vi


PREFACE<br />

vii<br />

PREFACE<br />

The Wyoming Game and Fish Department took great pride in hosting <strong>the</strong> <strong>Seventh</strong><br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, which was held in conjunction with <strong>the</strong> Thirty-Ninth North American<br />

Moose Conference and <strong>Workshop</strong> and <strong>the</strong> Fifth Western States and Provinces Deer and Elk<br />

<strong>Workshop</strong>. More than 190 people attended <strong>the</strong> mountain lion workshop representing 27 states, 3<br />

Canadian provinces, Mexico, Brazil, and <strong>the</strong> United Kingdom. Numerous state and federal<br />

agencies, tribal nations, private organizations, academia, and members <strong>of</strong> <strong>the</strong> general public were<br />

represented which attest to <strong>the</strong> varied and growing interest in mountain lions throughout North<br />

and South America.<br />

This workshop would not have been a success without <strong>the</strong> aid and cooperation <strong>of</strong> <strong>the</strong><br />

contributors and participants. Financial support, equipment, and manpower provided by <strong>the</strong><br />

Wyoming Chapter <strong>of</strong> <strong>the</strong> Wildlife Society, <strong>the</strong> Wyoming Game and Fish Department, and <strong>the</strong><br />

Wyoming Cooperative Fish and Wildlife Research Unit at <strong>the</strong> University <strong>of</strong> Wyoming made this<br />

workshop possible. A special thanks goes to <strong>the</strong> members <strong>of</strong> <strong>the</strong> organizing committee for <strong>the</strong>ir<br />

aid with all aspects <strong>of</strong> pre- and post-workshop activities, to <strong>the</strong> session chairs for keeping <strong>the</strong><br />

workshop moving in a timely fashion, and to <strong>the</strong> invited speakers who gave thoughtful insight<br />

into past, present, and future mountain lion management practices and research techniques.<br />

Many thanks to <strong>the</strong> Western Association <strong>of</strong> Fish and Wildlife Agencies (WAFWA) for<br />

sanctioning <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>; from this point forward, all mountain lion<br />

workshops will be sanctioned by WAFWA.<br />

Finally, we would like to thank all <strong>the</strong> presenters in <strong>the</strong> oral and poster sessions for <strong>the</strong><br />

depth <strong>of</strong> <strong>the</strong>ir research and <strong>the</strong> quality <strong>of</strong> <strong>the</strong>ir presentations. As a result <strong>of</strong> <strong>the</strong> efforts you all put<br />

forth, a standard has been set for presentations at future mountain lion workshops. Keep up <strong>the</strong><br />

great work!<br />

Scott Becker and Dave Moody<br />

<strong>Workshop</strong> Co-Chairs


IN MEMORY<br />

P. Ian Ross<br />

Born December 16, 1958 in Goderich, Ontario.<br />

Died June 29, 2003, age 44, near Nanyuki, Kenya.<br />

Ian was a true outdoorsman from <strong>the</strong> beginning, running a trapline while in high school in<br />

sou<strong>the</strong>rn Ontario. After graduating from <strong>the</strong> University <strong>of</strong> Guelph (1982), his first experiences<br />

with grizzly bears came in northwestern Alberta, where he studied <strong>the</strong> impacts <strong>of</strong> industrial<br />

development. It was <strong>the</strong> beginning <strong>of</strong> an illustrious 20-year career conducting research on large<br />

mammals in western Canada.<br />

He worked on cougars in southwestern Alberta from <strong>the</strong> early 1980’s until 1994. That project<br />

became one <strong>of</strong> <strong>the</strong> longest running research projects on Puma concolor in North America. The<br />

cougar project received national recognition on radio and television and Ian used that attention to<br />

foster a thoughtful and effective wildlife conservation message. He participated in <strong>the</strong> drafting<br />

<strong>of</strong> a management plan for cougars in Alberta as well as a conservation strategy for large<br />

carnivores in Canada. He was <strong>the</strong> senior author on 9 papers in peer-reviewed journals in<br />

addition to many o<strong>the</strong>r technical reports and popular articles.<br />

After <strong>the</strong> cougar project wrapped up, Ian conducted environmental impact studies in western and<br />

nor<strong>the</strong>rn Canada. He recently rewrote <strong>the</strong> grizzly bear status report for COSEWIC. He also<br />

worked tirelessly with The Wildlife Society-Alberta Chapter dealing with wildlife conservation<br />

issues. He served as President <strong>of</strong> <strong>the</strong> Chapter in 1997. Ian also continued to capture wildlife,<br />

including grizzly bears, for research projects, and in doing so assisted many graduate students<br />

with <strong>the</strong>ir research. He conducted his capture work using an exacting pr<strong>of</strong>essional approach<br />

while retaining an empathy for <strong>the</strong> wildlife he was pursuing. He cared for each individual and<br />

did his utmost to conduct captures in a humane manner.<br />

Ian was a committed and emotional friend and family man. Having no children <strong>of</strong> his own he<br />

was a hero to his young nieces, nephews and children <strong>of</strong> friends. He always remembered<br />

everyone’s birthdays. He hiked <strong>the</strong> foothills <strong>of</strong> <strong>the</strong> Rockies west <strong>of</strong> Calgary, as well as <strong>the</strong> U.S.<br />

desert southwest, <strong>the</strong> Canadian Arctic, Mexico and Africa. He loved to hunt and his dinner table<br />

was a testiment to his hunting prowess. His conservation ethic permeated all <strong>of</strong> his life. He did<br />

not consume needlessly and he encouraged all <strong>of</strong> us to do <strong>the</strong> same.<br />

In January 2003, Ian returned to field research when he joined Dr. Laurence Frank on <strong>the</strong><br />

Liakipia Predator Project, a project designed to find ways to allow for <strong>the</strong> coexistence <strong>of</strong> hyenas,<br />

lions, and leopards and people in <strong>the</strong> agricultural matrix that exists outside national parks in most<br />

<strong>of</strong> sou<strong>the</strong>rn Africa. Two days before his death he was on top <strong>of</strong> <strong>the</strong> world having collared his<br />

first leopard. On <strong>the</strong> evening he died Ian was tracking a radio-collared lion from a light aircraft.<br />

Searchers located its wreckage <strong>the</strong> next morning. Ian Ross died at <strong>the</strong> peak <strong>of</strong> his career, doing<br />

what he loved.<br />

By<br />

Martin Jalkotzy<br />

Arc Wildlife Services<br />

3527 - 35 Ave. SW<br />

Calgary, AB, T3E 1A2, Canada<br />

viii


ix<br />

IN MEMORY


STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA<br />

BRIAN F. WAKELING, Arizona Game and Fish Department, Game Branch, 2221 West<br />

Greenway Road, Phoenix, AZ 85023 USA<br />

Abstract: Arizona's mountain lion (Puma concolor) population numbers about 1,000-2,500 animals, and just over<br />

350 mountain lions were harvested through sport and depredation take in 5 <strong>of</strong> <strong>the</strong> last 6 years. Arizona bag limit is 1<br />

lion per person per year annually, except in a few units where multiple bag limits have been implemented; no<br />

multiple bag limit has been reached to date. Management for this big game animal is guided by strategic plan,<br />

species management guidelines, hunt guidelines, and a predation management policy. Management is currently<br />

under review by an internal team that is examining several predator species, including mountain lions. The internal<br />

review should be complete by <strong>the</strong> end <strong>of</strong> 2003. Public safety incident reports have increased substantially since<br />

1998.<br />

INTRODUCTION<br />

<strong>Mountain</strong> lion populations within<br />

Arizona remain robust and are currently<br />

estimated at 1,000-2,500 despite a prolonged<br />

drought throughout <strong>the</strong> southwestern United<br />

States. Portions <strong>of</strong> Arizona have received<br />

record low precipitation during 2002, and<br />

<strong>the</strong> decade <strong>of</strong> <strong>the</strong> 1990s was <strong>the</strong> driest on<br />

records for several portions <strong>of</strong> Arizona.<br />

Mule deer (Odocoileus hemionus)<br />

populations have declined, and in 2003 <strong>the</strong><br />

Arizona Game and Fish Commission<br />

authorized <strong>the</strong> lowest number <strong>of</strong> permits for<br />

deer hunting since <strong>the</strong> limited-draw permit<br />

system was established in Arizona.<br />

Figure 1. Arizona mountain lion harvest<br />

trends excluding tribal lands, 1984-2002.<br />

1<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

<strong>Mountain</strong> lion harvest has remained high, as<br />

annual statewide harvests have exceeded our<br />

strategic plan objectives (Arizona Game and<br />

Fish Department 2001) in 5 <strong>of</strong> <strong>the</strong> last 6<br />

years (Figure 1).<br />

<strong>Mountain</strong> lions are classified as big<br />

game by Arizona statute. Commission order<br />

has established <strong>the</strong> bag limit at 1 mountain<br />

lion per year, except in a few units.<br />

Successful hunters are required to report<br />

<strong>the</strong>ir harvest within 10 days and answer a<br />

series <strong>of</strong> standard questions. Beginning in<br />

July 2003, hunters will be asked to<br />

voluntarily provide a tooth, which may be<br />

used to estimate age through cementum<br />

annuli and determine gender using genetic<br />

techniques. The Department is investigating<br />

making <strong>the</strong> tooth submission mandatory.<br />

The management objectives for this species,<br />

as well as all big game species, are outlined<br />

in <strong>the</strong> agency strategic plan, Wildlife 2006<br />

(Arizona Game and Fish Department 2001)<br />

and species management guidelines. The<br />

strategic plan goals, objectives, and speciesspecific<br />

strategies for mountain lion<br />

management, that include:<br />

Objectives<br />

1. Maintain annual harvest at 250 to 300<br />

mountain lions (including depredation


2 STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA · Wakeling<br />

take).<br />

2. Provide recreational opportunity for<br />

3,000 to 6,000 hunters per year.<br />

3. Maintain existing occupied habitat and<br />

maintain <strong>the</strong> present range <strong>of</strong> mountain<br />

lions in Arizona.<br />

Species-Specific Strategies<br />

1. Maintain a complete database from all<br />

harvest sources, through a mandatory<br />

check-out system, including age, sex,<br />

kill location, etc. to index population<br />

trend.<br />

2. Conduct a hunter questionnaire<br />

biannually.<br />

3. Evaluate <strong>the</strong> management implications<br />

<strong>of</strong> population and relative density<br />

estimates.<br />

4. Implement hunt structures to increase<br />

and direct harvest emphasis toward<br />

areas with high lion populations, and<br />

where depredation complaints are<br />

substantiated, and evaluate <strong>the</strong><br />

effectiveness <strong>of</strong> <strong>the</strong>se efforts.<br />

5. Determine population numbers and<br />

characteristics on a hunt-area basis.<br />

6. Increase public awareness <strong>of</strong> mountain<br />

lions and <strong>the</strong>ir habits, to reduce<br />

conflicts with humans.<br />

7. Implement <strong>the</strong> Department’s Predation<br />

Management Policy.<br />

In addition, management direction is<br />

provided by species management guidelines<br />

and hunt guidelines. In October 2000, <strong>the</strong><br />

Arizona Game and Fish Commission<br />

approved <strong>the</strong> predation management policy<br />

that provides <strong>the</strong> agency guidance as to<br />

when and how to engage in predation<br />

management.<br />

<strong>Mountain</strong> lion management has changed<br />

as a direct result <strong>of</strong> biological investigations<br />

into predation effects. <strong>Mountain</strong> lion<br />

predation is being documented as a factor<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

that may be regulating prey populations<br />

(Ballard et al. 2001) in some areas <strong>of</strong><br />

Arizona, to include bighorn sheep (Ovis<br />

canadensis) (Kamler et al. 2002) and<br />

pronghorn (Antilocapra americana)<br />

(Ockenfels 1994a, b). These prey<br />

populations are at low levels, and reducing<br />

predator populations is likely to allow those<br />

prey populations to increase in number<br />

(Ballard et al. 2001). The standard bag limit<br />

for mountain lions has been altered in<br />

specific areas to allow for <strong>the</strong> harvest <strong>of</strong> 1<br />

mountain lion per day until a predetermined<br />

number <strong>of</strong> mountain lions are removed that<br />

equal about 50-75% <strong>of</strong> <strong>the</strong> estimated<br />

mountain lion population within that unit, at<br />

which time <strong>the</strong> bag limit reverts back to <strong>the</strong><br />

standard bag limit <strong>of</strong> 1 mountain lion per<br />

calendar year. The only exception to this is<br />

in <strong>the</strong> southwestern portion <strong>of</strong> <strong>the</strong> state<br />

where if even a single mountain lion is<br />

taken, <strong>the</strong> hunt area will be closed.<br />

Multiple bag limits were implemented in<br />

Units 13A and 13B in 1999, 16A South and<br />

18B South in 2001, 22 South in 1999, and<br />

Units 21 West, 28 South, and 37B North<br />

will be implemented this year. Research<br />

studies in Unit 22 South on bighorn sheep,<br />

that included investigations into nutrition,<br />

disease, and predation, indicate that <strong>the</strong><br />

multiple bag limit on mountain lions in that<br />

area, with increased effort by sportsmen to<br />

harvest mountain lions, seems to be<br />

positively influencing desert bighorn sheep<br />

recruitment and adult female survival. To<br />

implement a multiple bag limit on mountain<br />

lions, biologists must identify a prey species<br />

that has been reduced due to mountain lion<br />

predation (e.g., a declining population below<br />

management objectives) or a management<br />

action that is likely to be impacted by<br />

mountain lion predation (e.g., a planned<br />

translocation) to initiate and identify what<br />

management objectives must be met (e.g., 3<br />

years <strong>of</strong> 50:100 lamb:ewe ratios) before <strong>the</strong><br />

multiple bag limit is removed. Because this


is a relatively recent management approach<br />

in Arizona, refinements to implementation<br />

and new opportunities will undoubtedly<br />

develop. For instance, portions <strong>of</strong> Arizona<br />

have robust mountain lion populations that<br />

sustain large amounts <strong>of</strong> depredation<br />

removal (Cunningham et al. 1995) and may<br />

be able to provide recreational harvest at a<br />

higher level. These areas might provide<br />

opportunities to manage recreational harvest<br />

with multiple bag limits in <strong>the</strong> future, and<br />

attempt to transfer depredation take into<br />

recreational harvest.<br />

The Department has recently established<br />

an internal team to review management<br />

approaches for several predator species, to<br />

include mountain lions. This team will be<br />

reviewing social and biological issues and<br />

best management practices, and<br />

recommending possible changes to<br />

Arizona's management. This team will<br />

serve as an umbrella team for several<br />

subteams that will work on <strong>the</strong> biological<br />

basis for management, ga<strong>the</strong>r information on<br />

social acceptance, and conduct public<br />

outreach and education.<br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Mountain</strong> lions are distributed<br />

throughout most <strong>of</strong> Arizona, in varying<br />

densities (Figure 2). This distribution was<br />

reevaluated in 2002 by Department<br />

biologists and wildlife managers, and<br />

although subtle changes have been noted in<br />

<strong>the</strong> densities <strong>of</strong> lions, little change to <strong>the</strong><br />

distribution was identified. This map is still<br />

undergoing refinement and should be<br />

considered a draft. Additional information<br />

used by <strong>the</strong> Arizona Game and Fish<br />

Department in managing mountain lion<br />

population trends includes harvest,<br />

depredation reports, and age and gender<br />

from mandatory hunter reports.<br />

<strong>Mountain</strong> lion population estimates are<br />

based on density estimates developed from<br />

research studies, literature, and pr<strong>of</strong>essional<br />

experience within Arizona habitats. These<br />

STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA · Wakeling 3<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Figure 2. <strong>Mountain</strong> lion distribution and<br />

density estimates (draft) in Arizona<br />

excluding tribal lands, 2002.<br />

density estimates are reevaluated at<br />

infrequent intervals. Prior to 2002, <strong>the</strong> last<br />

reevaluation was conducted in 1993,<br />

although a few management units were<br />

reevaluated in 1998.<br />

HARVEST INFORMATION<br />

Licensed hunters may pursue mountain<br />

lions in Arizona if <strong>the</strong>y purchase a<br />

nonpermit tag prior to hunting. The annual<br />

bag limit is 1 lion, except for areas where a<br />

multiple bag limit is in place as discussed in<br />

<strong>the</strong> introduction. Strategic plan objectives<br />

for statewide harvests are based on historical<br />

harvest that removed about 10-15% <strong>of</strong> <strong>the</strong><br />

estimated statewide population. Recently,<br />

harvest combined with depredation removal<br />

has exceeded <strong>the</strong> strategic plan objective<br />

(Table 1). Phelps (2003) reported data on<br />

harvest prior to 1998. Still, statewide<br />

harvest is probably


4 STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA · Wakeling<br />

Table 1. Arizona mountain lion harvest summary excluding tribal lands, 1998-2002.<br />

Total Harvest<br />

Sport<br />

Harvest<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Gender <strong>of</strong> Sport<br />

Harvest<br />

Year<br />

Tags<br />

Issued Sport Depredation O<strong>the</strong>r b<br />

Using Dogs Male Female<br />

1998 6590<br />

1999 6885<br />

2000 7478<br />

2001 8109<br />

2002 7900 a<br />

289 52 1 192 150 136<br />

247 49 2 161 126 120<br />

276 53 0 193 133 141<br />

326 58 0 214 176 144<br />

263 50 5 175 154 115<br />

a<br />

2002 tags sold is preliminary.<br />

b<br />

Includes known kills o<strong>the</strong>r than sport or depredation (e.g., highway mortality, capture mortality, and illegal take).<br />

limits are established to take 50-75% <strong>of</strong> <strong>the</strong><br />

mountain lions that occupy an area when <strong>the</strong><br />

aforementioned criteria are met. To date,<br />

none <strong>of</strong> <strong>the</strong> multiple bag limits have been<br />

achieved.<br />

Arizona mountain lion seasons are<br />

currently open yearlong. About 7,900<br />

nonpermit tags were sold to hunters in 2002<br />

(Table 1). During 1998-2002, about 67%<br />

were taken with <strong>the</strong> aid <strong>of</strong> hounds, whereas<br />

24% were taken incidental to o<strong>the</strong>r<br />

activities. Currently, Arizona does not have<br />

a pursuit-only season.<br />

DEPREDATIONS AND HUMAN<br />

INTERACTIONS-CONFLICTS<br />

Complaints that come to <strong>the</strong> Arizona<br />

Game and Fish Department can take 1 <strong>of</strong> 2<br />

routes: nuisance wildlife or depredation.<br />

Nuisance complaints are dealt with through<br />

advice and education. Should a mountain<br />

lion pose a threat to public safety, <strong>the</strong><br />

Department will dispatch a wildlife manager<br />

to deal with <strong>the</strong> immediate situation,<br />

although we frequently contract with USDA<br />

APHIS Wildlife Services to conduct<br />

removal efforts. Between 1998 and 2002,<br />

312 public safety incidents have been<br />

reported involving mountain lions. The<br />

trend <strong>of</strong> <strong>the</strong>se reports over time has been<br />

steeply increasing (29 in 1998, 105 in 2002;<br />

Table 2). This increase in reports may be<br />

Table 2. Public incident reports that included<br />

mountain lions in Arizona excluding tribal<br />

lands, 1998-2002.<br />

Year<br />

1998<br />

1999<br />

2000<br />

2001<br />

2002<br />

Number <strong>of</strong> Incidents<br />

Reported<br />

29<br />

43<br />

46<br />

89<br />

105<br />

attributed to mountain lions pursuing prey<br />

near residential areas (which may be<br />

exacerbated by drought conditions),<br />

increasing residential development in<br />

mountain lion habitat, and <strong>the</strong> public's<br />

greater familiarity with <strong>the</strong> reporting<br />

process. During that time, few mountain<br />

lions (


objectives for mountain lions and ranges<br />

between 49 and 58 mountain lions annually<br />

(Table 1). The actual number <strong>of</strong> depredation<br />

incidents by year is difficult to accurately<br />

ascertain.<br />

ONGOING RESEARCH<br />

Arizona is fortunate to have a Research<br />

Branch within our Wildlife Management<br />

Division that may focus on issues<br />

surrounding wildlife management. In <strong>the</strong><br />

past, this has included many studies directly<br />

relating to mountain lions and that we<br />

currently base much <strong>of</strong> our mountain lion<br />

management. Currently, <strong>the</strong> Department<br />

does not have any ongoing research directly<br />

aimed at mountain lion management,<br />

although a study in Unit 22 that includes <strong>the</strong><br />

impacts <strong>of</strong> mountain lions on desert bighorn<br />

sheep is being completed. Studies by o<strong>the</strong>r<br />

organizations involving urban mountain<br />

lions are in <strong>the</strong> initial phases near Flagstaff<br />

and Tucson.<br />

LITERATURE CITED<br />

ARIZONA GAME AND FISH DEPARTMENT.<br />

2001. Wildlife 2006. Arizona Game<br />

and Fish Department, Phoenix.<br />

BALLARD, W.B., D.L. LUTZ, T.W. KEEGAN,<br />

L.H. CARPENTER, AND J.C. DEVOS, JR.<br />

2001. Deer-predator relationships: a<br />

review <strong>of</strong> recent North American studies<br />

with emphasis on mule and black-tailed<br />

STATUS OF MOUNTAIN LION POPULATIONS IN ARIZONA · Wakeling 5<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

deer. Wildlife Society Bulletin 29:99-<br />

115.<br />

CUNNINGHAM, S.C., L.A HAYNES, C.<br />

GUSTAVSON, AND D.D. HAYWOOD.<br />

1995. Evaluation <strong>of</strong> <strong>the</strong> interaction<br />

between mountain lions and cattle in <strong>the</strong><br />

Aravaipa-Klondyke area <strong>of</strong> sou<strong>the</strong>ast<br />

Arizona. Arizona Game and Fish<br />

Department Technical Report 17,<br />

Phoenix.<br />

KAMLER, J.F., R.M. LEE, J.C. DEVOS, JR.,<br />

W.B. BALLARD, AND H.A. WHITLAW.<br />

2002. Survival and cougar predation <strong>of</strong><br />

translocated bighorn sheep in Arizona.<br />

Journal <strong>of</strong> Wildlife Management<br />

66:1267-1272.<br />

OCKENFELS, R.A. 1994a. Factors affecting<br />

adult pronghorn mortality rates in central<br />

Arizona. Arizona Game and Fish<br />

Department Wildlife Digest Abstract 16,<br />

Phoenix.<br />

OCKENFELS, R.A. 1994b. <strong>Mountain</strong> lion<br />

predation on pronghorn in central<br />

Arizona. Southwestern Naturalist<br />

39:305-306.<br />

PHELPS, J. 2003. Status report on mountain<br />

lions in Arizona. Pages 8-10 in L. A.<br />

Harveson, P. M. Harveson, and R. W.<br />

Adams, eds. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> Sixth<br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, Texas Parks<br />

and Wildlife Department, Austin.


CALIFORNIA MOUNTAIN LION STATUS REPORT<br />

DOUG UPDIKE, Wildlife Programs Branch, California Department <strong>of</strong> Fish & Game, 1812 9 th<br />

Street, Sacramento, CA 95814, USA, email: dupdike@dfg.ca.gov<br />

INTRODUCTION<br />

California has a statewide mountain lion<br />

management plan. In 1990, mountain lions<br />

were legally classified as a “specially<br />

protected mammal” by <strong>the</strong> passage <strong>of</strong> a<br />

voter initiative (Proposition 117, June 1990<br />

ballot). Prior to that initiative, lions were<br />

classified as “game mammals.”<br />

The objectives for mountain lion<br />

management in California is to maintain<br />

healthy, wild populations <strong>of</strong> mountain lions<br />

for <strong>the</strong> benefit and enjoyment <strong>of</strong> <strong>the</strong> people<br />

in <strong>the</strong> State, to alleviate public safety<br />

incidents and reduce damage to private<br />

property (pets and livestock) by mountain<br />

lions. <strong>Mountain</strong> lions are not hunted in<br />

California, and <strong>the</strong>y may be killed only to<br />

preserve public safety, alleviate damage to<br />

private property or to protect listed bighorn<br />

sheep.<br />

Number<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

<strong>Mountain</strong> <strong>Lion</strong> Depredation Permits (1972 - 2002)<br />

6<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Lion</strong>s are currently distributed<br />

throughout all suitable habitats within<br />

California. <strong>Lion</strong> numbers appear to be<br />

stable at an estimated 4,000 to 6,000 adults.<br />

The number <strong>of</strong> lions in California is<br />

based upon extrapolating densities<br />

determined with <strong>the</strong> use <strong>of</strong> radio collars.<br />

These studies have been conducted in<br />

various locations <strong>of</strong> <strong>the</strong> State. The number<br />

<strong>of</strong> lions is determined by multiplying <strong>the</strong><br />

densities and <strong>the</strong> area represented by <strong>the</strong><br />

ecological province. The studies that<br />

provide local lion density data have been<br />

conducted over a period <strong>of</strong> a couple decades.<br />

Consequently, <strong>the</strong> Department recognizes<br />

<strong>the</strong> estimate has limited application.<br />

The Department issues depredation<br />

permits to property owners who have<br />

experienced damage from a mountain lion<br />

(Figure 1).<br />

1972<br />

1974<br />

1976<br />

1978<br />

1980<br />

1982<br />

1984<br />

1986<br />

1988<br />

1990<br />

1992<br />

1994<br />

1996<br />

1998<br />

2000<br />

2002<br />

Year<br />

Permits Issued<br />

<strong>Lion</strong>s Killed<br />

Figure 1. The number <strong>of</strong> mountain lion depredation permits issued and <strong>the</strong> number <strong>of</strong><br />

lions that have been killed as a result in California, 1972-2002.


HARVEST INFORMATION<br />

<strong>Mountain</strong> lion hunting is prohibited in<br />

California. No lions have been taken by<br />

licensed hunters since 1972. It is also illegal<br />

for lions that have been legally taken in<br />

o<strong>the</strong>r states to be imported into California.<br />

DEPREDATIONS AND HUMAN<br />

INTERACTIONS/CONFLICTS<br />

The Department’s Public Safety<br />

Guidelines are attached. This policy is<br />

intended to guide <strong>the</strong> actions and decisions<br />

<strong>of</strong> Department personnel who respond to<br />

mountain lion incidents.<br />

A summary <strong>of</strong> <strong>the</strong> number <strong>of</strong> human/lion<br />

incidents is provided in Table 1.<br />

We provide educational material to <strong>the</strong><br />

public to foster an understanding and<br />

appreciation <strong>of</strong> lions. A recent (May-June<br />

2000) issue <strong>of</strong> Outdoor California was<br />

devoted entirely to mountain lions. Most <strong>of</strong><br />

<strong>the</strong> articles are viewable at:<br />

http://www.dfg.ca.gov/coned/ocal/features.html<br />

In addition, we have produced a<br />

brochure, “Living with California <strong>Mountain</strong><br />

<strong>Lion</strong>s” which is viewable at:<br />

http://www.dfg.ca.gov/lion/index.html<br />

Depredation permits may be issued by<br />

<strong>the</strong> Department subject to <strong>the</strong> conditions<br />

found in Section 402, California Code <strong>of</strong><br />

Regulations, as follows:<br />

CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike 7<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

a. Revocable permits may be issued by <strong>the</strong><br />

department after receiving a report, from<br />

any owner or tenant or agent for <strong>the</strong>m, <strong>of</strong><br />

property being damaged or destroyed by<br />

mountain lion. The department shall<br />

conduct and complete an investigation<br />

within 48 hours <strong>of</strong> receiving such a<br />

report. Any mountain lion that is<br />

encountered in <strong>the</strong> act <strong>of</strong> inflicting injury<br />

to, molesting or killing livestock or<br />

domestic animals may be taken<br />

immediately if <strong>the</strong> taking is reported<br />

within 72 hours to <strong>the</strong> department and<br />

<strong>the</strong> carcass is made available to <strong>the</strong><br />

department. Whenever immediate<br />

action will assist in <strong>the</strong> pursuit <strong>of</strong> <strong>the</strong><br />

particular mountain lion believed to be<br />

responsible for damage to livestock or<br />

domestic animals, <strong>the</strong> department may<br />

orally authorize <strong>the</strong> pursuit and take <strong>of</strong> a<br />

mountain lion. The department shall<br />

investigate such incidents and, upon a<br />

finding that <strong>the</strong> requirements <strong>of</strong> this<br />

regulation have been met, issue a free<br />

permit for depredation purposes, and<br />

carcass tag to <strong>the</strong> person taking such<br />

mountain lion.<br />

b. Permittee may take mountain lion in <strong>the</strong><br />

manner specified in <strong>the</strong> permit, except<br />

that no mountain lion shall be taken by<br />

means <strong>of</strong> poison, leg-hold or metaljawed<br />

traps and snares.<br />

Table 1. Summary <strong>of</strong> <strong>the</strong> number <strong>of</strong> human/lion incidents in California, 1995-2002.<br />

1995 1996 1997 1998 1999 2000 2001 2002<br />

# <strong>of</strong> incidents 381 587 539 353 697 372 456 379<br />

# <strong>of</strong> safety<br />

incidents<br />

18 14 15 11 16 8 14 13<br />

take 9 7 11 12 10 7 11 13<br />

male 3 3 1 6 6 4 8 6<br />

female 3 1 6 6 3 3 3 5<br />

unknown 3 3 4 0 1 0 0 2<br />

# <strong>of</strong> sightings 191 346 340 214 382 174 240 224


8 CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike<br />

c. Permittee may take mountain lion in <strong>the</strong><br />

manner specified in <strong>the</strong> permit, except<br />

that no mountain lion shall be taken by<br />

means <strong>of</strong> poison, leg-hold or metaljawed<br />

traps and snares.<br />

d. Both males and females may be taken<br />

during <strong>the</strong> period <strong>of</strong> <strong>the</strong> permit<br />

irrespective <strong>of</strong> hours or seasons.<br />

e. The privilege granted in <strong>the</strong> permit may<br />

not be transferred, and only entitles <strong>the</strong><br />

permittee or <strong>the</strong> employee or agent <strong>of</strong><br />

<strong>the</strong> permittee to take mountain lion.<br />

Such person must be 21 years <strong>of</strong> age or<br />

over and eligible to purchase a<br />

California hunting license.<br />

f. Any person issued a permit pursuant to<br />

this section shall report by telephone<br />

within 24 hours <strong>the</strong> capturing, injuring<br />

or killing <strong>of</strong> any mountain lion to an<br />

<strong>of</strong>fice <strong>of</strong> <strong>the</strong> department or, if<br />

telephoning is not practical, in writing<br />

within five days after capturing, injuring<br />

or killing <strong>of</strong> <strong>the</strong> mountain lion. Any<br />

mountain lion killed under <strong>the</strong> permit<br />

must be tagged with <strong>the</strong> special tag<br />

furnished with <strong>the</strong> permit; both tags must<br />

be completely filled out and <strong>the</strong><br />

duplicate mailed to <strong>the</strong> Department <strong>of</strong><br />

Fish and Game, Sacramento, within 5<br />

days after taking any mountain lion.<br />

g. The entire carcass shall be transported<br />

within 5 days to a location agreed upon<br />

between <strong>the</strong> issuing <strong>of</strong>ficer and <strong>the</strong><br />

permittee, but in no case will a permittee<br />

be required to deliver a carcass beyond<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

<strong>the</strong> limits <strong>of</strong> his property unless he is<br />

willing to do so. The carcass <strong>of</strong><br />

mountain lions taken pursuant to this<br />

regulation shall become <strong>the</strong> property <strong>of</strong><br />

<strong>the</strong> state.<br />

h. Animals shall be taken in a humane<br />

manner so as to prevent any undue<br />

suffering to <strong>the</strong> animals<br />

i. The permittee shall take every<br />

reasonable precaution to prevent <strong>the</strong><br />

carcass from spoiling until disposed <strong>of</strong> in<br />

<strong>the</strong> manner agreed upon under<br />

subsection (f) <strong>of</strong> <strong>the</strong>se regulations<br />

j. The permit does not invalidate any city,<br />

county, or state firearm regulation.<br />

k. Permits shall be issued for a period <strong>of</strong> 10<br />

days. Permits may be renewed only<br />

after a finding by <strong>the</strong> department that<br />

fur<strong>the</strong>r damage has occurred or will<br />

occur unless such permits are renewed.<br />

The permittee may not begin pursuit <strong>of</strong> a<br />

lion more than one mile nor continue<br />

pursuit beyond a 10-mile radius from <strong>the</strong><br />

location <strong>of</strong> <strong>the</strong> reported damage.<br />

CURRENT RESEARCH<br />

a. Population genetics <strong>of</strong> lions<br />

b. <strong>Lion</strong>/deer/bighorn sheep predator prey<br />

relationships in Inyo/Mono counties and<br />

San Diego County<br />

c. <strong>Lion</strong> movements and corridors in Los<br />

Angeles/Ventura counties<br />

d. Impacts <strong>of</strong> habitat conversions and<br />

transportation corridors or lion<br />

movements and habitat use.


PUBLIC SAFETY WILDLIFE GUIDELINES – 2072<br />

Consistent with Section 1801 <strong>of</strong> <strong>the</strong> Fish<br />

and Game Code, <strong>the</strong>se Public Safety<br />

Wildlife Guidelines provide procedures to<br />

address public safety wildlife problems.<br />

<strong>Mountain</strong> lions, black bears, deer, coyotes,<br />

and large exotic carnivores that have<br />

threatened or Attacked humans are wildlife<br />

classified as public safety problems. Public<br />

safety wildlife incidents are classified into<br />

three types:<br />

A. Type Green (sighting)<br />

A report (confirmed or unconfirmed) <strong>of</strong><br />

an observation that is perceived to be a<br />

public safety wildlife problem. The mere<br />

presence <strong>of</strong> <strong>the</strong> wildlife species does not<br />

in itself constitute a threat.<br />

B. Type Yellow (threat)<br />

A report where <strong>the</strong> presence <strong>of</strong> <strong>the</strong><br />

public safety wildlife is confirmed by a<br />

field investigation, and <strong>the</strong> responding<br />

person (law enforcement <strong>of</strong>ficer or<br />

Department employee) perceives <strong>the</strong><br />

animal to be an imminent threat to<br />

public health or safety. Imminent threat<br />

means <strong>the</strong>re is a likelihood <strong>of</strong> human<br />

injury based on <strong>the</strong> totality <strong>of</strong> <strong>the</strong><br />

circumstances.<br />

C. Type Red (attack)<br />

An attack by a public safety wildlife<br />

species on a human resulting in physical<br />

contact, injury, or death.<br />

These guidelines are not intended to<br />

address orphaned, injured, or sick wildlife<br />

that have not threatened public safety. To<br />

achieve <strong>the</strong> intent <strong>of</strong> <strong>the</strong>se guidelines, <strong>the</strong><br />

following procedures shall be used.<br />

I. Wildlife Incident Report Form<br />

Fill out a Wildlife Incident Report<br />

Form (WMD-2) for all reports <strong>of</strong> public<br />

safety wildlife incidents. The nature <strong>of</strong><br />

<strong>the</strong> report will determine <strong>the</strong> response or<br />

CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike 9<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

investigative action to <strong>the</strong> public safety<br />

problem. For those reports that require a<br />

follow-up field investigation, <strong>the</strong><br />

Wildlife Incident Report Form will be<br />

completed by <strong>the</strong> field investigator. All<br />

completed Wildlife Incident Report<br />

Forms shall be forwarded through <strong>the</strong><br />

regional <strong>of</strong>fices to <strong>the</strong> Chief, WPB.<br />

II. Response to Public Safety Wildlife<br />

Problems<br />

The steps in responding to a public<br />

safety wildlife incident are diagramed<br />

below (Figure 3).<br />

Any reported imminent threats or<br />

attacks on humans by wildlife will<br />

require a follow-up field investigation.<br />

If a public safety wildlife species is<br />

outside its natural habitat and in an area<br />

where it could become a public safety<br />

problem, and if approved by <strong>the</strong> Deputy<br />

Director for <strong>the</strong> WIFD, it may be<br />

captured using restraint techniques<br />

approved by <strong>the</strong> Wildlife Investigations<br />

Laboratory (WIL). The disposition <strong>of</strong> <strong>the</strong><br />

captured wildlife may be coordinated<br />

with WIL.<br />

A. Type Green (sighting)<br />

If <strong>the</strong> investigator determines that no<br />

imminent threat to public safety exists,<br />

Figure 3. Steps in responding to a public<br />

safety wildlife incident.


10 CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike<br />

<strong>the</strong> incident is considered a Type Green.<br />

The appropriate action may include<br />

providing wildlife behavior information<br />

and mailing public educational materials<br />

to <strong>the</strong> reporting party.<br />

B. Type Yellow (threat)<br />

Once <strong>the</strong> field investigator finds<br />

evidence <strong>of</strong> <strong>the</strong> public safety wildlife<br />

and perceives <strong>the</strong> animal to be an<br />

imminent threat to public health or<br />

safety, <strong>the</strong> incident is considered a Type<br />

Yellow. In <strong>the</strong> event <strong>of</strong> threat to public<br />

safety, any Department employee<br />

responding to a reported public safety<br />

incident may take whatever action is<br />

deemed necessary within <strong>the</strong> scope <strong>of</strong><br />

<strong>the</strong> employee's authority to protect<br />

public safety. When evidence shows that<br />

a wild animal is an imminent threat to<br />

public safety, that wild animal shall be<br />

humanely euthanized (shot, killed,<br />

dispatched, destroyed, etc.). For Type<br />

Yellow incidents <strong>the</strong> following steps<br />

should be taken:<br />

1. Initiate <strong>the</strong> Incident Command<br />

System. The Incident Commander<br />

(IC) consults with <strong>the</strong> regional<br />

manager or designee to decide on <strong>the</strong><br />

notification process on a case-bycase<br />

basis. Full notification includes:<br />

<strong>the</strong> field investigator's supervisor, <strong>the</strong><br />

appropriate regional manager, <strong>the</strong><br />

Deputy Director, WIFD, Chief,<br />

Conservation Education and<br />

Enforcement Branch (CEEB), Chief,<br />

WPB, WIL, Wildlife Forensics Lab<br />

(WFL), <strong>the</strong> designated regional<br />

information <strong>of</strong>ficer, and <strong>the</strong> local law<br />

enforcement agency.<br />

2. If full notification is appropriate,<br />

notify Sacramento Dispatch at (916)<br />

445-0045. Dispatch shall notify <strong>the</strong><br />

above-mentioned personnel.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

3. Secure <strong>the</strong> scene as appropriate.<br />

Take all practical steps to preserve<br />

potential evidence. The IC holds<br />

initial responsibility and authority<br />

over <strong>the</strong> scene, locating <strong>the</strong> animal,<br />

its resultant carcass, and any o<strong>the</strong>r<br />

physical evidence from <strong>the</strong> attack.<br />

The IC will ensure proper transfer<br />

and disposition <strong>of</strong> all physical<br />

evidence.<br />

4. In most situations, it is important to<br />

locate <strong>the</strong> <strong>of</strong>fending animal as soon<br />

as practical. WIL may be <strong>of</strong><br />

assistance. The services <strong>of</strong> USDA,<br />

Wildlife Services (WS) can be<br />

arranged by <strong>the</strong> regional manager or<br />

designee contacting <strong>the</strong> local WS<br />

District Supervisor. If possible, avoid<br />

shooting <strong>the</strong> animal in <strong>the</strong> head to<br />

preserve evidence.<br />

5. If an animal is killed, <strong>the</strong> IC will<br />

decide on <strong>the</strong> notification process<br />

and notify Sacramento Dispatch if<br />

appropriate. Use clean protective<br />

gloves while handling <strong>the</strong> carcass.<br />

Place <strong>the</strong> carcass inside a protective<br />

durable body bag (avoid dragging<br />

<strong>the</strong> carcass, if possible).<br />

C. Type Red (attack)<br />

In <strong>the</strong> event <strong>of</strong> an attack, <strong>the</strong><br />

responding Department employee may<br />

take any action necessary that is within<br />

<strong>the</strong> scope <strong>of</strong> <strong>the</strong> employee's authority to<br />

protect public safety. When evidence<br />

shows that a wild animal has made an<br />

unprovoked attack on a human, that wild<br />

animal shall be humanely euthanized<br />

(shot, killed, dispatched, destroyed, etc.).<br />

For Type Red incidents <strong>the</strong> following<br />

steps should be taken:<br />

1. Ensure proper medical aid for <strong>the</strong><br />

victim. Identify <strong>the</strong> victim (obtain<br />

<strong>the</strong> following, but not limited to:<br />

name, address, phone number).


2. Notify Sacramento Dispatch at (916)<br />

445-0045. Dispatch shall notify <strong>the</strong><br />

field investigator's supervisor, <strong>the</strong><br />

appropriate regional manager, <strong>the</strong><br />

Deputy Director, WIFD, Chief,<br />

CEEB, Chief, WPB, WIL, WFL, <strong>the</strong><br />

designated regional information<br />

<strong>of</strong>ficer, and <strong>the</strong> local law<br />

enforcement agency.<br />

3. Initiate <strong>the</strong> Incident Command<br />

System. If a human death has<br />

occurred, an Enforcement Branch<br />

supervisor or specialist will respond<br />

to <strong>the</strong> Incident Command Post and<br />

assume <strong>the</strong> IC responsibilities. The<br />

IC holds initial responsibility and<br />

authority over <strong>the</strong> scene, locating <strong>the</strong><br />

animal, its resultant carcass, and any<br />

o<strong>the</strong>r physical evidence from <strong>the</strong><br />

attack. The IC will ensure proper<br />

transfer and disposition <strong>of</strong> all<br />

physical evidence.<br />

4. Secure <strong>the</strong> area as needed. Treat <strong>the</strong><br />

area as a crime scene. In order to<br />

expedite <strong>the</strong> capture <strong>of</strong> <strong>the</strong> <strong>of</strong>fending<br />

animal and preserve as much onscene<br />

evidence as possible, <strong>the</strong> area<br />

<strong>of</strong> <strong>the</strong> incident must be secured<br />

immediately by <strong>the</strong> initial responding<br />

<strong>of</strong>ficer. The area should be excluded<br />

from public access by use <strong>of</strong> flagging<br />

tape or similar tape (e.g., "Do Not<br />

Enter") utilized at crime scenes by<br />

local law enforcement agencies. One<br />

entry and exit port should be<br />

established. Only essential<br />

authorized personnel should be<br />

permitted in <strong>the</strong> excluded area. A<br />

second area outside <strong>the</strong> area <strong>of</strong> <strong>the</strong><br />

incident should be established as <strong>the</strong><br />

command post.<br />

5. In cases involving a human death,<br />

WFL personnel will direct <strong>the</strong><br />

ga<strong>the</strong>ring <strong>of</strong> evidence. Secure items<br />

such as clothing, tents, sleeping bags,<br />

CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike 11<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

objects used for defense during <strong>the</strong><br />

attack, objects chewed on by <strong>the</strong><br />

animal, or any o<strong>the</strong>r materials which<br />

may possess <strong>the</strong> attacking animal's<br />

saliva, hair, or blood.<br />

6. If <strong>the</strong> victim is alive, advise <strong>the</strong><br />

attending medical personnel about<br />

<strong>the</strong> Carnivore Attack-Victim<br />

Sampling Kit for collecting possible<br />

animal saliva stains or hair that<br />

might still be on <strong>the</strong> victim. If <strong>the</strong><br />

victim is dead, advise <strong>the</strong> medical<br />

examiner <strong>of</strong> this evidence need. This<br />

sampling kit may be obtained from<br />

<strong>the</strong> WFL.<br />

7. It is essential to locate <strong>the</strong> <strong>of</strong>fending<br />

animal as soon as practical. WIL<br />

may be <strong>of</strong> assistance. The services <strong>of</strong><br />

WS can be arranged by <strong>the</strong> regional<br />

manager or designee contacting <strong>the</strong><br />

local WS District Supervisor. If<br />

possible, avoid shooting <strong>the</strong> animal<br />

in <strong>the</strong> head to preserve evidence.<br />

8. If an animal is killed, <strong>the</strong> IC will<br />

notify Sacramento Dispatch. Treat<br />

<strong>the</strong> carcass as evidence. Use clean<br />

protective gloves and (if possible) a<br />

facemask while handling <strong>the</strong> carcass.<br />

Be guided by <strong>the</strong> need to protect <strong>the</strong><br />

animal's external body from: loss <strong>of</strong><br />

bloodstains or o<strong>the</strong>r such physical<br />

evidence originating from <strong>the</strong> victim;<br />

contamination by <strong>the</strong> animal's own<br />

blood; and contamination by <strong>the</strong><br />

human handler's hair, sweat, saliva,<br />

skin cells, etc. Tape paper bags over<br />

<strong>the</strong> head and paws, <strong>the</strong>n tape plastic<br />

bags over <strong>the</strong> paper bags. Plug<br />

wounds with tight gauze to minimize<br />

contamination <strong>of</strong> <strong>the</strong> animal with its<br />

own blood. Place <strong>the</strong> carcass inside a<br />

protective durable body bag (avoid<br />

dragging <strong>the</strong> carcass, if possible).<br />

9. WFL will receive from <strong>the</strong> IC and/or<br />

directly obtain all pertinent physical


12 CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike<br />

evidence concerning <strong>the</strong> primary<br />

questions <strong>of</strong> au<strong>the</strong>nticity <strong>of</strong> <strong>the</strong><br />

attack and identity <strong>of</strong> <strong>the</strong> <strong>of</strong>fending<br />

animal. WFL has first access and<br />

authority over <strong>the</strong> carcass after <strong>the</strong><br />

IC. WFL will immediately contact<br />

and coordinate with <strong>the</strong> county<br />

health department <strong>the</strong> acquisition <strong>of</strong><br />

appropriate samples for rabies<br />

testing. Once WFL has secured <strong>the</strong><br />

necessary forensic samples, <strong>the</strong>y will<br />

<strong>the</strong>n release authority over <strong>the</strong><br />

carcass to WIL for disease studies.<br />

10. An independent diagnostic<br />

laboratory approved by WIL will<br />

conduct necropsy and disease studies<br />

on <strong>the</strong> carcass. The WIL will retain<br />

primary authority over this aspect <strong>of</strong><br />

<strong>the</strong> carcass.<br />

D. Responsibilities <strong>of</strong> WIL<br />

WIL investigates wildlife disease<br />

problems statewide and provides<br />

information on <strong>the</strong> occurrence <strong>of</strong> both<br />

enzootic and epizootic disease in<br />

wildlife populations. Specimens<br />

involved in suspected disease problems<br />

are submitted to WIL for necropsy and<br />

disease studies. Most animals killed for<br />

public safety reasons will be necropsied<br />

to assess <strong>the</strong> status <strong>of</strong> health and whe<strong>the</strong>r<br />

<strong>the</strong> presence <strong>of</strong> disease may have caused<br />

<strong>the</strong> aggressive and/or unusual behavior.<br />

Type Yellow public safety animals<br />

killed may be necropsied by WIL or an<br />

independent diagnostic laboratory<br />

approved by WIL. Contact WIL<br />

immediately after a public safety animal<br />

is killed to determine where it will be<br />

necropsied. Arrangements are to be<br />

made directly with WIL prior to<br />

submission <strong>of</strong> <strong>the</strong> carcass to any<br />

laboratory.<br />

Type Red public safety animals<br />

killed will be necropsied by an<br />

independent diagnostic laboratory<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

approved by WIL. Contact WIL prior to<br />

submission <strong>of</strong> <strong>the</strong> carcass to any<br />

laboratory to allow <strong>the</strong> Department<br />

veterinarian to discuss <strong>the</strong> disease testing<br />

requirements with <strong>the</strong> attending<br />

pathologist. A disease testing protocol<br />

has been developed for use with Type<br />

Red public safety wildlife.<br />

E. Responsibilities <strong>of</strong> WFL<br />

WFL has <strong>the</strong> statewide responsibility<br />

to receive, collect, examine and analyze<br />

physical evidence, issue reports on<br />

evidence findings, and testify in court as<br />

to those results. WFL's primary<br />

functions in public safety incidents is to<br />

verify or refute <strong>the</strong> au<strong>the</strong>nticity <strong>of</strong> <strong>the</strong><br />

purported attack and to corroborate or<br />

refute <strong>the</strong> involvement <strong>of</strong> <strong>the</strong> suspected<br />

<strong>of</strong>fending animal.<br />

Type Yellow public safety animals<br />

killed may be examined by WFL<br />

personnel. The examination <strong>of</strong> <strong>the</strong><br />

carcass will be coordinated with WIL.<br />

All Type Red public safety animals<br />

killed must be examined by WFL<br />

personnel or a qualified person approved<br />

by WFL supervisor using specific<br />

procedures established by WFL.<br />

If a human death occurs,<br />

coordination <strong>of</strong> <strong>the</strong> autopsy between <strong>the</strong><br />

proper <strong>of</strong>ficials and WFL is important so<br />

that WFL personnel can be present<br />

during <strong>the</strong> autopsy for appropriate<br />

sampling and examination. In <strong>the</strong> event<br />

<strong>of</strong> human injury, it is important for WFL<br />

to ga<strong>the</strong>r any relevant physical evidence<br />

that may corroborate <strong>the</strong> au<strong>the</strong>nticity <strong>of</strong><br />

a wildlife attack, prior to <strong>the</strong> treatment <strong>of</strong><br />

injuries, if practical. If not practical,<br />

directions for sampling may be given<br />

over <strong>the</strong> telephone to <strong>the</strong> emergency<br />

room doctor by WFL.<br />

F. Media Contact<br />

Public safety wildlife incidents


attract significant media attention. Issues<br />

regarding site access, information<br />

dissemination, <strong>the</strong> public's safety,<br />

carcass viewing and requests to survey<br />

<strong>the</strong> scene can be handled by a designated<br />

employee. Each region shall designate<br />

an employee with necessary ICS training<br />

to respond as a regional information<br />

<strong>of</strong>ficer to public safety wildlife<br />

incidents.<br />

Type Yellow public safety wildlife<br />

incidents may require <strong>the</strong> notification <strong>of</strong><br />

a designated employee previously<br />

approved by <strong>the</strong> regional manager or<br />

designee to assist <strong>the</strong> IC in responding to<br />

CALIFORNIA MOUNTAIN LION STATUS REPORT · Updike 13<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

<strong>the</strong> media and disseminating<br />

information. The IC has <strong>the</strong> authority to<br />

decide if <strong>the</strong> designated employee<br />

should be dispatched to <strong>the</strong> site.<br />

All Type Red public safety wildlife<br />

incidents require that a designated<br />

employee, previously approved by <strong>the</strong><br />

regional manager or designee, to assist<br />

<strong>the</strong> IC in responding to <strong>the</strong> media and<br />

disseminating information, is called to<br />

<strong>the</strong> scene.<br />

The Department will develop and<br />

provide training for designated<br />

employees to serve as information<br />

<strong>of</strong>ficers for public safety wildlife<br />

incidents.


COLORADO MOUNTAIN LION STATUS REPORT<br />

JERRY A. APKER, Colorado Division <strong>of</strong> Wildlife, 0722 South Road 1 East, Monte Vista, CO<br />

81144, USA, email: jerry.apker@state.co.us<br />

MOUNTAIN LION CLASSIFICATION<br />

<strong>Mountain</strong> lion (Puma concolor) received<br />

no legal protection and were classified as a<br />

predator in Colorado from 1881 until 1965.<br />

During this time take <strong>of</strong> puma at any time,<br />

any place was encouraged by bounties and<br />

o<strong>the</strong>r laws. The first bounty was enacted in<br />

1881 at $10, in 1925 laws instructed game<br />

wardens to destroy predatory animals by<br />

trapping, poisoning, or hunting, and in 1929<br />

<strong>the</strong> bounty was increased to $50. For<br />

comparison <strong>the</strong> 1929 bounty, if <strong>of</strong>fered in<br />

2003 dollars, would be $540. The bounty<br />

was abolished in 1965, but some provisions<br />

for landowner take <strong>of</strong> a depredating puma<br />

remains in Colorado laws to this day. In<br />

1965, puma were reclassified as big game.<br />

Each Data Analysis Unit (DAU) within<br />

<strong>the</strong> State has a management plan developed<br />

with objectives for hunter harvest, game<br />

damage, and human-puma conflicts.<br />

Objectives are stated as <strong>the</strong> maximum level<br />

on a three-year running average.<br />

Implementation <strong>of</strong> DAU plans began in<br />

2001. Recent interest in annual puma kill<br />

revealed conflicting direction depending<br />

upon which objectives managers weighed<br />

most heavily. These conflicts pointed out a<br />

shortfall within <strong>the</strong> plans in that <strong>the</strong>y do not<br />

state a specific strategic goal for <strong>the</strong> DAU.<br />

Currently this must be inferred in <strong>the</strong> text <strong>of</strong><br />

<strong>the</strong> plan. Some DAUs are managed to<br />

suppress puma populations while o<strong>the</strong>rs are<br />

managed to maintain stable populations –<br />

recognizing <strong>the</strong> inherent difficulty in<br />

determining population changes. Within <strong>the</strong><br />

next year all management plans will be<br />

required to develop a strategic goal. We<br />

14<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

consider this an essential step for informing<br />

management decisions within a DAU about<br />

season structure and annual license<br />

allocation.<br />

In 1996 <strong>the</strong> Colorado Department <strong>of</strong><br />

Agriculture (CDA) was granted “exclusive<br />

jurisdiction over <strong>the</strong> control <strong>of</strong> depredating<br />

animals that pose a threat to an agricultural<br />

product or resource”. Thus, CDA has<br />

exclusive authority to determine <strong>the</strong><br />

disposition <strong>of</strong> an individual puma if it is<br />

depredating on livestock, while <strong>the</strong> Colorado<br />

Division <strong>of</strong> Wildlife (CDOW) retains<br />

authority to manage puma populations and<br />

all forms <strong>of</strong> recreational or scientific use.<br />

DISTRIBUTION, ABUNDANCE AND<br />

MONITORING<br />

The state is divided into 21 DAUs for<br />

<strong>the</strong> purposes <strong>of</strong> puma management (Figure<br />

1). DAUs are assemblages <strong>of</strong> Game<br />

Management Units (GMUs) within which<br />

Figure 1. Data analysis Units and relative<br />

abundance <strong>of</strong> puma within each DAU in<br />

Colorado.


Figure 2. Areas <strong>of</strong> puma occupancy in<br />

Colorado.<br />

puma occupancy has been mapped (Figure<br />

2).<br />

Colorado does not regularly estimate<br />

puma populations because no reliable, cost<br />

effective sample based population<br />

estimation technique currently exists. A<br />

projection <strong>of</strong> possible population has been<br />

made based on densities reported in<br />

literature for intensively studied populations.<br />

Low and high densities were selected from<br />

study areas that had habitat types most<br />

similar to Colorado. Densities were <strong>the</strong>n<br />

applied by biologists to area <strong>of</strong> puma habitat<br />

within DAUs. Areas not considered puma<br />

habitats, such as extreme high elevations,<br />

intensively farmed land, cities, highways, or<br />

reservoirs, were first deleted. Biologists<br />

were allowed to apply more constrained<br />

densities based upon <strong>the</strong>ir knowledge <strong>of</strong><br />

prey abundance or relative puma abundance.<br />

Finally, biologists were asked to pinpoint<br />

<strong>the</strong> puma density most applicable to DAUs<br />

within <strong>the</strong>ir management responsibility.<br />

These exercises resulted in a crude projected<br />

puma population <strong>of</strong> 3,000 to 7,000, with<br />

3,500 to 4,500 most probable. Based upon<br />

<strong>the</strong> foregoing, each DAU is assigned a<br />

relative abundance rating <strong>of</strong> high, moderate,<br />

or low with intergrades where estimated<br />

puma density is close to break points. High<br />

COLORADO MOUNTAIN LION STATUS REPORT · Apker 15<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

abundance is assigned at DAU densities <strong>of</strong><br />

over 3 puma/100 km 2 , moderate abundance<br />

at 2 to 3 puma/100 km 2 , and low abundance<br />

at anything less than 2 puma/100 km 2<br />

(Figure 1).<br />

Hunter harvest and total mortality is<br />

examined at <strong>the</strong> DAU level to monitor<br />

mortality for crude indications <strong>of</strong> population<br />

change. Puma mortality is documented<br />

through mandatory checks <strong>of</strong> hunter kill and<br />

mandatory reports for non-hunter mortality<br />

and is kept in a database. The database for<br />

hunter kill has been kept since 1980, and for<br />

non-hunter mortality since 1991. Mortality<br />

data is examined on three and ten year<br />

running averages due to relatively high<br />

annual variation. Data on depredation<br />

claims since is also maintained in a<br />

database.<br />

HARVEST AND HUNTING<br />

REGULATION<br />

Since 1972 a quota system has been used<br />

to manage hunter distribution and kill. From<br />

1992 <strong>the</strong> quota has increased from 459 to<br />

790 in 2002. However, <strong>the</strong> quota does not<br />

represent <strong>the</strong> harvest objective since <strong>the</strong><br />

quota is never achieved. Through<br />

compilation <strong>of</strong> DAU management plan<br />

objectives <strong>the</strong> harvest objective for <strong>the</strong> state<br />

is about 350 puma. Annual license sales<br />

have also increased since 1992 from about<br />

900 to just over 1,700 in 2002. While both<br />

quotas and license sales have increased over<br />

<strong>the</strong> past 10 years, percent <strong>of</strong> quota<br />

achievement and success relative to license<br />

sales have declined gradually (Figure 3).<br />

These trends are expected with increased<br />

available hunting opportunity toward a<br />

cryptic species. With more potential hunters<br />

<strong>the</strong>re is an increased likelihood that <strong>the</strong>re<br />

will be proportionately more hunters with<br />

less experience and less commitment or<br />

impetus to harvest an animal. Some have<br />

speculated that <strong>the</strong> trends indicate that overharvest<br />

has occurred, however <strong>the</strong> female


16 COLORADO MOUNTAIN LION STATUS REPORT · Apker<br />

Percent Q. Achievement & Success<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

1992 1993 1994 1995 1996 1997 1998 1999 2000 2001<br />

% Quota Achievement % Harvest Success by licenses sold<br />

Quota # Licenses Sold<br />

1700<br />

1500<br />

1300<br />

1100<br />

Figure 3. Colorado license sales, quota,<br />

percent success and percent quota<br />

achievement for puma.<br />

component <strong>of</strong> hunter harvest has not<br />

increased substantially which would be an<br />

indicator <strong>of</strong> over-harvest.<br />

Criteria used to guide quota setting are<br />

as follows:<br />

1. Strategic objective <strong>of</strong> <strong>the</strong> DAU or<br />

group <strong>of</strong> GMUs within a DAU. If<br />

management is directed at<br />

maintaining a stable population, <strong>the</strong>n<br />

<strong>the</strong> following also apply.<br />

2. Population for <strong>the</strong> DAU is projected<br />

based upon low and high density<br />

potential. Off-take should not<br />

exceed a bracketed range <strong>of</strong> 15% <strong>of</strong><br />

low-end population estimate and 8%<br />

<strong>of</strong> high-end population estimate.<br />

3. Short (3 year) and long-term trend<br />

(10 years) in proportion <strong>of</strong> females in<br />

mortality should be stable or<br />

downward and not over 50%.<br />

4. Damage claim amounts on 3-year<br />

average should not exceed DAU<br />

objective levels.<br />

5. Catch per unit effort indice (effort <strong>of</strong><br />

houndsmen to harvest).<br />

900<br />

700<br />

500<br />

300<br />

100<br />

Quota or License #'s<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Hunter harvest and total mortality<br />

figures for 2002 have not been completely<br />

tabulated at <strong>the</strong> time <strong>of</strong> this report. The<br />

average hunter harvest from 1992-1994 is<br />

308 with 41% female. The average hunter<br />

harvest from 1999-2001 is 365 with 45%<br />

female (Figure 4).<br />

Generally, from 1965 to <strong>the</strong> mid-late<br />

1970s seasons were mid fall through early<br />

spring. In <strong>the</strong> late 1970s through 1994<br />

seasons were liberalized, running almost<br />

continually through <strong>the</strong> year excluding late<br />

August – mid November deer or elk hunting<br />

seasons. Since 1995, seasons were revised<br />

to provide greater protection for pregnant<br />

females or females with dependent young,<br />

running on a calendar year basis from<br />

January 1 – March 31 and mid November –<br />

December 31. With a few exceptions <strong>the</strong><br />

bag limit has remained 1 per year <strong>of</strong> ei<strong>the</strong>r<br />

sex and some form <strong>of</strong> puma license has been<br />

required since 1965.<br />

Hunting with hounds is permitted with<br />

hunting pack size limited to 8 dogs. Almost<br />

all puma are harvested with <strong>the</strong> use <strong>of</strong><br />

hounds. There is no pursuit only season.<br />

With certain technical restrictions on each,<br />

legal weapons for take include rifle,<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

1992 1993 1994 1995 1996 1997 1998 1999 2000 2001<br />

Hunter Harvest<br />

- Male<br />

Hunter Harvest<br />

- Female<br />

Total<br />

Mortality<br />

Figure 4. Puma harvest and total mortality<br />

levels in Colorado.


handgun, shotgun, muzzleloading rifles,<br />

hand-held bows, and crossbows. It is illegal<br />

to kill a kitten or a female accompanied by<br />

kittens.<br />

DEPREDATION AND PUMA-HUMAN<br />

CONFLICT<br />

Colorado is liable for damage caused by<br />

big game, with certain limitations and<br />

restrictions. From 1972 until 2001 CDOW<br />

had to pay for damage by puma and black<br />

bear to any real or personal property. Black<br />

bear damage claims <strong>of</strong>ten included vehicles,<br />

buildings, appliances, etc., as well as<br />

livestock, but puma damage claims have<br />

been restricted to cattle, sheep, or o<strong>the</strong>r<br />

animals. Beginning in 2001, State liability<br />

was limited to agricultural products and<br />

property used in <strong>the</strong> production <strong>of</strong> raw<br />

agricultural products. Liability was also<br />

changed so that <strong>the</strong> State is not liable for<br />

more than $5,000 per animal.<br />

With <strong>the</strong> exception <strong>of</strong> 2000 <strong>the</strong> number<br />

<strong>of</strong> damage claims and <strong>the</strong> cost <strong>of</strong> damage<br />

have declined since 1997 (Figure 5). High<br />

damage costs in 2000 were mostly due to 6<br />

claims for <strong>the</strong> loss <strong>of</strong> 8 exotic domestic<br />

animals such as alpaca, llama, and<br />

250000<br />

225000<br />

200000<br />

175000<br />

150000<br />

125000<br />

100000<br />

75000<br />

50000<br />

25000<br />

0<br />

1979<br />

1981<br />

1983<br />

1985<br />

1987<br />

1989<br />

1991<br />

1993<br />

1995<br />

Sheep Cattle O<strong>the</strong>r Stock<br />

Figure 5. Amount paid on claims for<br />

depredation by puma in Colorado.<br />

1997<br />

1999<br />

2001<br />

COLORADO MOUNTAIN LION STATUS REPORT · Apker 17<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

commercially owned elk. Procedures for<br />

handling damage claims are governed by<br />

statute, regulations, and a game damage<br />

procedures manual.<br />

The State has no specific policy<br />

document providing direction for handling<br />

puma-human conflicts. However, following<br />

a human fatality in 1991, DOW staff<br />

developed procedures that have generally<br />

been adopted. Encounters involving puma<br />

are categorized as sightings, encounter<br />

involving pets, aggressive behavior toward<br />

humans, or attack on humans. Agency<br />

responses to <strong>the</strong>se types <strong>of</strong> encounters vary<br />

from providing education and information to<br />

pursue-kill <strong>the</strong> puma. In <strong>the</strong> past 5 years,<br />

fewer than 5-10 encounters beyond sightings<br />

are documented each year.<br />

On average over <strong>the</strong> past 5 years about<br />

20 puma per year are killed for reasons o<strong>the</strong>r<br />

than hunting. Most <strong>of</strong> <strong>the</strong>se, 12 per year, are<br />

control actions on depredating animals. The<br />

remainders are <strong>the</strong> result <strong>of</strong> road kills or<br />

illegal kills. Less than 1 per year on average<br />

are killed due to human safety concerns.<br />

PUMA RESEARCH PROGRAMS<br />

There are no current research<br />

investigations being conducted on puma.<br />

The Division <strong>of</strong> Wildlife is in <strong>the</strong> process <strong>of</strong><br />

hiring a research scientist specializing in<br />

carnivores with emphasis on puma initially.


FLORIDA FISH AND WILDLIFE CONSERVATION COMMISSION STATUS REPORT<br />

MARK LOTZ, Pan<strong>the</strong>r Section Biologist, Florida Fish and Wildlife Conservation Commission,<br />

566 Commercial Blvd., Naples, FL 34104-4709, USA, email: Mark.Lotz@fwc.state.fl.us<br />

E. DARRELL LAND, Pan<strong>the</strong>r Section Leader, Florida Fish and Wildlife Conservation<br />

Commission, 566 Commercial Blvd., Naples, FL 34104-4709, USA, email:<br />

Darrell.Land@fwc.state.fl.us<br />

INTRODUCTION<br />

The Florida pan<strong>the</strong>r (Puma concolor<br />

coryi) has been classified as endangered by<br />

<strong>the</strong> state <strong>of</strong> Florida since 1958 and by <strong>the</strong><br />

federal government since 1967. Formerly,<br />

pan<strong>the</strong>rs inhabited <strong>the</strong> sou<strong>the</strong>astern United<br />

States, ranging from sou<strong>the</strong>rn Florida to<br />

Arkansas and northward to Tennessee and<br />

South Carolina. Loss and fragmentation <strong>of</strong><br />

habitat coupled with unregulated killing<br />

over <strong>the</strong> past two centuries have reduced and<br />

isolated <strong>the</strong> pan<strong>the</strong>r to <strong>the</strong> point where only<br />

one population exists on approximately<br />

8,810 km 2 <strong>of</strong> habitat in south Florida (Maehr<br />

1990). The Florida Fish and Wildlife<br />

Conservation Commission (FWC) and <strong>the</strong><br />

U. S. Fish and Wildlife Service (USFWS)<br />

are <strong>the</strong> two lead authorities involved in all<br />

aspects <strong>of</strong> Florida pan<strong>the</strong>r recovery and<br />

protection. O<strong>the</strong>r agencies involved in<br />

pan<strong>the</strong>r recovery include <strong>the</strong> Florida<br />

Department <strong>of</strong> Environmental Protection,<br />

Florida Division <strong>of</strong> Forestry, National Park<br />

Service, South Florida Water Management<br />

District, as well as numerous nongovernmental<br />

organizations such as Florida<br />

Wildlife Federation, National Wildlife<br />

Federation, The Nature Conservancy, and<br />

<strong>the</strong> Florida Audubon Society. A recovery<br />

plan for <strong>the</strong> Florida pan<strong>the</strong>r was written in<br />

1981 with revisions in 1987 and 1995 with<br />

<strong>the</strong> objective <strong>of</strong> achieving three viable selfsustaining<br />

populations within <strong>the</strong> historic<br />

range. FWC initiated intensive research<br />

efforts in 1981 and <strong>the</strong>se studies continue<br />

18<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

today. By <strong>the</strong> end <strong>of</strong> 2002, FWC has<br />

handled 115 pan<strong>the</strong>rs for radio-telemetry<br />

studies and marked 142 neonate kittens at<br />

dens. FWC and many collaborators have<br />

published more than 200 papers and reports<br />

detailing pan<strong>the</strong>r life history, habitat use,<br />

mortality, dispersal, home range dynamics,<br />

biomedical findings, genetics, population<br />

modeling, and food habits.<br />

Florida pan<strong>the</strong>rs are threatened by<br />

demographic instability inherent in small,<br />

geographically isolated populations and<br />

erosion <strong>of</strong> genetic diversity from restricted<br />

gene flow and inbreeding. Genetic diversity<br />

is <strong>the</strong> basis for production <strong>of</strong> fit individuals<br />

as well as providing population elasticity in<br />

order to respond to changing environmental<br />

and habitat conditions. Historically, natural<br />

exchange <strong>of</strong> genetic material occurred<br />

among <strong>the</strong> Florida pan<strong>the</strong>r population in <strong>the</strong><br />

sou<strong>the</strong>astern United States and contiguous<br />

populations <strong>of</strong> P. c. cougar to <strong>the</strong> north, P.<br />

c. hippolestes to <strong>the</strong> northwest and P. c.<br />

stanleyana to <strong>the</strong> west (Young and Goldman<br />

1946). Genetic exchange between<br />

populations ceased as <strong>the</strong> coastal plain was<br />

gradually cleared and settled. Florida<br />

pan<strong>the</strong>rs steadily declined in abundance and<br />

distribution as a result. Inbreeding increased<br />

when potential breeders could no longer<br />

move among fragmented populations and<br />

<strong>the</strong> declining population size compounded<br />

demographic and genetic factors. A<br />

population viability analysis was conducted<br />

in 1992, which predicted <strong>the</strong> extinction <strong>of</strong>


<strong>the</strong> Florida pan<strong>the</strong>r within 24-63 years (Seal<br />

1992) and lead to <strong>the</strong> creation <strong>of</strong> A Plan for<br />

Genetic Restoration and Management <strong>of</strong> <strong>the</strong><br />

Florida Pan<strong>the</strong>r (Seal 1994).<br />

Genetic restoration <strong>of</strong> <strong>the</strong> Florida<br />

pan<strong>the</strong>r was implemented in 1995 with <strong>the</strong><br />

release <strong>of</strong> 8 female Texas cougars (P. c.<br />

stanleyana) into areas occupied by Florida<br />

pan<strong>the</strong>rs. Five <strong>of</strong> <strong>the</strong> 8 cougars produced a<br />

total <strong>of</strong> 20 <strong>of</strong>fspring and many <strong>of</strong> <strong>the</strong>se<br />

<strong>of</strong>fspring have survived and reproduced.<br />

The genetic restoration plan identified a goal<br />

<strong>of</strong> incorporating a 20% introgression <strong>of</strong><br />

Texas puma genes into <strong>the</strong> pan<strong>the</strong>r<br />

population and a preliminary assessment<br />

suggested that we may have achieved or<br />

slightly exceeded that level (Land and Lacy<br />

2000). As <strong>of</strong> January 2003, 5 <strong>of</strong> <strong>the</strong> original<br />

8 released Texas puma have since died and<br />

<strong>the</strong> remaining 3 females, thought to be<br />

reproductively senescent, were removed<br />

from <strong>the</strong> wild. We will continue monitoring<br />

pan<strong>the</strong>r genetic restoration by comparing<br />

reproductive performance, survival,<br />

phenotypic traits, and genetic characteristics<br />

among Texas and Florida descendants. Our<br />

goal is to develop a long-term management<br />

plan based on our study results to maintain<br />

genetic diversity, health, and long-term<br />

survival <strong>of</strong> <strong>the</strong> south Florida pan<strong>the</strong>r<br />

population.<br />

DISTRIBUTION AND ABUNDANCE<br />

Florida pan<strong>the</strong>rs occupy a core range in<br />

south Florida primarily in Collier, Hendry,<br />

Lee, and Dade counties. Major public lands<br />

include Big Cypress National Preserve,<br />

Everglades National Park, Florida Pan<strong>the</strong>r<br />

National Wildlife Refuge, Fakahatchee<br />

Strand State Preserve, Picayune Strand State<br />

Forest, and Okaloacoochee Slough State<br />

Forest. Large privately held ranches, used<br />

primarily for cattle and crop production, also<br />

constitute some <strong>of</strong> <strong>the</strong> most important<br />

habitat for pan<strong>the</strong>rs. Verified evidence,<br />

through road-kills, photos, or tracks, has<br />

also been found in Glades, Sarasota, and<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

FLORIDA STATUS REPORT · Lotz and Land 19<br />

Palm Beach Counties within <strong>the</strong> past 2 years<br />

(Land et al. 2002, Shindle et al. 2001).<br />

However, <strong>the</strong>se have all been dispersed or<br />

transient males. No females have been<br />

documented outside <strong>of</strong> <strong>the</strong> core range. One<br />

radio-collared male pan<strong>the</strong>r dispersed a<br />

straight-line distance <strong>of</strong> 224 km from his<br />

natal range (Maehr et al. 2002).<br />

The first Florida pan<strong>the</strong>r was radiocollared<br />

in 1981 by <strong>the</strong> Florida Game and<br />

Fresh Water Fish Commission (renamed to<br />

Florida Fish and Wildlife Conservation<br />

Commission in 2000). Throughout <strong>the</strong><br />

1980’s <strong>the</strong> population was estimated to be<br />

30-50 adults. The population has been<br />

increasing since about <strong>the</strong> mid 1990’s and<br />

today is estimated to be 80-100 adults. The<br />

release <strong>of</strong> Texas cougars for genetic<br />

restoration purposes in 1995 has contributed<br />

to this increase. Our population estimate is<br />

derived by counting currently radio-collared<br />

pan<strong>the</strong>rs and tallying observations <strong>of</strong><br />

uncollared pan<strong>the</strong>r sign encountered during<br />

yearly field activities.<br />

DEPREDATIONS AND HUMAN<br />

CONFLICTS<br />

FWC does not have a specific pan<strong>the</strong>r<br />

depredation or o<strong>the</strong>r human conflict protocol<br />

in place, but we do have a nuisance black<br />

bear policy that could provide guidance.<br />

The nuisance bear policy involves<br />

addressing <strong>the</strong> source <strong>of</strong> <strong>the</strong> problem,<br />

typically <strong>the</strong> removal or protection <strong>of</strong> bear<br />

attractants, prior to any stepwise progression<br />

<strong>of</strong> capture/handling <strong>of</strong> bears, removals, and<br />

ultimately, euthanasia. There have been no<br />

documented pan<strong>the</strong>r attacks on humans in<br />

Florida with only anecdotal accounts <strong>of</strong><br />

attacks prior to 1900 (Tinsley 1970). FWC<br />

regularly receives complaints about wildlife<br />

attacks on domestic livestock, many <strong>of</strong><br />

which are claimed to be pan<strong>the</strong>r<br />

depredations. However, upon investigation,<br />

<strong>the</strong> vast majority <strong>of</strong> <strong>the</strong>se incidents involve<br />

o<strong>the</strong>r predators including black bear, bobcat,<br />

fox, raccoon, opossum, coyote, and


20 FLORIDA STATUS REPORT · Lotz and Land<br />

domestic dog. We are aware <strong>of</strong> three valid<br />

pan<strong>the</strong>r depredations that were reported to<br />

FWC. The first involved a pan<strong>the</strong>r that<br />

seized a small dog by <strong>the</strong> head and<br />

subsequently dropped <strong>the</strong> dog alive after <strong>the</strong><br />

owner appeared at <strong>the</strong> door. A second<br />

depredation involved <strong>the</strong> killing <strong>of</strong> small<br />

goats from a rural homeowner’s yard in an<br />

area occupied by pan<strong>the</strong>rs. These<br />

complainants were given advice on how to<br />

protect <strong>the</strong>ir pets/livestock and to date no<br />

fur<strong>the</strong>r depredations have been reported.<br />

The last case was more complicated because<br />

it involved pan<strong>the</strong>rs that were taking<br />

advantage <strong>of</strong> a hunting preserve that was<br />

newly created by <strong>the</strong> Seminole Tribe on<br />

tribal lands. Non-native ungulates were<br />

stocked in an area known to be occupied by<br />

pan<strong>the</strong>rs and predictably, <strong>the</strong> pan<strong>the</strong>rs<br />

preyed upon <strong>the</strong>se ungulates. FWC and <strong>the</strong><br />

USFWS could do very little to address <strong>the</strong>se<br />

depredations because <strong>of</strong> <strong>the</strong> Endangered<br />

Species Act and because <strong>the</strong> preserve was<br />

developed on areas used by pan<strong>the</strong>rs.<br />

Although <strong>the</strong> tribe made a request for<br />

reimbursement <strong>of</strong> losses, no compensation<br />

was provided. Over time, <strong>the</strong> Seminole<br />

Tribe has adjusted <strong>the</strong> type <strong>of</strong> game animals<br />

that are stocked in <strong>the</strong> preserve, primarily<br />

stocking and selling wild hog hunts, and<br />

<strong>the</strong>se lower cost animals that are taken by<br />

pan<strong>the</strong>rs are less <strong>of</strong> a financial loss than <strong>the</strong><br />

various exotic deer species <strong>the</strong>y once<br />

stocked. Cattle ranchers apparently are<br />

unconcerned about potential pan<strong>the</strong>r<br />

depredations based on <strong>the</strong> lack <strong>of</strong><br />

complaints, and FWC food habits work has<br />

revealed that cattle are rarely taken by<br />

pan<strong>the</strong>rs. The presence <strong>of</strong> feral hogs on<br />

cattle ranches provide an abundant, easily<br />

taken prey base that may obviate <strong>the</strong> need<br />

for pan<strong>the</strong>rs to tackle cattle.<br />

RESEARCH AND PUBLICATIONS<br />

Current Research<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Florida Pan<strong>the</strong>r Genetic Restoration and<br />

Management<br />

This has been our focal study since 1995<br />

when 8 female Texas cougars were<br />

released to <strong>of</strong>fset <strong>the</strong> problems <strong>of</strong><br />

inbreeding. Genetic diversity and health<br />

<strong>of</strong> <strong>the</strong> Florida pan<strong>the</strong>r population needs<br />

to be restored to ensure survival, even<br />

with adequate habitat conservation and<br />

o<strong>the</strong>r enhancement measures. Genetic<br />

restoration is a direct and immediate<br />

action that will restore genetic variability<br />

and vitality for a healthier, more resilient<br />

population. The Plan for Genetic<br />

Restoration and Management <strong>of</strong> <strong>the</strong><br />

Florida Pan<strong>the</strong>r (Seal 1994) called for a<br />

20% introgression level <strong>of</strong> Texas genes<br />

throughout <strong>the</strong> population and<br />

preliminary analysis indicates we are on<br />

target. All Texas females have died or<br />

been removed. A minimum <strong>of</strong> 59<br />

intercross animals were produced and it<br />

is assumed that 44 still exist within <strong>the</strong><br />

population. Fifteen are radio-collared.<br />

This study was extended in order to<br />

collect and analyze critical samples from<br />

subsequent generations <strong>of</strong> Texas puma<br />

descendants. Our goal is to develop a<br />

long-term management plan based on<br />

our study results to maintain genetic<br />

diversity, health, and long-term survival<br />

<strong>of</strong> <strong>the</strong> south Florida pan<strong>the</strong>r population.<br />

A final report is anticipated next year.<br />

Feasibility <strong>of</strong> Using GPS Radio-collars<br />

on Florida Pan<strong>the</strong>rs<br />

The use <strong>of</strong> GPS technology in wildlife<br />

applications has garnered much interest<br />

in recent years but <strong>the</strong> current state <strong>of</strong><br />

<strong>the</strong> technology and its applicability to<br />

pan<strong>the</strong>rs has yet to be determined.<br />

Among <strong>the</strong> objectives <strong>of</strong> this study are to<br />

compare and evaluate GPS and aerial<br />

telemetry relocations, calculate <strong>the</strong><br />

percentage <strong>of</strong> successful GPS<br />

relocations, and evaluate <strong>the</strong> use <strong>of</strong> GPS<br />

collars on Florida pan<strong>the</strong>rs and make


ecommendations for future use. We<br />

placed 4 GPS collars from Telemetry<br />

Solutions (1130 Burnett Avenue, Suite J,<br />

Concord, CA 94520) on pan<strong>the</strong>rs during<br />

our 2001-2002 capture season. Two<br />

were Posrec collars that stored data on<br />

board until <strong>the</strong> collar was retrieved and<br />

<strong>the</strong> o<strong>the</strong>r two were Simplex units that<br />

had <strong>the</strong> ability to transmit data for<br />

remote downloads as well as store-onboard<br />

capabilities. Additionally, each<br />

collar was equipped with a VHF beacon<br />

in order to detect and recover carcasses,<br />

pinpoint and visit dens, and enable<br />

comparisons between GPS locations and<br />

aerial VHF relocations. Each pan<strong>the</strong>r<br />

equipped with a GPS collar was located<br />

thrice weekly during our regularly<br />

scheduled telemetry flights. All GPS<br />

collars have been recovered and we are<br />

currently evaluating data and<br />

performance. Two pan<strong>the</strong>rs wearing<br />

Posrec collars died 7 months after<br />

deployment, one Simplex model<br />

failed completely after only 4 months,<br />

and <strong>the</strong> remaining Simplex’s main<br />

battery failed after 6 months, disrupting<br />

GPS capabilities, but VHF function was<br />

maintained through <strong>the</strong> back-up battery.<br />

A final report is scheduled to be<br />

completed by <strong>the</strong> end <strong>of</strong> 2003.<br />

Feasibility <strong>of</strong> Using Remote Cameras to<br />

Survey Florida Pan<strong>the</strong>rs<br />

Most <strong>of</strong> what is known about Florida<br />

pan<strong>the</strong>rs, including population<br />

demographics, has come from radiotelemetry<br />

studies over <strong>the</strong> past 20 years.<br />

However, standardized survey<br />

techniques that estimate pan<strong>the</strong>r<br />

population parameters with associated<br />

measures <strong>of</strong> statistical confidence and<br />

that document significant changes in<br />

<strong>the</strong>se parameters over time have not<br />

been applied. The objective <strong>of</strong> this study<br />

is to assess whe<strong>the</strong>r infrared-triggered<br />

camera surveys for pan<strong>the</strong>rs provide<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

FLORIDA STATUS REPORT · Lotz and Land 21<br />

adequate data for inclusion into capturerecapture<br />

models based on <strong>the</strong> Lincoln-<br />

Peterson estimator. Remote camera<br />

surveys could complement existing labor<br />

and cost-intensive survey methodology<br />

to provide a more accurate estimate <strong>of</strong><br />

pan<strong>the</strong>r population parameters and<br />

document significant changes in <strong>the</strong>se<br />

parameters over time. Passive infrared<br />

cameras (Cam Trakker, CamTrak<br />

South Inc., Watkinsville, GA) were<br />

deployed systematically on two areas<br />

within <strong>the</strong> current occupied range <strong>of</strong><br />

Florida pan<strong>the</strong>rs. The Florida Pan<strong>the</strong>r<br />

National Wildlife Refuge provided an<br />

opportunity to assess camera survey<br />

methodology in a core area with a<br />

sample population <strong>of</strong> radio-collared and<br />

uncollared pan<strong>the</strong>rs. Long Pine Key<br />

within Everglades National Park<br />

provided an opportunity to assess<br />

camera survey methodology in a quasigeographically<br />

closed population <strong>of</strong><br />

radio-collared and uncollared pan<strong>the</strong>rs.<br />

Cameras were systematically placed in<br />

each study area and trials <strong>of</strong> 15 and 30<br />

days were run with 15 and 30 cameras<br />

per session. Field work was completed<br />

in 2002 and a final report is expected<br />

later this year.<br />

Feasibility <strong>of</strong> Extracting Florida Pan<strong>the</strong>r<br />

DNA from Scats<br />

Pan<strong>the</strong>r scats could potentially <strong>of</strong>fer <strong>the</strong><br />

safest and most cost effective tool for<br />

censussing numbers <strong>of</strong> pan<strong>the</strong>rs,<br />

measuring population genetic health, and<br />

identifying origins <strong>of</strong> Puma sign found<br />

outside <strong>of</strong> core pan<strong>the</strong>r areas. The<br />

purpose <strong>of</strong> this study is to evaluate <strong>the</strong><br />

use <strong>of</strong> pan<strong>the</strong>r scats as a source <strong>of</strong> DNA<br />

samples for on-going genetic<br />

monitoring. Existing tissue samples<br />

were used to calibrate and verify <strong>the</strong><br />

utility <strong>of</strong> extracting and analyzing DNA<br />

from scats. Scat collection routes were<br />

established along existing trails on four


22 FLORIDA STATUS REPORT · Lotz and Land<br />

areas and regularly surveyed by ATV.<br />

Over 400 km <strong>of</strong> trail were surveyed with<br />

scats encountered every 45 km on<br />

average. Additionally, scats were<br />

collected opportunistically during o<strong>the</strong>r<br />

field activities. Results are currently still<br />

being analyzed but microsatellite<br />

amplification <strong>of</strong> Florida pan<strong>the</strong>r DNA<br />

was successfully extracted in 60% <strong>of</strong> <strong>the</strong><br />

samples. Although collecting pan<strong>the</strong>r<br />

scat is labor intensive, utilizing DNA<br />

extracted from Florida pan<strong>the</strong>r scat holds<br />

promise as an unobtrusive technique to<br />

monitor <strong>the</strong> genetic health and individual<br />

makeup <strong>of</strong> <strong>the</strong> population. A final report<br />

will be completed later this year.<br />

Pan<strong>the</strong>r Peripheral Area Survey<br />

The only verified breeding population <strong>of</strong><br />

Florida pan<strong>the</strong>rs is in <strong>the</strong> sou<strong>the</strong>rn<br />

portion <strong>of</strong> <strong>the</strong> state, south <strong>of</strong> <strong>the</strong><br />

Caloosahatchee River and Lake<br />

Okeechobee, in <strong>the</strong> Big Cypress and<br />

Everglades physiographic regions. This<br />

population has been growing since <strong>the</strong><br />

mid 1990’s and so far 3 radio-collared<br />

pan<strong>the</strong>rs have crossed <strong>the</strong> river.<br />

Additionally, three uncollared pan<strong>the</strong>rs<br />

have been verified north <strong>of</strong> <strong>the</strong> river in<br />

recent years: two by tracks and/or<br />

photos, <strong>the</strong> o<strong>the</strong>r was road-killed. All <strong>of</strong><br />

<strong>the</strong>se have been males that dispersed<br />

from <strong>the</strong> core population. Only three are<br />

presumed to still be alive. Since resident<br />

male pan<strong>the</strong>rs typically encompass<br />

several females within <strong>the</strong>ir territory it is<br />

hypo<strong>the</strong>sized that searching <strong>the</strong>se ranges<br />

will afford <strong>the</strong> best opportunity <strong>of</strong><br />

finding o<strong>the</strong>r pan<strong>the</strong>rs if <strong>the</strong>y exist. This<br />

5-year study to determine <strong>the</strong> occurrence<br />

and status <strong>of</strong> pan<strong>the</strong>rs on peripheral<br />

areas <strong>of</strong> <strong>the</strong>ir presently known range has<br />

entered its final year. Systematic sign<br />

surveys have been conducted in areas<br />

where two male pan<strong>the</strong>rs had established<br />

territories. No sign <strong>of</strong> o<strong>the</strong>r pan<strong>the</strong>rs<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

was found. A final report is scheduled<br />

for 2003.<br />

Effect <strong>of</strong> Genetic Introgression on<br />

Prevalence and Intensity <strong>of</strong><br />

Gastrointestinal Helminth Infections in<br />

Florida Pan<strong>the</strong>rs<br />

The effects <strong>of</strong> genetic restoration <strong>of</strong> <strong>the</strong><br />

Florida pan<strong>the</strong>r are being examined on<br />

many fronts. Complementing o<strong>the</strong>r<br />

projects, this study will indirectly assess<br />

<strong>the</strong> suspected improvement in immune<br />

function in intergrades by comparing<br />

gastrointestinal tract parasite burdens to<br />

that seen in original Florida pan<strong>the</strong>rs.<br />

Concurrently we will assess <strong>the</strong> efficacy<br />

<strong>of</strong> field an<strong>the</strong>lmintic treatment <strong>of</strong><br />

pan<strong>the</strong>rs. Before introgression,<br />

gastrointestinal parasite burdens were<br />

assessed in 11original Florida pan<strong>the</strong>rs.<br />

Gastrointestinal tracts from a minimum<br />

<strong>of</strong> 7 pan<strong>the</strong>rs descended from Texas<br />

puma will be assessed by 2005 at which<br />

time a final report will be prepared.<br />

RECENT PUBLICATIONS<br />

CUNNINGHAM, M.W., M.R. DUNBAR, C.D.<br />

BUERGELT, B. HOMER, M. ROELKE-<br />

PARKER, S.K. TAYLOR, R. KING, S.B.<br />

CITINO, AND C. GLASS. 1999. Atrial<br />

septal defects in <strong>the</strong> Florida pan<strong>the</strong>r.<br />

Journal <strong>of</strong> Wildlife Diseases 35(3): 519-<br />

530.<br />

DEES, C.S., J.D. CLARK, AND F.T. VAN<br />

MANEN. 2001. Florida pan<strong>the</strong>r habitat<br />

use in response to prescribed fire.<br />

Journal <strong>of</strong> Wildlife Management 65:141-<br />

147.<br />

DUNBAR, M.R., M.W. CUNNINGHAM, AND<br />

S.T. LINDA. 1999. Vitamin A<br />

Concentrations in Serum and Liver from<br />

Florida Pan<strong>the</strong>rs. Journal <strong>of</strong> Wildlife<br />

Diseases 35(2): 171-177.<br />

JANIS, M.W. AND J.D. CLARK. 2002.<br />

Response <strong>of</strong> Florida pan<strong>the</strong>rs to<br />

recreational deer and hog hunting.


Journal <strong>of</strong> Wildlife Management.<br />

66:839-848.<br />

KRAMER, P.C. AND K.M. PORTIER. 2001.<br />

Modeling Florida pan<strong>the</strong>r movements in<br />

response to human attributes <strong>of</strong> <strong>the</strong><br />

landscape and ecological settings.<br />

Ecological Modeling 140:51-80.<br />

LAND, E.D., D.R. GARMIN, AND G.A. HOLT.<br />

1998. Monitoring female Florida<br />

pan<strong>the</strong>rs via cellular telephone. Wildlife<br />

Society Bulletin. 26(1): 29-31.<br />

LAND, E.D. AND R.C. LACY. 2000.<br />

Introgression level achieved through<br />

Florida pan<strong>the</strong>r genetic restoration.<br />

Endangered Species Update 17: 99-103.<br />

MAEHR, D.S. 1998. The Florida pan<strong>the</strong>r in<br />

modern mythology. Natural Areas<br />

Journal. 18(2): 179-184.<br />

MAEHR, D.S. AND J.P. DEASON. 2002.<br />

Wide-ranging carnivores and<br />

development permits: constructing a<br />

multi-scale model to evaluate impacts on<br />

<strong>the</strong> Florida pan<strong>the</strong>r. Clean Technologies<br />

and Environmental Policy. 3:398-406.<br />

MAEHR, D.S., R.C. LACY, E.D. LAND, O.L.<br />

BASS, JR., AND T.S. HOCTOR. 2002.<br />

Evolution <strong>of</strong> population viability<br />

assessments for <strong>the</strong> Florida pan<strong>the</strong>r: a<br />

multiperspective approach. Pages 284-<br />

311 in: Population Viability Analysis.<br />

University <strong>of</strong> Chicago Press, Chicago.<br />

MAEHR, D.S., E.D. LAND, D.B. SHINDLE,<br />

O.L. BASS, AND T.S. HOCTOR. 2002.<br />

Florida pan<strong>the</strong>r dispersal and<br />

conservation. Biological Conservation<br />

106:187-197.<br />

MANSFIELD, K.G. AND E.D. LAND. 2002.<br />

Cryptorchidism in Florida pan<strong>the</strong>rs:<br />

prevalence, features, and effects <strong>of</strong><br />

genetic restoration. Journal <strong>of</strong> Wildlife<br />

Diseases 38(4):693-698.<br />

ROTSTEIN, D.S., S. TAYLOR, J. HARVEY, J.<br />

BEAN. 1999. Hematologic effects <strong>of</strong><br />

Cytauxzoonosis in Florida Pan<strong>the</strong>rs and<br />

Texas Cougars in Florida. Journal <strong>of</strong><br />

Wildlife Diseases 35(3): 613-617.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

FLORIDA STATUS REPORT · Lotz and Land 23<br />

ROTSTEIN, D.S., R. THOMAS, K. HELMICK,<br />

S. CITINO, S. TAYLOR, M. DUNBAR.<br />

1999. Dermatophyte infections in freeranging<br />

Florida Pan<strong>the</strong>rs (Felis concolor<br />

coryi). Journal <strong>of</strong> Zoo and Wildlife<br />

Medicine 30(2): 281-284.<br />

ROTSTEIN, D.S., S.K. TAYLOR, J. BRADLEY,<br />

AND E.B. BREITSCHWERDT. 2000.<br />

Prevalence <strong>of</strong> Bartonella henselae<br />

antibody in Florida pan<strong>the</strong>rs. Journal <strong>of</strong><br />

Wildlife Diseases 36(1):157-160.<br />

ROTSTEIN, D.S., S.K. TAYLOR, A.<br />

BIRKENHAUER, M. ROELKE-PARKER,<br />

AND B.L. HOMER. 2002. Retrospective<br />

study <strong>of</strong> proliferative papillary vulvitis<br />

in Florida pan<strong>the</strong>rs. Journal <strong>of</strong> Wildlife<br />

Diseases 38:115-123.<br />

TAYLOR, S.K., E.D. LAND, M. LOTZ, M.<br />

ROELKE-PARKER, S.B. CITINO, AND D.<br />

ROTSTEIN. 1998. Anes<strong>the</strong>sia <strong>of</strong> freeranging<br />

Florida pan<strong>the</strong>rs, 1981-1998.<br />

<strong>Proceedings</strong> <strong>of</strong> American Association <strong>of</strong><br />

Zoo Veterinarians, Omaha, Nebraska.<br />

TAYLOR, S.K., C.D. BUERGELT, M.E.<br />

ROELKE-PARKER, B.L. HOMER, AND<br />

D.S. ROTSTEIN. 2002. Causes <strong>of</strong><br />

mortality <strong>of</strong> free-ranging Florida<br />

pan<strong>the</strong>rs. Journal <strong>of</strong> Wildlife Diseases<br />

38:107-114.<br />

LITERATURE CITED<br />

LAND, D., M. CUNNINGHAM, R. MCBRIDE,<br />

D. SHINDLE, AND M. LOTZ. 2002.<br />

Florida pan<strong>the</strong>r genetic restoration and<br />

management. Annual Report 2001-<br />

2002. Florida Fish and Wildlife<br />

Conservation Commission, Tallahassee.<br />

111pp.<br />

LAND, E.D. AND R.C. LACY. 2000.<br />

Introgression level achieved through<br />

Florida pan<strong>the</strong>r genetic restoration.<br />

Endangered Species Update 17: 99-103.<br />

MAEHR, D.S. 1990. The Florida pan<strong>the</strong>r<br />

and private lands. Conservation Biology<br />

4(2): 167-170.<br />

MAEHR, D.S., E.D. LAND, D.B. SHINDLE,<br />

O.L. BASS, AND T.S. HOCTOR. 2002.


24 FLORIDA STATUS REPORT · Lotz and Land<br />

Florida pan<strong>the</strong>r dispersal and<br />

conservation. Biological Conservation<br />

106:187-197.<br />

SEAL, U.S., AND WORKSHOP PARTICIPANTS.<br />

1992. Genetic management strategies<br />

and population viability <strong>of</strong> <strong>the</strong> Florida<br />

pan<strong>the</strong>r. Yulee, Florida: U.S. Fish and<br />

Wildlife Service.<br />

SEAL, U.S., editor. 1994. A plan for genetic<br />

restoration and management <strong>of</strong> <strong>the</strong><br />

Florida pan<strong>the</strong>r (Felis concolor coryi).<br />

Conservation Breeding Specialist Group,<br />

Apple Valley, MN. 24pp.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

SHINDLE, D., D. LAND, M. CUNNINGHAM,<br />

AND M. LOTZ. 2001. Florida pan<strong>the</strong>r<br />

genetic restoration and management.<br />

Annual Report 2000-2001. Florida Fish<br />

and Wildlife Conservation Commission,<br />

Tallahassee. 102pp.<br />

TINSLEY, J.B. 1970. The Florida Pan<strong>the</strong>r.<br />

Great Outdoors Publishing Company.<br />

St. Petersburg, FL.<br />

YOUNG, S.P. AND E.A. GOLDMAN. 1946.<br />

The puma – mysterious American cat.<br />

Dover Publications, Inc., New York.<br />

385pp.


IDAHO MOUNTAIN LION STATUS REPORT<br />

STEVE NADEAU, Wildlife Staff Biologist, Idaho Department <strong>of</strong> Fish and Game, 600 South<br />

Walnut, Box 25, Boise, Idaho 83707, USA, email: snadeau@idfg.state.id.us<br />

INTRODUCTION<br />

<strong>Lion</strong>s were classified as big game<br />

animals in 1972. The 1990 <strong>Mountain</strong> <strong>Lion</strong><br />

Management Plan, called for <strong>the</strong> reduction<br />

in harvest <strong>of</strong> female lions, and maintain a<br />

harvest <strong>of</strong> approximately 250 lions<br />

statewide. However, lion harvest peaked<br />

statewide in 1998 when 798 lions were<br />

harvested. Consequently, a new lion plan<br />

was developed to address <strong>the</strong> changes in <strong>the</strong><br />

populations and allow more hunting<br />

opportunity. Idaho completed <strong>the</strong> latest<br />

<strong>Mountain</strong> <strong>Lion</strong> Management Plan in 2002.<br />

The lion plan called for maintaining current<br />

lion distribution statewide as a goal.<br />

However, individual regions may adjust<br />

harvest to ei<strong>the</strong>r increase or decrease<br />

populations depending upon <strong>the</strong> objectives<br />

for that area. Seasons were made more<br />

lenient, running from August 30 – March 31<br />

in most units. In some areas, 2-lion bag<br />

limits were initiated. Hounds were allowed<br />

in most units, and non-resident hound<br />

hunting was expanded. Female quotas were<br />

still used in most <strong>of</strong> <strong>the</strong> sou<strong>the</strong>rn part <strong>of</strong> <strong>the</strong><br />

state.<br />

HISTORY<br />

The legal status and public perception <strong>of</strong><br />

mountain lions in Idaho has changed over<br />

time. In <strong>the</strong> late 1800’s and early 1900’s,<br />

mountain lions and o<strong>the</strong>r predators such as<br />

wolf, coyote, grizzly and black bears were<br />

perceived as significant threats to livestock<br />

and human interests and were systematically<br />

destroyed. Between 1915 and 1941, hunters<br />

employed cooperatively by <strong>the</strong> State,<br />

livestock associations, and <strong>the</strong> Federal<br />

25<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Government killed 251 mountain lions in<br />

Idaho; <strong>the</strong> take by private individuals is not<br />

known. During <strong>the</strong> period 1945-1958,<br />

bounties were paid for mountain lions in<br />

Idaho with an annual average <strong>of</strong> 80<br />

mountain lions turned in for payment<br />

(Figure 1). The 1953-54 winter period<br />

yielded <strong>the</strong> highest recorded bounty harvest<br />

<strong>of</strong> 144 mountain lions (Figure 1). Bounty<br />

payments ranged from $50 in <strong>the</strong> early<br />

1950’s to $25 per lion during <strong>the</strong> last 4 years<br />

<strong>of</strong> payments.<br />

<strong>Mountain</strong> lion sport harvest became<br />

increasingly popular after 1958. Average<br />

annual harvest was estimated at 142 lions<br />

from 1960 through 1971 (Figure 2). During<br />

this period <strong>the</strong>re were no restrictions or<br />

regulations on <strong>the</strong> harvest <strong>of</strong> mountain lions.<br />

An estimated 303 lions were harvested<br />

during <strong>the</strong> 1971-72 season.<br />

Research conducted by Maurice<br />

Hornocker in <strong>the</strong> Frank Church River <strong>of</strong> No-<br />

Number <strong>of</strong> <strong>Lion</strong>s Killed<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1950<br />

1951<br />

1952<br />

1953<br />

1954<br />

1955<br />

1956<br />

1957<br />

1958<br />

1959<br />

Figure 1. <strong>Mountain</strong> lion bounty records,<br />

1950 – 1959. From 1950-1954 bounty was $50<br />

per lion; 1955-1959 <strong>the</strong> bounty was $25 per<br />

lion.


26 IDAHO MOUNTAIN LION STATUS REPORT · Nadeau<br />

<strong>Lion</strong> Harvest<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

1960<br />

1962<br />

1964<br />

1966<br />

1968<br />

1970<br />

1972<br />

1974<br />

1976<br />

Unregulated Harvest Regulated<br />

1978<br />

1980<br />

Figure 2. Unregulated mountain lion harvest<br />

from 1960-71, and regulated harvest from<br />

1972 -1981.<br />

Return Wilderness from 1964-1973 added<br />

significantly to our knowledge. As a result<br />

<strong>of</strong> <strong>the</strong> research, <strong>the</strong> mountain lion was<br />

reclassified as a big game species in 1972.<br />

Harvest was <strong>the</strong>n able to be regulated and<br />

resulted in some closed units, bag limits, and<br />

shortened seasons. Mandatory reporting<br />

was started in 1973, and a tag has been<br />

required since 1975.<br />

Populations <strong>of</strong> elk and deer continued to<br />

increase across <strong>the</strong> state during <strong>the</strong> 1980’s<br />

and early 1990’s, and <strong>the</strong> resulting mountain<br />

lion population did as well. The apparent<br />

increase in lion populations allowed <strong>the</strong><br />

department to increase opportunity for<br />

harvest. Harvest continued to increase as a<br />

result <strong>of</strong> liberalized seasons and increased<br />

populations and peaked in 1997 (Figure 3).<br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Lion</strong>s were distributed across most <strong>of</strong> <strong>the</strong><br />

suitable habitat in <strong>the</strong> state (Figure 4).<br />

Management tended to keep lion<br />

populations at a low density in developed<br />

areas or areas with high road density.<br />

However, most <strong>of</strong> <strong>the</strong> areas that received<br />

high harvest lay adjacent to lightly roaded<br />

reservoir areas that seemed to continue to<br />

provide dispersing animals. Distribution<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

1982<br />

1984<br />

1986<br />

1988<br />

1990<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Harvest<br />

1992<br />

Year<br />

1994<br />

1996<br />

1998<br />

2000<br />

<strong>Lion</strong> Harvest<br />

Figure 3. Statewide mountain lion harvest.<br />

The year on <strong>the</strong> x-axis represents <strong>the</strong> date <strong>the</strong><br />

season started, i.e. seasons run from fall<br />

through spring.<br />

Figure 4. Statewide mountain lion harvest by<br />

management unit and lion DAU where<br />

rankings are based on lions harvested/100mi 2<br />

where very low=. 03, low=. 3-.5, moderate=.<br />

6-1.0, high=1.1-2.0, and very high=2.6-3.0.<br />

The shaded units have female lion quotas.


appeared to be somewhat stable, though<br />

overall abundance may be declining.<br />

Population estimates have not been<br />

made for Idaho in recent years, though some<br />

radio collaring mortality information in<br />

Idaho indicated a high rate <strong>of</strong> sustainable<br />

harvest in some areas. Given an estimated<br />

harvest rate statewide <strong>of</strong> approximately<br />

15%, we would estimate approximately<br />

4,600 lions (+ 2,000). Research has been<br />

ongoing to attempt to develop a population<br />

index, however, nothing has been finalized<br />

(Zager et al. 2002). All lions harvested must<br />

be reported. Pelts were tagged and a<br />

premolar was removed for aging. Prior to<br />

2000, lion ages were estimated using tooth<br />

drop measurements. Based on various tests,<br />

tooth sectioning replaced tooth drop as a<br />

more reliable estimate <strong>of</strong> age and has been<br />

used since 2000. For data analysis purposes,<br />

units were grouped by similar characteristics<br />

into Data Analysis Units (DAUs). Age data<br />

and harvest rates were used to attempt to<br />

identify population trends for a lion by<br />

DAU. Populations modeling using <strong>the</strong>se<br />

harvest data were used to estimate<br />

population demographics and relative<br />

abundance.<br />

<strong>Lion</strong> densities were highest in <strong>the</strong><br />

nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> state where white-tailed<br />

deer and elk were common. Harvest by<br />

DAU size was used to standardize and<br />

compare lion harvest rates and estimated<br />

lion abundance (Figure 4).<br />

HARVEST INFORMATION<br />

There were 99 big game management<br />

units in Idaho, which were grouped into 18<br />

mountain lion management DAUs (Figure<br />

4). The sou<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> state was<br />

predominantly managed under a female<br />

quota system, and <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong><br />

state was mostly general hunts with most<br />

seasons running from August 30 – March<br />

31. Quotas and seasons were set by unit or<br />

DAU, usually based on historical harvest<br />

rates, big game objectives, depredations,<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

IDAHO MOUNTAIN LION STATUS REPORT · Nadeau 27<br />

perceived lion population condition, lion<br />

hunter success rates and perceptions, public<br />

input, and commission desires.<br />

Biological objectives for lions were not<br />

well established by DAU, though age data<br />

were collected on all lions harvested. A<br />

minimum <strong>of</strong> 20% males 5+ years <strong>of</strong> age in<br />

<strong>the</strong> harvest was established as a test<br />

objective in some DAUs to adaptively<br />

manage populations by attempting to grow<br />

or reduce populations through harvest<br />

management, and monitor resultant age<br />

structures in <strong>the</strong> harvest. Regional wildlife<br />

managers in <strong>the</strong> state were given a great deal<br />

<strong>of</strong> flexibility to be able to set objectives for a<br />

given DAU. <strong>Lion</strong> harvest increased steadily<br />

through <strong>the</strong> 1980’s and 1990’s and peaked at<br />

798 mountain lions harvested in 1997. <strong>Lion</strong><br />

harvest declined in most areas <strong>of</strong> <strong>the</strong> state<br />

following <strong>the</strong> 1997 season despite a<br />

liberalized lion hunting season in most <strong>of</strong><br />

<strong>the</strong> state (Figure 3).<br />

Hunting with hounds accounted for<br />

about 80% <strong>of</strong> <strong>the</strong> annual lion harvest in<br />

Idaho. The rest <strong>of</strong> <strong>the</strong> harvest occurred<br />

incidentally to o<strong>the</strong>r big game hunting<br />

(13%), spot and stalk (5%), or predator<br />

calling (1%). The use <strong>of</strong> electronic calls<br />

was allowed in 2 management units where<br />

predation was a concern and access was<br />

limited. Dogs were prohibited through<br />

much <strong>of</strong> <strong>the</strong> general deer and elk rifle<br />

seasons. Pursuit with dogs was allowed in<br />

units with female quotas once <strong>the</strong> quota was<br />

reached. In a few <strong>of</strong> <strong>the</strong>se units, hunting for<br />

males was allowed once <strong>the</strong> female quota<br />

was reached.<br />

<strong>Mountain</strong> lion tag sales increased 25%<br />

from 1998–2002, and in 2002 were at an all<br />

time high <strong>of</strong> 20,640 total tags sold (Table 1).<br />

Reduced prices, increased nonresident sales<br />

<strong>of</strong> special tags, and liberalized seasons and<br />

nonresident hound hunter regulations all<br />

added to increased sales. Additionally, in<br />

some parts <strong>of</strong> <strong>the</strong> state outfitters were<br />

engaged to increase harvest <strong>of</strong> lions to help


28 IDAHO MOUNTAIN LION STATUS REPORT · Nadeau<br />

Table 1. <strong>Mountain</strong> lion tag sales in Idaho<br />

from 1998 through 2002.<br />

Year Resident<br />

Tags<br />

Nonresident<br />

Tags<br />

Total<br />

Tags<br />

Sold<br />

1998 16,196 351 16,547<br />

1999 17,072 813 17,885<br />

2000 18,369 961 19,330<br />

2001 18,561 888 19,449<br />

2002 19,757 883 20,640<br />

reduce predation problems on elk and<br />

bighorn sheep.<br />

DEPREDATIONS AND HUMAN<br />

CONFLICTS<br />

Currently, Idaho law allows for killing<br />

lions or bears that are in <strong>the</strong> act <strong>of</strong><br />

“molesting” livestock. This law also<br />

requires that lions killed in this fashion need<br />

to be reported to <strong>the</strong> Department. Idaho law<br />

also allows lions that are perceived as<br />

threats to human safety to be killed.<br />

Department policy provides that lions that<br />

have caused problems or have depredated<br />

should be captured and euthanized. Most<br />

depredations are reported to U.S. Wildlife<br />

Services and <strong>the</strong>y handle <strong>the</strong> removal.<br />

Policy also provides that lions that present a<br />

threat due to proximity to residential<br />

housing or o<strong>the</strong>r area <strong>of</strong> human habituation<br />

or activity should be moved or chased in a<br />

preemptive fashion. Depending on <strong>the</strong><br />

circumstance, if <strong>the</strong> animal has become<br />

habituated or caused problems, <strong>the</strong> lion can<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

be destroyed. Orphaned kittens are not<br />

rehabilitated for release back into <strong>the</strong> wild.<br />

Idaho averaged 3-4 safety related<br />

complaints annually from 1998-2002 and<br />

about 50% required capture or removal <strong>of</strong> a<br />

lion. There has been 1-recorded human<br />

injury in Idaho caused by lions, and that<br />

occurred in 1999 to a 13-year-old boy.<br />

<strong>Lion</strong> related depredations that required<br />

compensation averaged about 1-2 per year.<br />

Average annual compensation form 1998-<br />

2002 was $4717 for lion depredations on<br />

livestock. During that same time, 46 lions<br />

were removed due to depredation situations.<br />

RESEARCH<br />

The Department has been researching<br />

techniques for population monitoring in<br />

north central Idaho by conducting aerial<br />

track surveys (Gratson and Zager 2000), and<br />

a mark-recapture technique using rub<br />

stations and biopsy darts (Zager et al. 2002).<br />

These efforts are still preliminary in nature.<br />

LITERATURE CITED<br />

GRATSON, M.W., AND P. ZAGER. 2000. Elk<br />

ecology. Study IV. Factors influencing<br />

elk calf recruitment. Job No. 2. Calf<br />

mortality causes and rates. Federal Aid<br />

in Wildlife Restoration, Job Progress<br />

Report, W-160-R-26. Idaho Department<br />

<strong>of</strong> Fish and Game, Boise.<br />

ZAGER, P., M.W. GRATSON, AND C. WHITE.<br />

2002. Elk ecology. Study IV. Factors<br />

influencing elk calf recruitment. Job No.<br />

2. Calf mortality causes and rates.<br />

Federal Aid in Wildlife Restoration, Job<br />

Progress Report. W-160-R-29. Idaho<br />

Department <strong>of</strong> Fish and Game, Boise.


MONTANA MOUNTAIN LION STATUS REPORT<br />

RICH DeSIMONE, Montana Fish, Wildlife & Parks, 1420 East Sixth Avenue, Helena, MT<br />

59620, USA, email: rdesimone@state.mt.us<br />

ROSE JAFFE, Montana Fish, Wildlife & Parks, 1420 East Sixth Avenue, Helena, MT 59620,<br />

USA, email: rjaffe@state.mt.us<br />

INTRODUCTION<br />

<strong>Mountain</strong> lions in Montana are classified<br />

as a big game species. Overall management<br />

direction is provided in <strong>the</strong> Montana Fish,<br />

Wildlife & Parks’ (MFWP) 1996<br />

Environmental Impact Statement (EIS) –<br />

Management <strong>of</strong> <strong>Mountain</strong> <strong>Lion</strong>s in<br />

Montana. According to <strong>the</strong> EIS, objectives<br />

concerning lion management are “… to<br />

maintain mountain lion and prey<br />

populations, to maintain mountain lion<br />

populations at levels that are compatible<br />

with outdoor recreational desires, and to<br />

minimize human-lion conflicts and livestock<br />

depredation”.<br />

DISTRIBUTION AND ABUNDANCE<br />

Figure 1. Montana mountain lion hunting districts.<br />

29<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

<strong>Mountain</strong> lions are currently distributed<br />

over approximately 75% <strong>of</strong> <strong>the</strong> state. <strong>Lion</strong>s<br />

have filled habitats in western and central<br />

Montana and are continuing to expand in <strong>the</strong><br />

eastern part <strong>of</strong> <strong>the</strong> state. Montana does not<br />

estimate lion populations, however, trends<br />

are monitored through harvest/mortality<br />

data, tooth age information, damage/conflict<br />

reports, and information from houndsmen.<br />

HARVEST INFORMATION<br />

<strong>Lion</strong> harvest objectives are guided by<br />

balancing concern for human safety and<br />

demand for sport hunting. Montana’s 155<br />

deer and elk hunting districts are combined<br />

into 74 mountain lion hunting districts<br />

(Figure 1).


30 MONTANA MOUNTAIN LION STATUS REPORT · DeSimone and Jaffe<br />

Table 1. Montana lion hunting statistics, 1998-2002.<br />

1998 1999 2000 2001 2002<br />

License sales<br />

Resident 5421 5886 5138 5116 6337<br />

Non-resident 510 519 493 421 281<br />

Total 5931 6405 5631 5537 6618<br />

<strong>Lion</strong> Quota<br />

Harvest<br />

868 758 661 620 581<br />

Female 417 335 293 252 188<br />

Male 351 319 291 257 219<br />

Unknown 8 0 0 0 0<br />

Total 776 654 584 509 407<br />

Harvest is regulated through quotas and<br />

only one lion can be taken per hunter per<br />

year. Quotas include any lion, male and<br />

female, and female sub quotas. During <strong>the</strong><br />

fall hunting season (last week <strong>of</strong> Oct<br />

through Nov), <strong>the</strong> use <strong>of</strong> dogs is not<br />

allowed. Harvest during <strong>the</strong> fall season has<br />

been in affect for 4 years and less than 10<br />

lions were harvested each year (Table 1).<br />

Hunting with dogs is allowed during <strong>the</strong><br />

winter season (Dec 1 – Apr 14) and accounts<br />

for over 95% <strong>of</strong> <strong>the</strong> harvest. Licensed<br />

hunters are also allowed to chase lions<br />

during <strong>the</strong> winter season. Recent legislation<br />

will allow <strong>the</strong> purchase <strong>of</strong> non-harvest chase<br />

licenses.<br />

DEPREDATIONS AND HUMAN<br />

INTERACTION/CONFLICTS<br />

MFWP’s <strong>Mountain</strong> <strong>Lion</strong> Depredation<br />

Table 2. Montana mountain lion incidents and removals, 1998-2002.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

and Control Guidelines are used to deal with<br />

different types <strong>of</strong> incidents. Depending on<br />

<strong>the</strong> situation, management actions include<br />

education, relocation, and removal (Table<br />

2). Montana does not pay for losses<br />

attributed to lions.<br />

RESEARCH<br />

Garnet <strong>Mountain</strong>s – <strong>Mountain</strong> <strong>Lion</strong><br />

Research, 1998 – present.<br />

The goal is to document <strong>the</strong> influence <strong>of</strong><br />

hunting on population characteristics and<br />

evaluate <strong>the</strong> ability <strong>of</strong> various survey<br />

techniques to detect trends in lion<br />

abundance.<br />

1998 1999 2000 2001 2002<br />

Incidents 1<br />

Public safety 41 18 37 30 20<br />

Depredation 2 58 44 35 37 29<br />

Total<br />

Removals<br />

99 62 72 67 49<br />

Public safety 20 2 3 5 2<br />

Depredation 30 20 20 11 14<br />

Total 50 22 23 16 16<br />

1<br />

Incident: A conflict between a human and lion that may have serious results (i.e. a lion killing a dog or a lion that must be<br />

forced to back down).<br />

2<br />

Depredation: Includes death <strong>of</strong> pets and death and injury <strong>of</strong> livestock.


NEVADA MOUNTAIN LION STATUS REPORT<br />

RUSSELL WOOLSTENHULME, Nevada Department <strong>of</strong> Conservation & Natural Resources,<br />

1100 Valley View Road, Reno, NV 89512, USA, email: rwoolstenhulme@ndow.org<br />

INTRODUCTION<br />

The Nevada Division <strong>of</strong> Wildlife<br />

completed its Comprehensive <strong>Mountain</strong><br />

<strong>Lion</strong> Management Plan in January 1995.<br />

The Nevada Board <strong>of</strong> Wildlife<br />

Commissioners approved <strong>the</strong> plan in<br />

October <strong>of</strong> that year. The plan is scheduled<br />

for revision during 2003.<br />

The goals and objectives <strong>of</strong> <strong>the</strong> mountain<br />

lion plan are to maintain lion distribution in<br />

reasonable densities throughout Nevada, to<br />

control mountain lions creating a public<br />

safety hazard or causing property damage,<br />

and to provide recreational, educational and<br />

scientific use opportunities <strong>of</strong> <strong>the</strong> mountain<br />

lion resource. Additional goals include<br />

maintaining a balance between mountain<br />

lions and <strong>the</strong>ir prey, and finally to manage<br />

mountain lions as a metapopulation.<br />

The mountain lion’s legal classification<br />

in Nevada was changed by regulation from<br />

unprotected (predator) to game animal in<br />

1965. The change in classification resulted<br />

in <strong>the</strong> requirement <strong>of</strong> a valid hunting license<br />

to hunt mountain lion, along with some<br />

restrictions in <strong>the</strong> method <strong>of</strong> take. This<br />

provision precluded <strong>the</strong> taking <strong>of</strong> lions at<br />

any time o<strong>the</strong>r than from sunrise to sunset<br />

and it also defined legal weapons as<br />

shotgun, rifle, or bow and arrow. The<br />

season was defined as ei<strong>the</strong>r sex, yearround,<br />

and no limit was set nor was a tag<br />

required. <strong>Mountain</strong> lion harvest<br />

management has changed substantially from<br />

1965 to <strong>the</strong> present.<br />

In 1968, a tag requirement was<br />

instituted, and although no limits were<br />

established, it became possible to record<br />

31<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

sport hunter harvest. Ano<strong>the</strong>r major change<br />

occurred in 1970, when a limit <strong>of</strong> one lion<br />

per person was set, and a six-month season<br />

was established. During that year, <strong>the</strong><br />

requirement that all harvested lions be<br />

validated by a representative <strong>of</strong> <strong>the</strong><br />

Department within five days after <strong>the</strong> kill<br />

was also established. This regulation<br />

presented <strong>the</strong> Department <strong>the</strong> first real<br />

opportunity to collect biological data from<br />

<strong>the</strong> mountain lion.<br />

In 1972, <strong>the</strong> Nevada Department <strong>of</strong><br />

Wildlife initiated a study <strong>of</strong> <strong>the</strong> mountain<br />

lion as a part <strong>of</strong> <strong>the</strong> Ruby-Butte deer project<br />

in eastern Nevada. The objective was to<br />

determine <strong>the</strong> status <strong>of</strong> lion populations<br />

within this high-density deer area, and, to<br />

evaluate <strong>the</strong>m in relation to deer<br />

populations. Within two years, this<br />

objective was changed to: a) establish<br />

population estimates <strong>of</strong> mountain lions by<br />

mountain range or management area<br />

statewide, b) establish basic habitat<br />

requirements, c) establish a harvest<br />

management program. From that period on,<br />

increased emphasis was placed upon lion<br />

capture and marking with <strong>the</strong> more<br />

sophisticated telemetry devices which were<br />

being manufactured. This program involved<br />

lion monitoring from both land and air and<br />

was instrumental in expanding our life<br />

history information base, as well as<br />

providing an approach toward estimating <strong>the</strong><br />

annual population status in key mountain<br />

ranges. The findings from this study were<br />

<strong>the</strong>n utilized in formulating an approach<br />

toward estimating statewide lion<br />

populations. This ten (10) year study formed


32 NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme<br />

<strong>the</strong> basics for most management activities<br />

that have been implemented since<br />

publication <strong>of</strong> this study in 1983.<br />

In 1976, 26 mountain lion management<br />

areas were described statewide, and a<br />

harvest quota established for each to control<br />

<strong>the</strong> sport harvest. This “Controlled Quota<br />

Hunt” was <strong>the</strong> most restrictive season ever<br />

established for mountain lion in Nevada.<br />

In 1979, <strong>the</strong> “Controlled Quota Hunt”<br />

was modified utilizing six management<br />

areas whereby a harvest objective was<br />

established which allowed <strong>the</strong> hunting <strong>of</strong><br />

lions in each <strong>of</strong> <strong>the</strong> six areas until <strong>the</strong><br />

predetermined number <strong>of</strong> lion were taken.<br />

In 1981, <strong>the</strong> “Harvest Objective” hunting<br />

season concept was applied statewide.<br />

Initially this system required a hunter to<br />

obtain a free hunt permit for <strong>the</strong> opportunity<br />

to hunt in one (1) management area. In<br />

1994, hunters were allowed to obtain a free<br />

hunt permit that authorized <strong>the</strong> hunter to<br />

hunt in two (2) management areas until <strong>the</strong><br />

established harvest objective was reached.<br />

Both <strong>of</strong> <strong>the</strong>se permit systems allowed<br />

hunters to change management areas at will<br />

as long as <strong>the</strong> harvest objective had not been<br />

reached in <strong>the</strong> desired management area(s).<br />

In 1995, <strong>the</strong> hunt permit approach was<br />

modified to eliminate <strong>the</strong> physical issuance<br />

<strong>of</strong> a permit in favor <strong>of</strong> establishing a 1-800<br />

telephone number. This system allows<br />

hunters to hunt in any management area in<br />

which <strong>the</strong> harvest objective has not been<br />

reached. The hunter must, however, call <strong>the</strong><br />

1-800 number before starting to hunt to<br />

determine which management area(s) are<br />

still open to hunting.<br />

In 1997, changes were made to mountain<br />

lion regulations to increase mountain lion<br />

harvest, while maintaining <strong>the</strong> integrity <strong>of</strong><br />

<strong>the</strong> harvest objective limits system. Those<br />

changes included <strong>the</strong> reduction <strong>of</strong> tag fees,<br />

over-<strong>the</strong>-counter tag sales, increasing bag<br />

limits from one tag per hunter to two tags<br />

per hunter, and consolidation <strong>of</strong> some <strong>of</strong> <strong>the</strong><br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

harvest unit groups.<br />

In 1998, Nevada’s sou<strong>the</strong>rn region was<br />

modified to provide for a year-round hunting<br />

season on mountain lions. The entire state<br />

went to a year-round season in 2001.<br />

New changes were made again for <strong>the</strong><br />

2003 season. These changes modified<br />

harvest unit groups from 24 groups<br />

throughout <strong>the</strong> state to three statewide<br />

regions corresponding with <strong>the</strong> Division’s<br />

three management regions. The mountain<br />

lion season continues to be year-round but<br />

season dates were changed to March 1 st <strong>of</strong><br />

each year to <strong>the</strong> last day <strong>of</strong> February,<br />

corresponding with <strong>the</strong> dates on a Nevada<br />

hunting license.<br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Mountain</strong> lions seem well adapted to <strong>the</strong><br />

wide variety <strong>of</strong> habitat and environmental<br />

conditions that exist in Nevada. They have<br />

been observed to live or wander through<br />

almost every mountain range from <strong>the</strong><br />

Mojave Desert in extreme sou<strong>the</strong>rn Nevada<br />

to alpine forests at <strong>the</strong> highest elevations in<br />

<strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> state. Distribution<br />

appears to be primarily influenced by prey<br />

availability, and has remained fairly<br />

consistent through time.<br />

<strong>Mountain</strong> lion populations are estimated<br />

utilizing a life table model (retrospective<br />

harvest/ mortality). The model utilizes<br />

known harvest/ mortality rates and<br />

recruitment rates (as determined from markrecapture<br />

and telemetry studies) to calculate<br />

a retrospective estimate <strong>of</strong> minimum viable<br />

population size needed to sustain known<br />

harvest rates over <strong>the</strong> same time period.<br />

Although no defined confidence limit is<br />

used during this process, our confidence in<br />

this model is relatively high based on <strong>the</strong><br />

fact that harvest rates have continued over<br />

time at a constant rate without signs <strong>of</strong><br />

extirpation, reduced harvest rates, or<br />

increased average age <strong>of</strong> harvested lion.<br />

Based on our current estimation methods,


NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme 33<br />

Table 1. <strong>Mountain</strong> lion tag sales, sport hunter harvest, and hunter success by class <strong>of</strong> hunter.<br />

Tag Sales Harvest Hunter Success<br />

Year Resident Nonresident Total Resident Nonresident Total Resident Nonresident Total<br />

1998 643 124 767 73 67 140 11% 54% 18%<br />

1999 680 109 789 70 56 126 10% 51% 16%<br />

2000 883 169 1052 104 81 185 12% 48% 18%<br />

2001 838 98 936 103 58 161 12% 59% 17%<br />

2002 1030 202 1232 105 63 168 10% 31% 14%<br />

2003 1060 131 1191 89 39 128 8% 30% 11%<br />

Total 5,134 833 5,967 544 364 908 11% 44% 15%<br />

Average 856 139 995 91 61 151 11% 44% 15%<br />

lion populations within Nevada are between<br />

3000-4000 animals.<br />

HARVEST INFORMATION<br />

<strong>Mountain</strong> lions have been classified as a<br />

big game species since 1965. They have<br />

been hunted annually since that time. A<br />

Nevada resident mountain lion tag costs<br />

$25.00, and a Nevada nonresident mountain<br />

lion tag costs $100.00. On <strong>the</strong> average,<br />

nonresident hunters account for<br />

approximately 14% <strong>of</strong> tag sales, but harvest<br />

a greater proportion <strong>of</strong> lions than do resident<br />

hunters (Table 1). Total sport hunter harvest<br />

has averaged 151 lions per year for <strong>the</strong> last 6<br />

years (Table 2).<br />

The open season for hunting mountain<br />

lions in Nevada currently runs year-round<br />

(March 1 – last day <strong>of</strong> February) (Table 3).<br />

Table 2. <strong>Mountain</strong> lion harvest by harvest type and sex.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Any legal weapon may be used to harvest a<br />

mountain lion, and dogs may be used to hunt<br />

a mountain lion under <strong>the</strong> authority <strong>of</strong> a<br />

current State <strong>of</strong> Nevada hunting license and<br />

mountain lion tag. Because <strong>the</strong> mountain<br />

lion season is year-round no pursuit only<br />

season exists. A resident or a non-resident is<br />

eligible to obtain two mountain lion tags<br />

each year. A person who harvests a<br />

mountain lion in Nevada must, within 72<br />

hours after harvesting it, personally present<br />

<strong>the</strong> skull and hide to a representative <strong>of</strong> <strong>the</strong><br />

division for inspection. The representative<br />

shall affix a State <strong>of</strong> Nevada mountain lion<br />

seal permanently to <strong>the</strong> hide. A seal must be<br />

permanently affixed to <strong>the</strong> hide <strong>of</strong> a<br />

mountain lion before it can be possessed by<br />

an individual or removed from <strong>the</strong> state. It<br />

is unlawful to kill a female mountain lion<br />

Sport Hunter Harvest Depredation Take Total<br />

Year Male Female Total Male Female Total Male Female Total<br />

1998 85 55 140 12 8 20 97 63 160<br />

1999 77 49 126 12 10 22 89 59 148<br />

2000 102 83 185 8 3 11 110 86 196<br />

2001 95 66 161 8 8 16 103 74 177<br />

2002 99 69 168 10 16 26 109 85 194<br />

2003 77 51 128 7 8 15 84 59 143<br />

Average 89 62 151 10 9 18 99 71 170


34 NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme<br />

Table 3. Nevada <strong>Mountain</strong> <strong>Lion</strong> Units and Quotas 2003 – 2005.<br />

Unit Group<br />

UNIT 1 (Western Region)<br />

011 - 015, 021, 022, 031,<br />

032, 034, 035, 041 - 046,<br />

051, 181 – 184, 192, 194 -<br />

196, 201 - 206, 291<br />

2003-2004 Season<br />

Dates<br />

March 1, 2003 –<br />

Feb 29, 2004<br />

2003-2004<br />

Harvest Objectives<br />

114<br />

2004-2005 Season<br />

Dates<br />

March 1, 2004 –<br />

Feb 28, 2005<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

2004-2005<br />

Harvest Objectives<br />

033 Closed 0 Closed 0<br />

UNIT 2 (Eastern Region)<br />

061, 062, 064 – 068, 071 -<br />

078, 081, 101 – 108, 111 –<br />

115, 121, 131 – 134, 141 –<br />

145, 151, 152, 154, 155<br />

079*<br />

UNIT 3 (Sou<strong>the</strong>rn Region)<br />

161 - 164, 171 - 173, 211,<br />

212, 221 – 223, 231, 241 –<br />

244, 251 - 253, 261 - 268,<br />

271 – 272<br />

March 1, 2003 –<br />

Feb 29, 2004<br />

March 1, 2003 –<br />

Feb 29, 2004<br />

March 1, 2003 –<br />

Feb 29, 2004<br />

163<br />

4<br />

68<br />

March 1, 2004 –<br />

Feb 28, 2005<br />

March 1, 2004 –<br />

Feb 28, 2005<br />

March 1, 2004 –<br />

Feb 28, 2005<br />

280 – 284 Closed 0 Closed 0<br />

* Interstate hunt with Utah. Nevada and Utah hunters may hunt within open units in both states. Nevada hunters<br />

hunting in Utah must abide by Utah regulations.<br />

that is accompanied by a spotted kitten, or to<br />

kill or possess a spotted mountain lion<br />

kitten. It is also unlawful in Nevada to trap<br />

a mountain lion, if a mountain lion is<br />

accidentally trapped or killed, <strong>the</strong> person<br />

trapping or killing it shall report <strong>the</strong> trapping<br />

or killing within 48 hours to <strong>the</strong> division.<br />

The animal must be disposed <strong>of</strong> in<br />

accordance with state law.<br />

<strong>Mountain</strong> lion harvest objectives are<br />

calculated for each administrative region on<br />

a semi-annual basis using standardized<br />

methodology. Harvest objectives are<br />

calculated and recommended in order to<br />

achieve a specific management action over a<br />

short-term period (no more than two years).<br />

Management actions may be designed to<br />

increase, stabilize and maintain, or decrease<br />

mountain lion populations within each <strong>of</strong> <strong>the</strong><br />

three administrative regions in Nevada.<br />

114<br />

163<br />

Calculations <strong>of</strong> harvest objectives by<br />

administrative region incorporate <strong>the</strong> use <strong>of</strong><br />

scientific data to determine <strong>the</strong> current<br />

population trend and population density. A<br />

“political index” may be employed to adjust<br />

harvest objectives within smaller geographic<br />

areas (big game management areas) in order<br />

to achieve <strong>the</strong> desired management goal.<br />

Biologists make annual adjustments to<br />

harvest objective recommendations for each<br />

administrative region only after careful<br />

review <strong>of</strong> <strong>the</strong> following data and information<br />

that is collected, assembled and distributed<br />

by <strong>the</strong> Game Bureau by October <strong>of</strong> each<br />

year.<br />

A. Data used to assess population trend,<br />

including, but not limited to:<br />

4<br />

68<br />

1) The current regional population<br />

model.


2) Sex, weight and age data from<br />

harvested mountain lions for <strong>the</strong><br />

previous recording period (March 1<br />

- February 28).<br />

B. Data used to assess population density,<br />

including, but not limited to:<br />

1) The current regional population<br />

model.<br />

2) Data showing <strong>the</strong> unit <strong>of</strong> effort to<br />

observe or harvest mountain lions.<br />

3) Average weight information,<br />

comparing weights <strong>of</strong> harvested<br />

animals by sex and cohort group to<br />

<strong>the</strong> long-term data set (1968 -<br />

2003).<br />

C. Data to quantify “bio-political”<br />

considerations, including, but not<br />

limited to:<br />

1) A summary <strong>of</strong> <strong>the</strong> public safety<br />

complaint forms involving<br />

mountain lions as received by <strong>the</strong><br />

Bureau for <strong>the</strong> previous recording<br />

period.<br />

2) A report <strong>of</strong> damage to private<br />

property caused by mountain lions<br />

as annually prepared by ADC.<br />

3) A prey species accounting<br />

spreadsheet as prepared by <strong>the</strong><br />

region for <strong>the</strong> previous recording<br />

period. Adjustments from <strong>the</strong><br />

baseline harvest objective level for<br />

each administrative region will be<br />

recommended in order to achieve<br />

<strong>the</strong> short-term (two-year) goal <strong>of</strong><br />

maintaining, increasing, or<br />

decreasing mountain lion<br />

populations within <strong>the</strong> respective<br />

administrative region, utilizing<br />

harvest management as <strong>the</strong> primary<br />

tool to achieve <strong>the</strong> desired<br />

population goal.<br />

See Figure 1 for State <strong>of</strong> Nevada<br />

mountain lion hunt unit reference map.<br />

NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme 35<br />

Figure 1. Nevada mountain lion hunt unit<br />

reference map.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

DEPREDATIONS AND HUMAN<br />

INTERACTIONS/CONFLICTS<br />

The Nevada Division <strong>of</strong> Wildlife<br />

Comprehensive Management Plan<br />

specifically addresses policy and procedure<br />

for dealing with nuisance or problem<br />

mountain lions.<br />

The Division <strong>of</strong> Wildlife is responsible<br />

by statute for controlling wildlife causing<br />

damage to personal property or endangering<br />

personal safety. The Division also has a<br />

responsibility to provide sport-hunting<br />

opportunities to Nevada sportsmen. This<br />

protocol sets forth procedures to be followed<br />

in controlling and preventing lion damage,<br />

addressing public safety issues and<br />

responding to sport hunting opportunities.<br />

In carrying out this policy where mountain<br />

lion/human interactions are involved, agents<br />

shall have <strong>the</strong> discretion to choose <strong>the</strong> most


36 NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme<br />

applicable management option, following<br />

guidelines outlined in this protocol. All<br />

efforts will be directed at <strong>the</strong> individual lion<br />

causing <strong>the</strong> problem.<br />

<strong>Mountain</strong> lion/human interactions have<br />

increased throughout <strong>the</strong> West and in<br />

Nevada in <strong>the</strong> last several decades. During<br />

<strong>the</strong> same period, <strong>the</strong> number <strong>of</strong> depredation<br />

complaints and <strong>the</strong> number <strong>of</strong> lions taken on<br />

depredation complaints has also increased.<br />

The Division desires to reduce multiple<br />

depredations from <strong>the</strong> same animal and<br />

prevent harm to humans.<br />

The Division recognizes three distinct<br />

categories <strong>of</strong> mountain lions involved in<br />

human/lion interactions.<br />

A. Nuisance <strong>Lion</strong> - a lion involved in a<br />

direct meeting with a human but did<br />

not exhibit aggressive behavior toward<br />

<strong>the</strong> human, a lion repeatedly observed<br />

in an area, or a situation where<br />

personal property is at risk.<br />

B. Depredating <strong>Lion</strong> - a lion that has<br />

injured or killed livestock or domestic<br />

pets.<br />

C. Dangerous or Aggressive <strong>Lion</strong> - a<br />

lion that has exhibited aggressive<br />

behavior towards humans. A lion that<br />

has an unnatural interest in humans<br />

without provocation and is perceived<br />

to be a threat to public safety. A lion<br />

located in a place or situation where<br />

human safety is <strong>of</strong> concern may be<br />

considered dangerous.<br />

Various management options are<br />

available to Division employees when a<br />

mountain lion conflict arises. The Division<br />

employee responding to or assigned to<br />

handle a lion/human conflict will have <strong>the</strong><br />

primary responsibility to assess mountain<br />

lion involvement in an incident and conduct<br />

<strong>the</strong> necessary investigation. Agents may be<br />

required to make an assessment "on <strong>the</strong><br />

spot" or if time permits make an assessment<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

with consultation.<br />

At all opportunities, <strong>the</strong> Division will<br />

provide educational and informational<br />

materials to individuals concerned with lion<br />

management and people-lion conflict<br />

prevention. These materials will include<br />

options for pet and livestock protection and<br />

avoidance <strong>of</strong> dangerous encounters with<br />

mountain lions. Site-specific education and<br />

prevention efforts will be made in historic<br />

conflict areas.<br />

A field response by ei<strong>the</strong>r a Division<br />

employee or his/her designated agent is<br />

required for all lion/human interactions<br />

involving <strong>the</strong> categories <strong>of</strong> lions defined.<br />

1. Nuisance <strong>Mountain</strong> <strong>Lion</strong>s<br />

a. No management action combined<br />

with education effort.<br />

b. Deterrent methods combined with<br />

education effort.<br />

c. Capture, mark and relocate cougars<br />

if deterrent methods are<br />

unsuccessful or impractical. <strong>Lion</strong>s<br />

identified for relocation will be<br />

transported to <strong>the</strong> following release<br />

sites in priority order.<br />

1) Instate release locations within<br />

low conflict areas<br />

2) Out <strong>of</strong> state governmental<br />

agencies<br />

3) University or research facilities<br />

4) Zoological gardens or Zoos<br />

d. Nuisance lions will be destroyed if<br />

relocating or deterrent methods are<br />

unsuccessful or impractical.<br />

2. Depredation <strong>Mountain</strong> <strong>Lion</strong>s<br />

a. No management action combined<br />

with education effort.<br />

b. Deterrent methods including<br />

prevention materials (if applicable)<br />

combined with education effort.<br />

c. Capture, mark and relocate cougars<br />

if deterrent methods are<br />

unsuccessful or impractical. <strong>Lion</strong>s


identified for relocation will be<br />

transported to <strong>the</strong> following release<br />

sites in priority order.<br />

1) Instate release locations within<br />

low conflict areas<br />

2) Out <strong>of</strong> state governmental<br />

agencies<br />

3) University or research facilities<br />

4) Zoological gardens or Zoos<br />

d. Depredating lions will be destroyed<br />

if deterrent methods or live capture<br />

is unsuccessful or impractical.<br />

3. Aggressive (Dangerous) <strong>Mountain</strong><br />

<strong>Lion</strong>s<br />

a. If a lion is dangerous because <strong>of</strong> its<br />

location and not its behavior it may<br />

be trapped, marked and relocated.<br />

If a lion is frequenting a city or<br />

town, it may be destroyed if capture<br />

methods fail or are impractical.<br />

<strong>Lion</strong>s identified for relocation will<br />

be transported to <strong>the</strong> following<br />

release sites in priority order.<br />

1) Instate release locations within<br />

low conflict areas<br />

2) Out <strong>of</strong> state governmental<br />

agencies<br />

3) University or research facilities<br />

4) Zoological gardens or Zoos<br />

b. If <strong>the</strong> mountain lion is dangerous<br />

because it has exhibited aggressive<br />

behavior toward humans or is<br />

o<strong>the</strong>rwise perceived to be a threat to<br />

human safety, or if <strong>the</strong> lion is<br />

involved in an attack on a human,<br />

destroy and necropsy <strong>the</strong> lion.<br />

<strong>Lion</strong>s exhibiting aggressive<br />

behavior in remote areas should not<br />

be killed but instead an aggressive<br />

publicity and educational campaign<br />

should be made to alert people <strong>of</strong><br />

<strong>the</strong> danger in <strong>the</strong> remote area and<br />

promote human avoidance <strong>of</strong> <strong>the</strong><br />

area over <strong>the</strong> short-term.<br />

NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme 37<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

c. A detailed narrative report on each<br />

incident involving handling <strong>of</strong><br />

dangerous lions will be prepared by<br />

<strong>the</strong> agent in control <strong>of</strong> <strong>the</strong> incident<br />

and forwarded to <strong>the</strong> Supervising<br />

Regional Game Biologist.<br />

<strong>Mountain</strong> lion incidents involving<br />

attacks or injury to people will be<br />

immediately reported through <strong>the</strong><br />

chain <strong>of</strong> command to <strong>the</strong> Regional<br />

Manager, Administrator, Chief <strong>of</strong><br />

Game and Chief Game Warden.<br />

All lions destroyed will be reported<br />

on <strong>the</strong> 351-harvest form. A copy <strong>of</strong><br />

<strong>the</strong> detailed report, including any<br />

necropsy, coroner's report or o<strong>the</strong>r<br />

supporting information shall be sent<br />

to <strong>the</strong> Game bureau staff biologist<br />

responsible for mountain lions. A<br />

lion/human interaction form will be<br />

completed for each interaction.<br />

In those incidences where control<br />

becomes necessary, a regional list <strong>of</strong> persons<br />

who have requested consideration and are<br />

qualified to do control work, including<br />

private hunters/ trappers and outfitters/<br />

guides will be a source <strong>of</strong> control, as well as<br />

U.S.D.A. APHIS/ADC personnel.<br />

Hunters/trappers, outfitters/guides or<br />

U.S.D.A. agents will not initiate control<br />

unless requested to do so by <strong>the</strong> Division.<br />

Hunters or trappers may be authorized to<br />

control problem animals during open or<br />

closed seasons. The hunter or trapper will<br />

buy a license and tag for use during <strong>the</strong> open<br />

season until <strong>the</strong> hunter or trapper's tag is<br />

filled. The hunter may continue control<br />

work after <strong>the</strong> tag is filled only under <strong>the</strong><br />

authority <strong>of</strong> a depredation permit. Hunting<br />

during a closed season will be conducted<br />

only under <strong>the</strong> authority <strong>of</strong> a depredation<br />

permit. Depredation permits will only be<br />

issued to landowners/livestock owners for<br />

<strong>the</strong> control <strong>of</strong> specific depredating lions.<br />

Hunters or trappers may keep <strong>the</strong> lion if<br />

harvested under <strong>the</strong> authority <strong>of</strong> a valid


38 NEVADA MOUNTAIN LION STATUS REPORT · Woolstenhulme<br />

license and tag. All o<strong>the</strong>r lions become <strong>the</strong><br />

property <strong>of</strong> <strong>the</strong> State.<br />

The USDA, APHIS/ADC, may be<br />

contacted to do control work any time <strong>of</strong><br />

year. The APHIS/ADC agent shall attempt<br />

to control only <strong>the</strong> animal(s) causing<br />

damage. The agent will use discretion in <strong>the</strong><br />

control <strong>of</strong> young animals. All lions taken by<br />

APHIS/ADC are <strong>the</strong> property <strong>of</strong> <strong>the</strong> State.<br />

A mountain lion harvest report form is<br />

completed for all mountain lion mortalities.<br />

A mountain lion/human interaction form is<br />

completed for all lion/human interactions.<br />

Records <strong>of</strong> lion mortality and human/ lion<br />

interactions are kept in computer databases<br />

in Reno.<br />

RESEARCH AND PUBLICATIONS<br />

ERNEST, HOLLY B., WALTER M. BOYCE,<br />

VERNON C. BLEICH, BERNIE MAY, SAN<br />

J. STIVER, AND STEVEN G. TORRES. In<br />

Press. Genetic structure <strong>of</strong> mountain<br />

lion (Puma concolor) populations in<br />

California. Journal <strong>of</strong> Conservation<br />

Genetics.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

This paper used 412 samples from<br />

California and 19 samples collected in<br />

western Nevada within 50 km <strong>of</strong> California.<br />

The work helped define <strong>the</strong> geographic<br />

ranges <strong>of</strong> mountain lion populations in<br />

California. Population structure differed<br />

greatly by region - mountain lions in many<br />

California regions have significant barriers<br />

to genetic interchange and <strong>the</strong>refore are very<br />

different from one population to ano<strong>the</strong>r.<br />

This paper, plus <strong>the</strong> work done for <strong>the</strong><br />

Nevada DOW report indicate that<br />

populations in Nevada tend not to have as<br />

much obstruction to genetic interchange as<br />

those in most ecological regions <strong>of</strong><br />

California, in general. This study shows that<br />

mountain lion management and conservation<br />

efforts should be individualized according to<br />

region and incorporate landscape-level<br />

considerations to protect habitat<br />

connectivity.<br />

A follow up study by Dr. Holly Ernest,<br />

on mountain lion genetic variation and<br />

phylogeography in Nevada is currently<br />

being finalized for future publication.


NEW MEXICO MOUNTAIN LION STATUS REPORT<br />

RICK WINSLOW, Large Carnivore Biologist, New Mexico Department <strong>of</strong> Game and Fish, P.O.<br />

Box 25112, Santa Fe, New Mexico 87504, USA, email: Rwinslow@state.nm.us<br />

INTRODUCTION<br />

The New Mexico Department <strong>of</strong> Game<br />

and Fish has almost completed<br />

implementation <strong>of</strong> its cougar management<br />

plan. We are beginning to develop a new<br />

five-year plan.<br />

<strong>Mountain</strong> lions have been classified as<br />

protected big game animals in New Mexico<br />

since 1971 and are currently hunted<br />

throughout most <strong>of</strong> <strong>the</strong> occupied habitat in<br />

<strong>the</strong> state.<br />

<strong>Mountain</strong> lion management in New<br />

Mexico is multi-faceted. The department is<br />

attempting to develop a conservation<br />

strategy that allows both hunting and<br />

enjoyment <strong>of</strong> cougars by <strong>the</strong> non-hunting<br />

public. We also need to balance differing<br />

management issues: (1) depredation control<br />

to minimize economic losses to livestock<br />

operators, (2) minimizing human/cougar<br />

conflicts, (3) cougar removal where<br />

increasing deer and bighorn populations is<br />

<strong>the</strong> management priority.<br />

In 1999 we initiated a zone quota system<br />

for harvest management. Our zone quotas<br />

are based upon management decisions for<br />

ei<strong>the</strong>r increasing, maintaining stable, or<br />

decreasing lion populations.<br />

DISTRIBUTION AND ABUNDANCE<br />

Since <strong>the</strong>ir protection as a big game<br />

animal in 1971, mountain lions have steadily<br />

returned to suitable habitat throughout <strong>the</strong><br />

state. <strong>Mountain</strong> lions generally inhabit <strong>the</strong><br />

rougher country in New Mexico avoiding<br />

<strong>the</strong> low elevation desert areas and eastern<br />

plains. They do however occur in <strong>the</strong>se<br />

areas in conjunction with pockets <strong>of</strong> mule<br />

39<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

deer and areas <strong>of</strong> topographic diversity. Our<br />

current estimate <strong>of</strong> <strong>the</strong> cougar population in<br />

New Mexico is approximately 2150 cougars<br />

derived by multiplying density estimates by<br />

(Logan et al. 1996) <strong>the</strong> estimated amount <strong>of</strong><br />

mule deer habitat. For regional estimates,<br />

we use a population model based on rates <strong>of</strong><br />

recruitment and mortality from Logan et al.<br />

(1996), quantity <strong>of</strong> habitat and population<br />

density.<br />

Since 1979, successful hunters have<br />

been required to present <strong>the</strong>ir cougar to a<br />

Department <strong>of</strong>ficial within 5 days <strong>of</strong> <strong>the</strong>ir<br />

harvest to have <strong>the</strong> pelt tagged, a tooth<br />

collected for aging, sex verified, and o<strong>the</strong>r<br />

information ga<strong>the</strong>red. Reports <strong>of</strong> cougar<br />

depredation and damage are also kept.<br />

Harvest strategies have varied during <strong>the</strong><br />

32 years cougars have been classified as a<br />

game animal. In 1971 only <strong>the</strong> southwestern<br />

corner <strong>of</strong> New Mexico was open to cougar<br />

hunting with a bag limit <strong>of</strong> one cougar and a<br />

4-month season. More areas <strong>of</strong> <strong>the</strong> state<br />

were opened to cougar hunting and seasons<br />

leng<strong>the</strong>ned in subsequent years. From 1979<br />

to 1983 <strong>the</strong> season was 11 months long<br />

statewide with a bag limit <strong>of</strong> 2 cougars.<br />

In 1983 a bill was introduced to <strong>the</strong> New<br />

Mexico House <strong>of</strong> Representatives to return<br />

<strong>the</strong> cougar to its status prior to 1971 as a<br />

varmint. It was tabled but <strong>the</strong> legislature<br />

requested that <strong>the</strong> department ga<strong>the</strong>r more<br />

information on <strong>the</strong> cougar’s status. Evans<br />

(1983) investigated harvest trends and<br />

population estimates and determined that<br />

cougar populations had probably declined.<br />

His determination and public opinion


40 NEW MEXICO MOUNTAIN LION STATUS REPORT · Winslow<br />

resulted in more conservative harvest<br />

strategies.<br />

In 1984 <strong>the</strong> cougar season was shortened<br />

to 3 months in most <strong>of</strong> <strong>the</strong> state, with longer<br />

seasons in units that had high numbers <strong>of</strong><br />

depredation complaints. From 1985 until<br />

1999 <strong>the</strong> season was 4 months long<br />

throughout <strong>the</strong> state with a bag limit <strong>of</strong> one<br />

cougar. In 1999 <strong>the</strong> state instituted a zone<br />

management system with harvest objectives<br />

and quotas. The season was extended to 6<br />

months throughout <strong>the</strong> majority <strong>of</strong> <strong>the</strong> state<br />

with low-elevation bighorn sheep ranges<br />

open year round.<br />

In 2002, <strong>the</strong> cougar season remained at 6<br />

months with a 1 cougar bag limit throughout<br />

most <strong>of</strong> <strong>the</strong> state with <strong>the</strong> following<br />

exceptions: year around hunting in selected<br />

desert bighorn sheep areas and Rocky<br />

<strong>Mountain</strong> bighorn sheep areas in <strong>the</strong><br />

sou<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> state and on private<br />

property, and year round hunting in specific<br />

units in <strong>the</strong> sou<strong>the</strong>astern corner <strong>of</strong> <strong>the</strong> state<br />

that have historically suffered high<br />

depredation losses. The bag limit was<br />

increased to 2 cougars in bighorn areas.<br />

Currently <strong>the</strong> state has 15 cougar<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Figure 1. Cougar Harvest Management<br />

Zones in New Mexico during 2002-03.<br />

management zones (Figure 1). Each is<br />

managed through a quota system for<br />

increasing, decreasing or stable populations<br />

<strong>of</strong> cougar (Table 1). The ratio <strong>of</strong> males to<br />

females harvested generally equals 60:40<br />

Table 1. Cougar harvest objectives by management zone in New Mexico, 2002-03.<br />

Zone Game Management Units Included in Zone Harvest Objective<br />

A 2 and 7 14<br />

B 5 and 50-51 20<br />

C 43-46, 48-49, and 53-55 38<br />

D 41-42, 47, and 56-58 16<br />

E 9 and 10 16<br />

F 6 and 8 16<br />

G 13-14, and 17 17<br />

H 19, 20, and 28-29 3<br />

I 18, 30, 34, and 36-38 20<br />

J 15-16, 21, and 25 38<br />

K 22-24 22<br />

L 26-27 Unlimited<br />

M 31-33, and 39-40 5<br />

N 4 and 52 3<br />

O 12 3<br />

231 a<br />

a Not including unlimited areas.


(Table 2). Hunters tend to selectively<br />

harvest <strong>the</strong> larger male lions.<br />

Since sport hunting was implemented in<br />

1971, use <strong>of</strong> hounds has been allowed but<br />

cubs and females with cubs cannot be taken.<br />

At least 90% <strong>of</strong> <strong>the</strong> harvest is through hound<br />

hunting. There is no pursuit season.<br />

Approximately 2000 lion licenses are<br />

sold per year currently. This number has<br />

gone up during <strong>the</strong> past decade. There is no<br />

limit to <strong>the</strong> number <strong>of</strong> lion licenses sold per<br />

year.<br />

Since 1998 cougar depredation<br />

complaints have ranged from 28 to 45 per<br />

year with 1 to 20 cougars killed per year.<br />

<strong>Mountain</strong> lion depredation incidents are<br />

typically dealt with on a case-by-case basis<br />

NEW MEXICO MOUNTAIN LION STATUS REPORT · Winslow 41<br />

Table 2. Permits issued and cougars harvested in New Mexico, 1981-2003.<br />

Hunt Year<br />

Permits<br />

Issued<br />

Male Harvest<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

in New Mexico. Department policy is to<br />

resolve depredation and to minimize<br />

property damage, conflict and threat to<br />

human safety. When department or Wildlife<br />

Services investigation confirms a<br />

depredation incident, a depredation permit<br />

may be issued. Generally, ei<strong>the</strong>r snares or<br />

hounds are used to capture <strong>the</strong> <strong>of</strong>fending<br />

animal. <strong>Lion</strong>s involved in depredation<br />

incidents are destroyed. Landowners may<br />

also kill lions in defense <strong>of</strong> human safety or<br />

property. The sou<strong>the</strong>astern corner <strong>of</strong> <strong>the</strong><br />

state has a preventative control program,<br />

which is in effect in Unit 30 to reduce<br />

depredation on domestic sheep. The<br />

preventative control program destroyed 110<br />

mountain lions between 1989 and 1999 and<br />

Female<br />

Harvest<br />

Unknown Total Harvest<br />

1981-82 360 78 44 3 125<br />

1982-83 481 55 44 1 101<br />

1983-84 661 67 65 0 132<br />

1984-85 443 47 32 0 79<br />

1985-86 472 56 48 0 104<br />

1986-87 437 55 46 0 101<br />

1987-88 456 43 35 0 78<br />

1988-89 450 58 33 0 91<br />

1989-90 482 71 41 0 112<br />

1990-91 781 73 35 0 108<br />

1991-92 765 77 42 0 119<br />

1992-93 826 68 37 0 105<br />

1993-94 926 75 52 0 127<br />

1994-95 1145 87 61 2 150<br />

1995-96 842 74 45 0 119<br />

1996-97 980 114 62 1 177<br />

1997-98 974 108 58 2 168<br />

1998-99 1485 95 58 0 153<br />

1999-00 1702 98 58 0 156<br />

2000-01 NA 1 140 96 0 236<br />

2001-02 NA 1 127 91 1 219<br />

2002-03 NA 1 161 120 3 284 2<br />

1<br />

Not yet determined.<br />

2<br />

Numbers may not be complete.


42 NEW MEXICO MOUNTAIN LION STATUS REPORT · Winslow<br />

has continued in <strong>the</strong> years since.<br />

In situations where depredation cannot<br />

be confirmed, <strong>the</strong> district wildlife <strong>of</strong>ficer<br />

will <strong>of</strong>fer advice and suggestions as to how<br />

<strong>the</strong> complainant can avoid incidents with<br />

lions. <strong>Lion</strong>s captured for reasons o<strong>the</strong>r than<br />

depredation are relocated to ano<strong>the</strong>r area <strong>of</strong><br />

<strong>the</strong> state.<br />

Human safety incidents with lions are<br />

rare in New Mexico. Any lion involved in a<br />

human safety type <strong>of</strong> incident would be<br />

destroyed if caught.<br />

RESEARCH AND PUBLICATIONS<br />

Ligon (1926) conducted <strong>the</strong> first<br />

investigation on cougars in New Mexico and<br />

determined that <strong>the</strong>y were uncommon but<br />

preyed heavily upon domestic animals and<br />

deer. Hibben (1937) investigated lion<br />

biology in nor<strong>the</strong>rn and western portions <strong>of</strong><br />

<strong>the</strong> state. Prey use and movements in <strong>the</strong><br />

southwestern corner <strong>of</strong> <strong>the</strong> state were<br />

documented via radio telemetry in <strong>the</strong><br />

1970’s (Donaldson 1975, Johnson 1982).<br />

Cougar ecology in Carlsbad Caverns<br />

National Park, New Mexico and <strong>the</strong><br />

Guadalupe <strong>Mountain</strong>s National Park across<br />

<strong>the</strong> border in Texas was studied from 1982-<br />

85 (Smith et al. 1986). Ecology and<br />

population dynamics <strong>of</strong> cougars in <strong>the</strong> San<br />

Andres <strong>Mountain</strong>s <strong>of</strong> south central New<br />

Mexico were studied from 1985-95 (Logan<br />

et al. 1996). This was <strong>the</strong> most intensive<br />

investigation <strong>of</strong> desert-dwelling cougars<br />

ever conducted. Beausoleil (2001) reviewed<br />

historic and current status <strong>of</strong> mountain lions<br />

in New Mexico.<br />

LITERATURE CITED<br />

BEAUSOLEIL, R.A. 2001. Status <strong>of</strong> <strong>the</strong><br />

<strong>Mountain</strong> <strong>Lion</strong> in New Mexico, 1997-<br />

2000. New Mexico Naturalist’s Notes<br />

3(1) pp. 33-47.<br />

DONALDSON, B. 1975. <strong>Mountain</strong> lion<br />

research. Final Report, Pittman<br />

Robertson Project W-93-17, Work plan<br />

15, Job 1. New Mexico Department <strong>of</strong><br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Game and Fish, Santa Fe, New Mexico<br />

USA.<br />

EVANS, W. 1983. The cougar in New<br />

Mexico: biology, status, depredation <strong>of</strong><br />

livestock, and management<br />

recommendations. Response to House<br />

Memorial 42. New Mexico Department<br />

<strong>of</strong> Game and Fish, Santa Fe, New<br />

Mexico USA.<br />

HIBBEN, F.C. 1937. A preliminary study <strong>of</strong><br />

<strong>the</strong> mountain lion (Felis oregonenis<br />

spp.). University <strong>of</strong> New Mexico<br />

Bulletin, Biological Series 5(3) 5-59.<br />

JOHNSON, J. 1982. <strong>Mountain</strong> lion research.<br />

Final Report, Pittman Robertson Project<br />

W-124-R-4, Job 1. New Mexico<br />

Department <strong>of</strong> Game and Fish, Santa Fe,<br />

New Mexico USA.<br />

LIGON, J.S. 1927. Wild Life <strong>of</strong> New Mexico,<br />

Its Conservation and Management.<br />

Being a Report on <strong>the</strong> Game Survey <strong>of</strong><br />

<strong>the</strong> State, 1926 and 1927. State Game<br />

Commission Department <strong>of</strong> Game and<br />

Fish, Santa Fe, New Mexico.<br />

LOGAN, K.A., L.L. SWEANOR, T.K. RUTH,<br />

AND M.G. HORNOCKER. 1996. Cougars<br />

<strong>of</strong> <strong>the</strong> San Andres <strong>Mountain</strong>s, New<br />

Mexico. Federal Aid in Wildlife<br />

Restoration, Project W-128-R, for New<br />

Mexico Department <strong>of</strong> Game and Fish,<br />

Santa Fe, New Mexico USA.<br />

NEW MEXICO DEPARTMENT OF GAME AND<br />

FISH. 1997. Long range plan for <strong>the</strong><br />

management <strong>of</strong> cougar in New Mexico.<br />

Federal Aid in Wildlife Restoration<br />

Grant W-93-R-39, Project 1, Job 5. New<br />

Mexico Department <strong>of</strong> Game and Fish,<br />

Santa Fe, New Mexico USA.<br />

SMITH, T.E., R.R. DUKE, M.J. KUTILEK, AND<br />

H.T. HARVEY. 1986. <strong>Mountain</strong> lions<br />

(Felis Concolor) in <strong>the</strong> vicinity <strong>of</strong><br />

Carlsbad Caverns, New Mexico and<br />

Guadalupe <strong>Mountain</strong>s National Park,<br />

Texas. Harvey and Stanley Associates<br />

Incorporated, Alvisa, Texas USA.


STATE OF SOUTH DAKOTA MOUNTAIN LION STATUS REPORT<br />

MIKE KINTIGH, Regional Supervisor, South Dakota Game, Fish & Parks, 3305 West South St.,<br />

Rapid City, SD 57702, USA, email: Mike.Kintigh@state.sd.us<br />

INTRODUCTION<br />

South Dakota (SD) is currently<br />

developing a management plan for mountain<br />

lions. The second draft <strong>of</strong> this document is<br />

currently under review by South Dakota<br />

Department <strong>of</strong> Game, Fish and Parks (SD<br />

GFP) staff. This document is also available<br />

to <strong>the</strong> public and interested parties for<br />

review and comment. A copy can be<br />

obtained by contacting Regional Supervisor<br />

Mike Kintigh (listed above) or Dr. Larry<br />

Gigliotti at 605-773-4231. This summer/fall<br />

SD GFP will be taking fur<strong>the</strong>r steps to<br />

solicit public comments on <strong>the</strong> management<br />

plan.<br />

<strong>Mountain</strong> <strong>Lion</strong>s are currently classified<br />

as a State Threatened Species in SD.<br />

However, in July that classification will<br />

likely change significantly. Legislative<br />

action in January <strong>of</strong> 2003 closed a legal<br />

loophole by defining <strong>the</strong> lion as a big game<br />

animal. This action will facilitate <strong>the</strong><br />

removal <strong>of</strong> <strong>the</strong> lion from <strong>the</strong> State’s<br />

Threatened Species List. Game<br />

Commission action is still required to<br />

finalize <strong>the</strong> delisting and this is expected to<br />

occur in early June. Forty-five days is<br />

required for any commission finalization<br />

actions to take effect and this will occur in<br />

mid July, after <strong>the</strong> legislative action<br />

becomes law on July 1, 2003.<br />

It is important to note that while <strong>the</strong> lion<br />

will be taken <strong>of</strong>f <strong>the</strong> threatened species list<br />

in South Dakota, it will actually gain<br />

additional protection under law by being<br />

defined as a big game animal. Criminal<br />

penalties will increase from class 2<br />

misdemeanors to class 1 misdemeanors,<br />

43<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

carrying higher fines and longer jail<br />

sentences.<br />

A misconception exists in that by<br />

classifying <strong>the</strong> lion as a big game animal a<br />

hunting season will immediately be<br />

implemented. This is absolutely false! The<br />

lion will continue to be fully protected as a<br />

big game animal with a continuously closed<br />

season until at some undetermined point<br />

when additional management decisions are<br />

made.<br />

South Dakota has many objectives<br />

concerning mountain lion management.<br />

They are as follows:<br />

A) Evaluate <strong>the</strong> Legal status <strong>of</strong> <strong>the</strong><br />

<strong>Mountain</strong> <strong>Lion</strong> in SD by April 1,<br />

2003.<br />

B) Evaluate strategies for monitoring &<br />

censusing <strong>Mountain</strong> <strong>Lion</strong><br />

populations in SD by 2005.<br />

C) Maintain a statewide database <strong>of</strong> Mt.<br />

<strong>Lion</strong> activity including sightings,<br />

human interactions, depredation<br />

events and lion mortality.<br />

D) Develop a list <strong>of</strong> Mt. <strong>Lion</strong> research<br />

needs. Evaluate and prioritize<br />

annually.<br />

E) Develop Mt. <strong>Lion</strong> population<br />

management methods that are<br />

consistent with established goals and<br />

objectives.<br />

F) Identify and describe suitable habitat<br />

areas and parameters for Mt. <strong>Lion</strong>s<br />

in SD by Sept. 2003.<br />

G) Develop a comprehensive Public<br />

Education strategy for informing and<br />

educating <strong>the</strong> Staff, citizens and


44 SOUTH DAKOTA MOUNTAIN LION STATUS REPORT · Kintigh<br />

visitors about Mt. <strong>Lion</strong>s and personal<br />

safety while in Mt. <strong>Lion</strong> country.<br />

H) Develop a public involvement plan<br />

for implementation during 2003 and<br />

2004 for inclusion in our<br />

management planning process.<br />

Over <strong>the</strong> last 10 years South Dakota has<br />

not significantly changed <strong>the</strong> way we<br />

manage lions. During this period <strong>of</strong> time<br />

<strong>the</strong>y remained on <strong>the</strong> State’s Threatened<br />

Species List and very little was done to<br />

manage <strong>the</strong>m o<strong>the</strong>r than <strong>of</strong>fering <strong>the</strong>m full<br />

protection <strong>of</strong> <strong>the</strong> law. Our awareness <strong>of</strong><br />

lions did increase significantly during this<br />

time as we observed a steady apparent<br />

increase in <strong>the</strong>ir numbers. In recent years an<br />

Action Plan was developed to guide staff in<br />

dealing with problem lions. This Action<br />

Plan is currently under revision and will be<br />

included in <strong>the</strong> overall Management Plan.<br />

We also created a system for documenting<br />

and tracking lion activity. More significant<br />

changes are looming on <strong>the</strong> horizon as we<br />

remove <strong>the</strong> lion from <strong>the</strong> Threatened Species<br />

List and focus on concerted effort to manage<br />

our lions.<br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Lion</strong>s are currently distributed<br />

throughout <strong>the</strong> Black Hills, which contains<br />

<strong>the</strong> most suitable habitat in South Dakota.<br />

Some evidence <strong>of</strong> breeding populations also<br />

exists in <strong>the</strong> Custer National Forest in<br />

Harding County, <strong>the</strong> Badlands <strong>of</strong> eastern<br />

Pennington County and on <strong>the</strong> Pine Ridge<br />

Reservation <strong>of</strong> Shannon, Jackson and Bennett<br />

counties.<br />

Reports <strong>of</strong> lion activity have been<br />

received across all <strong>of</strong> South Dakota.<br />

Verification <strong>of</strong> reports outside <strong>of</strong> <strong>the</strong> Black<br />

Hills has proven to be very difficult,<br />

especially east <strong>of</strong> <strong>the</strong> Missouri River. Most<br />

occurrences outside <strong>of</strong> <strong>the</strong> Black Hills have<br />

been associated with river drainages that<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

provide marginal habitat.<br />

The lion population in South Dakota<br />

appears to be still growing slowly at this<br />

time. Some uncertainty exists as to what <strong>the</strong><br />

carrying capacity <strong>of</strong> <strong>the</strong> Black Hills for lions<br />

may be, though it is generally felt we are<br />

very close to that level now. Some evidence<br />

<strong>of</strong> dispersal from <strong>the</strong> Black Hills exists. To<br />

date we have only detected young males<br />

dispersing from <strong>the</strong> Black Hills. One young<br />

female was radio collared on <strong>the</strong> very edge<br />

<strong>of</strong> <strong>the</strong> Black Hills and some thought that she<br />

might disperse was expressed. She was<br />

poached before that determination was<br />

made.<br />

The cougar population in <strong>the</strong> Black Hills<br />

was estimated using program PUMA (Beier<br />

1993), incorporating parameters obtained<br />

from radio-collared cougars and habitat<br />

quality derived from a habitat-relation<br />

model. Annual home ranges were generated<br />

for 10 adult cougars monitored > 8 months,<br />

and spatial distribution <strong>of</strong> established males<br />

was analyzed using a home range overlap<br />

index. The area <strong>of</strong> <strong>the</strong> Black Hills was<br />

estimated at 8,400 km 2 , comprised <strong>of</strong><br />

6,702.9 km 2 <strong>of</strong> high quality and 1,697.1 km 2<br />

<strong>of</strong> lower quality habitat (based on a habitatrelation<br />

model developed for <strong>the</strong> species).<br />

Mean annual home range size <strong>of</strong> established<br />

adult male cougars (n = 3) was 809.2 km 2 ,<br />

and was significantly larger (P < 0.05) than<br />

that <strong>of</strong> adult females (n = 4), 182.3 km 2 .<br />

Based on sightings <strong>of</strong> family groups and<br />

radio-collared females, we documented up<br />

to 5 females occurring in established male<br />

ranges. Percent overlap for 3 established<br />

cougars averaged 33% (range = 18.0 -<br />

52.0%). Based on 5 population simulations,<br />

<strong>the</strong> total number <strong>of</strong> cougars in <strong>the</strong> Black<br />

Hills was estimated to be 127 to 149<br />

cougars; 46 to 49 adult females, 12 to 29<br />

adult males; 21 to 24 yearling females and<br />

males; and 45 to 48 female and male kittens.


HARVEST INFORMATION<br />

South Dakota has not had any form <strong>of</strong><br />

legalized mountain lion hunting since 1978.<br />

The future management <strong>of</strong> lions in South<br />

Dakota will include consideration <strong>of</strong> a<br />

hunting season as a management tool.<br />

Concerns about <strong>the</strong> impacts <strong>of</strong> hunting to <strong>the</strong><br />

stability <strong>of</strong> <strong>the</strong> population will weigh heavily<br />

when those decisions are made.<br />

DEPREDATIONS AND HUMAN<br />

INTERACTIONS/CONFLICTS<br />

South Dakota does operate with an<br />

“Action Plan For Managing <strong>Mountain</strong><br />

<strong>Lion</strong>/Human/Property Interactions.” An<br />

Action Plan was first developed in May <strong>of</strong><br />

1995 and has been revised since <strong>the</strong>n. This<br />

plan is included in <strong>the</strong> overall <strong>Mountain</strong><br />

<strong>Lion</strong> Management Plan, which is currently<br />

under development. For our agency,<br />

addressing “problem” lions is <strong>the</strong> most<br />

difficult aspect <strong>of</strong> maintaining a population<br />

<strong>of</strong> lions. Public emotions are strong and<br />

varied which results in many<br />

comments/opinions being expressed directly<br />

at <strong>the</strong> “Action Plan.”<br />

South Dakota’s Action Plan categorizes<br />

Human/<strong>Lion</strong> interactions into five types:<br />

1. Sighting - a visual observation <strong>of</strong> a<br />

lion or a report <strong>of</strong> lion tracks or o<strong>the</strong>r<br />

sign on unpopulated lands or rural<br />

areas within <strong>the</strong> Black Hills.<br />

2. Encounter - an unexpected direct<br />

neutral meeting between a human and<br />

SOUTH DAKOTA MOUNTAIN LION STATUS REPORT · Kintigh 45<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

a lion without incident (<strong>Mountain</strong> lion<br />

sightings in close proximity to homes,<br />

stables or livestock in rural areas and<br />

unpopulated lands outside <strong>of</strong> <strong>the</strong><br />

Black Hills). A mountain lion is<br />

observed for <strong>the</strong> first time in close<br />

proximity or within residential<br />

developments and occupied<br />

recreational area.<br />

3. Incident - a conflict between a human<br />

and lion that may have serious results<br />

(e.g. a lion that must be forced to back<br />

down). Recurring observations <strong>of</strong> a<br />

lion in close proximity or within<br />

residential developments and<br />

occupied recreational areas.<br />

Livestock is killed in rural areas.<br />

4. Substantial public threat - a mountain<br />

lion that is observed within a city near<br />

areas where children are regularly<br />

congregated, killing wildlife/pets<br />

residential developments or occupied<br />

recreational areas or repeatedly killing<br />

livestock.<br />

5. Attack - when a human is bodily<br />

injured or killed by contact with a<br />

mountain lion.<br />

Each occurrence requires an<br />

understanding <strong>of</strong> all <strong>the</strong> circumstances and<br />

any history involved before an action is<br />

decided upon. In general, with every report<br />

<strong>of</strong> a lion a field investigation is highly<br />

encouraged by agency personnel (Table 1).<br />

Verification is key to any response.<br />

Table 1. Public Safety reports and resulting lion removals in South Dakota, 1998 – 2002.<br />

Year Number<br />

Reports<br />

Number<br />

Incidents<br />

Number<br />

Encounters<br />

Threatening<br />

Encounters<br />

Number <strong>of</strong> Public<br />

Safety Incidents<br />

Number <strong>Lion</strong>s<br />

Removed<br />

1998 57 5 2 2 5 0<br />

1999 54 1 0 0 1 0<br />

2000 66 5 4 1 1 1<br />

2001 144 4 8 3 4 0<br />

2002 198 5 6 2 2 1


46 SOUTH DAKOTA MOUNTAIN LION STATUS REPORT · Kintigh<br />

Table 2. <strong>Mountain</strong> lion depredations, verified depredations, and resulting lion removals in South<br />

Dakota, 1998 – 2002.<br />

Year Number Depredations<br />

Number Depredations<br />

Verified<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Number <strong>Lion</strong>s<br />

Removed<br />

1998 1 1 0<br />

1999 0 0 0<br />

2000 2 1 0<br />

2001 3 2 1<br />

2002 4 2 0<br />

Note – one lion has been removed due to livestock depredation in 2003 already.<br />

Personnel are encouraged to take every<br />

opportunity to educate <strong>the</strong> public regarding<br />

all aspects <strong>of</strong> living with lions. Each lion<br />

reporting person receives an agency<br />

produced brochure on <strong>Mountain</strong> <strong>Lion</strong>s.<br />

Public education is emphasized at this time<br />

and every opportunity is taken.<br />

Keeping all options available to<br />

responding staff is very desirable to our<br />

agency. However, we will not pay for any<br />

damages incurred due to wildlife <strong>of</strong> any<br />

species.<br />

Relocation <strong>of</strong> problem lions was once<br />

considered, but, due to <strong>the</strong> geographically<br />

limited area <strong>of</strong> <strong>the</strong> Black Hills and <strong>the</strong><br />

existing lion population, it has been deemed<br />

an option that was unlikely to produce<br />

desirable results. Unusual circumstances<br />

may arise in which it may be attempted and<br />

<strong>the</strong> option has not been made totally<br />

unavailable.<br />

In rare cases, usually involving a single<br />

livestock producer, a permit has been issued<br />

for that individual to kill a lion that has been<br />

causing livestock depredation. Usually this<br />

only happens after agency efforts to remove<br />

<strong>the</strong> <strong>of</strong>fending lion have failed.<br />

Our agency is equipped with a trio <strong>of</strong><br />

trained lion hounds managed by an<br />

experienced houndsman. In most situations<br />

that necessitate a lion removal, <strong>the</strong> action is<br />

lead by our houndsman. Our state trappers<br />

are also equipped with leg snares, which are<br />

generally only set around livestock kills as<br />

<strong>the</strong> houndsman prepares to arrive on scene.<br />

On a few occasions, when a lion was a<br />

concern, but did not warrant removal we<br />

have chased <strong>the</strong> lion with hounds to haze <strong>the</strong><br />

lion. On at least one occasion <strong>the</strong> lion was<br />

treed and a radio collar was fitted to increase<br />

our knowledge <strong>of</strong> its activity.<br />

In regards to livestock depredation, we<br />

currently investigate every report <strong>of</strong> this but<br />

take slightly different approaches to<br />

resolution depending upon <strong>the</strong> location.<br />

Livestock kills within <strong>the</strong> Black Hills<br />

typically require multiple kills before action<br />

to remove <strong>the</strong> <strong>of</strong>fending lion is initiated.<br />

We are hesitant to remove lions from <strong>the</strong><br />

limited quality habitat available in South<br />

Dakota (Table 2). Livestock depredation<br />

complaints on <strong>the</strong> plains <strong>of</strong> South Dakota,<br />

where limited habitat and a strong<br />

agricultural industry exists, are addressed<br />

much more decisively and quickly.<br />

RESEARCH AND PUBLICATIONS<br />

The Department <strong>of</strong> Wildlife and<br />

Fisheries Sciences at South Dakota State<br />

University is currently completing a 5-year<br />

research project on cougars in <strong>the</strong> Black<br />

Hills. The main objectives <strong>of</strong> <strong>the</strong> research<br />

were to 1) develop and evaluate a cougar<br />

habitat-relation model to predict <strong>the</strong> current<br />

distribution 2) estimate <strong>the</strong> population size,<br />

and evaluate survey techniques to document


population trend. A digital habitat relation<br />

model was constructed for cougars that<br />

ranked land in <strong>the</strong> Black Hills National<br />

Forest according to its suitability to cougars.<br />

The model was based on <strong>the</strong> distribution <strong>of</strong><br />

prey (white-tailed deer and mule deer),<br />

stalking topography (slopes), concealment<br />

habitat (riparian habitat), and anthropogenic<br />

characteristics (high-density residential<br />

areas, presence <strong>of</strong> highways). During <strong>the</strong><br />

winters <strong>of</strong> 1998 – 2001, we captured, radiocollared,<br />

and obtained weekly locations <strong>of</strong><br />

12 cougars in <strong>the</strong> Black Hills; locations <strong>of</strong><br />

cougars were used to validate <strong>the</strong> habitatrelation<br />

model. The cougar population in<br />

<strong>the</strong> Black Hills was estimated using program<br />

PUMA, incorporating parameters obtained<br />

from radio-collared cougars and habitat<br />

quality derived from <strong>the</strong> habitat-relation<br />

model. The total number <strong>of</strong> cougars in <strong>the</strong><br />

Black Hills was estimated to be 127 to 149<br />

cougars.<br />

A 3-month pilot study, testing <strong>the</strong><br />

efficacy <strong>of</strong> detecting cougars using scent<br />

lures (skunk essence, Powder River cat call)<br />

and camera stations was conducted in<br />

cooperation with <strong>the</strong> University <strong>of</strong> North<br />

Dakota. The camera-scent-station survey<br />

was not effective at detecting cougar<br />

presence. Zero photos <strong>of</strong> cougars were<br />

recorded although o<strong>the</strong>r species (whitetailed<br />

deer, Odocoileus virginianus, mule<br />

deer, O. hemionus, raccoon, Procyon lotor,<br />

red squirrel, Tamiasciurus hudsonicus,<br />

turkey vulture, Cathartes aura, free-ranging<br />

cattle, feral dogs, and bobcat, Lynx rufus)<br />

were detected, and cougars were known to<br />

be in <strong>the</strong> area during <strong>the</strong> survey. A snowtracking<br />

helicopter survey (Vansickle and<br />

Lindzey 1991) using a probability sampling<br />

technique was attempted during <strong>the</strong> winter<br />

<strong>of</strong> 2001-2002. Although cougar tracks <strong>of</strong> a<br />

radio-collared female and her 2 kittens could<br />

clearly be identified, wea<strong>the</strong>r conditions<br />

(poor snow conditions) did not permit <strong>the</strong><br />

survey to be completed. However, a<br />

SOUTH DAKOTA MOUNTAIN LION STATUS REPORT · Kintigh 47<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

database <strong>of</strong> consecutive winter daily<br />

locations <strong>of</strong> 3 male and 3 female cougars<br />

was established to aid in analyses <strong>of</strong> any<br />

future helicopter surveys.<br />

During <strong>the</strong> Fall 2002, a second 5-year<br />

study was initiated. The objectives <strong>of</strong> <strong>the</strong><br />

research are 1) to estimate survival and<br />

document causes <strong>of</strong> mortality <strong>of</strong> cougar<br />

kittens, 2) Determine longevity <strong>of</strong><br />

established radio-collared cougars 3)<br />

Document dispersal distances, routes, and<br />

destinations <strong>of</strong> subadult cougars, and 4)<br />

conduct snow tracking helicopter population<br />

survey to document population trends.<br />

Currently, 12 cougars (6 females, 6 males)<br />

including 2 subadult males are being<br />

monitored weekly from fixed wing aircraft<br />

using aerial radio-telemetry techniques.<br />

PUBLICATIONS<br />

FECSKE, D.M., J.A. JENKS, AND F.G.<br />

LINDZEY. 2001. Characteristics <strong>of</strong><br />

mountain lion mortalities in <strong>the</strong> Black<br />

Hills, South Dakota. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong><br />

6th <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, San<br />

Antonio, Texas: In Press.<br />

FECSKE, D.M., AND J.A. JENKS. 2001. The<br />

mountain lion returns to South Dakota.<br />

South Dakota Conservation Digest<br />

68(4):3-5.<br />

FECSKE, D.M., AND J.A. JENKS. 2001.<br />

Status report <strong>of</strong> mountain lions in South<br />

Dakota. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> 6th<br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, San Antonio,<br />

TX. In Press.<br />

FECSKE, D.M., J.A. JENKS, AND F. G.<br />

LINDZEY. 2003. Mortality <strong>of</strong> an adult<br />

cougar due to a forest fire in <strong>the</strong> Black<br />

Hills. The Prairie Naturalist 00:<br />

Submitted.<br />

GIGLIOTTI, L.M., D.M. FECSKE, AND J.A.<br />

JENKS. 2002. <strong>Mountain</strong> lions in South<br />

Dakota: A public opinion survey. South<br />

Dakota Department <strong>of</strong> Game, Fish, and<br />

Parks, Pierre, SD. 182 pp.<br />

LONG, E.S., D.M. FECSKE, R.A. SWEITZER,<br />

J.A. JENKS, B.M. PIERCE, AND V.C.


48 SOUTH DAKOTA MOUNTAIN LION STATUS REPORT · Kintigh<br />

BLEICH. 2003. Efficacy <strong>of</strong> photographic<br />

scent stations to detect mountain lions.<br />

Western North American Naturalist 00:<br />

In Press.<br />

LITERATURE CITED<br />

BEIER, P. 1993. Puma: a population<br />

simulator for cougar conservation.<br />

Wildl. Soc. Bull. 21:356-357<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

VAN SICKLE, W.D., AND F.G. LINDZEY.<br />

1991. Evaluation <strong>of</strong> a cougar population<br />

estimator based on probability sampling.<br />

Journal <strong>of</strong> Wildlife Management 55:738-<br />

743.


MOUNTAIN LION STATUS REPORT FOR TEXAS<br />

JOHN YOUNG, Texas Parks and Wildlife Department, 3000 IH 35 South Suite 100, Austin, TX<br />

78612, USA, email: john.young@tpwd.state.tx.us<br />

Texas does not currently have a<br />

statewide management plan for mountain<br />

lions and <strong>the</strong> species is classified as nongame.<br />

Texas Parks and Wildlife<br />

Department (TPWD) non-game codes<br />

authorize <strong>the</strong> agency to establish hunting<br />

seasons, to close seasons, set bag limits,<br />

establish management zones, in o<strong>the</strong>r words,<br />

to utilize all <strong>of</strong> <strong>the</strong> management tools<br />

available for game species. With <strong>the</strong><br />

exception <strong>of</strong> a short list <strong>of</strong> non-game species<br />

<strong>of</strong> concern to TPWD, non-game species may<br />

be taken at any time <strong>of</strong> <strong>the</strong> year in any<br />

numbers, which is <strong>the</strong> case for mountain<br />

lions at <strong>the</strong> present time. TPWD’s objective<br />

for mountain lions is to maintain a viable<br />

population, while minimizing human<br />

conflicts. No changes in mountain lion<br />

status have occurred in <strong>the</strong> past decade.<br />

DISTRIBUTION AND ABUNDANCE<br />

Based on confirmed sightings and<br />

mortality records mountain lions are most<br />

common in <strong>the</strong> Trans Pecos and <strong>the</strong> brush<br />

country <strong>of</strong> South Texas. Mortality records<br />

49<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

over <strong>the</strong> last 20 years combined with photos<br />

confirm at least <strong>the</strong> occasional presence <strong>of</strong><br />

lions in all o<strong>the</strong>r sections <strong>of</strong> <strong>the</strong> state; more<br />

information is needed to determine<br />

population levels. Based on sightings, and<br />

voluntarily reported mortalities dating back<br />

to 1983, mountain lion populations appear<br />

stable. Table 1 presents mountain lion<br />

mortality information by ecological region<br />

for <strong>the</strong> time frame 1998/99 to 2001/02<br />

Texas does not currently estimate mountain<br />

lion populations, opting to monitor <strong>the</strong><br />

species using sightings and mortality<br />

reports. The lack <strong>of</strong> a satisfactory<br />

scientifically rigorous method to estimate<br />

mountain lions has been <strong>the</strong> primary reason<br />

TPWD has not attempted to do so. Texas<br />

has recently provided funding to a<br />

university-based scientist to estimate<br />

mountain lion population size, structure, and<br />

habitat factors utilizing new, highly credible<br />

molecular genetics. The study will be<br />

conducted over <strong>the</strong> next 2 years and will<br />

provide an estimate for Texas’ mountain<br />

lion population.<br />

Table 1. <strong>Mountain</strong> lion mortalities by ecological region, September 1998 through September 2002.<br />

Ecological Region 1998/99 1999/00 2000/01 20001/02<br />

Pineywoods 0 0 0 0<br />

Gulf Prairie & Marshes 0 3 0 0<br />

Post Oak Savannah 0 0 0 0<br />

Blackland Prairies 0 0 0 0<br />

Cross Timbers 0 0 0 1<br />

South Texas Plains 7 10 0 4<br />

Edwards Plateau 30 14 6 12<br />

Rolling Plains 0 0 0 0<br />

High Plains 0 0 0 0<br />

Trans-Pecos 92 60 64 48


50 TEXAS MOUNTAIN LION STATUS REPORT · Young<br />

HARVEST INFORMATION<br />

Texas relies primarily on hunters, private<br />

landowners, and trappers to voluntarily<br />

report mountain lion kills. Texas also<br />

obtains an annual report from Texas<br />

Wildlife Damage Management Services<br />

(Table 1). There is an open season on<br />

mountain lions in Texas year-round. TPWD<br />

does not set harvest guidelines or bag limits<br />

for this species. <strong>Mountain</strong> lions may be<br />

taken by trap, shooting, hunting with dogs,<br />

aerial hunting, or M44. Records on <strong>the</strong><br />

number <strong>of</strong> lions harvested by different<br />

methods are not collected.<br />

TPWD does not have a predator incident<br />

manual/policy/guideline for mountain lions<br />

although such has been developed for black<br />

bear. In <strong>the</strong> past 10 years <strong>the</strong>re are only 3<br />

known public safety incidents in Texas<br />

related to mountain lion. Due to <strong>the</strong>ir rarity,<br />

TPWD does not formally record/collect<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

information on public safety incidents<br />

involving mountain lion. Depredation<br />

complaints received at TPWD are referred to<br />

Texas Wildlife Damage Management<br />

Services (TWDMS). In 2001/02 a total <strong>of</strong><br />

53 lions were killed by TWDMS personnel.<br />

Information on cougars removed by<br />

TWDMS prior to 2001/02 had been<br />

combined with o<strong>the</strong>r mortalities and has not<br />

been available separately.<br />

Individuals wishing to report a sighting<br />

or a problem with mountain lions are<br />

encouraged to contact TPWD. The<br />

department provides individuals<br />

experiencing depredation problems with <strong>the</strong><br />

number for <strong>the</strong>ir local TWDMS <strong>of</strong>fice for<br />

action. Relocation <strong>of</strong> mountain lions is<br />

discouraged but may be conducted by<br />

private organizations if <strong>the</strong>y acquire <strong>the</strong><br />

appropriate permits.


UTAH MOUNTAIN LION STATUS REPORT<br />

CRAIG R. McLAUGHLIN, Utah Division <strong>of</strong> Wildlife Resources, 1594 W. North Temple, Salt<br />

Lake City, UT 84114, USA, email: craigmclaughlin@utah.gov<br />

Abstract: <strong>Mountain</strong> lions have been managed as a protected game species in Utah since 1967. In 1999 <strong>the</strong> Division<br />

<strong>of</strong> Wildlife Resources completed <strong>the</strong> Utah Cougar Management Plan, developed with <strong>the</strong> assistance <strong>of</strong> a publicbased<br />

Cougar Discussion Group that will guide management <strong>of</strong> cougars through 2009. Cougar harvests are<br />

managed under both harvest objective (quota) and limited entry strategies. The Division manages to sustain cougar<br />

densities on all management units except those that have approved predator management plans, where cougar<br />

harvests are increased to reduce cougar numbers and predation on big game. All cougar complaints are handled<br />

under <strong>the</strong> guidance <strong>of</strong> a Nuisance Cougar Complaints policy. Most cougar conflicts are handled through lethal<br />

control. Cougar habitat encompasses about 92,696 km 2 (35,790 mi 2 ). The statewide population was estimated at<br />

2,528-3,936 cougars in 1999 in conjunction with <strong>the</strong> Cougar Management Plan. Cougar harvests have ranged from<br />

492 to 373 annually since <strong>the</strong> 1997-1998 season. Both <strong>the</strong> hunting and pursuit seasons run from mid-December<br />

through June, although some units have extended or shortened seasons. Cougars have been implicated in 74-114<br />

separate depredation incidents per year since 1998, with livestock losses ranging from $53,700 to $97,700 per year.<br />

Harvest-based indicators <strong>of</strong> sustainable harvesting have not been met in recent years. Currently, management is<br />

operating on an individual-unit scale, where interpretation <strong>of</strong> harvest data is hampered by small sample sizes. In<br />

addition, <strong>the</strong> Division should develop a means to monitor both reproduction and survival. Harvest management<br />

should improve with understanding <strong>of</strong> cougar movements and dispersal, particularly between lightly hunted and<br />

heavily harvested cougar populations.<br />

51<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Key words: Cougar, livestock damage, harvest, management plan, mountain lion, Puma concolor<br />

INTRODUCTION<br />

<strong>Mountain</strong> lions (Puma concolor), or<br />

cougars, were persecuted as vermin in Utah<br />

from <strong>the</strong> time <strong>of</strong> European settlement (in<br />

1847) until 1966. In 1967 <strong>the</strong> Utah State<br />

Legislature changed <strong>the</strong> status <strong>of</strong> cougars to<br />

protected wildlife and since <strong>the</strong>n <strong>the</strong>y have<br />

been considered a game species with<br />

established hunting regulations. The Utah<br />

Division <strong>of</strong> Wildlife Resources (UDWR)<br />

developed <strong>the</strong> Utah Cougar Management<br />

Plan in 1999 (UDWR 1999b) with <strong>the</strong><br />

assistance <strong>of</strong> a Cougar Discussion Group<br />

composed <strong>of</strong> representatives <strong>of</strong> various<br />

public interest groups. This plan will guide<br />

cougar management in Utah through 2009.<br />

Its goal is to maintain a healthy cougar<br />

population within existing occupied habitat<br />

while considering human safety, economic<br />

concerns and o<strong>the</strong>r wildlife species.<br />

Management objectives include: 1)<br />

maintaining current (1999) cougar<br />

distribution, with a reasonable proportion <strong>of</strong><br />

older age animals and breeding females,<br />

balancing population numbers with o<strong>the</strong>r<br />

wildlife species; 2) minimizing <strong>the</strong> loss in<br />

quality and quantity <strong>of</strong> existing critical and<br />

high priority cougar habitat; 3) reducing <strong>the</strong><br />

risk <strong>of</strong> loss <strong>of</strong> human life and reducing<br />

chances <strong>of</strong> injury by cougar; 4) maintaining<br />

a downward trend in <strong>the</strong> number <strong>of</strong> livestock<br />

killed by cougar; and 5) maintaining quality<br />

recreational opportunity for a minimum <strong>of</strong><br />

800 persons per year through 2009.<br />

Utah’s cougar harvests are controlled on<br />

specific geographic areas, or management<br />

units (Figure 1), using two harvest<br />

strategies: harvest objective and limited<br />

entry. Under <strong>the</strong> harvest objective<br />

strategy, managers prescribe a quota, or<br />

number <strong>of</strong> cougars to be harvested on <strong>the</strong><br />

unit. An unlimited number <strong>of</strong> licensed


52 UTAH MOUNTAIN LION STATUS REPORT · McLaughlin<br />

Figure 1. Wildlife Management Units used<br />

by Utah Division <strong>of</strong> Wildlife Resources to<br />

manage cougar harvests. Some <strong>of</strong> <strong>the</strong> units<br />

have been subdivided for additional control<br />

<strong>of</strong> harvests.<br />

hunters are allowed to hunt during a season<br />

that is variable in length, as <strong>the</strong> hunting<br />

season closes as soon as <strong>the</strong> quota is filled or<br />

when <strong>the</strong> season end date is reached. Under<br />

<strong>the</strong> limited entry strategy, harvests are<br />

managed by limiting <strong>the</strong> number <strong>of</strong> hunters<br />

on a unit. The number <strong>of</strong> hunters is<br />

determined based upon an expectation <strong>of</strong><br />

hunting success and <strong>the</strong> desired harvest size.<br />

Individuals are usually selected for hunting<br />

on <strong>the</strong> unit through a random drawing<br />

process.<br />

In 1996 <strong>the</strong> Utah Wildlife Board<br />

approved a Predator Management Policy<br />

(UDWR 1996) that allows UDWR to<br />

increase cougar harvests on management<br />

units where big game populations are<br />

depressed, or where big game has recently<br />

been released to establish new populations.<br />

Most predator management plans directed at<br />

cougars have been designed to benefit mule<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

deer (Odocoileus hemionus) and bighorn<br />

sheep (Ovis canadensis). Cougar harvests<br />

have been liberalized where big game<br />

populations are far below objective (


Figure 2. Cougar habitat in Utah. All<br />

colored areas represent occupied cougar<br />

habitat.<br />

Wasatch <strong>Mountain</strong>s in nor<strong>the</strong>rn and central<br />

Utah.<br />

The last statewide cougar population<br />

estimates were developed in conjunction<br />

with <strong>the</strong> Utah Cougar Management Plan in<br />

1999 (UDWR 1999b). These estimates used<br />

extrapolations <strong>of</strong> cougar densities from<br />

published studies in <strong>the</strong> southwestern United<br />

States to: 1) <strong>the</strong> total area within all<br />

management units that comprise cougar<br />

range, and 2) <strong>the</strong> total amount <strong>of</strong> occupied<br />

cougar habitat within Utah. The habitat<br />

quality within each management unit was<br />

classified as ei<strong>the</strong>r high, medium or low<br />

based on vegetative characteristics, terrain<br />

ruggedness (following Riley 1998) and prey<br />

density. Cougar densities derived from<br />

research within Utah, California and New<br />

Mexico were associated with each habitat<br />

quality level (UDWR 1999b). High quality<br />

habitat was assigned a density range <strong>of</strong> 2.5-<br />

3.9 cougars/100 km 2 , medium quality<br />

UTAH MOUNTAIN LION STATUS REPORT · McLaughlin 53<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

habitat was assigned a density <strong>of</strong> 1.7-2.5<br />

cougars/100 km 2 and a density <strong>of</strong> 0.26-0.52<br />

cougar/100 km 2 was assigned to low quality<br />

habitat.<br />

The first statewide population estimate<br />

<strong>of</strong> 2,528-3,936 cougars resulted from<br />

summing unit population estimates. The<br />

number <strong>of</strong> cougars on each unit was<br />

estimated by first multiplying <strong>the</strong> total area<br />

contained within <strong>the</strong> unit by <strong>the</strong> highest<br />

density <strong>of</strong> <strong>the</strong> range assigned to it, and <strong>the</strong>n<br />

by <strong>the</strong> lowest density <strong>of</strong> <strong>the</strong> range assigned<br />

to it.<br />

For comparison, a second estimate <strong>of</strong><br />

2,927 cougars statewide was generated<br />

based upon mean cougar densities and total<br />

occupied cougar habitat within <strong>the</strong> state.<br />

Each management unit’s cougar population<br />

was estimated by extrapolating <strong>the</strong> mean<br />

cougar density assigned to <strong>the</strong> unit (based on<br />

<strong>the</strong> respective range indicated above) to <strong>the</strong><br />

amount <strong>of</strong> occupied cougar habitat within<br />

<strong>the</strong> unit, and unit estimates were summed to<br />

obtain <strong>the</strong> statewide figure. The two<br />

methods produced population estimates that<br />

show considerable agreement, but <strong>the</strong>y<br />

should be only viewed as general<br />

approximations <strong>of</strong> <strong>the</strong> statewide cougar<br />

population.<br />

Utah’s cougar population is monitored<br />

through mandatory reporting <strong>of</strong> all hunterharvested<br />

cougars, cougars that are killed on<br />

highways or in accidents and those taken by<br />

animal damage control programs (Table 1).<br />

Location <strong>of</strong> kill, sex and age (through a<br />

premolar for age estimation) are recorded<br />

for every cougar killed, and provide <strong>the</strong> data<br />

used to assess management performance in<br />

relation to established target values that<br />

serve as indicators <strong>of</strong> population status.<br />

“Rules <strong>of</strong> thumb”, expressed as threshold<br />

values <strong>of</strong> 1) a minimum percentage <strong>of</strong> older<br />

aged animals in <strong>the</strong> harvest, 2) a maximum<br />

percentage <strong>of</strong> females in <strong>the</strong> harvest, and 3)<br />

minimum adult survival were set to ensure<br />

that cougar densities are maintained within


PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Table 1. Utah cougar harvests, 1989-1990 thru 2001-2002.<br />

Total Percent Percent<br />

Percent Treed<br />

Hunters Harvest Adult Adult Sub-adult Sub-adult Sport Average Quota Percent Female+ ADC O<strong>the</strong>r Total Adult > per<br />

6 years Pursuit<br />

Year Afield Permits Objective Males Females Males Females Harvest Age Success Filled Females Sub-adult Harvest Mortality Mortality Survival old Day<br />

1989-90 478 527 123 44 23 27 217 41.2% 32.7% 43.3% 48 10 275 0.41<br />

1990-91 480 525 144 46 40 35 265 50.5% 30.6% 45.7% 38 22 325 0.49<br />

1991-92 485 525 128 51 32 30 241 45.9% 33.6% 46.9% 34 22 297 0.45<br />

1992-93 598 591 206 64 54 48 372 62.9% 30.1% 44.6% 53 42 467 0.49<br />

1993-94 575 659 165 87 51 49 352 53.4% 38.6% 53.1% 53 10 415 0.57<br />

1994-95 656 791 205 103 57 66 431 54.5% 39.2% 52.4% 54 24 509<br />

1995-96 787 872 160 105 109 78 452 3.5 51.8% 40.5% 64.6% 33 39 524 0.67 16.7% 0.48<br />

1996-97 1376 595 275 172 172 125 107 576 3.8 56.0% 88.3% 48.4% 70.1% 40 50 666 0.67 20.0% 0.33<br />

1997-98 1370 509 270 204 159 57 72 492 3.2 54.4% 79.6% 47.0% 58.5% 27 23 542 0.63 14.5% 0.36<br />

1998-99 1201 446 230 156 100 50 67 373 3.1 49.0% 64.0% 44.8% 58.2% 13 1 387 0.62 10.1% 0.29<br />

1999-00 817 343 304 194 106 64 71 435 2.9 60.0% 81.0% 40.7% 55.4% 25 9 469 0.57 9.7% 0.28<br />

2000-01 272 371 165 127 77 80 449 3.3 52.0% 35.4% 46.1% 63.3% 73 20 542 0.63 12.8% 0.37<br />

2001-02 258 339 159 108 55 71 393 2.9 45.5% 59.5% 12 7 412 0.61 9.0%<br />

Total 2181 1272 794 801 5048<br />

Average 802.1 531.8 298.2 167.8 97.8 61.1 61.6 388.3 3.2 52.6% 69.7% 39.8% 55.1% 38.7 21.5 448.5 62.7% 13.3% 0.41<br />

Performance Targets: 40.0% 65.0% 15.0% 0.38<br />

54 UTAH MOUNTAIN LION STATUS REPORT · McLaughlin


UTAH MOUNTAIN LION STATUS REPORT · McLaughlin 55<br />

Table 2. Confirmed livestock losses due to cougar depredation in Utah, FY1992 to FY2002.<br />

Total Cougar<br />

Fiscal Year<br />

Number <strong>of</strong> Confirmed Losses:<br />

Incidents Ewes Lambs Bucks Calf Goat O<strong>the</strong>r<br />

Confirmed<br />

Losses<br />

Value<br />

Losses<br />

Taken by<br />

WS<br />

1992 103 175 745 0 4 0 922 34<br />

1993 114 263 722 1 2 0 988 $94,644.00 53<br />

1994 115 258 646 5 6 0 915 $120,615.00 53<br />

1995 152 335 760 24 12 0 1130 $111,495.00 54<br />

1996 112 257 621 2 6 0 878 $79,277.00 33<br />

1997 110 375 531 20 11 0 937 $106,210.00 46<br />

1998 114 253 506 19 13 0 805 $97,703.00 27<br />

1999 69 244 406 18 4 0 730 $92,945.00 11<br />

2000 82 160 371 2 15 0 548 $60,750.00 22<br />

2001 74 136 361 12 3 1 587 $61,395.00 18<br />

2002 95 167 453 18 11 2 1 652 $53,748.42 74<br />

TOTAL 1140 2623 6122 121 87 3 1 8957 $825,034.00 351<br />

all management units, except where predator<br />

management plans are in place. Threshold<br />

values <strong>of</strong> <strong>the</strong> harvest criteria were obtained<br />

from <strong>the</strong> literature and from past evaluations<br />

<strong>of</strong> cougar population dynamics in Utah. This<br />

approach is likely conservative, but it is<br />

justified based upon our limited knowledge<br />

<strong>of</strong> <strong>the</strong> abundance <strong>of</strong> deer and alternate prey<br />

in Utah (UDWR 1999b). Ongoing research<br />

on 2 study sites, under <strong>the</strong> direction <strong>of</strong> Dr.<br />

Michael Wolfe (Utah State University), is<br />

supplying comparative data on <strong>the</strong> dynamics<br />

<strong>of</strong> cougars subjected to varying levels <strong>of</strong><br />

hunting harvest. This information should<br />

help <strong>the</strong> Division refine management criteria<br />

in <strong>the</strong> near future. The Division also<br />

monitors trends in numbers <strong>of</strong> cougar<br />

incident reports, which have fluctuated in<br />

recent years (Table 2). Attempts to reduce<br />

<strong>the</strong> number <strong>of</strong> cougar management units that<br />

are subject to predator management plans<br />

have met with little success, mostly due to<br />

continued drought and deteriorating range<br />

conditions.<br />

HARVEST INFORMATION<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Cougar hunting in Utah is regulated on a<br />

management-unit basis to address<br />

differences in cougar densities, hunter<br />

access and management objectives.<br />

Annually, <strong>the</strong> composition <strong>of</strong> each unit’s<br />

harvest is compared to performance targets<br />

that were selected to maintain cougar<br />

densities: 1) maintain an average <strong>of</strong> 15% or<br />

greater <strong>of</strong> <strong>the</strong> harvest in older age classes<br />

(>6 years <strong>of</strong> age); 2) maintain total adult<br />

survival at or above 65%; 3) restrict <strong>the</strong><br />

female component to


56 UTAH MOUNTAIN LION STATUS REPORT · McLaughlin<br />

The harvest objective strategy is <strong>of</strong>ten<br />

used on units where managers want to<br />

ensure a substantial harvest. This strategy<br />

can result in hunter crowding and less hunter<br />

selectivity toward males, as many hunters<br />

take <strong>the</strong> first cougar <strong>the</strong>y encounter.<br />

Consequently, <strong>the</strong> harvest may be weighted<br />

toward young animals and females.<br />

Conversely, limited entry hunts allow<br />

managers to spread hunting effort over a<br />

longer time period and shift harvesting<br />

pressure toward adult males. This strategy<br />

is commonly used on management units that<br />

are readily accessible to hunters to minimize<br />

crowding and promote hunter selectivity for<br />

adult males.<br />

Since 2001, a few units have been<br />

harvested under a hybrid strategy, where<br />

both harvest objective and limited entry<br />

hunts are held. This approach attempts to<br />

produce a large harvest while encouraging<br />

some hunter selectivity. Under <strong>the</strong> hybrid<br />

strategy, a limited entry hunt is opened<br />

early, followed by a harvest objective hunt<br />

that is delayed until mid-winter. In <strong>the</strong> past,<br />

managers have used female sub quotas in<br />

conjunction with harvest objective strategies<br />

to protect females in <strong>the</strong> face <strong>of</strong> increased<br />

harvest pressure. This strategy has been<br />

discontinued because it biased <strong>the</strong> harvest<br />

sex composition toward females (through<br />

early closure when <strong>the</strong> sub quota was<br />

attained) and prevented meaningful<br />

evaluations <strong>of</strong> harvest sex composition<br />

under criterion 3 above.<br />

Each year, regional wildlife managers<br />

review <strong>the</strong> size and composition <strong>of</strong> harvests<br />

from individual units in relation to<br />

management rules <strong>of</strong> thumb and <strong>the</strong>n make<br />

recommendations for <strong>the</strong> forthcoming<br />

season. Often, <strong>the</strong>ir evaluations result in<br />

changes in <strong>the</strong> number <strong>of</strong> permits allocated,<br />

<strong>the</strong> size <strong>of</strong> quotas and/or changes in harvest<br />

strategy. These regulation changes <strong>of</strong>ten<br />

result in year-to-year fluctuation in harvest<br />

strategy and hence harvest pressure. As a<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

result, variances in harvest size and<br />

composition are difficult to interpret. Total<br />

harvest has varied from 492 to 373 since <strong>the</strong><br />

1997-1998 season, with no definite trend<br />

(Table 1).<br />

Nearly all cougars harvested in Utah are<br />

taken with <strong>the</strong> aid <strong>of</strong> dogs. An individual<br />

hunter is restricted to holding ei<strong>the</strong>r a<br />

limited entry permit or a harvest objective<br />

permit per season, and must wait 3 years to<br />

reapply once he/she acquires a permit. The<br />

bag limit is 1 cougar per season and kittens<br />

and females accompanied by young are<br />

protected from harvest. Currently <strong>the</strong><br />

cougar-hunting season runs from December<br />

14, 2002 through June 1, 2003 on both<br />

limited entry and harvest objective units.<br />

However, some units are open year-round<br />

and some have earlier or later opening dates.<br />

Because harvest objective units close as<br />

soon as <strong>the</strong> objective (quota) is reached,<br />

hunters must call a toll-free number daily to<br />

ensure that <strong>the</strong> season in <strong>the</strong>ir hunt unit is<br />

still open.<br />

Pursuit (chase or no-kill) seasons<br />

provide additional recreational opportunities<br />

over most <strong>of</strong> <strong>the</strong> State. The pursuit season<br />

generally runs December 14, 2002 through<br />

June 1, 2003, but specific units have yearround<br />

pursuit and a few units are closed to<br />

pursuit hunting. In recent years, <strong>the</strong> Division<br />

has sold about 600-700 cougar pursuit<br />

permits annually (Table 3).<br />

The Division began managing cougar<br />

harvests through statewide limited entry<br />

hunting in 1990 and increased numbers <strong>of</strong><br />

Table 3. Number <strong>of</strong> cougar pursuit permits<br />

sold in Utah, 1999-2002.<br />

Year Resident Non-Resident Total<br />

1999-2000 572 49 621<br />

2000-2001 595 59 654<br />

2001-2002 621 84 705<br />

Combined 1788 192 1980


UTAH MOUNTAIN LION STATUS REPORT · McLaughlin 57<br />

Table 4. Comparison <strong>of</strong> harvest characteristics for Utah management units that have predator<br />

management plans (designed to reduce cougar numbers) and units that are managed to sustain<br />

cougar populations.<br />

Criteria (Threshold for<br />

sustaining population)<br />

Predator Management Plan in Place<br />

1999-2000 2000-2001 2001-2002<br />

% Females ( 6 years (>15) 9.7 9.8 10<br />

Adult Survival (>0.65) 0.60 0.61 0.52<br />

Cougar treed/day (0.38) 0.24 0.16<br />

permits through 1995-1996 (Table 1). In<br />

1996-1997, additional harvest pressure was<br />

added by switching some management units<br />

to <strong>the</strong> harvest objective (quota) system and a<br />

record high <strong>of</strong> 1,376 hunters was afield<br />

(Table 1). Since <strong>the</strong>n, <strong>the</strong> number <strong>of</strong> hunters<br />

afield has declined nearly one-third. The<br />

hunting harvest has declined over <strong>the</strong> same<br />

period (Table 1).<br />

Units with predator management plans<br />

designed to reduce cougar densities produce<br />

harvests <strong>of</strong> similar composition to areas<br />

where <strong>the</strong> management objective is to<br />

sustain higher population densities (Table<br />

4). Throughout <strong>the</strong> State, <strong>the</strong> proportion <strong>of</strong><br />

harvest comprised <strong>of</strong> females has usually<br />

been above <strong>the</strong> prescribed threshold for<br />

maintaining cougar densities, <strong>the</strong> percent <strong>of</strong><br />

older aged cougars in <strong>the</strong> harvest has<br />

remained below <strong>the</strong> desired threshold level,<br />

adult survival is below <strong>the</strong> desired level, and<br />

<strong>the</strong> cougar treeing rate is below <strong>the</strong> value<br />

ascribed as an indicator <strong>of</strong> secure population<br />

abundance. Given <strong>the</strong> relative abundance <strong>of</strong><br />

de facto refugia for cougars in Utah<br />

(National Parks, wilderness and inaccessible<br />

tracts) and <strong>the</strong> species’ propensity to<br />

disperse long distances, current harvest<br />

prescriptions may not prove effective for<br />

attaining ei<strong>the</strong>r <strong>of</strong> <strong>the</strong> State’s management<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

No Predator Management Plan<br />

1999-2000 2000-2001 2001-2002<br />

38 46 47<br />

7.6 12.3 9<br />

0.59 0.61 0.62<br />

0.30 0.24<br />

objectives (maintenance <strong>of</strong> population<br />

density, or substantial reduction in<br />

population density).<br />

Evaluation <strong>of</strong> Harvest Information<br />

The harvest-based criteria used in Utah’s<br />

cougar management system are based upon<br />

published research, and represent <strong>the</strong><br />

expectation <strong>of</strong> harvest statistics that are<br />

associated with sustained population<br />

densities. However, managers have not<br />

been able to fully meet all threshold values<br />

since <strong>the</strong> Cougar Management Plan was<br />

adopted in 1999. There may be several<br />

explanations for this difficulty, including <strong>the</strong><br />

geographic scale <strong>of</strong> management actions and<br />

differences in <strong>the</strong> vital rates <strong>of</strong> cougar<br />

populations within Utah.<br />

The proportion <strong>of</strong> mature (>6 years <strong>of</strong><br />

age) cougars in <strong>the</strong> harvest is used as an<br />

index <strong>of</strong> <strong>the</strong> presence <strong>of</strong> mature cougars in<br />

<strong>the</strong> underlying population. If this proportion<br />

declines below 15%, <strong>the</strong> management plan<br />

assumes that <strong>the</strong> harvest rate is<br />

unsustainable. However, scarcity <strong>of</strong> olderaged<br />

cougars in harvests could also result<br />

from light (sustainable) harvesting <strong>of</strong> a<br />

productive cougar population by<br />

nonselective hunters, where relatively few<br />

cougars are taken and <strong>the</strong> harvest is


58 UTAH MOUNTAIN LION STATUS REPORT · McLaughlin<br />

composed <strong>of</strong> mostly subadults and youngeraged<br />

adults.<br />

The proportion <strong>of</strong> adult females in <strong>the</strong><br />

harvest is assumed to increase with<br />

increasing harvest pressure, and <strong>the</strong><br />

threshold level chosen for sustainability in<br />

Utah (>40%) is based upon research from<br />

several western states. However, managers<br />

are evaluating small management units,<br />

some containing


understanding with UDWR. Their reports<br />

are compiled on a fiscal year basis (and<br />

<strong>the</strong>refore numbers/year differ from those<br />

reported in Table 1), and confirm livestock<br />

losses ranging from $53,700 to $97,700 per<br />

year since 1998 (Table 2). Cougars were<br />

implicated in 74-114 separate depredation<br />

incidents per year during this period, killing<br />

548-805 sheep, cattle and goats annually<br />

(Table 2).<br />

RESEARCH AND PUBLICATIONS<br />

UDWR is funding research conducted<br />

through <strong>the</strong> Utah State University, under <strong>the</strong><br />

direction <strong>of</strong> Dr. Michael Wolfe. This<br />

research has been ongoing on two study<br />

sites since 1995, and is directed at<br />

determining means <strong>of</strong> quantifying cougar<br />

populations and evaluating <strong>the</strong> effects <strong>of</strong><br />

harvesting on <strong>the</strong>m. Field research is<br />

currently underway by David Stoner, MS<br />

candidate.<br />

Recent Publications<br />

MAXFIELD, BRIAN D. 2002. Utah cougar<br />

harvest report 1998-1999. Annual<br />

UTAH MOUNTAIN LION STATUS REPORT · McLaughlin 59<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

performance report, Fed. Aid Project No.<br />

W-150-R-8. Publ. No. 02-07, Utah Div.<br />

Wildlife Res., Salt Lake City. 38 pp.<br />

MAXFIELD, BRIAN D. 2002. Utah cougar<br />

harvest report 1999-2000. Annual<br />

performance report, Fed. Aid Project No.<br />

W-150-R-8. Publ. No. 02-08, Utah Div.<br />

Wildlife Res., Salt Lake City. 41 pp.<br />

LITERATURE CITED<br />

RILEY, S.J. 1998. Integration <strong>of</strong><br />

environmental, biological, and human<br />

dimensions for management <strong>of</strong> mountain<br />

lions (Puma concolor) in Montaina. Ph.<br />

D. Diss., Cornell Univ. 158 pp.<br />

UDWR. 1996. Predator Management<br />

Policy. Utah Div. <strong>of</strong> Wildlife Res. Salt<br />

Lake City. UDWR. 1999a. Nuisance<br />

Cougar Complaints. Policy No.<br />

W5WLD-5. Utah Div. <strong>of</strong> Wildlife Res. 4<br />

pp.<br />

UDWR. 1999B. UTAH COUGAR MANAGEMENT<br />

PLAN. UTAH DIV. OF WILDLIFE RES. SALT<br />

LAKE CITY. 60 PP.


WASHINGTON COUGAR STATUS REPORT<br />

RICHARD A. BEAUSOLEIL, Bear / Cougar Specialist, Washington Department <strong>of</strong> Fish and<br />

Wildlife, 3515 Chelan Highway, Wenatchee, Washington, 98801, USA<br />

DONALD A. MARTORELLO, Bear, Cougar, and Special Species Section Manager,<br />

Department <strong>of</strong> Fish and Wildlife, 600 Capitol Way North, Olympia, Washington, 98501,<br />

USA<br />

ROCKY D. SPENCER, Dangerous Wildlife Specialist, Washington Department <strong>of</strong> Fish and<br />

Wildlife, 42404 North Bend Way SE, North Bend, Washington, 98045, USA<br />

INTRODUCTION<br />

Cougar (Puma concolor) occur<br />

throughout most <strong>of</strong> <strong>the</strong> forested regions <strong>of</strong><br />

Washington State, encompassing<br />

approximately 88,497 km 2 or 51% <strong>of</strong> <strong>the</strong><br />

State (Figure 1). Cougar became a protected<br />

big game species in 1966 and hunting<br />

seasons and harvest limits were established<br />

under <strong>the</strong> management authority <strong>of</strong><br />

Washington Department <strong>of</strong> Fish and<br />

Wildlife (WFDW). In 1967, <strong>the</strong><br />

Washington State Legislature passed a bill<br />

establishing a tag system in Washington. In<br />

1970, WDFW began mandatory reporting <strong>of</strong><br />

cougar kills and in 1979 inspection and<br />

sealing <strong>of</strong> cougar pelts was required for data<br />

collection. In <strong>the</strong> mid-1980’s WDFW began<br />

collecting cougar teeth for age analysis.<br />

Figure 1. Distribution <strong>of</strong> cougars (gray) and<br />

cougar management units in Washington.<br />

60<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Currently, <strong>the</strong> statewide cougar management<br />

goal is to maintain healthy, self-sustaining<br />

cougar populations within each cougar<br />

management unit (CMU), except CMU 9,<br />

while minimizing <strong>the</strong> number <strong>of</strong> negative<br />

human-cougar interactions.<br />

HUNTING SEASONS AND HARVEST<br />

TRENDS<br />

Cougar seasons have changed<br />

significantly over <strong>the</strong> last several years<br />

(Figure 2). During <strong>the</strong> November 1996<br />

general election, Washington voters passed<br />

Initiative 655 (I-655) that banned <strong>the</strong> use <strong>of</strong><br />

hounds for hunting cougar and bobcat, and<br />

<strong>the</strong> use <strong>of</strong> bait and hounds for hunting black<br />

bear. In an effort to mitigate <strong>the</strong> anticipated<br />

decrease in cougar harvest (i.e., post I-655),<br />

permit-only seasons were replaced with<br />

general seasons, cougar seasons were<br />

leng<strong>the</strong>ned from approximately 6 weeks to 7<br />

and one-half months, and bag limit was<br />

increased from 1 to 2 cougar/year.<br />

Legislation was also passed that provided<br />

<strong>the</strong> authority to <strong>the</strong> Fish and Wildlife<br />

Commission to establish reduced costs for<br />

cougar and black bear transport tags, which<br />

<strong>the</strong>y did from $24 to $5 in 1996 (cougar tags<br />

can also be purchased as part <strong>of</strong> a big game<br />

package). The outcome <strong>of</strong> <strong>the</strong>se strategies is<br />

that <strong>the</strong> number <strong>of</strong> hunters purchasing a<br />

cougar tag in Washington has increased<br />

from 1,000 to 59,000. As a result, annual


350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

General Seasons<br />

Dogs Allowed<br />

Kill report required<br />

WASHINGTON COUGAR STATUS REPORT · Beausoleil et al. 61<br />

1979 - 1986<br />

1987 - 1995 1996 - 2002<br />

Figure 2. Cougar season structure and harvest in Washington, 1979-2002.<br />

cougar harvest during post I-655 years has<br />

increased slightly; however, <strong>the</strong> composition<br />

<strong>of</strong> <strong>the</strong> harvest has changed dramatically.<br />

The majority <strong>of</strong> cougar harvested pre-I 655<br />

was done so with <strong>the</strong> aid <strong>of</strong> dogs, thus<br />

mostly males and older animals were taken.<br />

Since 1996, <strong>the</strong> majority <strong>of</strong> cougars are<br />

harvested ei<strong>the</strong>r as opportunistic encounters<br />

by deer/elk and cougar hunters, or by using<br />

tracking and calling techniques. These<br />

harvest methods are not as selective as using<br />

dogs. Therefore, since 1996, hunters have<br />

harvested more females and younger<br />

cougars (see oral presentation titled Cougar<br />

Harvest Characteristics With and Without<br />

<strong>the</strong> Use <strong>of</strong> Dogs in this proceedings).<br />

POPULATION STATUS AND TREND<br />

ANALYSIS<br />

The status <strong>of</strong> cougar populations is<br />

currently estimated through computer<br />

population simulation models, harvest<br />

characteristics, and, to a lesser degree,<br />

trends in human-cougar interactions.<br />

Based on population reconstruction<br />

models, harvest age data, and statewide<br />

cougar habitat estimates (using GAP<br />

analysis), <strong>the</strong> cougar population in<br />

Permit Seasons<br />

Dogs Allowed<br />

I 655<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

General Seasons<br />

Dogs Banned<br />

Washington is likely between 2,400–4,000<br />

animals, and cougar population size is likely<br />

declining in a few areas <strong>of</strong> <strong>the</strong> state.<br />

Typically, <strong>the</strong> status <strong>of</strong> local or regional<br />

cougar populations are monitored via hunter<br />

effort and success, median age data, and<br />

percentage <strong>of</strong> females in <strong>the</strong> harvest; but<br />

only when viewed over several years with<br />

consistent harvest methods. Due to <strong>the</strong><br />

changes in harvest methods during <strong>the</strong> last<br />

several years (predominantly hound hunters<br />

during pre I-655 years versus entirely spotstalk<br />

hunters during post I-655 years), no<br />

reliable trend data exist to accurately assess<br />

regional cougar populations or exploitation<br />

levels. As such, new population monitoring<br />

efforts are beginning in 2003, where cougar<br />

density and adult female survival will be<br />

evaluated and monitored in key areas <strong>of</strong> <strong>the</strong><br />

State.<br />

HUMAN CONFLICT<br />

Human-cougar interactions are managed<br />

through public education, capture-removal,<br />

depredation permits, and public safety<br />

cougar removals. Since 1995, WDFW has<br />

recorded information on human-cougar<br />

interactions. Of particular concern is <strong>the</strong>


62 WASHINGTON COUGAR STATUS REPORT · Beausoleil et al.<br />

To address human safety<br />

To protect threatened and<br />

endangered species<br />

To prevent loss <strong>of</strong> domestic<br />

animals<br />

To increase game populations<br />

0 20 40 60 80 100<br />

Figure 3. During a general public opinion<br />

survey, <strong>the</strong> percent <strong>of</strong> Washington<br />

respondents that supported reducing<br />

predator numbers for specific purposes<br />

(Duda et al. 2002).<br />

increasing trend in human safety incidents,<br />

and pet and livestock depredations. When<br />

Washington citizens were asked about <strong>the</strong>ir<br />

attitudes regarding cougars, over 80%<br />

responded that reducing predator numbers<br />

for public safety is acceptable (Figure 3).<br />

Recognizing <strong>the</strong> widespread scope <strong>of</strong> <strong>the</strong><br />

issue and its importance to cougars and<br />

people in <strong>the</strong> future, current cougar<br />

management goals include maintaining<br />

sustainable cougar populations and reducing<br />

human-cougar interactions. In some cases,<br />

reducing cougar populations to a lower, but<br />

sustainable level may help achieve both <strong>of</strong><br />

<strong>the</strong>se goals (Table 1). Given <strong>the</strong> recent<br />

Confirmed complaints<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

0<br />

247<br />

495<br />

563<br />

927<br />

694<br />

936<br />

498<br />

378<br />

1995 1996 1997 1998 1999 2000 2001 2002<br />

Year<br />

Figure 4. Total confirmed cougar complaints<br />

in Washington, 1995-2002 (includes human<br />

safety and pet/livestock incidents).<br />

history <strong>of</strong> high human-cougar interactions,<br />

WDFW developed a special cougar removal<br />

process to address cougar densities in areas<br />

with high levels <strong>of</strong> human-cougar<br />

interactions. Under rules adopted by <strong>the</strong><br />

Fish and Wildlife Commission, public safety<br />

cougar removals occurred in 17 Game<br />

Management Units from Dec 15 – Mar. 15,<br />

in both <strong>the</strong> 2001-2002 and 2002-03 seasons;<br />

in those seasons 109 and 76 cougar were<br />

identified for removal and licensed hunters<br />

Table 1. Cougar population objectives for each cougar management unit in Washington, 2002.<br />

CMU Geographic Area Population Objective<br />

1 Coastal Maintain a stable cougar population<br />

2 Puget Sound Reduce * cougar population to enhance public safety and protection <strong>of</strong> property<br />

3 North Cascades Maintain a stable cougar population<br />

4 South Cascades Maintain a stable cougar population<br />

5 East Cascades North Reduce * cougar population to enhance public safety and protection <strong>of</strong> property<br />

6 East Cascades South Maintain a stable cougar population<br />

7 Nor<strong>the</strong>astern Reduce * cougar population to enhance public safety and protection <strong>of</strong> property<br />

8 Blue <strong>Mountain</strong>s Maintain a stable cougar population<br />

9 Columbia Basin Unsustainable; not considered suitable cougar habitat<br />

* Implement cougar population reductions over a 3-year period and monitor annually.


emoved 67 and 54 animals, respectively<br />

(61% and 71% success rate, respectively).<br />

Confirmed human-cougar incidents<br />

decreased by 47% during <strong>the</strong> 2001 calendar<br />

year from 936 in 2000 to 498 and an<br />

additional 24% in 2002 to 378 (Figure 4).<br />

MANAGEMENT CONCLUSIONS<br />

The statewide cougar population appears<br />

to be declining at this time due to increased<br />

female harvest and objectives to address<br />

public safety and protection <strong>of</strong> property.<br />

Given <strong>the</strong> distribution <strong>of</strong> cougars in<br />

Washington and <strong>the</strong> projected growth <strong>of</strong><br />

human populations, interactions between<br />

humans and cougars will likely continue.<br />

WASHINGTON COUGAR STATUS REPORT · Beausoleil et al. 63<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

As such, <strong>the</strong> long-term future <strong>of</strong> cougar in<br />

Washington ultimately rests in our ability to<br />

co-exist. Therefore, management efforts<br />

should continue to look for ways to<br />

minimize human-cougar interactions,<br />

particularly at <strong>the</strong> local population level.<br />

LITERATURE CITED<br />

DUDA, M.D., P.E. DE MICHELE, M. JONES,<br />

W. TESTERMAN, C. ZURAWSKI, J.<br />

DEHOFF, A. LANIER, S.J. BISSELL, P.<br />

WANG, AND J.B. HERRICK. 2002.<br />

Washington residents’ opinions on and<br />

attitudes toward hunting and game<br />

species management. Harrisonburg,<br />

Virginia, USA.


WYOMING MOUNTAIN LION STATUS REPORT<br />

SCOTT A. BECKER, Trophy Game Section, Wyoming Game and Fish Department, 260 Buena<br />

Vista, Lander, WY 82520, USA, email: Scott.Becker@wgf.state.wy.us<br />

DANIEL D. BJORNLIE, Trophy Game Biologist, Wyoming Game and Fish Department, 260<br />

Buena Vista, Lander, WY 82520, USA, email: Dan.Bjornlie@wgf.state.wy.us<br />

DAVID S. MOODY, Trophy Game Section Coordinator, Wyoming Game and Fish Department,<br />

260 Buena Vista, Lander, WY 82520, USA, email: Dave.Moody@wgf.state.wy.us<br />

INTRODUCTION<br />

Management <strong>of</strong> mountain lions (Puma<br />

concolor) has changed markedly since <strong>the</strong><br />

nineteenth century. In 1882, <strong>the</strong> Wyoming<br />

Territorial government enacted legislation<br />

placing a bounty on mountain lions and<br />

o<strong>the</strong>r predators. This allowed for lion<br />

hunting throughout <strong>the</strong> year and no bag<br />

limits were enforced. In 1973, <strong>the</strong> mountain<br />

lion was reclassified as a trophy game<br />

animal, which made <strong>the</strong> Wyoming Game<br />

and Fish Department (WGFD) fiscally liable<br />

for confirmed livestock losses. The<br />

following year, <strong>the</strong> first hunting season was<br />

established that included <strong>the</strong> entire state as a<br />

single hunt area, a bag limit <strong>of</strong> 1 lion per<br />

year was enacted, kittens and females with<br />

kittens at side were protected, and hunters<br />

were required to present skulls and pelts <strong>of</strong><br />

harvested lions to <strong>the</strong> nearest WGFD<br />

District Office or local game warden.<br />

In 1997, <strong>the</strong> WGFD prepared a draft<br />

management plan for mountain lions, but <strong>the</strong><br />

plan has yet to be finalized. However, six<br />

main objectives outlined in <strong>the</strong> draft<br />

management plan continue to guide lion<br />

management objectives for <strong>the</strong> state <strong>of</strong><br />

Wyoming, <strong>the</strong>y are: 1) maintain mountain<br />

lion populations within suitable habitat<br />

throughout Wyoming; 2) provide mountain<br />

lion-related recreational opportunities; 3)<br />

minimize female lion harvest in areas where<br />

population stability or increase is desirable;<br />

4) minimize mountain lion depredation and<br />

64<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

lion/human interactions; 5) tailor<br />

management objectives to conditions present<br />

within each <strong>Mountain</strong> <strong>Lion</strong> Management<br />

Unit (MLMU) where possible; and 6)<br />

implement more specific, quantifiable<br />

objectives within each MLMU as<br />

information on <strong>the</strong> state’s lion population<br />

allows. Using <strong>the</strong>se objectives as<br />

guidelines, <strong>the</strong> WGFD attempts to balance<br />

recreational demand and harvest with <strong>the</strong><br />

biological needs <strong>of</strong> lion populations<br />

throughout <strong>the</strong> state.<br />

DISTRIBUTION AND ABUNDANCE<br />

<strong>Mountain</strong> lions are distributed<br />

throughout nearly all habitats in Wyoming<br />

although densities are not uniform. <strong>Lion</strong><br />

densities are thought to be highest in <strong>the</strong><br />

Bighorn, Owl Creek, and Laramie mountain<br />

ranges (Wyoming Game and Fish<br />

Department 1997), while some <strong>of</strong> <strong>the</strong> lowest<br />

densities may be found in <strong>the</strong> grasslands <strong>of</strong><br />

nor<strong>the</strong>astern Wyoming. In <strong>the</strong> Bighorn<br />

<strong>Mountain</strong>s, Logan and Irwin (1985) found<br />

that mixed conifer and curl leaf mountain<br />

mahogany habitats were used most in<br />

relation to availability, whereas sagebrush<br />

grass habitat types were generally avoided.<br />

In <strong>the</strong> Snowy Range <strong>Mountain</strong>s <strong>of</strong><br />

sou<strong>the</strong>astern Wyoming, lions were found at<br />

lower elevations during <strong>the</strong> winter and<br />

concentrated <strong>the</strong>ir use near <strong>the</strong> timber/prairie<br />

interface (Chuck Anderson, personal<br />

communication).


Figure 1. <strong>Mountain</strong> lion management units<br />

and hunt areas in Wyoming, 2002.<br />

WYOMING MOUNTAIN LION STATUS REPORT · Becker et al. 65<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

HARVEST INFORMATION<br />

Data on mountain lions are ga<strong>the</strong>red<br />

annually among 28 hunt areas that are<br />

grouped into 5 MLMUs (Figure 1), <strong>the</strong><br />

boundaries <strong>of</strong> which encompass large areas<br />

with contiguous topographic features and<br />

are believed to encompass population<br />

centers. Each hunt area has a maximum<br />

annual mortality quota that varies from 2 –<br />

34, with 5 areas also having a maximum<br />

female mortality quota (Table 1). If ei<strong>the</strong>r<br />

quota is filled, <strong>the</strong> hunting season in that<br />

hunt area automatically closes. Currently,<br />

hunting seasons open on September 1 and<br />

close on March 31 for all hunt areas except<br />

Table 1. Wyoming mountain lion management units, hunt areas, season dates, and quotas for<br />

harvest year 2002.<br />

<strong>Mountain</strong> <strong>Lion</strong><br />

Management Unit<br />

Nor<strong>the</strong>ast<br />

Sou<strong>the</strong>ast<br />

Southwest<br />

North-Central<br />

West<br />

Hunt Area Season Dates<br />

Annual Mortality<br />

Quota<br />

Annual Female<br />

Mortality Quota<br />

1 Sept. 1-Mar. 31 7<br />

24 Sept. 1-Mar. 31 2<br />

5 Sept. 1-Mar. 31 12<br />

6 Sept. 1-Mar. 31 34<br />

7 Sept. 1-Mar. 31 6<br />

8 Sept. 1-Mar. 31 8<br />

9 Sept. 1-Mar. 31 3<br />

25 Sept. 1-Mar. 31 3<br />

27 Sept. 1-Aug. 31 20<br />

10 Sept. 1-Mar. 31 6<br />

11 Sept. 1-Mar. 31 2<br />

12 Sept. 1-Mar. 31 6 3<br />

13 Sept. 1-Mar. 31 3<br />

16 Sept. 1-Mar. 31 6<br />

15 Sept. 1-Aug. 31 25<br />

21 Sept. 1-Mar. 31 25<br />

22 Sept. 1-Aug. 31 15<br />

23 Sept. 1-Mar. 31 15 8<br />

2 Sept. 1-Mar. 31 12 6<br />

3 Sept. 1-Mar. 31 8 4<br />

4 Sept. 1-Mar. 31 4<br />

14 Sept. 1-Mar. 31 9<br />

17 Sept. 1-Mar. 31 5<br />

18 Sept. 1-Mar. 31 12<br />

19 Sept. 1-Mar. 31 20<br />

20 Sept. 1-Mar. 31 15<br />

26 Sept. 1-Mar. 31 12 7<br />

28 Sept. 1-Mar. 31 3


66 WYOMING MOUNTAIN LION STATUS REPORT · Becker et al.<br />

15, 22, and 27, in which year round seasons<br />

exist. Quotas begin at <strong>the</strong> start <strong>of</strong> each<br />

hunting season and include all legal and<br />

illegal hunting mortalities.<br />

<strong>Mountain</strong> lion data in Wyoming are<br />

limited to information obtained annually<br />

from harvest or o<strong>the</strong>r documented forms <strong>of</strong><br />

mortality. Since 1974, hunters have been<br />

required to present <strong>the</strong> skull and pelt <strong>of</strong><br />

harvested lions to a district game warden or<br />

biologist at <strong>the</strong> nearest WGFD regional<br />

<strong>of</strong>fice within 72 hours after <strong>the</strong> harvest.<br />

Information collected during <strong>the</strong>se<br />

inspections include: harvest date, location,<br />

sex, lactation status, estimated age, number<br />

<strong>of</strong> days spent hunting, whe<strong>the</strong>r or not dogs<br />

were used, and number <strong>of</strong> lions observed<br />

while hunting. Skulls and pelts must be<br />

presented in an unfrozen condition so teeth<br />

can be removed. Evidence <strong>of</strong> sex must<br />

remain naturally attached to <strong>the</strong> pelt for<br />

accurate identification.<br />

Legal shooting hours are from one-half<br />

hour before sunrise to one-half hour after<br />

sunset. The individual bag limit for lions is<br />

1 lion per hunter per calendar year, except<br />

for 1 hunt area in central Wyoming, where 1<br />

additional lion may be taken each calendar<br />

year. Kittens (


Total <strong>Lion</strong> Harvest<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

78<br />

1993<br />

1994<br />

95 110<br />

145 144<br />

206<br />

214<br />

201<br />

186<br />

172<br />

1995<br />

1996<br />

1997<br />

1998<br />

1999<br />

2000<br />

2001<br />

2002<br />

Figure 2. Total Wyoming mountain lion<br />

harvest, 1993-2002.<br />

information into mountain lion harvest<br />

analyses in order to better assess mountain<br />

lion population trends. This will eventually<br />

aid in adjusting population objectives and,<br />

thus quotas, to ensure sustainable lion<br />

populations statewide.<br />

There has been a steady increase in<br />

harvest since 1993, which has leveled <strong>of</strong>f in<br />

recent years at around 200 (Figure 2). Since<br />

1993, <strong>the</strong> average percent <strong>of</strong> females in <strong>the</strong><br />

harvest has been 43%, ranging from 32% in<br />

1993 to 51% in 2000 (Figure 3). The<br />

percent <strong>of</strong> adults in <strong>the</strong> female harvest has<br />

steadily declined in <strong>the</strong> past 10 years, falling<br />

from around 70% adult females in 1993 and<br />

1994 to around 40% adults in 2001 and 2002<br />

(Figure 4). This decline in <strong>the</strong> past two<br />

years is likely due in part to a change in <strong>the</strong><br />

criteria used to classify adults and juveniles<br />

prior to <strong>the</strong> 2001 hunting season. Since<br />

1993, hunter effort has ranged from 3.3 to<br />

5.8 days per lion for an average <strong>of</strong> 3.9 days<br />

per lion. Ninety-two percent <strong>of</strong> all<br />

successful hunters in Wyoming harvested<br />

lions with <strong>the</strong> aid <strong>of</strong> dogs from 1993 – 2002.<br />

DEPEDATIONS AND HUMAN-LION<br />

INTERACTIONS/CONFLICTS<br />

Currently, Wyoming uses a statewide<br />

protocol for managing trophy game<br />

WYOMING MOUNTAIN LION STATUS REPORT · Becker et al. 67<br />

Percent<br />

100%<br />

75%<br />

50%<br />

25%<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

0%<br />

1993<br />

1994<br />

1995<br />

1996<br />

1997<br />

1998<br />

1999<br />

2000<br />

2001<br />

2002<br />

Percent Females Percent Males<br />

Figure 3. Percent male and female mountain<br />

lion harvest in Wyoming, 1993-2002.<br />

depredations and interactions with humans.<br />

A depredating lion is defined as a lion that<br />

injures or kills livestock or domestic pets.<br />

In addition, 4 types <strong>of</strong> human/mountain lion<br />

interactions are defined by <strong>the</strong> WGFD, <strong>the</strong>y<br />

are 1) recurring sighting – repeated sightings<br />

<strong>of</strong> a particular lion; 2) encounter – an<br />

unexpected meeting between a human and a<br />

lion without incident; 3) incident – an<br />

account <strong>of</strong> abnormal lion behavior that could<br />

have more serious results in <strong>the</strong> future (e.g.,<br />

a lion attacking a pet, or a lion exhibiting<br />

aggressive behavior, without attack, toward<br />

Percent<br />

100%<br />

75%<br />

50%<br />

25%<br />

0%<br />

1993<br />

1994<br />

1995<br />

1996<br />

1997<br />

1998<br />

1999<br />

2000<br />

Adult Females Juvenile Females<br />

2001<br />

2002<br />

Figure 4. Percent adult and juvenile female<br />

mountain lion harvest in <strong>the</strong> total female<br />

harvest in Wyoming, 1993-2002.


68 WYOMING MOUNTAIN LION STATUS REPORT · Becker et al.<br />

humans); and 4) attack – human injury or<br />

death resulting from a lion attack. Each<br />

incident is handled on a case-by-case basis<br />

and is dealt with accordingly based on <strong>the</strong><br />

location <strong>of</strong> <strong>the</strong> incident, <strong>the</strong> threat to human<br />

safety, <strong>the</strong> severity <strong>of</strong> <strong>the</strong> incident, and <strong>the</strong><br />

number <strong>of</strong> incidents <strong>the</strong> animal has been<br />

involved in. Every effort is made to prevent<br />

unnecessary escalation <strong>of</strong> incidents through<br />

an ascending order <strong>of</strong> options and<br />

responsibilities:<br />

1) No Management Action Taken<br />

- Informational packets are provided<br />

to <strong>the</strong> reporting party that describe<br />

mountain lion natural history and<br />

behavior, damage prevention tips,<br />

and what to do in <strong>the</strong> event <strong>of</strong> an<br />

encounter.<br />

2) Deterrent Methods<br />

- Removal or securing <strong>of</strong> attractant<br />

- Removal <strong>of</strong> depredated carcass<br />

- Removal or protection <strong>of</strong> livestock<br />

3) Aversive Conditioning<br />

- Use <strong>of</strong> rubber bullets<br />

- Use <strong>of</strong> pepper spray<br />

- Use <strong>of</strong> noise making devices or<br />

flashing lights<br />

- Informational packets provided to<br />

<strong>the</strong> reporting party<br />

4) Trapping and Relocation<br />

- If <strong>the</strong> above efforts do not deter <strong>the</strong><br />

lion from <strong>the</strong> area, if public safety<br />

is compromised, if it is a first<br />

<strong>of</strong>fense, <strong>of</strong> if it has been a lengthy<br />

span <strong>of</strong> time between <strong>of</strong>fenses<br />

- Informational packets provided to<br />

<strong>the</strong> reporting party<br />

5) Lethal Removal <strong>of</strong> <strong>the</strong> Animal by <strong>the</strong><br />

WGFD<br />

- If <strong>the</strong> above methods do not deter<br />

<strong>the</strong> lion, if public safety is<br />

compromised, or if <strong>the</strong> <strong>of</strong>fending<br />

lion has been involved in multiple<br />

incidents in a short span <strong>of</strong> time<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

- Wyoming statute 23-3-115 allows<br />

property owners or <strong>the</strong>ir employees<br />

and lessees to kill mountain lions<br />

damaging private property, given<br />

that <strong>the</strong>y immediately notify <strong>the</strong><br />

nearest game warden <strong>of</strong> <strong>the</strong><br />

incident<br />

- <strong>Lion</strong>s that have been removed from<br />

<strong>the</strong> population will be used for<br />

educational purposes<br />

- Informational packets provided to<br />

<strong>the</strong> reporting party<br />

Education is a very important aspect <strong>of</strong><br />

human and mountain lion interaction<br />

prevention. Therefore, <strong>the</strong> WGFD works<br />

closely with hunters, outfitters,<br />

recreationalists, livestock operators, and<br />

homeowners in an attempt to minimize<br />

conflicts with trophy game animals. Every<br />

spring, <strong>the</strong> WGFD hosts bear and lion<br />

workshops throughout <strong>the</strong> state to inform <strong>the</strong><br />

public about bear and lion biology, front and<br />

back-country food storage techniques, and<br />

what to do in <strong>the</strong> event <strong>of</strong> an encounter with<br />

a bear or lion. In addition, numerous<br />

presentations are given throughout <strong>the</strong> year<br />

to civic, private, and school groups. Media<br />

outlets are also used to inform, and in rare<br />

incidents warn, <strong>the</strong> general public about bear<br />

and lion safety issues and any recent<br />

sightings.<br />

Even with all <strong>the</strong> educational efforts<br />

undertaken by <strong>the</strong> WGFD and preventive<br />

measures taken by <strong>the</strong> public, conflicts with<br />

mountain lions do occur. The number <strong>of</strong><br />

mountain lion conflicts have ranged from a<br />

low <strong>of</strong> 13 reported incidents in 2002 to a<br />

high <strong>of</strong> 64 reported incidents in 1997. There<br />

have been a total <strong>of</strong> 40 mountain lion/human<br />

interactions in Wyoming since 1996 with no<br />

major injuries or deaths reported.<br />

Wyoming statute 23-1-901 provides<br />

monetary compensation for confirmed<br />

livestock damage caused by mountain lions.<br />

The number <strong>of</strong> damage claims for <strong>the</strong> last 10<br />

years range from 11 in 1995 to 28 in 1998,


and payments made to claimants range from<br />

a low <strong>of</strong> $22,627 paid in 1999 to a high <strong>of</strong><br />

$44,071 paid in 1998 (Table 2). One<br />

hundred percent <strong>of</strong> <strong>the</strong> mountain lion<br />

damage claims paid in 2002 was for sheep<br />

depredations. From 1996 to 2002, 84% <strong>of</strong><br />

reported lion depredations in Wyoming have<br />

involved sheep, 6% have involved horses,<br />

6% unknown livestock species, and 4% have<br />

involved cattle. An average <strong>of</strong> 4.9 nuisance<br />

lions were removed annually in <strong>the</strong> last 10<br />

years while an average <strong>of</strong> 1 lion was<br />

translocated annually from 1996 – 2002 (no<br />

translocation data available prior to 1996).<br />

PUBLIC ATTITUDES<br />

In 1995, <strong>the</strong> WGFD contracted with <strong>the</strong><br />

Survey Research Center at <strong>the</strong> University <strong>of</strong><br />

Wyoming to determine attitudes and<br />

knowledge <strong>of</strong> Wyoming residents on<br />

mountain lions and mountain lion<br />

management (Gasson and Moody 1995).<br />

Over 71% <strong>of</strong> <strong>the</strong> approximately 500<br />

WYOMING MOUNTAIN LION STATUS REPORT · Becker et al. 69<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

respondents believed lions were a benefit to<br />

Wyoming. Attitudes toward mountain lion<br />

hunting were generally supportive, with<br />

49.6% agreeing or strongly agreeing that<br />

mountain lion hunting should continue and<br />

29.3% disagreeing or strongly disagreeing.<br />

The remaining respondents were ei<strong>the</strong>r<br />

neutral or did not answer. However, most<br />

(57%) disagreed or strongly disagreed that<br />

hunting lions with dogs should continue as a<br />

legal method <strong>of</strong> take. Only 25.3% <strong>of</strong><br />

respondents agreed or strongly agreed, while<br />

<strong>the</strong> remaining respondents were neutral or<br />

did not respond to <strong>the</strong> question. A large<br />

majority <strong>of</strong> respondents (80.7%) agreed or<br />

strongly agreed that mountain lion hunting<br />

seasons should be modified to avoid<br />

harvesting kittens or running females with<br />

kittens. A large majority <strong>of</strong> respondents<br />

(71%) were also opposed to <strong>the</strong> use <strong>of</strong> dogs<br />

to run and tree lions during non-harvest,<br />

chase seasons.<br />

Table 2. Wyoming ten-year mountain lion damage claim and translocation/removal history (all<br />

causes).<br />

Year # Claims $ Claimed $ Paid Translocations Removals<br />

1993 29 33,214.56 30,002.53<br />

0<br />

1994 26 30,498.51 24,646.00<br />

a<br />

5<br />

1995 11 40,634.67 34,594.67<br />

a<br />

4<br />

1996 14 28,540.96 24,947.95 0 6<br />

1997 20 28,935.16 28,761.50 1 10<br />

1998 28 56,171.39 44,070.79 2 5<br />

1999 21 32,307.63 22,627.43 2 6<br />

2000 20 42,352.69 30,773.59 0 5<br />

2001 15 38,322.79 25,592.46 1 6<br />

2002 13 35,870.99 32,075.05 0 2<br />

Mean 19.7 36,686.74 29,809.20 0.86 4.9<br />

a<br />

No data available.<br />

a


70 WYOMING MOUNTAIN LION STATUS REPORT · Becker et al.<br />

RESEARCH AND PUBLICATIONS<br />

ANDERSON, C.R., JR., AND F.G. LINDZEY.<br />

2003. Estimating cougar predation rates<br />

from GPS location clusters. Journal <strong>of</strong><br />

Wildlife Management 67(2): 307-316.<br />

ANDERSON, C.R., JR., F.G. LINDZEY, AND<br />

N.P. NIBBELINK. In review. Estimating<br />

cougar abundance using probability<br />

sampling: an evaluation <strong>of</strong> transect<br />

versus block design. Journal <strong>of</strong> Wildlife<br />

Management 00(0): 000-000.<br />

ANDERSON, C.R., JR., AND F.G. LINDZEY.<br />

In press. Monitoring changes in cougar<br />

sex/age structure with changes in<br />

abundance as an index to population<br />

trend.<br />

ANDERSON, C.R., JR., F.G. LINDZEY, AND<br />

D.B. MCDONALD. In press. Genetic<br />

structure <strong>of</strong> cougar populations across<br />

<strong>the</strong> Wyoming Basin: metapopulation or<br />

megapopulation.<br />

ANDERSON, C.R., JR., AND F.G. LINDZEY.<br />

2000. A guide to estimating mountain<br />

lion age classes. Wyoming Cooperative<br />

Fish and Wildlife Research Unit,<br />

Laramie, Wyoming.<br />

GASSON, W., AND D. MOODY. 1995.<br />

Attitudes <strong>of</strong> Wyoming residents on<br />

mountain lion management. Planning<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

rep. #40, Wyoming Game and Fish<br />

Department, Cheyenne. 7 pp.<br />

WYOMING GAME AND FISH DEPARTMENT.<br />

1997. <strong>Mountain</strong> <strong>Lion</strong> Management<br />

Plan. Wyoming Game and Fish<br />

Department. 30 pp.<br />

WYOMING GAME AND FISH DEPARTMENT.<br />

1999. Protocol for managing aggressive<br />

wildlife/human interactions. Wyoming<br />

Game and Fish Department. 17 pp.<br />

WYOMING GAME AND FISH DEPARTMENT.<br />

2003. Annual mountain lion mortality<br />

summary: harvest year 2002. Trophy<br />

Game Section, Lander, Wyoming. 22<br />

pp.<br />

LITERATURE CITED<br />

GASSON W. AND D. MOODY. 1995.<br />

Attitudes <strong>of</strong> Wyoming residents on<br />

mountain lion management. Planning<br />

rep. #40, Wyoming Game and Fish<br />

Department, Cheyenne, 7 pp.<br />

LOGAN, K.A. AND L.L. IRWIN. 1985.<br />

<strong>Mountain</strong> lion habitats in <strong>the</strong> Bighorn<br />

<strong>Mountain</strong>s, Wyoming. Wildlife Society<br />

Bulletin 13: 257-262.<br />

WYOMING GAME AND FISH DEPARTMENT.<br />

1997. <strong>Mountain</strong> <strong>Lion</strong> Management<br />

Plan. Wyoming Game and Fish<br />

Department. 30 pp.


CRYPTIC COUGARS - PERSPECTIVES ON THE PUMA IN THE EASTERN,<br />

MIDWESTERN, AND GREAT PLAINS REGIONS OF NORTH AMERICA<br />

JAY W. TISCHENDORF DVM, Director, American Ecological Research Institute (AERIE),<br />

Post Office Box 1826, Great Falls, MT 59403, USA, email: TischendorfJ@Hotmail.com<br />

Abstract: The subject <strong>of</strong> cougars in eastern North America continues to intrigue and perplex wildlife biologists,<br />

managers, and nature enthusiasts. Almost uniformly considered extirpated throughout states and provinces in<br />

eastern and midwestern North America over a century ago, growing numbers <strong>of</strong> reports, some accompanied by<br />

incontrovertible evidence such as full specimens, blood, scat, track, or film documentation, suggest that Puma<br />

concolor is re-establishing, or has re-established, itself in some areas <strong>of</strong> this vast region. Similar evidence exists for<br />

<strong>the</strong> Great Plains. This paper, while probably raising more questions than it answers, examines <strong>the</strong> best and most<br />

current evidence for <strong>the</strong> occurrence <strong>of</strong> cougars in <strong>the</strong> East, Midwest, and Great Plains; discusses <strong>the</strong> <strong>of</strong>ficial status <strong>of</strong><br />

<strong>the</strong> species; and provides a perspective on <strong>the</strong> scientific, social, and political opportunities and challenges posed by<br />

this fascinating and compelling situation.<br />

71<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Key words: cougar, recovery, East, Midwest, Great Plains, prairie, North America, Puma concolor<br />

INTRODUCTION AND OBJECTIVES<br />

The possible existence <strong>of</strong> <strong>the</strong> puma<br />

(Puma concolor) in eastern and midwestern<br />

North America today, approximately 100<br />

years since its supposed extirpation from <strong>the</strong><br />

region, is among <strong>the</strong> most provocative and<br />

exciting mysteries in <strong>the</strong> modern realms <strong>of</strong><br />

natural history, ecology, wildlife<br />

management, and conservation biology.<br />

Importantly, <strong>the</strong> story <strong>of</strong> <strong>the</strong> cougar in <strong>the</strong><br />

East, <strong>the</strong> ghost <strong>of</strong> North America, as it was<br />

dubbed by Bruce Wright, an early champion<br />

for its recovery, has far-reaching, global<br />

implications for carnivore conservation,<br />

continental ecological equilibrium, and<br />

perhaps most <strong>of</strong> all, our own fulfillment as<br />

stewards <strong>of</strong> <strong>the</strong> planet (Wright 1959). To<br />

understand this yet unfolding story, several<br />

fundamental concepts need review:<br />

1. Throughout North America from <strong>the</strong><br />

Great Plains eastward, with <strong>the</strong><br />

exception <strong>of</strong> Florida, <strong>the</strong> puma was<br />

generally considered extirpated by <strong>the</strong><br />

early 1900s (Young and Goldman<br />

1946).<br />

2. Since that time, in virtually every state<br />

and every province across this vast<br />

region, scores <strong>of</strong> people, including<br />

pr<strong>of</strong>essional scientists, biologists,<br />

naturalists, and foresters, have been<br />

reporting observations <strong>of</strong> cougars or<br />

<strong>the</strong>ir sign (Wright 1972, Tischendorf<br />

and Henderson 1994).<br />

3. While many <strong>of</strong> <strong>the</strong>se reports are<br />

unverifiable or erroneous, a surprising<br />

number have been confirmed, <strong>of</strong>ten<br />

with <strong>the</strong> details published in peerreviewed<br />

literature. This history <strong>of</strong><br />

confirmed reports since <strong>the</strong> time <strong>of</strong><br />

supposed extirpation suggests, at a<br />

minimum, <strong>the</strong> periodic presence <strong>of</strong> freeranging<br />

cougars in <strong>the</strong> region.<br />

4. Several plausible explanations exist for<br />

<strong>the</strong>se cryptic cats: 1) continued<br />

existence <strong>of</strong> native pumas; 2)<br />

immigration <strong>of</strong> western cats; 3)<br />

presence <strong>of</strong> feral escaped or released<br />

captives (FERCs); or 4) combinations<br />

<strong>of</strong> any or all <strong>of</strong> <strong>the</strong>se (Nowak 1976,<br />

Downing 1984).


72 CRYPTIC COUGARS · Tischendorf<br />

From ecological, social, and political<br />

standpoints <strong>the</strong>re are three main questions<br />

that this paper seeks to answer. One, are<br />

<strong>the</strong>re cougars in <strong>the</strong> aforementioned region<br />

today? Two, if pumas are present, do <strong>the</strong>y<br />

represent a breeding population(s)? Finally,<br />

what is <strong>the</strong> future <strong>of</strong> Puma concolor east <strong>of</strong><br />

<strong>the</strong> Rocky <strong>Mountain</strong>s? What truly does <strong>the</strong><br />

public want when it comes to large<br />

carnivore recovery or restoration in <strong>the</strong><br />

East? Possibilities here include active<br />

recovery, passive recovery (i.e., <strong>the</strong> animals<br />

establish viable populations on <strong>the</strong>ir own<br />

without active, direct human intervention),<br />

or overt efforts to preclude recovery.<br />

DISCUSSION<br />

Puma Presence, Populations, and <strong>the</strong> Big<br />

Picture Perspective<br />

To effectively understand <strong>the</strong> cryptic<br />

cougar situation, it is critical to maintain a<br />

big picture perspective (Tischendorf 1992c,<br />

1996a, b). Among <strong>the</strong> many who have<br />

commented on <strong>the</strong> subject over <strong>the</strong> years,<br />

and especially among those skeptical <strong>of</strong><br />

cougar presence or recovery east <strong>of</strong> <strong>the</strong><br />

Rockies, this perspective, “from Nova<br />

Scotia to Nebraska” (Tischendorf 1996a:43),<br />

has <strong>of</strong>ten been lacking (Tischendorf 1992c;<br />

1996a, b). Such a perspective was,<br />

however, utilized by Bruce Wright and,<br />

more recently, by United States Fish and<br />

Wildlife Service (USFWS) researcher<br />

Robert Downing. Downing authored <strong>the</strong><br />

eastern cougar recovery plan and speculated<br />

on <strong>the</strong> presence <strong>of</strong> an extremely low density,<br />

widely dispersed puma population in <strong>the</strong><br />

eastern United States (USA) (Downing<br />

1981, 1984; United States Fish and Wildlife<br />

Service 1982).<br />

Downing’s views, coupled with updated<br />

range information syn<strong>the</strong>sized by Allen<br />

Anderson and intensive independent review<br />

<strong>of</strong> 100 years’ worth <strong>of</strong> reports and<br />

documentation, led one author to<br />

subsequently suggest that <strong>the</strong>re were<br />

actually upwards <strong>of</strong> four loosely interrelated<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

puma populations in <strong>the</strong> East and Midwest<br />

(Downing 1981, 1984; Anderson 1983;<br />

Tischendorf 1993c). Each <strong>of</strong> <strong>the</strong>se lowdensity<br />

puma populations was believed to<br />

consist <strong>of</strong> widely dispersed, widely roaming,<br />

and perhaps transient animals (Tischendorf<br />

1993c). These populations were believed to<br />

have <strong>the</strong>ir epicenters in <strong>the</strong> Canadian<br />

Maritimes-New England region, <strong>the</strong> Great<br />

Lakes-nor<strong>the</strong>rn Midwest region, <strong>the</strong><br />

Missouri-Arkansas-Oklahoma area, and <strong>the</strong><br />

Sou<strong>the</strong>ast.<br />

This <strong>the</strong>ory <strong>of</strong> course remains unproven,<br />

although it was revisited at a previous<br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong> by several <strong>of</strong> this<br />

author’s colleagues similarly associated with<br />

<strong>the</strong> West Virginia-based Eastern Cougar<br />

Foundation (ECF) (Bolgiano et al. 2000).<br />

Members <strong>of</strong> <strong>the</strong> ECF, formed in 1999, are<br />

utilizing automatic cameras in an attempt to<br />

document consistent cougar presence or<br />

family groups that could support <strong>the</strong> above<br />

hypo<strong>the</strong>sis. The ECF (website at<br />

www.easterncougar.org) is notable in that it<br />

has been able to positively partner with<br />

several governmental agencies and share in<br />

<strong>the</strong> efforts to recover pumas in <strong>the</strong> East.<br />

Such critical cooperation is also<br />

demonstrated with <strong>the</strong> Eastern Cougar<br />

Network (ECN). This group’s website,<br />

www.easterncougarnet.org, is a nonadvocacy<br />

amalgamation <strong>of</strong> peer-reviewed<br />

contributions on <strong>the</strong> subject <strong>of</strong> cougars from<br />

essentially every state and provincial<br />

resource agency from <strong>the</strong> Great Plains<br />

eastward. The site thus serves effectively as<br />

a real-time source <strong>of</strong> scientifically based<br />

status information on <strong>the</strong> cat, and perhaps<br />

one day o<strong>the</strong>r predators including gray<br />

wolves (Canis lupus), black bears (Ursus<br />

americana), lynx (Lynx canadensis), and<br />

wolverines (Gulo gulo) east <strong>of</strong> <strong>the</strong> Rockies.<br />

Seemingly integral to <strong>the</strong> subject <strong>of</strong><br />

cougars east <strong>of</strong> <strong>the</strong> Rockies is <strong>the</strong> question<br />

<strong>of</strong> whe<strong>the</strong>r <strong>the</strong> species persisted in its native<br />

state beyond <strong>the</strong> days <strong>of</strong> its supposed


extirpation. In <strong>the</strong> big picture, however, if<br />

free-ranging pumas are present and<br />

behaving in wild puma ways, <strong>the</strong>n <strong>the</strong>ir<br />

origin, whe<strong>the</strong>r from native eastern or<br />

western stock or sanctioned or unsanctioned<br />

releases, should not alter <strong>the</strong>ir proper<br />

management and may be irrelevant. While<br />

<strong>the</strong> cats in many confirmed puma reports are<br />

written <strong>of</strong>f as FERCs and denied<br />

consideration as legitimate ecological<br />

entities, <strong>the</strong> North American continent teems<br />

with a host <strong>of</strong> wildlife populations having<br />

<strong>the</strong>ir origins in captivity. These span <strong>the</strong><br />

spectrum from critically endangered species<br />

to non-native exotics raised like barnyard<br />

fowl and annually introduced solely for<br />

sporting opportunities. Yet <strong>the</strong>se former<br />

captives continue to benefit from <strong>of</strong>ficial<br />

recognition, management, and protection.<br />

Should mountain lions that happen to show<br />

up in areas where <strong>the</strong>ir presence is<br />

considered improbable be any different?<br />

Having said this, <strong>the</strong> historically<br />

consistent pattern <strong>of</strong> sightings and periodic<br />

confirmations, while circumstantial,<br />

suggests native pumas did persist in many<br />

areas <strong>of</strong> <strong>the</strong>ir former midwestern and eastern<br />

range at least into <strong>the</strong> 1940s and 1950s.<br />

After World War II, however, ownership <strong>of</strong><br />

cougars and o<strong>the</strong>r wild, exotic, or novelty<br />

animals became part <strong>of</strong> mainstream<br />

Americana and some captive cougars likely<br />

ended up as FERCs. Unfortunately this<br />

phenomenon continues today and is not<br />

necessarily limited to <strong>the</strong> eastern USA. As a<br />

result, <strong>the</strong> ultimate origin <strong>of</strong> almost any freeranging<br />

puma today, even with genetic<br />

testing, may truly be indeterminate.<br />

Summary <strong>of</strong> Occurrence Records<br />

In keeping with a big picture perspective<br />

<strong>of</strong> <strong>the</strong> cryptic cougar subject, it is useful to<br />

review a sampling <strong>of</strong> bonafide puma reports.<br />

Examples <strong>of</strong> confirmed or highly credible<br />

reports, mostly peer-reviewed, follow.<br />

“Confirmed kill” indicates that a puma was<br />

killed and <strong>the</strong> incident documented both<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 73<br />

photographically and by written or verbal<br />

elaboration <strong>of</strong> substantial details, or without<br />

photos but with written or verbal elaboration<br />

<strong>of</strong> substantial details by a pr<strong>of</strong>essional<br />

scientist or wildlife manager associated with<br />

or employed by a governmental natural<br />

resource agency or academic institution.<br />

“Reported kill” involves highly credible<br />

documentation by a natural resource<br />

pr<strong>of</strong>essional <strong>of</strong> a mountain lion being killed,<br />

but reflects a lack <strong>of</strong> substantial details.<br />

“Confirmed tracks” indicates that a track or<br />

tracks consistent with those <strong>of</strong> a puma were<br />

located and documented via measurements<br />

and/or photographs subsequently published<br />

in mainstream scientific or popular literature<br />

and thus widely available for independent<br />

scrutiny and au<strong>the</strong>ntication.<br />

Reported kill - Williston, North Dakota,<br />

1902 (Bailey 1926)<br />

Reported kill - Bears Paw <strong>Mountain</strong>s,<br />

Montana, 1910 (White 1967)<br />

Reported kill - Fontana Village area,<br />

Tennessee, 1920 (Linzey and Linzey<br />

1971)<br />

Confirmed kill - Mundleville, New<br />

Brunswick, 1932 (Wright 1972)<br />

Confirmed kill - Little Saint John Lake,<br />

Maine-Quebec border, 1938 (Wright<br />

1972)<br />

Confirmed kill - Madison, Saskatchewan,<br />

1939 (Clarke 1942)<br />

Confirmed kill - Pasquia Hills,<br />

Saskatchewan, 1948 (White 1963)<br />

Confirmed kill - Asheville, Alabama, 1948<br />

(Anonymous 1948)<br />

Confirmed kill - Mena, Arkansas, 1948<br />

(Lewis 1969, Nowak 1976)<br />

Confirmed kill - Sims, Arkansas, 1949<br />

(Sealander 1951)<br />

Confirmed kill - Black Hills, South Dakota,<br />

1958 (Mann 1959)<br />

Reported kills (2) - Newcastle, Wyoming,<br />

ca 1950s-1960s (Roop 1971)<br />

Reported kills (2) - Van Tassell, Wyoming,<br />

ca 1959-1960 (Roop 1971)


74 CRYPTIC COUGARS · Tischendorf<br />

Confirmed kill - Keithville, Louisiana,<br />

1965 (Goertz and Abegg 1966)<br />

Confirmed kill - Edinboro, Pennsylvania,<br />

1967 (Doutt 1969)<br />

Confirmed carcass - Checotah, Oklahoma,<br />

1968 (Lewis 1969)<br />

Confirmed kill - Hamburg, Arkansas, 1969<br />

(Noble 1971)<br />

Reported kill - Ekalaka, Montana, ca 1970<br />

(Nowak 1976)<br />

Confirmed kill - Pikeville, Tennessee, 1971<br />

(Nowak 1976)<br />

Confirmed kill - Stead, Manitoba, 1973<br />

(Nero and Wrigley 1977)<br />

Confirmed kill - Cutknife, Saskatchewan,<br />

1975 (White 1976)<br />

Cougar reportedly trapped - Baca County,<br />

Colorado, 1976 (Boddicker 1980)<br />

Confirmed hematological evidence -<br />

Menominee County, Michigan, 1984<br />

(Bill Adrian, Colorado Division <strong>of</strong><br />

Wildlife, personal communication)<br />

Puma trapped, radio-collared, translocated<br />

to Black Hills - central South Dakota,<br />

1990 or 1992 (Ted Benzon and Ron<br />

Sieg, South Dakota Department <strong>of</strong><br />

Game, Fish and Parks, personal<br />

communication; Tischendorf and<br />

Henderson 1994) (Note: This cat was<br />

killed in <strong>the</strong> Black Hills in 1996 [Ron<br />

Sieg, South Dakota Department <strong>of</strong><br />

Game, Fish and Parks, personal<br />

communication])<br />

Confirmed kill - Golden Valley County,<br />

North Dakota, 1991 (Tischendorf and<br />

Henderson 1994)<br />

Confirmed kill - Pine Ridge area, Nebraska,<br />

1991 (Tischendorf 1992a, Tischendorf<br />

and Henderson 1994)<br />

Cougar trapped and translocated to<br />

Colorado - Worthington, Minnesota,<br />

1991 (Tischendorf 1992a, b)<br />

Confirmed kill - Lowery, South Dakota,<br />

1992 (Tischendorf and Henderson<br />

1994)<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Confirmed kill - Lake Abitibi, Quebec,<br />

1992 (Tischendorf 1993a)<br />

Confirmed tracks - McKiel Lake, New<br />

Brunswick, 1992 (Tischendorf 1993b,<br />

Cumberland and Dempsey 1994)<br />

Confirmed kill - Texas County, Missouri,<br />

1994 (Hardin 1996, Bolgiano et al.<br />

2000)<br />

Confirmed kill - Mitchell, Nebraska, 1996<br />

(Frank Andelt, Nebraska Game and<br />

Parks Commission, personal<br />

communication)<br />

Confirmed kill - Floyd County, Kentucky,<br />

1997 (Bolgiano 2001)<br />

Confirmed kill - Randolph County, Illinois,<br />

2000 (Clark et al 2002)<br />

Confirmed kill - Duluth, Minnesota, 2001<br />

(Anonymous 2002)<br />

Confirmed kill - Harlan, Iowa, 2001<br />

(Anonymous 2002, Clark et al 2002)<br />

Confirmed kill - Callaway County,<br />

Missouri, 2003 (Graham 2003)<br />

Almost 30 years ago Nowak (1976:143-<br />

144) commented, “The sum <strong>of</strong> evidence<br />

suggests that native cougar populations have<br />

maintained <strong>the</strong>mselves in sou<strong>the</strong>astern<br />

Canada, within <strong>the</strong> former range <strong>of</strong> F. c.<br />

cougar (sic: should be couguar), and in <strong>the</strong><br />

Ozark Plateau and adjoining forests <strong>of</strong><br />

Arkansas, sou<strong>the</strong>rn Missouri, eastern<br />

Oklahoma, and nor<strong>the</strong>rn Louisiana.”<br />

Indeed, even if ecologically significant<br />

populations did not persist, <strong>the</strong> above list<br />

suggests it is doubtful that <strong>the</strong>se furtive<br />

felids were ever totally extirpated from <strong>the</strong><br />

vast, and in many cases relatively<br />

inaccessible, environs <strong>of</strong> this area.<br />

Relatively pristine areas within New<br />

Brunswick, Quebec, Ontario, and Manitoba,<br />

for instance, could possibly have sustained<br />

individual pumas or even vestigial, remnant<br />

populations <strong>of</strong> <strong>the</strong>se cats through <strong>the</strong> “Dark<br />

Age” <strong>of</strong> wildlife and habitat management<br />

late in <strong>the</strong> late 19 th and early 20 th centuries.<br />

In <strong>the</strong> USA, a number <strong>of</strong> areas could<br />

also have served as similar refugia. As late


as <strong>the</strong> mid-1940s, for instance, noted<br />

mammalogist Victor Cahalane<br />

acknowledged cougar presence in<br />

Shenandoah National Park and adjacent<br />

Blue Ridge country <strong>of</strong> Virginia (Cahalane<br />

1948). O<strong>the</strong>r plausible refuges include<br />

nor<strong>the</strong>rn Maine, <strong>the</strong> Adirondacks, <strong>the</strong><br />

Quabbin Reservoir area in Massachusetts,<br />

nor<strong>the</strong>rn Pennsylvania, <strong>the</strong> impenetrable<br />

sou<strong>the</strong>astern and sou<strong>the</strong>rn coastal swamps,<br />

<strong>the</strong> Great Smoky <strong>Mountain</strong>s area, <strong>the</strong><br />

Tennessee-Virginia-Kentucky-West<br />

Virginia border country, <strong>the</strong> dense<br />

northwoods <strong>of</strong> Michigan and Minnesota, <strong>the</strong><br />

Ozark and Ouachita <strong>Mountain</strong> complex <strong>of</strong><br />

Missouri, Arkansas, and Oklahoma, and <strong>the</strong><br />

Black Hills <strong>of</strong> South Dakota and Wyoming.<br />

White-tailed deer (Odocoileus<br />

virginianus) and mule deer (Odocoileus<br />

hemionus) irruptions in many <strong>of</strong> <strong>the</strong>se same<br />

areas were identified circa 1940, confirming<br />

that within only a few generations <strong>of</strong> <strong>the</strong><br />

puma’s supposed demise <strong>the</strong>se sites had an<br />

adequate prey base to sustain or attract <strong>the</strong><br />

species (Leopold et al. 1947). While an<br />

alternative argument is <strong>of</strong> course that <strong>the</strong><br />

irruptions resulted from lack <strong>of</strong> predators,<br />

irruptions were also noted in <strong>the</strong> Rocky<br />

<strong>Mountain</strong>s, where historically we know<br />

mountain lion populations may have been<br />

depleted but were never decimated.<br />

Indeed, in seminal biological<br />

publications as late as <strong>the</strong> mid-1970s and<br />

1980s several noted mammalogists and<br />

wildlife scientists postulated <strong>the</strong> continued<br />

presence <strong>of</strong> puma populations in <strong>the</strong>se very<br />

same areas (Cahalane 1964, Burt and<br />

Grossenheider 1976, Nowak 1976, Deems<br />

and Pursley 1978, Russell 1978, Hall 1981,<br />

Anderson 1983).<br />

None<strong>the</strong>less, such reports seemingly<br />

generated merely passing interest from<br />

mainstream science and remained largely <strong>of</strong>f<br />

<strong>the</strong> radar screen <strong>of</strong> wildlife biologists.<br />

Today, with <strong>the</strong> growing groundswell in<br />

conservation biology and corridor ecology,<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 75<br />

scientists are more inclined to recognize <strong>the</strong><br />

importance and implications <strong>of</strong> <strong>the</strong> presence<br />

<strong>of</strong> small, insular predator populations or<br />

even remnant, wandering, or dispersing<br />

individuals. Indeed, such enclaves or<br />

individuals may provide <strong>the</strong> necessary seed<br />

for recolonization or recovery.<br />

Natural resource agencies responsible<br />

for some <strong>of</strong> <strong>the</strong>se above-mentioned areas<br />

today, such as <strong>the</strong> Black Hills, have in fact<br />

confirmed extant puma populations<br />

(Tischendorf and Henderson 1994).<br />

Additionally, for quite some time o<strong>the</strong>r<br />

states such as North Dakota, Minnesota,<br />

Wisconsin, Missouri, and Arkansas have<br />

acknowledged at least limited and sporadic<br />

puma presence (Sealander and Gipson 1973,<br />

Gerson 1988, Tischendorf and Henderson<br />

1994, Clark et al. 2002, Graham 2003,<br />

Heisel 2003). Michigan Department <strong>of</strong><br />

Natural Resources <strong>of</strong>ficials acknowledge <strong>the</strong><br />

presence <strong>of</strong> pumas in <strong>the</strong> Great Lakes State<br />

as well, and ongoing work <strong>the</strong>re at least<br />

suggests <strong>the</strong> possibility <strong>of</strong> a resident,<br />

breeding population (Johnson 2002,<br />

Zuidema 2002, Mike Zuidema, Michigan<br />

Department <strong>of</strong> Forestry, retired, personal<br />

communication). Given <strong>the</strong> habitat, cover,<br />

prey base, and presence <strong>of</strong> a thriving<br />

carnivore guild that includes populations <strong>of</strong><br />

wolves, bears, coyotes (Canis latrans),<br />

bobcats (Felis rufus), fishers (Martes<br />

pennanti), and probably an occasional lynx,<br />

it would perhaps be more surprising if<br />

Michigan did not have a resident puma<br />

population.<br />

The situation in <strong>the</strong> Prairie Provinces <strong>of</strong><br />

Canada, with <strong>the</strong>ir seemingly less<br />

sensational and less controversial approach<br />

to <strong>the</strong> cat, is similar, if not even more<br />

definitive. In Saskatchewan, <strong>the</strong> late Tom<br />

White documented <strong>the</strong> presence <strong>of</strong> a small<br />

population <strong>of</strong> pumas filtering among <strong>the</strong><br />

coulees and more rugged reaches <strong>of</strong> this<br />

sprawling province (White 1982). The<br />

Yukon and Northwest Territories, and


76 CRYPTIC COUGARS · Tischendorf<br />

Alaska as well, have a consistent history <strong>of</strong><br />

credible puma reports, suggesting occasional<br />

dispersal, while Manitoba Conservation<br />

continues to recognize a stable and perhaps<br />

growing puma population in that province<br />

(Cahalane 1964; Weddle 1965; Kuyt 1971;<br />

White 1982; Wrigley and Nero 1982; Robert<br />

W. Nero, Manitoba Museum <strong>of</strong> Man and<br />

Nature, retired, personal communication).<br />

Cougar Comeback<br />

Some researchers believe that pumas, as<br />

wolves did in <strong>the</strong> nor<strong>the</strong>rn Rocky <strong>Mountain</strong>s<br />

in <strong>the</strong> 1980s, are in fact re-colonizing many<br />

areas in <strong>the</strong> Great Plains and central<br />

mountains eastward (Tischendorf and<br />

Henderson 1994). As is true for <strong>the</strong> Dakotas<br />

and Minnesota, most <strong>of</strong> <strong>the</strong> prairie states,<br />

including Nebraska, Kansas, Oklahoma, and<br />

Iowa acknowledge, if not resident <strong>the</strong>n<br />

transient occurrences <strong>of</strong> pumas (Tischendorf<br />

and Henderson 1994; Johnson 1998, 2000).<br />

The same is true for <strong>the</strong> eastern portions <strong>of</strong><br />

Montana, Wyoming, Colorado, and Texas,<br />

where, in some cases, sporadic puma<br />

presence has been noted for years but where<br />

documented occurrences <strong>of</strong> <strong>the</strong>se “prairie<br />

pan<strong>the</strong>rs” are clearly on <strong>the</strong> increase<br />

(Boddicker 1980; Berg et al. 1983; Johnson<br />

1998, 2000; Riley 1991; Roop 1971; Russ<br />

1997).<br />

Deer-rich riparian zones along river<br />

systems such as <strong>the</strong> Yellowstone, Missouri,<br />

North and South Platte, Arkansas, Canadian,<br />

Red, and Colorado River in Texas, can<br />

undoubtedly serve as effective corridors for<br />

puma immigration all <strong>the</strong> way to <strong>the</strong><br />

sou<strong>the</strong>astern Texas coast, Mississippi River,<br />

and beyond. Additionally, <strong>the</strong><br />

documentation <strong>of</strong> puma deaths along<br />

railroad tracks in Nebraska and Illinois<br />

suggests <strong>the</strong> possibility that railroad right<strong>of</strong>-ways<br />

and associated brush belts may also<br />

be effective pathways for pumas (Frank<br />

Andelt, Nebraska Game and Parks<br />

Commission, personal communication;<br />

Clark et al. 2002).<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

The same pattern <strong>of</strong> puma recolonization<br />

discussed above could be<br />

occurring from <strong>the</strong> mid-continent’s nor<strong>the</strong>rn<br />

reaches south and eastward. For instance,<br />

Manitoba’s puma population may be linked<br />

with Ontario to <strong>the</strong> east and nor<strong>the</strong>rn<br />

Minnesota, Wisconsin, and Michigan to <strong>the</strong><br />

south. Conversely, if low numbers <strong>of</strong> pumas<br />

have in fact inhabited some <strong>of</strong> <strong>the</strong>se areas all<br />

along, <strong>the</strong>ir acknowledged presence today<br />

may be a function <strong>of</strong> both immigration and<br />

numerical local growth.<br />

In <strong>the</strong> Nor<strong>the</strong>ast, a similar phenomenon<br />

<strong>of</strong> range reestablishment may be taking<br />

place. This sentiment was first voiced by<br />

Canadian biologist Bruce Wright, famed<br />

World War II frogman-commando, Leopold<br />

student, early champion for <strong>the</strong> eastern<br />

cougar, and a strong advocate for eastern<br />

carnivore recovery (Wright 1959, 1972;<br />

Tischendorf 1996a; Allardyce 2001). It was<br />

Wright’s belief that throughout European<br />

man’s settlement <strong>of</strong> <strong>the</strong> region pan<strong>the</strong>rs<br />

persisted in <strong>the</strong> central highlands <strong>of</strong> New<br />

Brunswick and by <strong>the</strong> mid-1900s were, like<br />

<strong>the</strong> spokes <strong>of</strong> a wheel, re-populating and<br />

reclaiming <strong>the</strong>ir former range in <strong>the</strong> East.<br />

This belief, while perennially difficult to<br />

reconcile with <strong>the</strong> lack <strong>of</strong> confirmed puma<br />

populations in New Brunswick, or anywhere<br />

else in <strong>the</strong> East outside <strong>of</strong> Florida, is<br />

exemplified by growing numbers <strong>of</strong> not<br />

simply puma reports, but <strong>of</strong> highly credible<br />

or even verified puma reports (Gerson 1988,<br />

Cumberland and Dempsey 1994, Snow<br />

1994, Stocek 1995, Bolgiano et al. 2000).<br />

These include specimens, scats, tracks, and<br />

videotapes depicting <strong>the</strong>se cats across a wide<br />

geographical zone extending essentially<br />

from Ontario to Newfoundland and<br />

southward to Georgia.<br />

In <strong>the</strong> nor<strong>the</strong>ast USA, Maine and New<br />

York are perhaps <strong>the</strong> most promising in<br />

terms <strong>of</strong> numbers <strong>of</strong> credible puma reports.<br />

One ra<strong>the</strong>r compelling report from Maine<br />

involved a shaken hunter who, at extremely


close range, came upon <strong>the</strong> gripping scene<br />

<strong>of</strong> what he described as a puma mortally<br />

ravaging a bobcat (Tischendorf 1994a). A<br />

sampling <strong>of</strong> o<strong>the</strong>r data from Maine on file at<br />

<strong>the</strong> American Ecological Research Institute<br />

(AERIE) includes a hair sample<br />

confirmation from 1995, a track photograph<br />

from <strong>the</strong> mid-1990s, and a credible 2000<br />

report <strong>of</strong> what was thought to be a female<br />

puma and kitten. This author has also seen<br />

puma track photos taken by biologist George<br />

Matula <strong>of</strong> <strong>the</strong> Maine Department <strong>of</strong> Inland<br />

Fisheries and Wildlife at a deer yard during<br />

routine winter surveys circa 1984 or 1985.<br />

Credible New York puma reports on file<br />

with AERIE include <strong>the</strong> killing <strong>of</strong> a puma<br />

kitten by a bobcat hunter in <strong>the</strong> early 1990s<br />

and three believable sightings <strong>of</strong> individual<br />

pumas by pr<strong>of</strong>essional natural resource<br />

workers. All <strong>of</strong> <strong>the</strong>se events stem from <strong>the</strong><br />

vast Adirondack Park area and occurred<br />

during <strong>the</strong> 1990s.<br />

O<strong>the</strong>r areas <strong>of</strong> <strong>the</strong> Nor<strong>the</strong>ast are not<br />

without <strong>the</strong>ir own intriguing data. In 1994,<br />

for instance, a group <strong>of</strong> 3 pumas was<br />

observed and tracked near <strong>the</strong> community <strong>of</strong><br />

Craftsbury, Vermont (Theodore Reed,<br />

Friends <strong>of</strong> <strong>the</strong> Eastern Pan<strong>the</strong>r, personal<br />

communication). Presumably an adult<br />

female with 2 kittens, a scat deposited by <strong>the</strong><br />

group was collected; subsequent analysis<br />

confirmed presence <strong>of</strong> puma hairs (Bonnie<br />

Yates, USFWS Wildlife Forensic<br />

Laboratory, personal communication). A<br />

hair sample from <strong>the</strong> remote and<br />

untrammeled Gaspe’ Peninsula in nor<strong>the</strong>rn<br />

Quebec was also recently confirmed as that<br />

<strong>of</strong> a puma by Marc Gauthier and his<br />

Canadian research team (Mark Dowling,<br />

Eastern Cougar Network, personal<br />

communication).<br />

Evidence <strong>of</strong> Breeding and Validity <strong>of</strong><br />

Sighting Reports<br />

The questions <strong>of</strong> confirmed puma<br />

breeding and actual puma numbers are<br />

problematic. In <strong>the</strong> absence <strong>of</strong> systematic,<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 77<br />

scientifically objective, replicable, and<br />

typically expensive multi-year studies, such<br />

as those involving mark and recapture<br />

techniques or radio telemetry, it is difficult<br />

to extrapolate population-level<br />

characteristics <strong>of</strong> any animal. And, as<br />

several speakers at this conference have<br />

noted, even with robust, million dollar<br />

studies, it is difficult to quantify puma<br />

populations. Even more difficult and more<br />

expensive is monitoring a puma population<br />

over substantial periods <strong>of</strong> time. What does<br />

this bode for eastern and midwestern<br />

resource agencies trying to decode <strong>the</strong> issue<br />

<strong>of</strong> cryptic cats that many seem to report but<br />

few can verify?<br />

Complicating <strong>the</strong> issue is <strong>the</strong> fact that<br />

agencies and <strong>the</strong>ir human constituency east<br />

<strong>of</strong> <strong>the</strong> Rocky <strong>Mountain</strong>s have limited<br />

exposure to large carnivores and <strong>the</strong>ir<br />

management. In this geographic area <strong>the</strong>re<br />

truly is a different mindset and comfort level<br />

towards research and management involving<br />

<strong>the</strong>se animals, especially those capable <strong>of</strong><br />

attacking and killing people. In <strong>the</strong> Black<br />

Hills <strong>of</strong> South Dakota, for instance, radiotracking<br />

<strong>of</strong> a young male puma in <strong>the</strong> early<br />

1990s was discontinued after only a short<br />

time due to concerns over liability if <strong>the</strong><br />

animal were implicated in damage to a<br />

human or to human property (Tischendorf<br />

and Henderson 1994).<br />

Again due to concerns over liability,<br />

Missouri <strong>of</strong>ficials are reluctant to approve<br />

any studies involving handling or marking<br />

<strong>of</strong> black bears, which are apparently<br />

repopulating <strong>the</strong> Show-Me-State (Lynn<br />

Robbins, Southwest Missouri State<br />

University, personal communication).<br />

Confounding <strong>the</strong> matter fur<strong>the</strong>r are <strong>the</strong><br />

controversial predatory and wide-ranging<br />

characteristics <strong>of</strong> <strong>the</strong> animal and local<br />

uncertainty regarding its actual status as an<br />

endangered species versus a FERC. Not<br />

surprisingly <strong>the</strong>n, in <strong>the</strong> case <strong>of</strong> <strong>the</strong> puma in<br />

<strong>the</strong> East, Midwest, or prairies where it is


78 CRYPTIC COUGARS · Tischendorf<br />

frequently perceived by natural resource<br />

agencies as a “species non grata”, few<br />

intensive prospective documentation efforts<br />

have ever been undertaken (Van Dyke1983,<br />

McGinnis 1988, Humphreys 1994). To this<br />

author’s knowledge, east <strong>of</strong> <strong>the</strong> Black Hills<br />

and north <strong>of</strong> Florida and sou<strong>the</strong>rn Georgia<br />

no free-ranging puma has ever been radioinstrumented<br />

or o<strong>the</strong>rwise marked and<br />

tracked.<br />

What we are left with in <strong>the</strong> case <strong>of</strong><br />

pumas east <strong>of</strong> <strong>the</strong> Rockies are largely<br />

sighting reports and compilations <strong>of</strong> sighting<br />

reports. Such data are <strong>of</strong>ten met with<br />

incredulity, yet historically <strong>the</strong> scientific<br />

literature, particularly that related to<br />

carnivores, is replete with papers involving<br />

nothing more than sighting reports. Articles<br />

by Berg et al. (1983), involving pumas, and<br />

Quinn (1995), who worked with urban<br />

coyotes, are but two <strong>of</strong> many peer-reviewed<br />

examples <strong>of</strong> which this author is aware.<br />

Where people are reporting unknown or<br />

unsuspected animals, it <strong>of</strong>ten seems one is<br />

eventually killed and populations are<br />

subsequently substantiated, vindicating<br />

those who originally reported sightings.<br />

Fur<strong>the</strong>rmore, <strong>the</strong> survival <strong>of</strong> rare animals<br />

<strong>of</strong>ten depends on timely and critical<br />

decision-making. If sighting reports are <strong>the</strong><br />

best with which a researcher has to work,<br />

and particularly if some level <strong>of</strong> scientific<br />

rigor can be applied to <strong>the</strong>ir interpretation,<br />

as demonstrated by Quinn’s coyote research<br />

involving sighting reports in Washington, D.<br />

C., <strong>the</strong>n it is unscientific and ill advised to<br />

carte blanche ignore such reports (Quinn<br />

1995).<br />

In <strong>the</strong> case <strong>of</strong> colonization or repopulation,<br />

by definition <strong>the</strong>re is a temporal<br />

continuum <strong>of</strong> occurrence. Early in <strong>the</strong><br />

process, <strong>the</strong> animals in question are few.<br />

Colonization, re-colonization, or repopulation<br />

ends, if successful, with a selfsustaining<br />

population. The puma<br />

phenomenon east <strong>of</strong> <strong>the</strong> Rocky <strong>Mountain</strong>s is<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

presumably somewhere along this<br />

continuum. Evidence presented at this<br />

conference suggests puma presence in some<br />

areas <strong>of</strong> <strong>the</strong> West, for instance east-central<br />

and eastern New Mexico, is on <strong>the</strong> same<br />

continuum (Rick Winslow, New Mexico<br />

Game and Fish Department, personal<br />

communication).<br />

The question <strong>of</strong> breeding is, <strong>of</strong> course,<br />

pivotal in <strong>the</strong> discussion <strong>of</strong> purported puma<br />

populations and presence. Here again,<br />

however, o<strong>the</strong>r than efforts to collect<br />

sighting reports, <strong>the</strong>re has been little formal,<br />

proactive modern research on <strong>the</strong> species in<br />

eastern North America so information on<br />

this topic is limited. None<strong>the</strong>less, some<br />

useful information can be derived from <strong>the</strong><br />

available data. In <strong>the</strong> 1970s or early 1980s,<br />

a string <strong>of</strong> credible eyewitness reports<br />

suggested <strong>the</strong> presence <strong>of</strong> an adult puma and<br />

kitten(s) along <strong>the</strong> Blue Ridge Parkway<br />

(Robert Downing, USFWS, retired, personal<br />

communication). A carnivore biologist<br />

claims to have observed a family group <strong>of</strong><br />

pumas in nor<strong>the</strong>rn Minnesota in <strong>the</strong> 1970s<br />

(Bill Berg, Minnesota Department <strong>of</strong><br />

Natural Resources, retired, personal<br />

communication). More recently in<br />

Minnesota, breeding was also implied in <strong>the</strong><br />

case <strong>of</strong> <strong>the</strong> female puma killed outside<br />

Duluth in 2001 (Anonymous 2002). In her<br />

company were two kittens, both later<br />

captured and placed into captivity (Mark<br />

Dowling, Eastern Cougar Network, personal<br />

communication). The cougar killed in Floyd<br />

County, Kentucky in 1997, cited earlier, had<br />

spotted pelage and was in <strong>the</strong> company <strong>of</strong> at<br />

least one o<strong>the</strong>r, larger cat, when it was<br />

struck by a car (Bolgiano 2001). A<br />

biological scientist observed a puma and<br />

several kittens in Missouri in <strong>the</strong> early<br />

1990s. This is one <strong>of</strong> several episodes,<br />

including <strong>the</strong> poaching <strong>of</strong> a cougar (Texas<br />

County, cited earlier) and <strong>the</strong> videotaping <strong>of</strong><br />

a puma at a deer kill that triggered a<br />

substantial increase in public and agency


awareness <strong>of</strong> <strong>the</strong> puma in <strong>the</strong> state (Lynn<br />

Robbins, Southwest Missouri State<br />

University, personal communication). As<br />

noted above, recent credible reports <strong>of</strong><br />

females with kittens have also originated<br />

from both Vermont and Maine. Many o<strong>the</strong>r<br />

instances <strong>of</strong> apparent puma breeding in <strong>the</strong><br />

East were discussed by Wright (1972). Such<br />

isolated incidents are certainly not<br />

unequivocal pro<strong>of</strong> <strong>of</strong> a puma population or<br />

breeding, but in <strong>the</strong> big picture <strong>the</strong>y do tend<br />

to support <strong>the</strong> belief that at least a few<br />

pumas are present and sporadic reproduction<br />

is occurring.<br />

Predator Parallels - Bobcat, Black Bear,<br />

Jaguar, and Coyote<br />

Many <strong>of</strong> those skeptical <strong>of</strong> puma<br />

presence in eastern North America cite <strong>the</strong><br />

vast suburbanization and urbanization <strong>of</strong> <strong>the</strong><br />

region as an effective limiting factor. Yet, if<br />

populations <strong>of</strong> <strong>the</strong> versatile puma can exist<br />

in human dense areas <strong>of</strong> California,<br />

Colorado, and Florida, <strong>the</strong> species could<br />

surely inhabit portions <strong>of</strong> <strong>the</strong> East and<br />

Midwest, especially given <strong>the</strong> high<br />

populations <strong>of</strong> deer, feral hogs, and o<strong>the</strong>r,<br />

mid-sized and smaller game found<br />

throughout <strong>the</strong> region. Fur<strong>the</strong>rmore, o<strong>the</strong>r<br />

eastern carnivores including bobcats, black<br />

bears, and wolves are apparently acclimating<br />

to evolving habitats and human presence and<br />

expanding <strong>the</strong>ir populations and/or ranges<br />

(Stoll 1996). Benchmarking with <strong>the</strong>se<br />

species supports <strong>the</strong> contention that <strong>the</strong> even<br />

more elastic puma can do <strong>the</strong> same. A<br />

similar comparison can be made with <strong>the</strong><br />

jaguar (Pan<strong>the</strong>ra onca) in <strong>the</strong> West, which,<br />

while noted in <strong>the</strong> region only rarely for<br />

decades, has been probing borderland<br />

Mexico-Arizona-New Mexico habitat with<br />

increasing frequency in recent years (Brown<br />

and Lopez-Gonzalez 2000).<br />

The coyote provides an additional case<br />

study in relation to <strong>the</strong> possible existence <strong>of</strong><br />

<strong>the</strong> puma east <strong>of</strong> <strong>the</strong> Rocky <strong>Mountain</strong>s<br />

(Tischendorf 1994b). This adaptable, mid-<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 79<br />

sized, quintessentially western carnivore<br />

arrived on <strong>the</strong> midwestern and eastern<br />

landscape as a veritable newcomer in<br />

approximately <strong>the</strong> 1960s and 70s (Gerry<br />

Parker, Canadian Wildlife Service, retired,<br />

personal communication). Similar to what<br />

transpires today with many puma reports,<br />

coyote presence was initially refuted or<br />

attributed to feral individuals or<br />

hybridization with dogs. In retrospect, <strong>the</strong>se<br />

assessments were not entirely correct. It is a<br />

convincing reflection <strong>of</strong> <strong>the</strong> regional habitat<br />

quality and prey base that so successfully<br />

has <strong>the</strong> species colonized <strong>the</strong> eastern half <strong>of</strong><br />

<strong>the</strong> continent it is today a thriving, legally<br />

trapped, hunted, and pursued game animal.<br />

If <strong>the</strong> puma is as adaptable as its history<br />

suggests, <strong>the</strong>n intuitively it is simply a<br />

matter <strong>of</strong> time before it follows <strong>the</strong> coyote’s<br />

lead.<br />

Biolegal Issues<br />

While current <strong>the</strong>ories support <strong>the</strong><br />

contention that North American pumas are<br />

genetically panmictic, <strong>the</strong> Endangered<br />

Species Act (ESA) specifically identifies<br />

only <strong>the</strong> subspecies P. c. couguar (<strong>the</strong><br />

supposed true “eastern puma”) and P. c.<br />

coryi (<strong>the</strong> “Florida pan<strong>the</strong>r” or more<br />

correctly “sou<strong>the</strong>rn puma” [Greenwell<br />

1996:18, 36]) as endangered (Florida<br />

Pan<strong>the</strong>r Interagency Committee 1993,<br />

Greenwell 1996, Culver et al 2000).<br />

Florida, with its mongrel mix <strong>of</strong> native,<br />

Texan, “Piper”, and illicitly released<br />

animals, has overcome this issue by working<br />

with <strong>the</strong> federal government to enact<br />

similarity <strong>of</strong> appearance protection laws for<br />

all <strong>of</strong> its pumas. Thus, Florida’s pumas,<br />

regardless <strong>of</strong> origin, now all fall under <strong>the</strong><br />

convenient, albeit taxonomically outdated<br />

umbrella moniker <strong>of</strong> “Florida pan<strong>the</strong>r”<br />

(Alvarez 1993).<br />

Elsewhere in <strong>the</strong> East and <strong>the</strong> Midwest<br />

<strong>the</strong>re exists much confusion and feline<br />

filibustering about what constitutes a puma<br />

meriting ESA protection versus one that can


80 CRYPTIC COUGARS · Tischendorf<br />

be considered a western wanderer or<br />

escaped or released captive (Tischendorf<br />

1994b, Cardoza and Langlois 2002).<br />

Clouding <strong>the</strong> issue is <strong>the</strong> fact that much <strong>of</strong><br />

<strong>the</strong> Midwest was considered original range<br />

<strong>of</strong> F. c. schorgeri, <strong>the</strong> supposed “Wisconsin<br />

puma” which technically fits nei<strong>the</strong>r into <strong>the</strong><br />

ESA nor <strong>the</strong> eastern cougar recovery plan <strong>of</strong><br />

1982, which in any case has never been<br />

implemented (Nowak 1976, Hall 1981,<br />

USFWS 1982).<br />

As Albert Einstein reportedly said,<br />

“Perfection <strong>of</strong> means and confusion <strong>of</strong> goals<br />

seem, in my opinion, to characterize our<br />

age.” More simply, sometimes <strong>the</strong> process<br />

can get in <strong>the</strong> way <strong>of</strong> <strong>the</strong> purpose. Clearly,<br />

<strong>the</strong> Endangered Species Act is meant to<br />

protect rare animals. Equally apparent, <strong>the</strong><br />

puma from <strong>the</strong> Great Plains eastward is rare.<br />

Should <strong>the</strong> ESA unequivocally and<br />

ultimately serve as a tool to protect this li<strong>the</strong><br />

and golden ghost as it reestablishes itself<br />

across its original range? Or can <strong>the</strong> case be<br />

made that federal delisting, in concert with<br />

state or multi-state management plans and<br />

agreements, unencumbered by federal<br />

oversight, may more favorably serve <strong>the</strong><br />

puma <strong>of</strong> <strong>the</strong> Great Plains, Midwest, and<br />

East?<br />

The Jaguar Conservation Team<br />

(JAGCT), a diverse coalition <strong>of</strong> agencies<br />

and individuals working toge<strong>the</strong>r to develop<br />

and implement a sound plan for protecting<br />

and conserving jaguars in <strong>the</strong> Southwest,<br />

may serve as a template organization for<br />

those involved with puma recovery east <strong>of</strong><br />

<strong>the</strong> Rocky <strong>Mountain</strong>s. Formed in 1997,<br />

prior to listing <strong>of</strong> <strong>the</strong> jaguar as an<br />

endangered species north <strong>of</strong> Mexico by <strong>the</strong><br />

USFWS, <strong>the</strong> JAGCT operates under a<br />

formal conservation agreement with <strong>the</strong><br />

USFWS and today functions as an ad hoc<br />

jaguar recovery team (Bill Van Pelt, Arizona<br />

Game and Fish Department, personal<br />

communication). In essence, <strong>the</strong> JAGCT<br />

attempts to preempt <strong>the</strong> potential for heavy-<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

handedness and long-distance directives <strong>of</strong><br />

<strong>the</strong> ESA by working locally with all its<br />

stakeholders to proactively find effective<br />

solutions to conflict (Bill Van Pelt, Arizona<br />

Game and Fish Department, personal<br />

communication).<br />

For now <strong>the</strong> future <strong>of</strong> eastern or<br />

midwestern pumas is largely tied to <strong>the</strong><br />

ESA. Certainly similarity <strong>of</strong> appearance<br />

semantics related to pumas and <strong>the</strong> ESA in<br />

Florida have symbiotically done much to<br />

pave <strong>the</strong> way for puma recovery elsewhere<br />

east <strong>of</strong> <strong>the</strong> Rockies. Still, if <strong>the</strong> ESA is to<br />

play a pivotal role, it clearly requires<br />

modification to recognize P. concolor in<br />

terms <strong>of</strong> individuals and populations,<br />

without regard to an obsolete concept <strong>of</strong><br />

subspecies, as <strong>the</strong> rare animal that east <strong>of</strong> <strong>the</strong><br />

Rocky <strong>Mountain</strong>s it truly is (Tischendorf<br />

1994c, 1995). Such modification would<br />

thus effectively resolve <strong>the</strong> unintentional but<br />

critical double standard for recovery that<br />

exists for pumas in Florida versus those<br />

elsewhere in <strong>the</strong> East and Midwest. Based<br />

on <strong>the</strong> powerful sou<strong>the</strong>rn precedent,<br />

similarity <strong>of</strong> appearance protection for all<br />

free-ranging non-nuisance pumas and<br />

potential puma populations today and<br />

tomorrow, not only in Florida, but those<br />

from <strong>the</strong> Great Plains eastward, would be a<br />

simple, logical, and consistent step<br />

(Tischendorf 1994c, 1995; Cardoza and<br />

Langlois 2002).<br />

In <strong>the</strong> nor<strong>the</strong>rn Rockies with <strong>the</strong> wolf,<br />

and in Florida with <strong>the</strong> puma, recovery has<br />

been facilitated by formal restoration efforts.<br />

It is questionable whe<strong>the</strong>r such<br />

governmentally sanctioned activities will be<br />

carried out to enhance natural cougar<br />

recovery in eastern or midwestern North<br />

America. Economic issues certainly exist,<br />

witness <strong>the</strong> budgets necessary for not only<br />

gray wolf and pan<strong>the</strong>r restoration, but those<br />

for o<strong>the</strong>r high-pr<strong>of</strong>ile endangered species<br />

like <strong>the</strong> red wolf (Canis rufus), black-footed<br />

ferret (Mustela nigripes), whooping crane


(Grus americana), California condor<br />

(Gymnogyps californianus), and peregrine<br />

falcon (Falco peregrinus).<br />

Additionally, given that <strong>the</strong> species in<br />

question is not just endangered but large,<br />

carnivorous, potentially hazardous to<br />

humans, and an effective ecological<br />

regulator <strong>of</strong> ungulates that figure<br />

prominently in a deeply entrenched hunting<br />

tradition, <strong>the</strong>re is certainly potential for<br />

controversy and conflict among various<br />

public constituencies.<br />

Still, <strong>the</strong>re are intriguing possibilities for<br />

<strong>the</strong> future <strong>of</strong> <strong>the</strong> puma in eastern North<br />

America. Even despite <strong>the</strong> limitations <strong>of</strong> <strong>the</strong><br />

ESA, given nothing more than appropriate<br />

deer management, can <strong>the</strong> once ubiquitous<br />

puma survive, re-populate, and thrive in <strong>the</strong><br />

East, Midwest, and Great Plains? Evidence<br />

presented in this paper tends to support this<br />

scenario.<br />

What truly are <strong>the</strong> public attitudes<br />

toward this widely ranging and exquisitely<br />

adaptable carnivore? Do agency attitudes<br />

mirror public sentiment? These critical<br />

human dimension wildlife topics would<br />

make excellent subject matter for a graduate<br />

student project.<br />

As pumas return to <strong>the</strong> Great Plains,<br />

Midwest, and East, <strong>the</strong>re will inevitably be<br />

conflict, as occurs in <strong>the</strong> West, with<br />

agricultural and suburban interests.<br />

Exemplifying this, uncannily, as this paper<br />

was being revised in October 2003, a freeranging<br />

young male puma, presumably<br />

dispersing from <strong>the</strong> west, was captured and<br />

placed into captivity after causing public<br />

unrest and alarm in urban Omaha, Nebraska.<br />

Meanwhile, in Iowa ano<strong>the</strong>r young male<br />

puma was shot and killed by a farmer.<br />

Can we as wildlife pr<strong>of</strong>essionals devise a<br />

new ESA or state-level paradigm for<br />

carnivore management, and specifically <strong>the</strong><br />

phenomenon <strong>of</strong> novel puma presence, that<br />

provides effective oversight for scientific,<br />

evidence-based decisions while allowing for<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 81<br />

sensible managerial flexibility and creativity<br />

at <strong>the</strong> key stakeholders’ state and local<br />

level? Does <strong>the</strong> JAGCT, with its formal and<br />

proactive conservation agreement with <strong>the</strong><br />

USFWS, provide a workable model for this?<br />

Are <strong>the</strong>re in fact several low density,<br />

highly mobile puma populations and<br />

breeding foci in eastern and mid-western<br />

North America? This question remains<br />

unresolved, but <strong>the</strong> growing body <strong>of</strong><br />

evidence discussed herein suggests that this<br />

possibility should not be discounted.<br />

Finally, despite <strong>the</strong> substantial evidence<br />

to <strong>the</strong> contrary, if pumas are in fact absent<br />

from <strong>the</strong> landscapes <strong>of</strong> <strong>the</strong> prairies,<br />

Midwest, and East, what is <strong>the</strong> prognosis for<br />

<strong>the</strong>ir human-aided restoration in seemingly<br />

viable ecosystems like <strong>the</strong> Adirondack<br />

<strong>Mountain</strong>s, <strong>the</strong> prey-rich Alleghenies, <strong>the</strong><br />

Ozark or Ouachita <strong>Mountain</strong>s, or <strong>the</strong><br />

expansive nor<strong>the</strong>rn forests <strong>of</strong> Minnesota,<br />

Michigan, and Maine?<br />

Many <strong>of</strong> <strong>the</strong>se and o<strong>the</strong>r questions were<br />

raised and addressed in detail at <strong>the</strong> historic<br />

“Eastern Cougar Conference, 1994” held in<br />

Erie, Pennsylvania in June 1994<br />

(Tischendorf and Ropski 1996). This was<br />

<strong>the</strong> first, and remains <strong>the</strong> only, formal<br />

conference ever devoted entirely to <strong>the</strong><br />

subject <strong>of</strong> pumas in eastern North America.<br />

A second conference is planned for<br />

Morgantown, West Virginia, in April 2004.<br />

This conference will roughly mark <strong>the</strong> tenth<br />

anniversary <strong>of</strong> <strong>the</strong> first ga<strong>the</strong>ring, providing<br />

a centralized forum for ongoing discussions<br />

and updates on research related to this<br />

intriguing subject.<br />

CONCLUSION<br />

In conclusion, 4 key points:<br />

1. Adaptable animals adapt and <strong>the</strong><br />

puma is one <strong>of</strong> <strong>the</strong> most adaptable<br />

land mammals in <strong>the</strong> New World.<br />

2. There are no ecological reasons why<br />

<strong>the</strong> puma could not exist in eastern


82 CRYPTIC COUGARS · Tischendorf<br />

North America today. Across much<br />

<strong>of</strong> <strong>the</strong> Great Plains, midwestern, and<br />

eastern portions <strong>of</strong> <strong>the</strong> continent <strong>the</strong><br />

evidence suggests that in low<br />

densities it does.<br />

3. Using <strong>the</strong> model <strong>of</strong> <strong>the</strong> JAGCT, a<br />

diverse but integrated Conservation<br />

Team should be formed as soon as<br />

possible to promulgate appropriate<br />

changes to <strong>the</strong> ESA (including even<br />

possible delisting) and to critically<br />

evaluate <strong>the</strong> long dormant recovery<br />

plan for cougars in <strong>the</strong> East, which<br />

requires updating to reflect recent<br />

knowledge related to <strong>the</strong> puma east <strong>of</strong><br />

<strong>the</strong> Rocky <strong>Mountain</strong>s (Tischendorf<br />

1996b, Cardoza and Langlois 2002).<br />

4. As mankind enters this new<br />

millennium, return <strong>of</strong> <strong>the</strong> puma to its<br />

former range in <strong>the</strong> Great Plains,<br />

Midwest, and East provides an<br />

opportunity for wildlife pr<strong>of</strong>essionals<br />

with limited firsthand experience with<br />

large carnivores to demonstrate <strong>the</strong>ir<br />

expertise in scientifically and<br />

sociologically managing this<br />

relatively secretive predator on <strong>the</strong><br />

many-faceted modern ecological<br />

interface <strong>of</strong> private and public lands,<br />

politics, and public opinion.<br />

In today’s anthropocentric world, <strong>the</strong><br />

puma, as is <strong>the</strong> case for large carnivores<br />

everywhere, is unfortunately a victim <strong>of</strong> its<br />

own three “E’s” - its evolution, its ecology,<br />

and its ethology. Widely ranging, oblivious<br />

to human-delineated boundaries, a large and<br />

potentially dangerous predator that preys<br />

effectively and efficiently on ungulates both<br />

wild and domestic, <strong>the</strong> puma is an<br />

irrefutable, anachronistic, and controversial<br />

symbol <strong>of</strong> our primeval wild.<br />

In <strong>the</strong> end, <strong>the</strong> message that emerges<br />

today for <strong>the</strong> puma in <strong>the</strong> East is this: even<br />

at its highest densities, few people will ever<br />

see a living, wild puma. Certainly we can<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

live without this great cat. Even more<br />

certainly, it can live without us. Enmeshed<br />

in controversy, entrenched in folklore,<br />

history, and legend, endangered across a<br />

huge portion <strong>of</strong> its historic range, <strong>the</strong> largely<br />

secretive puma presents us with <strong>the</strong> great<br />

challenge <strong>of</strong>, and <strong>the</strong> magnificent<br />

opportunity for, harmonious coexistence.<br />

Hopefully mankind will rise to this rare<br />

occasion to ensure that <strong>the</strong> puma is again a<br />

celebrated and wisely managed part <strong>of</strong> our<br />

Great Plains, midwestern, and eastern<br />

wildlife heritage.<br />

ACKNOWLEDGEMENTS<br />

This paper is dedicated to <strong>the</strong> memory <strong>of</strong><br />

Frank C. Craighead, Jr., <strong>of</strong> Moose,<br />

Wyoming, whose life and career were an<br />

inspiration to a generation <strong>of</strong> wildlife<br />

biologists. Kerry Murphy and Randy<br />

Matchett served knowledgeably and capably<br />

as manuscript referees. A special thank you<br />

is extended to <strong>the</strong> Wyoming Game and Fish<br />

Department, coordinator and host for this<br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>.<br />

LITERATURE CITED<br />

ALLARDYCE, G. 2001. On <strong>the</strong> track <strong>of</strong> <strong>the</strong><br />

New Brunswick pan<strong>the</strong>r - <strong>the</strong> story <strong>of</strong><br />

Bruce Wright and <strong>the</strong> eastern pan<strong>the</strong>r.<br />

Privately published, Fredericton, New<br />

Brunswick. 145 pages.<br />

ALVAREZ, K. 1993. Twilight <strong>of</strong> <strong>the</strong> pan<strong>the</strong>r<br />

- biology, bureaucracy, and failure in an<br />

endangered species program. Myakka<br />

River Publishing, Sarasota, FL. 501<br />

pages.<br />

ANDERSON, A. E. 1983. A critical review <strong>of</strong><br />

literature on puma (Felis concolor).<br />

Colorado Division <strong>of</strong> Wildlife, Denver.<br />

91 pages.<br />

ANONYMOUS. 1948. <strong>Mountain</strong> lion killed<br />

in St. Clair County. Alabama<br />

Conservationist. April issue. Page 11.<br />

ANONYMOUS. 2002. Cougar killed in<br />

Minnesota. Page 4 in Eastern Cougar


Foundation Newsletter, winter issue. 6<br />

pages.<br />

BAILEY. V. 1926. A biological survey <strong>of</strong><br />

North Dakota. North American Fauna<br />

#49. Washington D. C. Pages 144-146.<br />

BERG, R. L., L. L. MCDONALD, AND M. D.<br />

STRICKLAND. 1983. Distribution <strong>of</strong><br />

mountain lions in Wyoming as<br />

determined by mail questionnaire.<br />

Wildlife Society Bulletin 11(3):265-268.<br />

BODDICKER, M. L. 1980. <strong>Mountain</strong> lion.<br />

Prevention and control <strong>of</strong> wildlife<br />

damage. Great Plains Agricultural<br />

Council, Cooperative Extension Service,<br />

Kansas State University, Manhattan. 6<br />

pages.<br />

BOLGIANO, C., T. LESTER, AND D. MAEHR.<br />

2000. Field evidence <strong>of</strong> cougars in<br />

eastern North America. In L. A.<br />

Harveson, editor, <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong><br />

Sixth <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, San<br />

Antonio, TX. In press.<br />

BOLGIANO, C. 2001. Important<br />

confirmation: Spotted cougar kitten<br />

killed in Kentucky. Pages 1-2 in Eastern<br />

Cougar Foundation Newsletter, summer<br />

issue. 6 pages.<br />

BROWN, D. E. AND C. A. LOPEZ-GONZALEZ.<br />

2000. Search for el tigre. Defenders<br />

75(2):8-13.<br />

BURT, W. H. AND R. P. GROSSENHEIDER.<br />

1976. A field guide to <strong>the</strong> mammals. 3 rd<br />

edition. Houghton Mifflin Company,<br />

Boston. 289 pages.<br />

CAHALANE, V. H. 1948. The status <strong>of</strong><br />

mammals in <strong>the</strong> U. S. National Park<br />

system, 1947. Journal <strong>of</strong> Mammalogy<br />

29(3):247-259.<br />

CAHALANE, V. H. 1964. A preliminary<br />

study <strong>of</strong> <strong>the</strong> distribution and numbers <strong>of</strong><br />

cougar, grizzly and wolf in North<br />

America. New York Zoological Society,<br />

New York. 12 pages.<br />

CARDOZA, J. E. AND S. A. LANGLOIS. 2002.<br />

The eastern cougar: A management<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 83<br />

failure? Wildlife Society Bulletin<br />

30(1):265-273.<br />

CLARK, D. W., S. C. WHITE, A. K. BOWERS,<br />

L. D. LUCIO, AND G. A. HEIDT. 2002. A<br />

survey <strong>of</strong> recent accounts <strong>of</strong> <strong>the</strong><br />

mountain lion (Puma concolor) in<br />

Arkansas. Sou<strong>the</strong>astern Naturalist<br />

1(3):269-278.<br />

CLARKE, C. H. D. 1942. Cougar in<br />

Saskatchewan. Canadian Field-<br />

Naturalist 56(1):45.<br />

CULVER, M., W. E. JOHNSON, J. PECON-<br />

SLATTERY, AND S. J. O’BRIEN. 2000.<br />

Genomic ancestry <strong>of</strong> <strong>the</strong> American puma<br />

(Puma concolor). Journal <strong>of</strong> Heredity<br />

91(3):186-197.<br />

CUMBERLAND, R. E. AND J. A. DEMPSEY.<br />

1994. Recent confirmation <strong>of</strong> a cougar,<br />

Felis concolor, in New Brunswick.<br />

Canadian Field-Naturalist 108(2):224-<br />

226.<br />

DEEMS, E. F., JR. AND D. PURSLEY, editors.<br />

1978. North American furbearers:<br />

Their management, research and harvest<br />

status in 1976. International Association<br />

<strong>of</strong> Fish and Wildlife Agencies,<br />

Washington, D.C. 177 pages.<br />

DOUTT, J. K. 1969. <strong>Mountain</strong> lions in<br />

Pennsylvania? American Midland<br />

Naturalist 82(1):281-285.<br />

DOWNING, R. L. 1981. The current status<br />

<strong>of</strong> <strong>the</strong> cougar in <strong>the</strong> sou<strong>the</strong>rn<br />

Appalachian. Pages 142-151 in<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> Nongame and<br />

Endangered Wildlife Symposium, 13-14<br />

August, A<strong>the</strong>ns, GA.<br />

DOWNING, R. L. 1984. The search for<br />

cougars in <strong>the</strong> eastern United States.<br />

Cryptozoology 3:31-49.<br />

FLORIDA PANTHER INTERAGENCY<br />

COMMITTEE. 1993. Florida pan<strong>the</strong>rs<br />

getting a shot <strong>of</strong> fresh bloodlines. Fur-<br />

Fish-Game 90(12):10.<br />

GERSON, H. B. 1988. Cougar, Felis<br />

concolor, sightings in Ontario.


84 CRYPTIC COUGARS · Tischendorf<br />

Canadian Field-Naturalist 102(3):419-<br />

423.<br />

GOERTZ, J. W. AND R. ABEGG. 1966.<br />

Pumas in Louisiana. Journal <strong>of</strong><br />

Mammalogy 47(4):727.<br />

GRAHAM, B. 2003. Cougar killed Monday<br />

in Missouri is more pro<strong>of</strong> big cats roam<br />

state, experts say. The Kansas City Star,<br />

13 August.<br />

GREENWELL, J. R. 1996. The place <strong>of</strong> <strong>the</strong><br />

eastern puma in <strong>the</strong> natural history <strong>of</strong> <strong>the</strong><br />

larger felids. Pages 9-37 in J. W.<br />

Tischendorf and S. J. Ropski, editors,<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> Eastern Cougar<br />

Conference, 1994. American Ecological<br />

Research Institute, Fort Collins, CO.<br />

245 pages.<br />

HALL, E. R. 1981. The mammals <strong>of</strong> North<br />

America. 2 nd edition. John Wiley and<br />

Sons, New York. 1171 pages.<br />

HARDIN, S. E. 1996. The status <strong>of</strong> <strong>the</strong><br />

puma (Puma concolor) in Missouri,<br />

based on sightings. Unpublished M.S.<br />

<strong>the</strong>sis, Southwest Missouri State<br />

University, Springfield, MO.<br />

HEISEL, E. J. 2003. Cougars in Missouri -<br />

here for now, but future uncertain. Page<br />

1 in Eastern Cougar Foundation<br />

Newsletter, Spring issue. 6 pages.<br />

HUMPHREYS, C. R. 1994. Pan<strong>the</strong>rs <strong>of</strong> <strong>the</strong><br />

coastal plain. The Fig Leaf Press,<br />

Wilmington, NC. 200 pages.<br />

JOHNSON, K. 1998. Island habitats: A<br />

stronghold <strong>of</strong> carnivore biodiversity in<br />

agriculturally modified environments.<br />

Endangered Species Update 15(2):21-24.<br />

JOHNSON, K. 2000. Prairie lions.<br />

Wyoming Wildlife, February issue.<br />

Pages 13-19.<br />

JOHNSON, K. 2002. The mountain lions <strong>of</strong><br />

Michigan. Endangered Species Update<br />

19(2):27-31.<br />

KUYT, E. 1971. Possible occurrence <strong>of</strong><br />

cougar near Fort Smith, N.W.T. Blue<br />

Jay 29(3):142-143.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LEOPOLD, A, L. K. SOWLS, AND D. L.<br />

SPENCER. 1947. A survey <strong>of</strong> <strong>the</strong> overpopulated<br />

deer ranges in <strong>the</strong> United<br />

States. Journal <strong>of</strong> Wildlife Management<br />

11(2):162-177.<br />

LEWIS, J. C. 1969. Evidence <strong>of</strong> mountain<br />

lions in <strong>the</strong> Ozarks and adjacent areas,<br />

1948-1968. Journal <strong>of</strong> Mammalogy<br />

50(2):371-372.<br />

LINZEY, A. V. AND D. W. LINZEY. 1971.<br />

Mammals <strong>of</strong> Great Smoky <strong>Mountain</strong>s<br />

National Park. University <strong>of</strong> Tennessee<br />

Press, Knoxville. 114 pages.<br />

MANN, T. 1959. The “phantom” <strong>of</strong> Elk<br />

<strong>Mountain</strong>. South Dakota Conservation<br />

Digest 26(1):3-5.<br />

MCGINNIS, H. J. 1988. Search for puma<br />

(Felis concolor subsp.) tracks on dirt<br />

roads in <strong>the</strong> National Space Technology<br />

Laboratories and Mississippi Army<br />

Ammunition Plant sites and surrounding<br />

restrictive easement area in Hancock<br />

County, Mississippi. Final report to <strong>the</strong><br />

Missisippi Wildlife Heritage Fund. 34<br />

pages.<br />

NERO, R. W. AND R. E. WRIGLEY. 1977.<br />

Status and habits <strong>of</strong> <strong>the</strong> cougar in<br />

Manitoba. Canadian Field-Naturalist<br />

91(1):28-40.<br />

NOBLE, R. E. 1971. A recent record <strong>of</strong> <strong>the</strong><br />

puma (Felis concolor) in Arkansas.<br />

Southwestern Naturalist 16(2):209.<br />

NOWAK, R. M. 1976. The cougar in <strong>the</strong><br />

United States and Canada. United States<br />

Fish and Wildlife Service, Washington<br />

D. C. and New York Zoological Society,<br />

New York. 190 pages.<br />

QUINN, T. 1995. Using public sighting<br />

information to investigate coyote use <strong>of</strong><br />

urban habitat. Journal <strong>of</strong> Wildlife<br />

Management 59(2):238-245.<br />

RILEY, S. J. 1991. The cat with many<br />

names. Montana Outdoors 22(6):15-21.<br />

ROOP, L. 1971. The Wyoming lion<br />

situation. Wyoming Wildlife 35(12):16-<br />

21.


RUSS, W. B. 1997. The status <strong>of</strong> mountain<br />

lions in Texas. Pages 69-73 in W. D.<br />

Padley, editor, <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> Fifth<br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>. Total pages<br />

not provided in proceedings.<br />

RUSSELL, K. R. 1978. <strong>Mountain</strong> lion.<br />

Pages 207-225 in J. L. Schmidt and D.<br />

L. Gilbert, editors, Big game <strong>of</strong> North<br />

America - ecology and management.<br />

Stackpole Books, Harrisburg, PA. 494<br />

pages.<br />

SEALANDER, J. A. 1951. <strong>Mountain</strong> lion in<br />

Arkansas. Journal <strong>of</strong> Mammalogy<br />

32(3):364.<br />

SEALANDER, J. A. AND P. S. GIPSON. 1973.<br />

Status <strong>of</strong> <strong>the</strong> mountain lion in Arkansas.<br />

Arkansas Academy <strong>of</strong> Science<br />

<strong>Proceedings</strong> 27:38-41.<br />

SNOW, D. 1994. A report <strong>of</strong> a cougar near<br />

Port Rexton, NF. Osprey 25(1):44-46.<br />

STENLUND, M. 1985. Popple leaves and<br />

boot oil - a wildlife biologist in nor<strong>the</strong>rn<br />

Minnesota. Heritage North, Grand<br />

Rapids, MN. 126 pages.<br />

STOCEK, R. 1995. The cougar, Felis<br />

concolor, in <strong>the</strong> Maritime Provinces.<br />

Canadian Field-Naturalist 109(1):19-22.<br />

STOLL, B. 1996. What - black bears in<br />

Ohio? Wild Ohio, spring issue. Pages<br />

2-3.<br />

TISCHENDORF, J. W. 1992a. O<strong>the</strong>r cryptic<br />

cougars. Page 3 in Eastern Pan<strong>the</strong>r<br />

Update #1. American Ecological<br />

Research Institute, Fort Collins, CO. 6<br />

pages.<br />

TISCHENDORF, J. W. 1992b. Cougars in <strong>the</strong><br />

news. Pages 2-3 in Eastern Pan<strong>the</strong>r<br />

Update #2. American Ecological<br />

Research Institute, Ft. Collins, CO. 6<br />

pages.<br />

TISCHENDORF, J. W. 1992c. Commentary:<br />

Cougars, cooperation, communication…<br />

Pages 3-4 in Eastern Pan<strong>the</strong>r Update #2.<br />

American Ecological Research Institute,<br />

Ft. Collins, CO. 6 pages.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

CRYPTIC COUGARS · Tischendorf 85<br />

TISCHENDORF, J. W. 1993a. Pan<strong>the</strong>rs<br />

elsewhere. Page 2-3 in Eastern Pan<strong>the</strong>r<br />

Update #3. American Ecological<br />

Research Institute, Ft. Collins, CO. 8<br />

pages.<br />

TISCHENDORF, JAY W. 1993b. Pan<strong>the</strong>r<br />

status update. Pages 1-2 in Eastern<br />

Pan<strong>the</strong>r Update #3. American<br />

Ecological Research Institute, Ft.<br />

Collins, CO. 8 pages.<br />

TISCHENDORF, J. W. 1993c. Commentary:<br />

The changing perspective - what, or<br />

where, next? Pages 3-5 in Eastern<br />

Pan<strong>the</strong>r Update #3. American<br />

Ecological Research Institute, Ft.<br />

Collins, CO. 8 pages.<br />

TISCHENDORF, J W. 1994a. Are cougars<br />

back in <strong>the</strong> Nor<strong>the</strong>ast? AMC Outdoors<br />

60(10):21-23.<br />

TISCHENDORF, J. W. 1994b. A comment on<br />

<strong>the</strong> growing number <strong>of</strong> puma reports.<br />

Pages 2-3 in Eastern Pan<strong>the</strong>r Update #6.<br />

American Ecological Research Institute,<br />

Ft. Collins, CO. 7 pages.<br />

TISCHENDORF, J. W. 1994c. Eastern<br />

Cougar Conference, 1994. Page 3 in<br />

Eastern Pan<strong>the</strong>r Update #6. American<br />

Ecological Research Institute, Ft.<br />

Collins, CO. 7 pages.<br />

TISCHENDORF, JAY W. 1995. Texas cats<br />

join <strong>the</strong> sou<strong>the</strong>rn pan<strong>the</strong>r. Page 2 in<br />

Eastern Pan<strong>the</strong>r Update #7, American<br />

Ecological Research Institute, Ft.<br />

Collins, CO. 5 pages.<br />

TISCHENDORF, J. W. 1996a. Bruce Wright,<br />

<strong>the</strong> ghost cat, and o<strong>the</strong>r players. Pages<br />

39-45 in J. W. Tischendorf and S. J.<br />

Ropski, editors, <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong><br />

Eastern Cougar Conference, 1994.<br />

American Ecological Research Institute,<br />

Fort Collins, CO. 245 pages.<br />

TISCHENDORF, J. W. 1996b. Cats and<br />

dogma - search for <strong>the</strong> eastern puma.<br />

Predator Project Newsletter 7(4):8-9.<br />

TISCHENDORF, J. W. AND F. R. HENDERSON.<br />

1994. The puma in <strong>the</strong> central


86 CRYPTIC COUGARS · Tischendorf<br />

mountains and Great Plains - a synopsis.<br />

Blue Jay 52(4):218-223.<br />

TISCHENDORF, J. W. AND S. J. ROPSKI,<br />

editors. 1996. <strong>Proceedings</strong> <strong>of</strong> <strong>the</strong><br />

Eastern Cougar Conference, 1994.<br />

American Ecological Research Institute,<br />

Ft. Collins, CO. 245 pages.<br />

UNITED STATES FISH AND WILDLIFE<br />

SERVICE. 1982. Eastern Cougar<br />

Recovery Plan. United States Fish and<br />

Wildlife Service, Atlanta, GA. 17 pages.<br />

VAN DYKE, F. 1983. A western study <strong>of</strong><br />

cougar track surveys and environmental<br />

disturbances affecting cougars related to<br />

<strong>the</strong> status <strong>of</strong> <strong>the</strong> eastern cougar (Felis<br />

concolor couguar). Ph.D. <strong>the</strong>sis, State<br />

University <strong>of</strong> New York, Syracuse, NY.<br />

245 pages.<br />

WEDDLE, F. 1965. The ghost cats <strong>of</strong> <strong>the</strong><br />

Yukon. Defenders <strong>of</strong> Wildlife News<br />

40(5):53.<br />

WHITE, T. 1963. Cougars in Saskatchewan.<br />

Blue Jay 21(1):32-34.<br />

WHITE, T. 1967. History <strong>of</strong> <strong>the</strong> cougar in<br />

Saskatchewan. Blue Jay 25(2):84-89.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

WHITE, T. 1976. Cougar shot at Cutknife,<br />

Saskatchewan. Blue Jay 34(3):181.<br />

WHITE, T. 1982. Saskatchewan cougar -<br />

elusive cat. Special Publication Number<br />

14, Saskatchewan Natural History<br />

Society, Regina. 80 pages.<br />

WRIGHT, B. S. 1959. The ghost <strong>of</strong> North<br />

America - <strong>the</strong> story <strong>of</strong> <strong>the</strong> eastern<br />

pan<strong>the</strong>r. Vantage Press, New York. 140<br />

pages.<br />

WRIGHT, B. S. 1972. The eastern pan<strong>the</strong>r -<br />

a question <strong>of</strong> survival. Clarke, Irwin,<br />

and Company Limited, Toronto. 180<br />

pages.<br />

WRIGLEY, R. E. AND R. W. NERO. 1982.<br />

Manitoba’s big cat. Manitoba Museum<br />

<strong>of</strong> Man and Nature, Winnipeg. 68<br />

pages.<br />

YOUNG, S. P. AND E. A. GOLDMAN. 1946.<br />

The puma - mysterious American cat.<br />

The American Wildlife Institute,<br />

Washington, D. C. 358 pages.<br />

ZUIDEMA, M. 2002. Are <strong>the</strong>re mountain<br />

lions in Michigan? Fur-Fish-Game<br />

99(3):11-13.


MOUNTAIN LION STATUS REPORT: BRITISH COLUMBIA<br />

MATT AUSTIN, Large Carnivore Specialist, Ministry <strong>of</strong> Water, Land and Air Protection, PO<br />

Box STN PROV GOVT, Victoria BC V8W 9M4, Canada, email:<br />

Matt.Austin@gems7.gov.bc.ca<br />

Abstract: <strong>Mountain</strong> lions are classified as “Big Game” in British Columbia under <strong>the</strong> provincial<br />

Wildlife Act. There is no provincial mountain lion management plan, however, <strong>the</strong>re is a species<br />

account within <strong>the</strong> provincial Wildlife Harvest Strategy. The harvest management goal for<br />

mountain lions is “to optimize population sustainability within ecosystems while allowing for<br />

options and opportunities associated with hunting and viewing.” <strong>Mountain</strong> lions occupy all<br />

suitable habitats within BC. The distribution <strong>of</strong> mountain lions has been expanding northward in<br />

recent years due deer population increases resulting from less severe winters. The current<br />

provincial mountain lion population estimate is 4,000-6,000 and likely declining after peaking in<br />

<strong>the</strong> late 1990s. Declines are related to <strong>the</strong> severe winter in 1996/97 that reduced deer<br />

populations. Population estimates are based on <strong>the</strong> “best guesses” <strong>of</strong> regional biologists based<br />

on anecdotal and harvest/conflict information. Confidence in <strong>the</strong> population estimate is low but<br />

we have greater confidence in <strong>the</strong> trend estimate. <strong>Mountain</strong> lion hunting is allowed under<br />

General Open Seasons in all but two nor<strong>the</strong>rn regions with negligible populations. There are no<br />

explicit harvest criteria or objectives aside from quotas for female harvest in one region. Both<br />

harvest and mortalities resulting from conflicts increased from 1985 until 1996 and <strong>the</strong>n declined<br />

through 2002. Conservation Officers respond to conflicts with mountain lions through<br />

education, translocation or destruction; compensation is not provided for losses. Known<br />

mountain lion attacks increased during <strong>the</strong> 20 th century, peaking in <strong>the</strong> 1990s.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

87


88<br />

IMPROVING OUR UNDERSTANDING OF MOUNTAIN LION MANAGEMENT<br />

TRENDS: THE VALUE OF CONSISTENT MULT-STATE RECORD KEEPING<br />

CHRISTOPHER M. PAPOUCHIS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95814, USA, email: cpapouchis@mountainlion.org<br />

LYNN MICHELLE CULLENS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95814, USA, email: cullens@mountainlion.org<br />

Abstract: The sound management and conservation <strong>of</strong> mountain lions relies on comprehensive<br />

scientific data. Yet <strong>the</strong> cost <strong>of</strong> mountain lion research can be prohibitive and <strong>the</strong> results are <strong>of</strong>ten<br />

difficult if not impossible to extrapolate. Wildlife managers, field researchers, and conservation<br />

organizations would benefit from more complete and consistent records <strong>of</strong> validated mountain<br />

lion sightings, hunting mortalities, depredation incidents, and road kills. Scientists who have<br />

mined such data in <strong>the</strong> past have isolated important variables, generated important hypo<strong>the</strong>ses,<br />

and targeted future research. But <strong>the</strong>ir work is usually limited by funding, academic or agency<br />

agendas. Fur<strong>the</strong>r, <strong>the</strong>re is no long-term multi-state repository for mountain lion data. The task<br />

<strong>of</strong> data collection is made more difficult because <strong>the</strong>re is no multi-state standard, and <strong>the</strong>refore<br />

states collect and store data inconsistently. This presentation explores <strong>the</strong> potential for<br />

developing a multi-state database, and examines <strong>the</strong> existing state data sets in order to identify<br />

<strong>the</strong> essential variables that might be included.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


LESSENING THE IMPACT OF A PUMA ATTACK ON A HUMAN<br />

E. LEE FITZHUGH, Department <strong>of</strong> Wildlife, Fish, and Conservation Biology, University <strong>of</strong><br />

California, Davis, One Shields Avenue, Davis, CA 95616-8751, USA, email:<br />

elfitzhugh@ucdavis.edu<br />

SABINE SCHMID-HOLMES, Department <strong>of</strong> Wildlife, Fish, and Conservation Biology,<br />

University <strong>of</strong> California, Davis, One Shields Avenue, Davis, CA 95616-8751, USA<br />

MARC W. KENYON, Department <strong>of</strong> Wildlife, Fish, and Conservation Biology, University <strong>of</strong><br />

California, Davis, One Shields Avenue, Davis, CA 95616-8751, USA<br />

KATHY ETLING, 6830 St. Tropez Circle, Osage Beach, MO 65065, USA<br />

Abstract: We reviewed current data on puma (Puma concolor) attacks and near-attacks on humans to identify better<br />

ways for people to protect <strong>the</strong>mselves. Not since Paul Beier’s paper in 1991 has anyone documented, established<br />

criteria for validity, and analyzed puma attacks on humans, and much more data are now available. In attempting to<br />

examine human-puma behavioral interactions to 2003, <strong>the</strong> authors have collected accounts <strong>of</strong> 16 fatal and 92 nonfatal<br />

attacks that meet Beier’s criteria. In addition, we have an additional 32 fatal and 84 non-fatal attacks that failed<br />

to meet Beier’s criteria, ei<strong>the</strong>r for lack <strong>of</strong> physical contact, lack <strong>of</strong> verification, occurrence in Latin America,<br />

occurrence prior to 1890, or because <strong>the</strong>y were attacks on hunters. We also have accumulated 155 accounts <strong>of</strong><br />

behavioral interactions between pumas and humans at close proximity that did not result in an attack. We contrasted<br />

<strong>the</strong>se with incidents that resulted in an attack. We analyzed <strong>the</strong> use <strong>of</strong> Beier’s fatal:non-fatal attack ratio to predict<br />

missing incidents, and suspect that <strong>the</strong> criterion <strong>of</strong> validation may bias data for attacks prior to 1950. However,<br />

most <strong>of</strong> Beier’s statements and conclusions are confirmed. While <strong>the</strong> analysis is yet incomplete, this presentation<br />

includes highlights <strong>of</strong> our tentative analysis concerning common questions about puma attacks, illustrated by stories<br />

<strong>of</strong> real situations. Being aggressive and making loud noises helps protect people from a possible puma attack.<br />

Warning gunshots are much less effective than is yelling. Charging <strong>the</strong> puma seems to make it run away, but may<br />

result in some injury to <strong>the</strong> person who is charging. Groups <strong>of</strong> 5 people or more are relatively safe, but children in<br />

those groups may still be attacked. Hunters imitating animal sounds or smells may attract pumas, but <strong>the</strong>se<br />

situations usually do not result in serious injuries. People attacked while sleeping on <strong>the</strong> ground <strong>of</strong>ten receive only<br />

minor injuries because <strong>the</strong> puma runs away when <strong>the</strong> person or companions awake, yell, and resist. The strategies<br />

will usually work, but not always, because pumas have different personalities and seem to react differently to <strong>the</strong><br />

same situation.<br />

89<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Key words: Animal damage, attacks on humans, conflict with wildlife, cougar, human dimensions <strong>of</strong> wildlife<br />

management, mountain lion, pest control, predation, Puma concolor<br />

INTRODUCTION<br />

In this paper we use <strong>the</strong> common name<br />

“puma” to describe Puma concolor.<br />

Occasionally, when we quote o<strong>the</strong>r people,<br />

we retain <strong>the</strong>ir terminology, and <strong>the</strong>y <strong>of</strong>ten<br />

use “cougar” instead <strong>of</strong> “puma.” Beier<br />

(1991) analyzed 9 fatal and 44 non-fatal<br />

attacks by pumas on humans that occurred<br />

between 1890 and 1990 in <strong>the</strong> United States<br />

and Canada. In order to include an attack in<br />

his analysis, it must have been published,<br />

included statements from agency or medical<br />

personnel, and involved contact in which <strong>the</strong><br />

human was bitten, clawed, or knocked down<br />

by <strong>the</strong> puma. Excluded were situations<br />

involving captive pumas and in which<br />

people deliberately approached or harassed a<br />

puma. He found that 64% <strong>of</strong> victims were<br />

children, and only 13 <strong>of</strong> 37 (35%) <strong>of</strong> <strong>the</strong>se<br />

were alone, while 11 <strong>of</strong> 17 (65%) adult<br />

victims were alone when attacked. He also<br />

found that an aggressive response might<br />

avert and/or repel an attack. Yearlings and<br />

underweight cougars were most likely to


90 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

attack humans. Beier believed he had<br />

discovered all fatal attacks since 1890 that<br />

met his criteria, and all non-fatal attacks<br />

since 1970. Based on <strong>the</strong> ratio <strong>of</strong> fatal to<br />

non-fatal attacks during 1970-1990, he<br />

estimated that he had failed to identify about<br />

12 non-fatal attacks between 1890 and 1970.<br />

Beier documented an increase in frequency<br />

<strong>of</strong> attacks from <strong>the</strong> 1890-1969 period to<br />

1970-1990. While Beier did not tabulate<br />

“near-attacks,” he did analyze <strong>the</strong> victim’s<br />

actions that may have served to prevent <strong>the</strong><br />

attack. “Fighting back” and shouting loudly<br />

were actions that seemed to avert or repel<br />

attacks, as did waving arms, poking or<br />

hitting with sticks, throwing rocks, etc.<br />

Beier also reported an attempt at aversive<br />

conditioning <strong>of</strong> one puma, but it failed to<br />

prevent future aggression. We substantiated<br />

most <strong>of</strong> Beier’s findings and, because we<br />

have more data, we produced some<br />

additional tentative findings.<br />

We have accounts <strong>of</strong> a total <strong>of</strong> 224<br />

attacks by pumas on humans and 155<br />

behavioral interactions that did not result in<br />

an actual attack. The number <strong>of</strong> accounts<br />

through April 2003 that have information<br />

useful for analyzing any specific question is<br />

variable, but only 108 accounts meet Beier’s<br />

(1991) criteria. Of <strong>the</strong> 116 attacks that failed<br />

Beier’s criteria, 32 were fatal and 84 were<br />

non-fatal. Reasons for failure include lack <strong>of</strong><br />

physical contact, lack <strong>of</strong> verification,<br />

occurrence in Latin America, occurrence<br />

prior to 1890, or because <strong>the</strong>y were attacks<br />

on hunters. Beier’s strict criteria avoid<br />

errors <strong>of</strong> commission, but allow for many<br />

omissions. Because we are interested in<br />

behavior more than just counting attacks, we<br />

decided to relax <strong>the</strong> Beier limitations as long<br />

as accounts seemed plausible and contained<br />

useful information. Our intent is to analyze<br />

human and puma behavior in <strong>the</strong>se<br />

situations where data are available and<br />

compare attacks with encounters to try to<br />

provide better advice for people who come<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

face-to-face with a puma, or who want to<br />

prepare for that eventuality. We are still<br />

organizing <strong>the</strong> data, so this report is not <strong>the</strong><br />

final one, but we do have a few<br />

recommendations to make at this time, and<br />

we will make some observations about<br />

reliability <strong>of</strong> reports and frequency <strong>of</strong><br />

attacks in general.<br />

METHODS<br />

We defined “non-attack encounters” as<br />

behavioral interactions between pumas and<br />

humans at close proximity that do not result<br />

in an attack. We purposely did not define<br />

“close proximity” to place emphasis on<br />

“behavioral interactions.” We excluded<br />

incidents in which <strong>the</strong> puma was sighted and<br />

<strong>the</strong>n left, and included incidents in which <strong>the</strong><br />

puma and human exchanged multiple<br />

behaviors. “Close proximity” is necessary<br />

for this to occur, but <strong>the</strong> distance may vary,<br />

and we were more flexible regarding<br />

distance criteria than for behavioral criteria.<br />

If we believed we could learn from <strong>the</strong><br />

interaction, we included it. Most <strong>of</strong> our data<br />

are from published popular accounts,<br />

sometimes substantiated by an agency<br />

incident report. Etling (2001), in particular,<br />

solicited personal accounts from individuals<br />

in <strong>the</strong>ir own words, both before and after<br />

publication <strong>of</strong> her book. We categorized<br />

incidents in several ways to better analyze<br />

and evaluate <strong>the</strong> data. One mentioned in<br />

this paper is a category we called “attacks<br />

terminated by humans.” These are incidents<br />

in which a puma was shot while charging, or<br />

at least clearly intent on creeping up very<br />

close to a human in spite <strong>of</strong> <strong>the</strong> person’s<br />

efforts to discourage such behavior. We<br />

have 20 such accounts, 10 <strong>of</strong> which are from<br />

hunters. In only 3 <strong>of</strong> <strong>the</strong> hunting accounts<br />

was it clear that <strong>the</strong> hunter was doing<br />

something that might attract a puma (e.g.,<br />

using deer scent, calling turkeys, etc.). Six<br />

accounts were <strong>of</strong> agency employees<br />

investigating previous encounters between<br />

humans and pumas.


We defined a child as a person under 13<br />

years <strong>of</strong> age, whereas Beier (1991) defined a<br />

child as being under 16. We differed from<br />

Beier because we believed that <strong>the</strong> younger<br />

age better represented when girls and boys<br />

reach adult size and behavior.<br />

We entered data into a spreadsheet and<br />

made inferences where <strong>the</strong>y were defensible.<br />

For example, when <strong>the</strong> victim fired a shot,<br />

and <strong>the</strong> attack was fatal to <strong>the</strong> victim, we<br />

inferred that <strong>the</strong> shot did not deter <strong>the</strong><br />

attacking puma, even though <strong>the</strong> account did<br />

not specifically say so. The matrix was<br />

organized with individual incidents in rows,<br />

separated into various categories such as:<br />

Beier fatal, Beier non-fatal, non-verified<br />

(o<strong>the</strong>rwise meeting Beier’s criteria), fatal<br />

prior to 1890, nonfatal prior to 1890, Latin<br />

American incidents, close encounters,<br />

provoked attacks, encounters while hunting,<br />

and 10 o<strong>the</strong>r categories. Columns included<br />

raw data and data coded into categories for<br />

analysis, all <strong>of</strong> which made 193 columns,<br />

including 147 columns <strong>of</strong> original data.<br />

Broad categories <strong>of</strong> data in columns<br />

included descriptions <strong>of</strong> <strong>the</strong> habitat and<br />

setting, identification and descriptions <strong>of</strong> <strong>the</strong><br />

victims, and detailed descriptions <strong>of</strong> <strong>the</strong><br />

incident, including a written description and<br />

data parsed into separate columns.<br />

Information sources, previous reported<br />

puma activities in <strong>the</strong> area, necropsy results,<br />

and injuries sustained by <strong>the</strong> victim(s) also<br />

were entered. An example <strong>of</strong> data sought<br />

and entered is in <strong>the</strong> partial questionnaire<br />

mentioned below.<br />

On a few recent occasions, when it was<br />

possible, we sent a 4-page questionnaire to<br />

<strong>the</strong> witness <strong>of</strong> an attack to get more detailed<br />

information. The questionnaire was generic<br />

and designed to help people remember detail<br />

without leading <strong>the</strong>m to specific answers.<br />

Thus, we did not ask whe<strong>the</strong>r <strong>the</strong> puma was<br />

pumping its rear feet up and down, but<br />

instead we asked what <strong>the</strong> puma was doing<br />

with its feet. Although <strong>the</strong> situations were<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 91<br />

serious, <strong>the</strong> responses sometimes interacted<br />

with <strong>the</strong> generic nature <strong>of</strong> <strong>the</strong> questions in a<br />

humorous manner. A real example <strong>of</strong> a<br />

questionnaire we recently received from a<br />

man who gave permission to use it will<br />

illustrate both <strong>the</strong> detail <strong>of</strong> <strong>the</strong> questions and<br />

some <strong>of</strong> <strong>the</strong> humor that occasionally occurs.<br />

It also illustrates an incident that is not<br />

classified as an attack under <strong>the</strong> Beier<br />

(1991) criteria, because no contact was<br />

made. We have called it a “terminated<br />

attack.”<br />

To appreciate <strong>the</strong> humor, it helps to<br />

imagine <strong>the</strong> victim’s perspective on <strong>the</strong><br />

attack and his response to <strong>the</strong> “ivory-tower”<br />

nature <strong>of</strong> <strong>the</strong> questions that came from some<br />

stranger at a far-<strong>of</strong>f university. The victim<br />

was hunting deer in a remote area when he<br />

was charged from behind. The puma<br />

vocalized with a “growl-hiss” sound, which<br />

alerted <strong>the</strong> victim to <strong>the</strong> charge. He killed<br />

<strong>the</strong> puma during its charge, after missing his<br />

first shot. After <strong>the</strong> second shot <strong>the</strong> puma<br />

fell only 5.2 m (17 feet) from him. Here are<br />

a few <strong>of</strong> <strong>the</strong> questions and <strong>the</strong> victim’s<br />

responses:<br />

How was <strong>the</strong> puma identified (what<br />

evidence or characteristics)?<br />

The cougar died, not much question that<br />

it was a cougar.<br />

Condition <strong>of</strong> teeth:<br />

Perfect teeth, no fillings.<br />

Condition <strong>of</strong> claws:<br />

Damned sharp.<br />

Did attack involve a fatality?<br />

Yes, <strong>the</strong> cougar.<br />

Puma posture and position <strong>of</strong> ears at time <strong>of</strong><br />

first sighting:<br />

The cat was charging me. I later<br />

measured <strong>the</strong> distance from where it<br />

started <strong>the</strong> charge, which was 86 feet. I<br />

don’t recall what position <strong>the</strong> ears were<br />

in. [86 feet is 26.2 m].<br />

What was puma doing with eyes and tail at<br />

time <strong>of</strong> first sighting?<br />

Tail seemed to be floating out behind <strong>the</strong>


92 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

cat.<br />

What was puma doing with its feet at time<br />

<strong>of</strong> first sighting?<br />

Bounding toward me.<br />

Victim behavior just after first sighting?<br />

Putting rifle to shoulder and firing.<br />

Were <strong>the</strong>re signs <strong>of</strong> aggression by puma?<br />

The cougar was charging me, full speed<br />

ahead, which seemed pretty aggressive to<br />

me at <strong>the</strong> time.<br />

Did victim fight back?<br />

Yes.<br />

How?<br />

I shot <strong>the</strong> puma in <strong>the</strong> throat/chest.<br />

Puma response:<br />

Puma rolled and died.<br />

Was puma injured by victim?<br />

Yes, severely.<br />

Quality <strong>of</strong> Data<br />

Some accounts are not included in this<br />

analysis. The 224 attacks and 155<br />

encounters we analyze do not include 14<br />

incidents for which we believe additional<br />

investigation is needed to validate <strong>the</strong>ir<br />

accuracy. Nor do we include 8 o<strong>the</strong>r<br />

incidents we suspect, but cannot prove, are<br />

duplicates <strong>of</strong> incidents included in <strong>the</strong> tally.<br />

In addition, we have 10 more reports that we<br />

decided not to use because <strong>the</strong>y included too<br />

little information or were <strong>of</strong> doubtful<br />

validity. Our data do include incidents that<br />

do not meet Beier’s criteria, but we kept<br />

those separate in order not to invalidate<br />

comparison with Beier’s (1991) findings. As<br />

we analyze data more completely, more <strong>of</strong><br />

<strong>the</strong> incidents may be excluded, primarily for<br />

lack <strong>of</strong> information. Our files also include<br />

several accounts that have recently come to<br />

our attention that are not yet entered in <strong>the</strong><br />

database and are not included in this paper.<br />

These latter data, and additional details from<br />

<strong>the</strong> accounts we have analyzed will be<br />

included in a later, more complete treatment.<br />

As already mentioned, we included all<br />

<strong>the</strong> accounts we have discovered, and we<br />

tried to estimate <strong>the</strong> validity <strong>of</strong> each. The<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

editor <strong>of</strong> Outdoor Life, from 1897 to 1925,<br />

asked his readers to provide accounts <strong>of</strong><br />

puma attacks on humans. He <strong>the</strong>n tried to<br />

verify <strong>the</strong> accuracy <strong>of</strong> each account,<br />

generally without success. In each case, he<br />

was ei<strong>the</strong>r unable to locate <strong>the</strong> respondent,<br />

or <strong>the</strong> knowledgeable people from <strong>the</strong> area<br />

where <strong>the</strong> attack was supposed to have<br />

occurred claimed no knowledge <strong>of</strong> it<br />

(Anonymous 1925, cited by Beier 1991).<br />

During this same period, Forest and Stream,<br />

which became Field and Stream, also<br />

printed numerous personal accounts <strong>of</strong><br />

encounters with pumas. At this time, we<br />

have been able to locate only 1 reference to<br />

incidents that may have been confirmed<br />

(Marsh 1917), or failed confirmation, by<br />

Outdoor Life (Anonymous 1917). We are<br />

aware <strong>of</strong> one, and perhaps three fraudulent<br />

accounts in recent years, and we also<br />

questioned <strong>the</strong> validity <strong>of</strong> one unusual<br />

account that we later found had been<br />

confirmed by an agency. We recently tried,<br />

unsuccessfully, to obtain agency<br />

confirmation <strong>of</strong> an account, only later to find<br />

that <strong>the</strong> confirmation had been provided to<br />

Etling several years earlier. Therefore, <strong>the</strong><br />

verifications <strong>the</strong>mselves can be erroneous in<br />

ei<strong>the</strong>r direction. It is possible that we have<br />

analyzed a few spurious reports, but if so,<br />

<strong>the</strong>ir effect on our findings should be minor.<br />

We have placed our 379 useful incidents<br />

in categories <strong>of</strong> similar types <strong>of</strong> incidents<br />

and levels <strong>of</strong> reliability. The 108 fatal and<br />

non-fatal incidents that meet Beier’s (1991)<br />

criteria may be considered to be a complete<br />

count <strong>of</strong> well-defined and verified attacks.<br />

(See a more complete defense <strong>of</strong> this<br />

assumption in <strong>the</strong> results and discussion<br />

section). The few verified non-fatal<br />

incidents we may have missed would not<br />

affect group values in an important way.<br />

The 116 o<strong>the</strong>r attacks and 155 encounters<br />

represent nei<strong>the</strong>r a total count nor a<br />

statistical sample, nor do we know anything<br />

about <strong>the</strong> underlying statistical distribution


except that attacks by pumas on humans are<br />

rare. We know nothing <strong>of</strong> bias caused by<br />

missing data. We can speculate that missing<br />

cases may not have been considered<br />

important enough to report; <strong>the</strong>re may have<br />

been nobody to report to; fatal incidents may<br />

have been undiscovered; <strong>the</strong>y may have<br />

been reported and <strong>the</strong> records lost; <strong>the</strong>y may<br />

have been printed in obscure references that<br />

we did not find; we may have wrongly<br />

discarded some incidents recorded from<br />

word-<strong>of</strong>-mouth accounts, etc. Thus, we<br />

have a core <strong>of</strong> strictly defined data that we<br />

treat as a total count. These data are<br />

restricted by <strong>the</strong> verification criterion in<br />

such a way that <strong>the</strong> passage <strong>of</strong> time reduces<br />

opportunity for verification. The core exists<br />

in a matrix <strong>of</strong> less well-defined incidents,<br />

<strong>the</strong> statistical properties <strong>of</strong> which are<br />

unknown.<br />

At a finer level, even <strong>the</strong> welldocumented<br />

cases have many blank cells in<br />

<strong>the</strong> data matrix because specific items were<br />

unknown or not reported, and <strong>the</strong>re was no<br />

way to infer <strong>the</strong> information. In <strong>the</strong>se<br />

situations, we usually cannot assume <strong>the</strong><br />

nature <strong>of</strong> possible biases caused by missing<br />

data, if <strong>the</strong>y exist.<br />

To summarize, we have nearly total<br />

counts <strong>of</strong> incidents in Beier-quality<br />

categories, missing incidents in o<strong>the</strong>rs, and<br />

missing data in all incidents. The categories<br />

in which we can assume total counts may be<br />

subject to a time-related bias that may affect<br />

<strong>the</strong> statistical distribution <strong>of</strong> <strong>the</strong> data. We<br />

have little a priori information to guide us.<br />

Therefore, we are restricted mostly to<br />

descriptive statistics and forming hypo<strong>the</strong>ses<br />

that may later be independently verified.<br />

We do use a Chi-square test to explore <strong>the</strong><br />

similarity <strong>of</strong> <strong>the</strong> distributions in three<br />

different categories, two <strong>of</strong> which we<br />

assume are total counts.<br />

Statistical Methods<br />

We have no a priori models or<br />

hypo<strong>the</strong>ses. We had believed that most, if<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 93<br />

not all, puma attacks were predatory, but<br />

information provided by Sweanor et al.<br />

(personal communication) contradicts that<br />

belief. Thus, our analysis is exploratory,<br />

examining <strong>the</strong> data to find hypo<strong>the</strong>ses and<br />

descriptive models that may later be<br />

subjected to data collection and statistical<br />

interpretation. In <strong>the</strong> few cases where<br />

statistical testing was warranted, we report<br />

<strong>the</strong> test used along with <strong>the</strong> results, but for<br />

<strong>the</strong> majority <strong>of</strong> situations, descriptive<br />

statistics are <strong>the</strong> only analysis used.<br />

Never<strong>the</strong>less, we feel secure in drawing<br />

some conclusions about how to reduce risk<br />

<strong>of</strong> serious injury during a puma encounter.<br />

RESULTS AND DISCUSSION<br />

What Can We Tell About <strong>the</strong> Data?<br />

With respect to counting attacks and<br />

comparing frequencies, we can be a little<br />

more specific about statistical qualities <strong>of</strong><br />

<strong>the</strong> data. Like Beier (1991), we believe we<br />

have a near-complete count <strong>of</strong> verified fatal<br />

attacks from 1890 to 2003 in <strong>the</strong> U. S. and<br />

Canada. Authors <strong>of</strong> new books (Danz,<br />

1999, Deurbrouck and Miller 2001, and<br />

Etling 2001) did extensive new searches,<br />

and failed to find any fatal attacks that meet<br />

Beier’s criteria that were not included by<br />

Beier in his original list, or else occurred<br />

after his publication. The 108 attacks we<br />

analyze that meet Beier’s criteria include 7<br />

fatal and 38 non-fatal attacks that occurred<br />

after Beier published his list, and 9 fatal and<br />

54 non-fatal attacks that meet Beier’s<br />

criteria and occurred between 1890 and<br />

1991. On <strong>the</strong> basis <strong>of</strong> additional information<br />

(personal communication, Dale Elliot to<br />

Etling, July 2000), we moved one <strong>of</strong> Beier’s<br />

non-fatal attacks (Bird and Sieh, Nevada,<br />

1971) to <strong>the</strong> “provoked attack” category and<br />

added 11 new non-fatal attacks between<br />

1890 and 1991. It is possible that attacks,<br />

even fatal ones, occurred in <strong>the</strong> U. S. that<br />

were never known, especially during <strong>the</strong><br />

depression years <strong>of</strong> <strong>the</strong> 1930s and <strong>the</strong><br />

various gold rushes in localities in <strong>the</strong>


94 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

Table 1. Calculations <strong>of</strong> non-fatal attacks by pumas in <strong>the</strong> U.S. and Canada between 1890 and 1969<br />

that might not have been detected.<br />

Source<br />

Beier (1991)<br />

1970-1990<br />

Our data<br />

1970-2001<br />

Ratio NF/F<br />

from 1970<br />

onward<br />

× Fatal<br />

attacks 1890-<br />

1969<br />

= Calculated<br />

non-fatal<br />

attacks<br />

− Attacks<br />

detected<br />

1890-1969<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

= Calculated<br />

attacks not<br />

detected<br />

31÷5 = 6.2 4 25 13 12<br />

69÷12 = 5.75 4 23 21 2<br />

western U.S. It is certain that attacks<br />

occurred in Latin America for which no<br />

records are available. Written accounts <strong>of</strong><br />

attacks occur in some obscure publications,<br />

not susceptible to easy location. One<br />

example is <strong>the</strong> killing <strong>of</strong> Henry Ramsey in<br />

1876 (Hunter 1922:110). That account was<br />

found by scanning <strong>the</strong> table <strong>of</strong> contents <strong>of</strong> a<br />

county history that was in a fund-raising<br />

auction <strong>of</strong> “white elephant” donations. All<br />

<strong>of</strong> <strong>the</strong>se missing accounts would, <strong>of</strong> course,<br />

fail Beier’s criterion <strong>of</strong> verification, and<br />

perhaps o<strong>the</strong>r criteria as well. We believe<br />

we can use <strong>the</strong> 16 fatal attacks that meet<br />

Beier’s criteria as a complete count, not<br />

requiring statistical measures <strong>of</strong> variability.<br />

The 92 Beier-quality non-fatal attacks<br />

probably include a large majority <strong>of</strong> all nonfatal<br />

attacks (Table 1). They should be<br />

representative, and probably can be treated<br />

as a complete count. However, <strong>the</strong>y were<br />

not sampled according to a statistical design.<br />

The remaining 116 attacks and 155<br />

encounters have unknown statistical<br />

properties, with variable report quality and<br />

amounts <strong>of</strong> information.<br />

Beier (1991) estimated that he might<br />

have missed finding 12 non-fatal accounts<br />

between 1970 and 1990. He did this by<br />

assuming that he found all <strong>the</strong> accounts from<br />

1970 to 1990, and multiplying <strong>the</strong> ratio <strong>of</strong><br />

non-fatal to fatal attacks during that period<br />

by <strong>the</strong> number <strong>of</strong> fatal attacks from 1890 to<br />

1969. We discovered 8 <strong>of</strong> <strong>the</strong> 12 missing<br />

accounts, and also 3 more between 1970 and<br />

1990, so we recalculated <strong>the</strong> potentially<br />

missing non-fatal accounts (Table 1). These<br />

were calculated through 2001, as <strong>the</strong> 2002<br />

data are yet incomplete. The analysis in<br />

Table 1 assumes a constant rate <strong>of</strong> attacks<br />

across years. Puma populations, prey<br />

populations, and <strong>the</strong> number, age, sex, and<br />

group size <strong>of</strong> people at risk may have<br />

changed considerably since 1890. These<br />

factors may affect <strong>the</strong> attack rate. Thus, <strong>the</strong><br />

calculation may be invalid to <strong>the</strong> degree that<br />

<strong>the</strong>se parameters have changed.<br />

Even if <strong>the</strong> analysis in Table 1 is valid,<br />

<strong>the</strong> number <strong>of</strong> fatal attacks is small enough<br />

that a change in even one attack can alter <strong>the</strong><br />

calculation <strong>of</strong> non-detected non-fatal<br />

attacks. Because <strong>the</strong>re is a chance that we<br />

may have missed some fatal attacks prior to<br />

1970 (and especially prior to 1950 as<br />

discussed later), this is a tentative<br />

calculation that serves only to illustrate that<br />

<strong>the</strong>re likely are some incidents we have not<br />

found, but that number is relatively small.<br />

We will return to this topic later with respect<br />

to possible bias caused by Beier’s<br />

verification criterion.<br />

Comparing Verified and Unverified Data<br />

Figure 1 shows <strong>the</strong> relationship through<br />

time between <strong>the</strong> 15 Beier-quality fatal<br />

attacks, <strong>the</strong> 86 non-fatal attacks, and <strong>the</strong> 27<br />

non-verified, non-fatal attacks, 1890-1999.<br />

The Beier-quality non-fatal attack curve


Figure 1. A comparison <strong>of</strong> patterns <strong>of</strong> fatal<br />

and non-fatal attacks that conform to Beier’s<br />

(1991) criteria with non-verified non-fatal<br />

attacks.<br />

diverges from <strong>the</strong> o<strong>the</strong>r two beginning in<br />

1950. All three types <strong>of</strong> data have been<br />

subject to <strong>the</strong> same bias from a conscious<br />

increase in collecting attack reports<br />

beginning with Barnes (1960), our effort<br />

from 1984 (Fitzhugh and Gorenzel 1986),<br />

and intensive searches beginning about 1990<br />

(Beier 1991) and increasing in 1998-2001<br />

(Danz 1999, Deuerbrouck and Miller 2001,<br />

Etling 2001).The Beier-quality fatal attacks<br />

curve (Figure 1) began to exceed past levels<br />

in <strong>the</strong> 1970s, and increased even more in <strong>the</strong><br />

1990s. We confirmed that Beier<br />

documented all <strong>the</strong> verified fatal attacks<br />

since 1890, and Figure 1 shows that <strong>the</strong> nonverified,<br />

non-fatal attacks coincide closely<br />

with Beier-quality fatal attacks. The<br />

difference in <strong>the</strong> shape <strong>of</strong> <strong>the</strong> curves in<br />

Figure 1 may be partly attributable to an<br />

increase in agency funding, staffing, and<br />

attention to puma incidents beginning in<br />

about 1950, allowing for more verification<br />

and recording <strong>of</strong> non-fatal incidents (Harley<br />

Shaw, personal communication). Thus, <strong>the</strong><br />

post-1950 non-verified data may be<br />

depressed because a greater proportion <strong>of</strong><br />

those incidents were verified than happened<br />

pre-1950. The post-1950 verified non-fatal<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 95<br />

Figure 2. Non-fatal to fatal attack ratios.<br />

Bars represent a 20-year running average<br />

beginning with 1890-1909, 1900-1919, etc.,<br />

except that 1990 represents only 1990-1999.<br />

Includes only data that conform to Beier’s<br />

(1991) criteria. The zero value at 1890<br />

represents a 20-year period, 1890-1909, with<br />

no verified non-fatal attacks. The zero value<br />

at 1950 represents a 20-year period, 1950-<br />

1969, without fatal attacks. (N = 101; 15 =<br />

fatal, 86 = non-fatal.)<br />

data would have increased by <strong>the</strong> same<br />

amount.<br />

The proportional change in <strong>the</strong> Beierquality<br />

non-fatal attack curve after 1949,<br />

applied to higher numbers <strong>of</strong> non-fatal<br />

incidents compared with fatal incidents,<br />

magnifies <strong>the</strong> visual comparison between<br />

<strong>the</strong> curves, although prior to 1950 <strong>the</strong> nonfatal<br />

curve is only slightly higher than <strong>the</strong><br />

fatal curve. The effect <strong>of</strong> magnification can<br />

be removed by examining proportions<br />

directly, using non-fatal to fatal ratios by 20year<br />

periods (Figure 2). The 1960-1990<br />

ratios seem consistent, with an average<br />

value <strong>of</strong> 6.5. The 1890-1930 average is less,<br />

at 2.4 (including 1890-1909, when <strong>the</strong>re<br />

were no recorded non-fatal attacks), but is<br />

more variable. Only <strong>the</strong> 1940-1959 ratio<br />

seems unusually high, created by <strong>the</strong> first<br />

big increase in non-fatal attacks, which<br />

occurred during <strong>the</strong> 1950s, while no fatal<br />

attacks occurred 1950-1959. While non-


96 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

Table 2. The effect <strong>of</strong> including non-verified puma attack incidents on <strong>the</strong> non-fatal:fatal attack<br />

ratios before and after 1950, U. S. and Canada.<br />

Data 1890-1949 1950-1999 Difference<br />

Beier-quality only 6 ÷ 4 = 1.5 86 ÷ 12 = 7.2 5.7<br />

Beier-quality & all non-verified 16 ÷ 7 = 2.3 96 ÷ 13 = 7.3 5.0<br />

Beier-quality & non-verified non-fatal only 16 ÷ 4 = 4.0 96 ÷ 12 = 8.0 4.0<br />

fatal attacks did not decrease after that, fatal<br />

attacks increased starting in 1970, bringing<br />

<strong>the</strong> ratios down. Some <strong>of</strong> <strong>the</strong> variation in<br />

Figure 2 is caused by zero values. The nonverified,<br />

non-fatal data (Figure 1) are<br />

approximately <strong>of</strong> <strong>the</strong> same value as <strong>the</strong><br />

Beier-quality fatal data, and <strong>the</strong> curves are<br />

very similar, so we may be justified in<br />

combining <strong>the</strong> non-verified non-fatal data (n<br />

= 23) with <strong>the</strong> Beier-quality non- fatal data<br />

(n = 86). Four non-verified fatal attacks also<br />

were included with <strong>the</strong> 15 verified fatal<br />

attacks to be consistent. Ratios for <strong>the</strong><br />

periods before and after 1950 (Table 2)<br />

show that adding all non-verified incidents<br />

(fatal and non-fatal) increased <strong>the</strong> ratios<br />

slightly; ratios were even greater when only<br />

non-verified, non fatal incidents were added.<br />

The non-verified data also reduced <strong>the</strong><br />

differences between <strong>the</strong> earlier and later<br />

periods. If we assume that <strong>the</strong> underlying<br />

ratio <strong>of</strong> non-fatal to fatal attacks is<br />

consistent across years, it appears that<br />

excluding non-verified data changes <strong>the</strong><br />

ratios. The changes represent bias if we are<br />

justified in using <strong>the</strong> non-verified data.<br />

Beier’s (1991) calculation <strong>of</strong> ratios from<br />

1970-1990, to estimate missing non-fatal<br />

attacks prior to 1970, assumed a constant<br />

relationship. It seems logical that we have<br />

detected a larger proportion <strong>of</strong> <strong>the</strong> actual<br />

non-fatal attacks in recent years, and <strong>the</strong><br />

data seem to indicate that this is so (Figures<br />

1-3). However, some biological,<br />

demographic, and cultural differences may<br />

have caused a change in ratios not related to<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

reporting frequency. These changes have to<br />

do with changes in persecution <strong>of</strong> pumas,<br />

especially before, during, and after World<br />

Wars I and II, <strong>the</strong> number <strong>of</strong> people using<br />

puma habitat, changes in <strong>the</strong> degree <strong>of</strong> puma<br />

habituation to humans, changes in <strong>the</strong><br />

proportion <strong>of</strong> children versus adults exposed<br />

to pumas, and changes in <strong>the</strong> inclination or<br />

ability <strong>of</strong> people to report incidents. We<br />

believe <strong>the</strong> underlying non-fatal to fatal ratio<br />

may have changed about <strong>the</strong> middle <strong>of</strong> <strong>the</strong><br />

20 th century, but measurement <strong>of</strong> this<br />

potential change is confused by changes in<br />

Figure 3. Non-fatal to fatal attack ratios.<br />

Bars represent a 20-year running average<br />

beginning with 1890-1909, 1900-1919, etc.,<br />

except that 1990 represents only 1990-1999.<br />

Included are data that conform to Beier’s<br />

(1991) criteria and non-verified data, both<br />

fatal and non-fatal. The zero value at 1950<br />

represents a 20-year period, 1950-1969,<br />

without fatal attacks. (n = 128; 19 = fatal, 109<br />

= non-fatal.)


<strong>the</strong> rate <strong>of</strong> detection <strong>of</strong> attacks and changes<br />

in verification <strong>of</strong> detected attacks.<br />

We did subject <strong>the</strong> Beier-quality data to<br />

2 Chi-square comparisons with 27 <strong>of</strong> <strong>the</strong><br />

accounts that were not verified, but met<br />

o<strong>the</strong>r Beier criteria. These are <strong>the</strong> same data<br />

shown in Figure 1. We used <strong>the</strong> Beierquality<br />

data as <strong>the</strong> observed value and <strong>the</strong><br />

non-verified non-fatal data, paired with <strong>the</strong><br />

Beier-quality data by decades, as <strong>the</strong><br />

expected value. The Beier- quality fatal data<br />

were not different from <strong>the</strong> non-verified,<br />

non-fatal data (χ 2 = 13.7, 10 df, P =


98 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

Figure 4. Puma responses to noise, including<br />

lethal shots from firearms (n = 133).<br />

to draw out <strong>the</strong> duration <strong>of</strong> <strong>the</strong> sound. For<br />

example, a puma was recently stalking<br />

chickens in <strong>the</strong> yard <strong>of</strong> a lighthouse<br />

compound, with people around and active,<br />

in broad daylight. One man banged doors,<br />

without effect, <strong>the</strong>n got a .22 rifle and fired a<br />

shot into <strong>the</strong> ground without even causing<br />

<strong>the</strong> puma to flinch or look up. It was fixated<br />

on <strong>the</strong> chickens. The man <strong>the</strong>n fired 7-8<br />

shots rapidly into <strong>the</strong> ground, upon which<br />

<strong>the</strong> puma looked up and walked into <strong>the</strong><br />

nearby brush, but did not leave <strong>the</strong> area. It<br />

stayed on a nearby high area and watched<br />

while people put <strong>the</strong> chickens into a pen<br />

(Robert Hansen, Pacific Rim National Park,<br />

Vancouver Island, B.C., personal<br />

communication, 8 May, 2003). This<br />

indicates that <strong>the</strong> puma did not react to a<br />

single shot or to slamming a door, but did<br />

react to a subsequent rapid series <strong>of</strong> shots.<br />

We conclude that noise is effective, but<br />

<strong>the</strong> kind <strong>of</strong> noise makes a difference. The<br />

best deterrent in <strong>the</strong> event <strong>of</strong> a puma<br />

encounter is to yell or scream as loudly as<br />

possible. If you are going to shoot a gun,<br />

you should fire in rapid succession to<br />

frighten <strong>the</strong> animal away or shoot to kill <strong>the</strong><br />

puma.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

What if You Charge <strong>the</strong> Puma?<br />

Our data include 6 accounts in which <strong>the</strong><br />

primary victim ei<strong>the</strong>r charged <strong>the</strong> puma and<br />

fought with it, or engaged in “mock lunges”<br />

toward <strong>the</strong> puma. In 3 cases in which<br />

people actually charged and made contact<br />

with <strong>the</strong> puma, <strong>the</strong> puma left <strong>the</strong> area,<br />

sometimes after a brief scuffle in which <strong>the</strong><br />

human suffered light degrees <strong>of</strong> injury. Two<br />

examples follow: a man heard a commotion<br />

in his back yard and went to investigate. He<br />

thought his Scottie dog was being attacked<br />

by a large German shepherd. It was really a<br />

mountain lion, but he didn’t realize it until<br />

he had jumped onto <strong>the</strong> attackers back.<br />

When <strong>the</strong> man realized it was a puma, he let<br />

go after a brief scuffle and <strong>the</strong> puma ran <strong>of</strong>f.<br />

The man received stitches for cuts behind<br />

his ear (Colorado Division <strong>of</strong> Wildlife<br />

2002). In <strong>the</strong> o<strong>the</strong>r case, a man noticed a<br />

puma eating his daughter’s house cat, and<br />

decided to save <strong>the</strong> cat by wrestling with <strong>the</strong><br />

puma. The puma swatted <strong>the</strong> man in <strong>the</strong><br />

face, and <strong>the</strong> man <strong>the</strong>n decided to let go. The<br />

puma left with <strong>the</strong> house cat in its mouth<br />

(The New York Times 2002).<br />

Two cases involved repeated “mock<br />

lunges” by people causing <strong>the</strong> pumas to<br />

leave <strong>the</strong> area without attacking. In <strong>the</strong> first<br />

case, a woman came upon a puma crouched<br />

about 1.8 m (6 feet) away. It began to move<br />

toward her in a crouched position, growling.<br />

She lunged forward, holding arms wide and<br />

growled back at it. It retreated a bit, began<br />

to approach again. She growled and lunged<br />

again; <strong>the</strong> puma retreated again, not as<br />

startled as it was <strong>the</strong> first time. Then <strong>the</strong><br />

puma took a last glance and turned into <strong>the</strong><br />

forest. The woman walked backwards<br />

awhile, <strong>the</strong>n turned around and ran (Personal<br />

correspondence to K. Etling on 4 Dec, 2001,<br />

from K. Hogland).<br />

In <strong>the</strong> second example, a puma<br />

confronted 2 biology students ga<strong>the</strong>ring data<br />

in Alum Rock Park, San Jose, California.<br />

The women yelled and made <strong>the</strong>mselves


look bigger, but <strong>the</strong> cat continued to<br />

advance. One <strong>of</strong> <strong>the</strong> women snarled like a<br />

dog and “mock lunged,” and <strong>the</strong> puma ran<br />

into some bushes. A nearby rancher<br />

approached on horseback, accompanied by 2<br />

dogs. When <strong>the</strong>y directed <strong>the</strong>ir attention<br />

toward where <strong>the</strong> girls thought <strong>the</strong> puma<br />

was, it bounded <strong>of</strong>f (Linda Lewis, web site:<br />

, citing personal communications<br />

with Jessie Dickson, April 18-19, 2001).<br />

In probably <strong>the</strong> most dramatic example<br />

demonstrating puma behavior following<br />

human aggressiveness, a deer hunter and a<br />

puma were stalking <strong>the</strong> same deer when <strong>the</strong><br />

deer detected <strong>the</strong> puma and fled. From 27<br />

m (30 yards) away, <strong>the</strong> puma transferred its<br />

stalk to <strong>the</strong> hunter. The hunter hid behind a<br />

tree while <strong>the</strong> puma approached, crouching.<br />

As <strong>the</strong> puma got close <strong>the</strong> hunter jumped out<br />

and yelled. That puma left running (Ford<br />

1994). The puma obviously knew <strong>the</strong> hunter<br />

was behind <strong>the</strong> tree, but <strong>the</strong> hunter’s actions<br />

probably appeared to <strong>the</strong> puma as an attack<br />

coming from a hidden (ambush) position.<br />

The action successfully interrupted <strong>the</strong><br />

predatory stalking behavior and instigated a<br />

flight behavior.<br />

Is It Safer to Hike in Groups?<br />

Solitary people are 3 times as likely to<br />

be attacked or to have an encounter as<br />

people in pairs or larger groups (Figure 5).<br />

However, only groups <strong>of</strong> 5 or more seem<br />

fairly secure against attack. We were much<br />

less likely to find data on non-attack<br />

encounters than on attacks. Thus, <strong>the</strong><br />

relative levels <strong>of</strong> <strong>the</strong> paired bars in Figure 5<br />

cannot be used to compare attack and nonattack<br />

encounters. We assumed that <strong>the</strong><br />

reporting rates for attacks and non-attacks<br />

are different but are not affected differently<br />

by group size. Figure 5 shows that <strong>the</strong><br />

relationship (but not absolute proportions) <strong>of</strong><br />

attack and non-attack encounters is similar,<br />

regardless <strong>of</strong> group size. The similar<br />

percentages within group size categories<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 99<br />

Figure 5. Relationship <strong>of</strong> human group size<br />

and age composition with type <strong>of</strong> encounter<br />

(n = 379).<br />

and <strong>the</strong> consistent pattern among <strong>the</strong>m<br />

indicates to us that <strong>the</strong> tendency <strong>of</strong> a puma<br />

to approach humans (or for humans to come<br />

close to pumas) is related to group size and<br />

is independent <strong>of</strong> whe<strong>the</strong>r an attack occurs.<br />

It seems to indicate that once a puma is in<br />

close proximity to humans, whe<strong>the</strong>r an<br />

attack occurs or not may be explained,<br />

statistically speaking, as a random or<br />

systematic decision, affecting all group sizes<br />

to <strong>the</strong> same extent once <strong>the</strong> initial approach<br />

is made. Such a mechanism could be<br />

created ei<strong>the</strong>r by <strong>the</strong><br />

physiological/behavioral state <strong>of</strong> <strong>the</strong> puma<br />

or size and behavior <strong>of</strong> <strong>the</strong> human(s), or<br />

both interacting. If we could detect and<br />

record non-attack encounters as thoroughly<br />

as we do attack encounters, we might be<br />

able to create more hypo<strong>the</strong>ses based on <strong>the</strong><br />

ratios and timing <strong>of</strong> one to <strong>the</strong> o<strong>the</strong>r. Data<br />

presented by Sweanor et al., at <strong>the</strong> <strong>Seventh</strong><br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, help<br />

considerably in this direction. We<br />

encourage all who study radio collared<br />

pumas to record and publish similar data.<br />

In Figure 5, adults strongly predominate<br />

in <strong>the</strong> single person attack and non-attack


100 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

Figure 6. Proportions <strong>of</strong> human age classes<br />

in different types <strong>of</strong> incidents, as affected by<br />

human group size (n = 379).<br />

categories and in <strong>the</strong> two-person non-attack<br />

category. This may reflect <strong>the</strong> relative use<br />

<strong>of</strong> wildland trails by single and paired adults<br />

versus children. However, <strong>the</strong> categories <strong>of</strong><br />

attacks on two people and on 3-5 people<br />

show increasing proportions <strong>of</strong> groups<br />

mostly composed <strong>of</strong> children, while <strong>the</strong> nonattack<br />

categories for <strong>the</strong> same size groups<br />

still show a predominance <strong>of</strong> groups mostly<br />

composed <strong>of</strong> adults. This indicates that,<br />

while attacks on groups <strong>of</strong> two or more<br />

people are much fewer than attacks on<br />

individuals, children in <strong>the</strong>se larger groups<br />

are at only slightly less risk than children<br />

alone.<br />

Ano<strong>the</strong>r way <strong>of</strong> viewing <strong>the</strong> same data is<br />

to scale each column separately and<br />

proportion <strong>the</strong> age groups within columns<br />

ra<strong>the</strong>r than being based on total incidents<br />

(Figure 6). When we do this, <strong>the</strong> increased<br />

proportion <strong>of</strong> attacks on mostly-children<br />

groups in <strong>the</strong> larger group sizes becomes<br />

more noticeable, although sample size is<br />

small (n = 16 for children in groups <strong>of</strong> >2<br />

people; n = 17 for adults in <strong>the</strong> same size<br />

groups). Among <strong>the</strong> 17 groups <strong>of</strong> 3 or more<br />

people composed mostly <strong>of</strong> adults, children<br />

were attacked in 11 <strong>of</strong> those groups. For<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

example, a group <strong>of</strong> 2 or 3 children (11-12<br />

years <strong>of</strong> age) plus 5-6 adults were on a<br />

kayak tour, camped on Compton Island in<br />

Johnstone Strait, V.I., B.C. recently. The<br />

group had just gotten out <strong>of</strong> <strong>the</strong> kayaks and<br />

was on <strong>the</strong> beach, standing in a group with<br />

<strong>the</strong> children in <strong>the</strong> middle, when a 100-lb.<br />

puma charged out <strong>of</strong> <strong>the</strong> brush, grabbed a<br />

12-year-girl and dragged her <strong>of</strong>f. Shouting<br />

and noise made by <strong>the</strong> adults made <strong>the</strong> cat<br />

drop <strong>the</strong> girl, who survived <strong>the</strong> incident<br />

(Robert Hansen, Pacific Rim National Park,<br />

Vancouver Island, B.C., personal<br />

communication, 8 May, 2003).<br />

Is It Safe to Sleep on <strong>the</strong> Ground?<br />

We know <strong>of</strong> 12 victims who were<br />

accosted in <strong>the</strong>ir sleep (Figure 7). Only one,<br />

a man in Argentina in 1898, was seriously<br />

injured, and details <strong>of</strong> that account are<br />

lacking (Pritchard 1902; Roosevelt 1914:29;<br />

Barnes 1960:119,122). Seven victims were<br />

uninjured, and 4 o<strong>the</strong>rs suffered minor<br />

injuries. For example, a boy was sleeping<br />

on a mat at night when a puma walked up,<br />

pawed him, and tried to drag <strong>the</strong> mat away.<br />

His fa<strong>the</strong>r rescued him. The boy suffered<br />

scratches and had 10 stitches in his ear.<br />

(Phoenix Gazette and Tucson Citizen 1994;<br />

Danz 1999: 276-277; personal<br />

Figure 7. Injuries that occurred when victim<br />

was attacked while sleeping (n = 12).


communication with Kevin Bergersen,<br />

Arizona Game and Fish Department, 15 Jul<br />

2002.)<br />

We speculate that in most <strong>of</strong> <strong>the</strong>se<br />

situations <strong>the</strong> innate attack behavior ei<strong>the</strong>r<br />

never was initiated, or was satisfied when<br />

<strong>the</strong> puma determined that <strong>the</strong> prey already<br />

was moribund, and only needed dragging to<br />

a protected location to be fed upon. The<br />

puma simply examines <strong>the</strong> sleeper, and <strong>the</strong>n<br />

drags <strong>the</strong> sleeping bag or mat away from <strong>the</strong><br />

site. When <strong>the</strong> victim or companions awake<br />

and begin making noise, <strong>the</strong> puma is<br />

frightened and leaves. Perhaps <strong>the</strong>re is an<br />

element <strong>of</strong> surprise or ambush by humans;<br />

perhaps <strong>the</strong> attack behavior is not initiated<br />

and <strong>the</strong>refore does not need to be fulfilled.<br />

We found, in addition to <strong>the</strong> 12 mentioned<br />

above, several oral history accounts (not<br />

included in our data)<strong>of</strong> pumas covering<br />

sleeping people with debris (e.g., Seton<br />

1929:98-99).<br />

Are You Safer Riding Horseback?<br />

Of 9 incidents involving riders on<br />

horseback, none was seriously injured and<br />

most were unharmed. In 8 <strong>of</strong> <strong>the</strong>se cases,<br />

<strong>the</strong> puma failed in <strong>the</strong> initial attack and <strong>the</strong><br />

horse outran it. In <strong>the</strong> o<strong>the</strong>r case, <strong>the</strong> rider<br />

killed <strong>the</strong> puma. If <strong>the</strong> rider dismounted,<br />

accidentally or o<strong>the</strong>rwise, <strong>the</strong> situation<br />

became more similar to a puma attacking a<br />

person afoot.<br />

Risk <strong>of</strong> Attacks<br />

Puma attacks on humans are rare by<br />

almost any measure. Bear attacks are much<br />

more common. But rarity may be a matter<br />

<strong>of</strong> scale, proximity and individuality. Risk<br />

depends on where people are and what <strong>the</strong><br />

conditions are in that area. Statistics that<br />

compare <strong>the</strong> frequency <strong>of</strong> puma attacks with<br />

that <strong>of</strong> some o<strong>the</strong>r phenomenon, such as<br />

lightning strikes, are misleading. Such data<br />

normally fail to account for risk levels. For<br />

example, <strong>the</strong> risk <strong>of</strong> being attacked by a wild<br />

puma in a large metropolitan area is close to<br />

zero, but <strong>the</strong> large urban areas include most<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 101<br />

<strong>of</strong> <strong>the</strong> people who make up <strong>the</strong> denominator<br />

in <strong>the</strong> attack rate. The risk <strong>of</strong> being attacked<br />

in a puma’s natural territory is higher, only<br />

because <strong>the</strong> puma may be present. The<br />

same logic applies to bites by domestic<br />

dogs. Mail carriers are more at risk from<br />

dog bites than are o<strong>the</strong>r people because <strong>the</strong>y<br />

periodically enter yards that <strong>the</strong> dogs<br />

consider to be <strong>the</strong>ir territory. In fact, <strong>the</strong><br />

mail carriers’ risk level increases<br />

dramatically <strong>the</strong> instant <strong>the</strong>y enter a yard<br />

containing a dog.<br />

The situation is similar with puma<br />

attacks. Most <strong>of</strong> us have almost no risk, but<br />

under certain conditions <strong>the</strong> risk rises<br />

considerably. Data presented by Mattson et<br />

al, at <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>,<br />

was an effort to predict conceptually some<br />

<strong>of</strong> <strong>the</strong>se risk factors. Our presentation, on<br />

<strong>the</strong> o<strong>the</strong>r hand, is mostly an effort to help<br />

people manage an attack situation. Data<br />

presented by Sweanor et al., at <strong>the</strong> <strong>Seventh</strong><br />

<strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>, is a strong<br />

beginning toward collecting data to identify<br />

risk factors.<br />

A person in an encounter should always<br />

judge <strong>the</strong> puma’s responses and adjust<br />

defensive measures according to <strong>the</strong> puma<br />

response. Each puma has its own<br />

personality, as <strong>the</strong> following incident<br />

illustrates. A United States Forest Service<br />

employee was in a remote area away from<br />

roads doing silvicultural analysis. While in<br />

her sleeping bag in camp she noticed a puma<br />

at her feet near <strong>the</strong> edge <strong>of</strong> her sleeping bag.<br />

She calmly said, ”excuse me,” whereupon<br />

<strong>the</strong> puma moved 4.6 m (15 feet) away and<br />

sat. It seemed to be a healthy, curious puma.<br />

If she got upset and tried to make it leave, it<br />

got aggressive; if she lay <strong>the</strong>re quietly, it<br />

calmed down, reclined, and licked itself. If<br />

she yelled at it, it would pin its ears and<br />

charge. The puma started to circle her, so<br />

she threw sticks and rocks at it to keep it at<br />

bay and lit <strong>the</strong> fire. She called in coordinates<br />

via radio to her supervisor. The puma


102 REDUCING PUMA ATTACKS · Fitzhugh et al.<br />

watched all night and finally left <strong>the</strong> area as<br />

rescuers arrived. (Personal communication<br />

from Kathleen Kavalok to K. Etling).<br />

The incident above illustrates <strong>the</strong><br />

presence <strong>of</strong> mind needed when confronted<br />

by a puma. This person did everything just<br />

right, but <strong>the</strong> puma responded almost <strong>the</strong><br />

opposite way from what was expected. We<br />

believe <strong>the</strong> information we have provided is<br />

correct in a statistical sense. It is very<br />

important to remember what to do, but also<br />

be prepared to adapt to <strong>the</strong> puma’s behavior.<br />

With luck and aggressiveness on your part,<br />

you may avoid an attack and also teach <strong>the</strong><br />

puma that humans may not be food after all.<br />

ACKNOWLEDGEMENTS<br />

We thank S. C. Reed for helping build<br />

<strong>the</strong> spreadsheet, verify data, review files,<br />

and enter new data. J. Schmidt helped with<br />

<strong>the</strong> original compilation in 1985-86. R.W.<br />

Riley, K. J. Stahle, S. E. Gordon, M. A.<br />

Whittaker, B. R. Campos, M. E. Jackson, E.<br />

Chen, A. M. White helped with data<br />

organization and entry. E. L. Blake<br />

provided some new ideas for analysis.<br />

Details <strong>of</strong> incidents were provided by P.<br />

Swift, California Department <strong>of</strong> Fish and<br />

Game, R. Beausoleil, formerly <strong>of</strong> <strong>the</strong> New<br />

Mexico Department <strong>of</strong> Game and Fish, K.<br />

Bergersen, Arizona Game and Fish<br />

Department, T. R. Collom, W. Castillo, and<br />

D. Whittaker, <strong>of</strong> <strong>the</strong> Oregon Department <strong>of</strong><br />

Fish and Wildlife, M. Austin and B. Guiltner<br />

<strong>of</strong> <strong>the</strong> B.C. Ministry <strong>of</strong> Environment, M.<br />

Gillett, R. Skiles E. Myers, A. Davis, J.<br />

Case, K. McKinlay-Jones, <strong>of</strong> <strong>the</strong> National<br />

Park Service, A. Barton, M. Shuey, L.<br />

Lewis, and S. Galentine. We made<br />

extensive use <strong>of</strong> new information in books<br />

by K. Etling (2001), by J. Deurbrock and D.<br />

Miller (2001), and by H. P. Danz (1999).<br />

We thank <strong>the</strong> many who have provided<br />

information about <strong>the</strong>ir personal<br />

experiences, ei<strong>the</strong>r to us or to previous<br />

authors. P. Beier (1991) provided <strong>the</strong> first<br />

scientific analysis <strong>of</strong> puma attack data, and<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

inspired us to continue our work. Finally,<br />

<strong>the</strong> very foundation, without which this<br />

effort may not have ever begun, is C. T.<br />

Barnes’ (1960) book. His book stimulated,<br />

and to a large extent, enabled this project<br />

during its infancy. R. G. Coss, H. G. Shaw,<br />

L. L. Sweanor, W. F. Laudenslayer, W. E.<br />

Howard, and R. E. Marsh improved <strong>the</strong><br />

paper with reviews <strong>of</strong> <strong>the</strong> draft manuscript,<br />

but errors remain <strong>the</strong> author’s responsibility.<br />

LITERATURE CITED<br />

ANONYMOUS. 1917. Untitled comment by<br />

editor following Marsh 1917. Outdoor<br />

Life 38:194.<br />

ANONYMOUS. 1925. An old question.<br />

Outdoor Life 46:113.<br />

ANONYMOUS. 1994. Phoenix Gazette and<br />

Tucson Citizen. 20 July 1994.<br />

ANONYMOUS. 2002. The New York Times.<br />

12 November 2002.<br />

BARNES, CLAUDE T. 1960. "The Cougar or<br />

<strong>Mountain</strong> <strong>Lion</strong>." The Ralton Co. Salt<br />

Lake City, Utah, USA. 175 pp.<br />

BEIER, P. 1991. Cougar attacks on humans<br />

in <strong>the</strong> United States and Canada.<br />

Wildlife Society Bulletin 19:403-412.<br />

COLORADO DIVISION OF WILDLIFE. 2002.<br />

News Report. 8 January 2002.<br />

DANZ, HAROLD P. 1999. Cougar! Swallow<br />

Press/Ohio University Press. A<strong>the</strong>ns,<br />

Ohio, USA. 310 pp.<br />

DEURBROCK, JO AND DEAN MILLER. 2001.<br />

Cat attacks: true stories and hard lessons<br />

from cougar country. Sasquatch Books.<br />

Seattle, Washington, USA. 221 pp.<br />

ETLING, KATHY. 2001. Cougar attacks:<br />

encounters <strong>of</strong> <strong>the</strong> worst kind. The Lyons<br />

Press/Globe Pequot Press. Guilford,<br />

Connecticut, USA. 246 pp.<br />

FITZHUGH, E. LEE AND W.P. GORENZEL.<br />

1986. Biological status <strong>of</strong> mountain<br />

lions in California. pp. 336-346 in<br />

<strong>Proceedings</strong>, Twelfth Vertebrate Pest<br />

Conference. T. P. Salmon (ed.).<br />

Vertebrate Pest Council <strong>of</strong> <strong>the</strong>


Vertebrate Pest Conference. University<br />

<strong>of</strong> California, Davis, California, USA.<br />

FORD, PHIL. 1994. Burney bowhunter<br />

stalked by fearless mountain lion.<br />

Fishing and Hunting News. September<br />

15-29, 1994.<br />

HUNTER, J. MARVIN. 1922. Pioneer history<br />

<strong>of</strong> Bandera County; seventy-five years <strong>of</strong><br />

intrepid history. Hunter’s Printing<br />

House. Bandera, Texas, USA. 241 pp.<br />

MARSH, CHARLES. 1917. Children attacked<br />

by cougar. Outdoor Life 38:193-194.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

REDUCING PUMA ATTACKS · Fitzhugh et al. 103<br />

PRITCHARD, C. HESKETH. 1902. Through<br />

<strong>the</strong> heart <strong>of</strong> Patagonia. New York, USA.<br />

346 pp.<br />

ROOSEVELT, THEODORE. 1914:29. Through<br />

<strong>the</strong> Brazilian wilderness. New York,<br />

USA.<br />

SETON, E.T. 1929. Lives <strong>of</strong> game animals.<br />

Doubleday, Doran & Company, Inc.<br />

Garden City, New York, USA. Vol I,<br />

Part I.


104<br />

A CONCEPTUAL MODEL AND APPRAISAL OF EXISTING RESEARCH RELATED<br />

TO INTERACTIONS BETWEEN HUMANS AND PUMAS<br />

DAVID J. MATTSON, USGS Southwest Biological Science Center, Colorado Plateau Field<br />

Station, P.O. Box 5614, Nor<strong>the</strong>rn Arizona University, Flagstaff, AZ 86011-5614, USA,<br />

email: David.Mattson@nau.edu<br />

JAN V. HART, USGS Southwest Biological Science Center, Colorado Plateau Field Station,<br />

P.O. Box 5614, Nor<strong>the</strong>rn Arizona University, Flagstaff, AZ 86011-5614, USA, email:<br />

Jan.Hart@nau.edu<br />

PAUL BEIER, School <strong>of</strong> Forestry, Nor<strong>the</strong>rn Arizona University, Flagstaff, AZ 86011-5018,<br />

USA, email: Paul.Beier@nau.edu<br />

JESSE MILLEN-JOHNSON, Bates College, Lewiston, ME 04240, USA, email:<br />

jmillenj@bates.edu<br />

Abstract: Recorded encounters between humans and pumas have been increasing throughout <strong>the</strong><br />

western contiguous U.S., as have puma-caused human injuries and deaths. We developed a<br />

conceptual model <strong>of</strong> interactions between humans and pumas to aid <strong>the</strong> design <strong>of</strong> a study in <strong>the</strong><br />

Flagstaff uplands <strong>of</strong> Arizona, USA, and to appraise <strong>the</strong> scope and strength <strong>of</strong> existing related<br />

research. The model represents contact and resulting human injuries as <strong>the</strong> outcome <strong>of</strong> 2<br />

processes: (1) <strong>the</strong> frequency <strong>of</strong> encounter between humans and pumas, and (2), given an<br />

encounter, <strong>the</strong> probability that it will turn injurious to a human. Conceptually, different suites <strong>of</strong><br />

factors govern <strong>the</strong>se 2 phenomena. The model representing frequency <strong>of</strong> encounter includes 15<br />

putative explanatory variables, 7 <strong>of</strong> which relate directly to pumas. The model representing<br />

probability <strong>of</strong> injury also includes 15 explanatory variables, 7 <strong>of</strong> which pertain directly to pumas.<br />

The remaining variables in both models relate directly or indirectly to human presence or<br />

behavior. Of <strong>the</strong> 44 identified relations among <strong>the</strong>se variables, 6 have been well studied and an<br />

additional 18 have been subject to some level <strong>of</strong> systematic analysis. The remaining 20<br />

relations, including many plausibly critical ones, are currently informed only by speculation,<br />

anecdote, or deduction. Much research yet needs to be done before <strong>the</strong> level and nature <strong>of</strong><br />

contact between humans and pumas can be adequately explained and predicted. Moreover,<br />

much <strong>of</strong> this additional research needs to address human behavior and factors related to<br />

distributions and numbers <strong>of</strong> humans. Of <strong>the</strong> uninvestigated factors with plausibly major effects,<br />

habituation <strong>of</strong> cougars promises to be <strong>the</strong> most difficult to study. O<strong>the</strong>rwise, numbers and<br />

distributions <strong>of</strong> human facilities (including roads and trails), puma population sizes, human<br />

behavior, and human knowledge <strong>of</strong> pumas are potentially important explanatory factors<br />

amenable to inquiry. We argue that all <strong>of</strong> <strong>the</strong>se putative effects should be considered before<br />

reaching conclusions about <strong>the</strong> “causes” <strong>of</strong> human-puma encounters and puma-caused human<br />

injuries, whe<strong>the</strong>r for a region or a given study area.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


RELATIONSHIPS BETWEEN LAND TENURE SYSTEM, MOUNTAIN LION<br />

PROTECTION STATUS, AND LIVESTOCK DEPREDATION RATE<br />

MARCELO MAZZOLLI, Projeto Puma - R. Liberato Carioni 24, Lagoa, 88062-005,<br />

Florianópolis – SC, Brazil, email: marcelo_puma@yahoo.com<br />

Abstract: <strong>Mountain</strong> lion depredation impact on managed livestock in ranching-dominated<br />

landscapes was compared with depredation in forestry-dominated landscapes. In forestrydominated<br />

landscapes, twenty-one depredation incidents were recorded at three ranches,<br />

resulting in <strong>the</strong> loss <strong>of</strong> 58 sheep and goat. An additional 11 head were killed during an unknown<br />

number <strong>of</strong> attacks. These losses amounted to 14 to 60 percent <strong>of</strong> total stock per year (in number<br />

<strong>of</strong> animals). Confining or corralling flocks during <strong>the</strong> night at first provoked a reduction <strong>of</strong><br />

livestock depredation, but subsequently depredation begun to occur during daylight. In <strong>the</strong><br />

ranching-dominated landscape, on <strong>the</strong> o<strong>the</strong>r hand, depredation losses were not reported on flocks<br />

corralled during <strong>the</strong> night, but reached 85% <strong>of</strong> <strong>the</strong> total stock when free-ranging. I hypo<strong>the</strong>size,<br />

based on field data and on available information in literature, that mountain lion depredation in<br />

forestry areas during <strong>the</strong> day and higher depredation on corralled livestock during <strong>the</strong> night may<br />

result from lower hunting pressure on mountain lions than in ranching areas. Understanding<br />

mountain lion predation behavior may help to modify livestock husbandry, allowing wildlife<br />

managers and ranchers to minimize depredation without direct persecution <strong>of</strong> mountain lions.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

105


106<br />

MOUNTAIN LION MOVEMENTS AND PERSISTENCE IN A FRAGMENTED, URBAN<br />

LANDSCAPE IN SOUTHERN CALIFORNIA<br />

SETH P.D. RILEY, Santa Monica <strong>Mountain</strong> National Recreation Area, 401 W. Hillcrest Dr.,<br />

Thousand Oaks, CA 91360, USA, email: seth_riley@nps.gov<br />

RAYMOND M. SAUVAJOT, Santa Monica <strong>Mountain</strong> National Recreation Area, 401 W.<br />

Hillcrest Dr., Thousand Oaks, CA 91360, USA<br />

ERIC C. YORK, Santa Monica <strong>Mountain</strong> National Recreation Area, 401 W. Hillcrest Dr.,<br />

Thousand Oaks, CA 91360, USA, email: eric_york@nps.gov<br />

Abstract: As natural habitat is increasingly eliminated and fragmented by human land uses <strong>the</strong><br />

long-term prospects for conservation <strong>of</strong> carnivore populations become correspondingly worse.<br />

This is especially true for larger carnivores such as mountain lions, which require significant<br />

amounts <strong>of</strong> both space and prey. In rapidly urbanizing sou<strong>the</strong>rn California, conservation <strong>of</strong><br />

carnivores in general, and <strong>of</strong> mountain lions in particular, is particularly challenging. In <strong>the</strong><br />

Santa Monica <strong>Mountain</strong>s and surrounding areas, we have begun a project using GPS collars to<br />

determine mountain lion movement and space use in a fragmented landscape. Our goal is to<br />

determine whe<strong>the</strong>r lions are successfully traversing freeways and o<strong>the</strong>r human-made barriers<br />

between large areas <strong>of</strong> natural habitat. Ultimately, we hope to determine whe<strong>the</strong>r enough natural<br />

habitat can be preserved, and enough connectivity maintained between core habitat areas, to<br />

maintain lion populations in such a landscape. We have collared lions already in <strong>the</strong> Santa<br />

Monica <strong>Mountain</strong>s, and determined that one large male is using <strong>the</strong> entire mountain range (home<br />

range <strong>of</strong> 394 km 2 ), from a major freeway to <strong>the</strong> east to a developed agricultural valley to <strong>the</strong><br />

west, and from <strong>the</strong> Pacific Ocean to <strong>the</strong> south to a major freeway to <strong>the</strong> north. Given <strong>the</strong> small<br />

number <strong>of</strong> lions likely persisting in <strong>the</strong> Santa Monica <strong>Mountain</strong>s, connectivity is as important, if<br />

not more important, than we anticipated. We continue to collar o<strong>the</strong>r lions in <strong>the</strong> study region to<br />

evaluate whe<strong>the</strong>r any exchange occurs across barriers created by freeways and urban<br />

development. While both <strong>the</strong> male and <strong>the</strong> female in <strong>the</strong> <strong>Mountain</strong>s have approached <strong>the</strong><br />

freeway to <strong>the</strong> north, nei<strong>the</strong>r one has crossed it in <strong>the</strong> 9-12 months that we have been following<br />

<strong>the</strong>m. We are also investigating kill sites to determine kill rates, species <strong>of</strong> kills, and whe<strong>the</strong>r<br />

lions are preying on any domestic animals. So far in <strong>the</strong> Santa Monica <strong>Mountain</strong>s our data<br />

indicate that lions are killing 3-4 deer/month <strong>of</strong> all different age/sex classes, and an occasional<br />

coyote or raccoon. The collared animals are almost never seen by anyone, including <strong>the</strong><br />

researchers tracking <strong>the</strong>m, even though <strong>the</strong>y cross numerous roads and trails and sometimes<br />

venture close to residential areas.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


PUMA RESPONSES TO CLOSE ENCOUNTERS WITH RESEARCHERS<br />

LINDA L. SWEANOR, Wildlife Research Biologist, Wildlife Health Center, University <strong>of</strong><br />

California at Davis, Sou<strong>the</strong>rn California Puma Project Field Station, P.O. Box 1114,<br />

Julian, CA 92036-1114, USA, email: lsweanor@mindspring.com<br />

KENNETH A. LOGAN, Wildlife Research Biologist, Wildlife Health Center, University <strong>of</strong><br />

California at Davis, Sou<strong>the</strong>rn California Puma Project Field Station, P.O. Box 1114,<br />

Julian, CA 92036-1114, USA<br />

MAURICE G. HORNOCKER, Senior Scientist, Wildlife Conservation Society, Box 929,<br />

Bellevue, ID 83313, USA<br />

Abstract: Recent books and articles have provided information on relatively rare, but violent<br />

attacks where people were injured or killed by pumas. However, <strong>the</strong>re is a paucity <strong>of</strong> data on <strong>the</strong><br />

type and variation in behavior wild pumas exhibit when approached by humans. During a 10year<br />

puma study in New Mexico, we approached pumas and visually observed <strong>the</strong>ir behavior on<br />

262 occasions. The study area was remote and was closed to most human activity; consequently<br />

<strong>the</strong> pumas living <strong>the</strong>re had rare opportunities for contact with people. We categorized <strong>the</strong><br />

approach based on <strong>the</strong> status <strong>of</strong> <strong>the</strong> puma, <strong>the</strong> number <strong>of</strong> people involved, <strong>the</strong> distance between<br />

<strong>the</strong> puma(s) and people, and <strong>the</strong> puma’s subsequent response to <strong>the</strong> approach. Pumas we<br />

approached included adult females with nursing (


108<br />

PRELIMINARY RESULTS OF FLORIDA PANTHER GENETIC ANALYSES<br />

WARREN E. JOHNSON, National Cancer Institute, Frederick, MD 21702-1201, USA, email:<br />

johnsonw@ncifcrf.gov<br />

DARRELL LAND, Florida Fish and Wildlife Conservation Commission, 566 Commercial<br />

Blvd., Naples, FL 34104, USA, email: darrell.land@fwc.state.fl.us<br />

JAN MORTENSON, National Cancer Institute, Frederick, MD 21702-1201, USA, email:<br />

martenoj@ncifcrf.gov<br />

MELODY ROELKE-PARKER, National Cancer Institute, Frederick, MD 21702-1201, USA,<br />

email: roelkem@ncifcrf.gov<br />

STEPHEN J. O’BRIEN, National Cancer Institute, Frederick, MD 21702-1201, USA, email:<br />

obriens@ncifcrf.gov<br />

Abstract: Previous genetic analyses showed that Florida pan<strong>the</strong>rs (Puma concolor coryi) had <strong>the</strong><br />

lowest genetic diversity among all North American puma and subsequent modeling suggested<br />

that fur<strong>the</strong>r declines could increase <strong>the</strong> probability <strong>of</strong> extinction. Currently, <strong>the</strong>re are fewer than<br />

100 pan<strong>the</strong>rs in south Florida. Although on-going habitat conservation strategies may provide<br />

long-term stability for today’s population extents, <strong>the</strong>se same strategies are unlikely to allow <strong>the</strong><br />

population to grow to 500 or more individuals whereby genetic viability is more assured. As a<br />

result, a plan for Florida pan<strong>the</strong>r genetic restoration was created in 1994 and implementation<br />

began in <strong>the</strong> spring <strong>of</strong> 1995 with <strong>the</strong> release <strong>of</strong> 8 female Texas puma into areas occupied by<br />

pan<strong>the</strong>rs. Our objectives were to monitor <strong>the</strong> effectiveness <strong>of</strong> genetic restoration by developing<br />

an array <strong>of</strong> molecular genetic markers that characterized <strong>the</strong> status <strong>of</strong> current and past<br />

populations, to construct a pedigree among Florida pan<strong>the</strong>rs to follow inheritance patterns, to<br />

infer degrees <strong>of</strong> relatedness among individuals, and to help predict <strong>the</strong> future viability <strong>of</strong> <strong>the</strong><br />

population. We have completed genotyping over 175 samples from Florida pan<strong>the</strong>rs at 23<br />

microsatellite loci and <strong>the</strong>se included individuals from canonical Florida pan<strong>the</strong>rs, <strong>the</strong><br />

Everglades subpopulation (Piper stock), released Texas puma, crosses among all stocks, and<br />

captive animals <strong>of</strong> unknown ancestry from <strong>the</strong> early 1970’s to <strong>the</strong> present. Genetic restoration<br />

has increased hetereozygocity within <strong>the</strong> population, but we have documented <strong>the</strong> loss <strong>of</strong> some<br />

pan<strong>the</strong>r matrilines. Certain morphological traits such as cryptorchidism, kinked tails, cowlicks,<br />

and atrial septal defects observed in canonical pan<strong>the</strong>rs are not present in <strong>the</strong> Texas puma<br />

descendants. We have identified several subgroups within our population and <strong>the</strong>se subgroups<br />

seem to be partially <strong>the</strong> product <strong>of</strong> philopatric tendencies among dispersing female <strong>of</strong>fspring.<br />

Male pan<strong>the</strong>rs may be physically and behaviorally capable <strong>of</strong> siring <strong>of</strong>fspring earlier than<br />

suggested by radiotelemetry work and resident and resident males are not siring all litters with<br />

females within <strong>the</strong> respective males’ home ranges. Intraspecific aggression, a common mortality<br />

agent for young male pan<strong>the</strong>rs, may not be removing pan<strong>the</strong>rs prior to producing <strong>of</strong>fspring.<br />

Future monitoring should ensure sampling across all pan<strong>the</strong>r subgroups in order to adequately<br />

estimate total population genetic characteristics.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


GENETIC STRUCTURE OF COUGAR POPULATIONS ACROSS THE WYOMING<br />

BASIN: METAPOPULATION OR MEGAPOPULATION<br />

CHUCK R. ANDERSON, JR., Zoology and Physiology Department, University <strong>of</strong> Wyoming,<br />

Box 3166, University Station, Laramie, WY 82071, USA, email: cander@uwyo.edu<br />

FRED G. LINDZEY, Wyoming Cooperative Fish and Wildlife Research Unit, Box 3166,<br />

University Station, Laramie, WY 82071, USA, email: flindzey@uwyo.edu<br />

DAVE B. McDONALD, Zoology and Physiology Department, University <strong>of</strong> Wyoming,<br />

Bioscience Room 413, University Station, Laramie, WY 82071, USA, email:<br />

dbmcd@uwyo.edu<br />

Abstract: Using microsatellite DNA analyses at 9 loci, we examined genetic structure <strong>of</strong> 5<br />

geographically distinct cougar (Puma concolor) populations separated by <strong>the</strong> Wyoming Basin<br />

and a distant cougar population from southwest Colorado. Observed heterozygosity was similar<br />

among populations (Hobs = 0.49-0.59) and intermediate to that <strong>of</strong> o<strong>the</strong>r large carnivores.<br />

Estimates <strong>of</strong> genetic structure (FST = 0.029, RST = 0.028) and number <strong>of</strong> migrants per generation<br />

(Nem) suggested high gene flow across <strong>the</strong> central Rocky <strong>Mountain</strong>s. Estimates <strong>of</strong> <strong>the</strong> number <strong>of</strong><br />

migrants per generation were lowest between <strong>the</strong> southwest Colorado cougar population and<br />

cougar populations north <strong>of</strong> <strong>the</strong> Wyoming Basin (northwest WY, north-central WY, and <strong>the</strong><br />

Black Hills, SD; Nem = 2.9-3.0) and highest among cougar populations from adjacent mountain<br />

ranges (Nem = 10.2-30.2), suggesting an effect <strong>of</strong> both isolation by distance and <strong>of</strong> habitat<br />

matrix. We applied a model-based clustering method to infer population structure from<br />

individual genotypes and noted that both males and females from throughout <strong>the</strong> region were<br />

best described as a single panmictic population. Estimates <strong>of</strong> relatedness (rxy) did not differ (P ><br />

0.05) between males and females. Estimated relative effective population size did not differ<br />

significantly among populations (P > 0.05), but <strong>the</strong> higher estimates were from contiguous<br />

mountain ranges (i.e., northwest WY, southwest WY, and southwest CO) and lower estimates<br />

were from less contiguous terminal mountain ranges (i.e., north-central WY and Snowy Range<br />

WY). Based on measures <strong>of</strong> gene flow we examined, extinction risk in <strong>the</strong> near future appears<br />

extremely low, even for <strong>the</strong> relatively isolated Black Hills cougar population. Cougars in <strong>the</strong><br />

central Rocky <strong>Mountain</strong>s appear to constitute a large panmictic population ra<strong>the</strong>r than a<br />

metapopulation. Estimated effective population size for cougars in <strong>the</strong> central Rocky <strong>Mountain</strong>s<br />

ranged from 1,797 to 4,532 depending on analysis method and <strong>the</strong> mutation model assumed.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

109


110<br />

ECOLOGICAL SIGNIFICANCE AND EVOLUTION OF A COMMON COUGAR<br />

RETROVIRUS<br />

ROMAN BIEK, Fish and Wildlife Biology Program, University <strong>of</strong> Montana, Missoula, MT<br />

59812, USA, email: rbiek@selway.umt.edu<br />

MARY POSS, Fish and Wildlife Biology Program and Division <strong>of</strong> Biological Sciences,<br />

University <strong>of</strong> Montana, Missoula, MT 59812, USA, email: mposs@selway.umt.edu<br />

Abstract: As for most wildlife species, little is known about <strong>the</strong> organisms that infect cougars in<br />

<strong>the</strong> wild. In an ongoing project, we are studying a retrovirus related to domestic cat-FIV that is<br />

naturally found in North American cougars with <strong>the</strong> aim <strong>of</strong> assessing <strong>the</strong> virus’ possible<br />

demographic consequences on <strong>the</strong> cougar host as well as its epidemiology and short-term<br />

evolution. Tests for possible effects on survival and reproduction as well as secondary exposure<br />

to o<strong>the</strong>r pathogens in infected individuals are conducted based on a large data set compiled from<br />

several intensively studied cougar populations. In addition, DNA sequences <strong>of</strong> virus obtained<br />

from infected individuals are used to determine <strong>the</strong> genetic population structure <strong>of</strong> cougar-FIV in<br />

<strong>the</strong> Rocky <strong>Mountain</strong> region. We determined that that this virus is changing its genetic<br />

composition within a matter <strong>of</strong> decades. Because restrictions <strong>of</strong> cougar movement will be<br />

reflected in <strong>the</strong> distribution <strong>of</strong> closely related viruses, distributional data for <strong>the</strong> virus are thus<br />

likely to contain information about current patterns <strong>of</strong> connectivity among cougar populations. A<br />

preliminary analysis <strong>of</strong> <strong>the</strong>se data indicates that spread <strong>of</strong> <strong>the</strong> virus occurs mainly locally but also<br />

showed evidence for recent transmission events over distances > 300 km. These results show that<br />

studying <strong>the</strong> ecology <strong>of</strong> cougar-FIV can provide important insights into <strong>the</strong> ecology <strong>of</strong> <strong>the</strong> cougar<br />

host even beyond immediate disease impacts.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


CHARACTERISTICS OF MOUNTAIN LION BED, CACHE AND KILL SITES IN<br />

NORTHEASTERN OREGON<br />

JAMES J. AKENSON, Taylor Ranch Field Station, HC 83 Box 8070, Cascade ID 83611, USA,<br />

email: tayranch@directpc.com<br />

M. CATHY NOWAK, Cat Tracks Wildlife Consulting, P.O. Box 195, Union, OR 97883, USA,<br />

email: mcnowak@eoni.com<br />

MARK G. HENJUM, Oregon Department <strong>of</strong> Fish and Wildlife, 107 20 th Street, La Grande, OR<br />

97850, USA<br />

GARY W. WITMER, USDA APHIS National Wildlife Research Center, 4101 La Porte Avenue,<br />

Fort Collins, CO 80521, USA<br />

Abstract: We described mountain lion (Puma concolor) habitat characteristics during two studies in <strong>the</strong> same area <strong>of</strong><br />

nor<strong>the</strong>astern Oregon during <strong>the</strong> 1990s. In <strong>the</strong> first study (1992-1994) we evaluated micro-habitat features associated<br />

with 61 diurnal bed sites that were not associated with kills. We used similar techniques in <strong>the</strong> second study (1996-<br />

1998) to evaluate habitat features at 79 cache sites near lion-killed prey. A dog was used to find 93% <strong>of</strong> <strong>the</strong> diurnal<br />

bed sites. Radio telemetry triangulation was used in <strong>the</strong> second study. Characteristics <strong>of</strong> diurnal bed sites and cache<br />

sites were compared with random habitat plots. Rock structure and downed logs were identified as important habitat<br />

components at diurnal bed sites. Canopy cover at cache sites was significantly higher than at random sites. Cache<br />

sites also were associated with rock structure, but not to <strong>the</strong> same degree as diurnal bed sites. In both studies<br />

mountain lions used sites in close proximity to habitat edges more frequently than expected based on random plots.<br />

Understanding <strong>the</strong> similarities and differences <strong>of</strong> habitat use at diurnal bed, cache and kill sites sheds light on <strong>the</strong><br />

ecological adaptation <strong>of</strong> mountain lions to <strong>the</strong> multiple environmental influences and disturbances <strong>of</strong> managed<br />

forests.<br />

111<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Key words: Puma concolor, microhabitat use, diurnal bed site, cache site, kill site, habitat edge, forest management<br />

<strong>Mountain</strong> lion recovery has been one <strong>of</strong><br />

<strong>the</strong> great wildlife conservation success<br />

stories <strong>of</strong> <strong>the</strong> 20 th century. As we move into<br />

<strong>the</strong> 21 st century, <strong>the</strong> challenges for mountain<br />

lion conservation are less related to species<br />

persecution, and more related to concerns<br />

with habitat fragmentation and issues <strong>of</strong><br />

human-lion coexistence on <strong>the</strong> expanding<br />

fringe <strong>of</strong> urbanization. The interface<br />

between human resource development and<br />

mountain lion habitat use has persisted for<br />

centuries in North America. Historically,<br />

mountain lions have occupied most habitats<br />

occurring on this continent. <strong>Mountain</strong> lions<br />

have typically been associated with <strong>the</strong><br />

rugged, rocky, forested terrain <strong>of</strong> <strong>the</strong> Rocky<br />

<strong>Mountain</strong>s in <strong>the</strong> western United States;<br />

however, this species is so adaptable it can<br />

thrive in deserts, swamps, tropical jungles,<br />

and sub-alpine forests (Hornocker 1976).<br />

The lion has come into conflict with humans<br />

on several fronts. In <strong>the</strong> past, <strong>the</strong> majority<br />

<strong>of</strong> interactions between humans and<br />

mountain lions were associated with<br />

settlement and agricultural practices (Young<br />

1946). With increasing human population<br />

and urban sprawl, <strong>the</strong> zone <strong>of</strong> conflict has<br />

shifted to <strong>the</strong> urban-wildland interface<br />

(Beier 1995).<br />

Habitat fragmentation can take a more<br />

subtle form than <strong>the</strong> direct effect imparted<br />

by urbanization. Across much <strong>of</strong> <strong>the</strong><br />

mountain lion’s range, logging has occurred<br />

at various intensities. Studies in Utah and<br />

Arizona, found that mountain lions ei<strong>the</strong>r<br />

avoided active timber sale areas (Van Dyke


112 MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al.<br />

et al. 1986) or adjusted <strong>the</strong>ir activity pattern<br />

from <strong>the</strong> norm (Ackerman 1982), to<br />

maximize night-time concealment from<br />

human contact. Timber sale size, relative to<br />

a resident mountain lion’s home range, was<br />

a big factor on <strong>the</strong> degree <strong>of</strong> disturbance and<br />

influence on a lion’s willingness to maintain<br />

its home range (Van Dyke et al. 1986).<br />

Small-area logging operations were less <strong>of</strong> a<br />

negative factor for resident adults. Van<br />

Dyke et al. (1986) also concluded that<br />

dispersing young animals were more<br />

adversely affected by logging and road<br />

system development than were established<br />

adults. By comparison, Gagliuso (1991) did<br />

not find avoidance by radio-collared lions to<br />

ei<strong>the</strong>r recent logging or high road densities<br />

in his southwestern Oregon study area.<br />

Differences in his findings from Van Dyke<br />

et al. were related to under-story density and<br />

rapid recovery <strong>of</strong> brush in newly logged<br />

areas. The southwest Oregon study area had<br />

more than twice <strong>the</strong> precipitation <strong>of</strong> <strong>the</strong><br />

Arizona and Utah studies.<br />

We compare <strong>the</strong> results <strong>of</strong> our studies<br />

within <strong>the</strong> same nor<strong>the</strong>ast Oregon study area<br />

and discuss similarities and differences with<br />

studies in Utah, Arizona and southwest<br />

Oregon (Van Dyke et al. 1986, Gagliuso<br />

1991). Our studies in nor<strong>the</strong>ast Oregon<br />

were conducted in a climatological,<br />

geographical, and anthropogenic situation<br />

somewhere in-between those areas described<br />

by Van Dyke et al., and Gagliuso. The<br />

objectives <strong>of</strong> this paper are to: 1) connect 2<br />

habitat investigations to gain a more<br />

complete understanding <strong>of</strong> microhabitat use<br />

relative to mountain lion life history, and 2)<br />

compare mountain lion microhabitat use in<br />

nor<strong>the</strong>ast Oregon with similar work in o<strong>the</strong>r<br />

regions <strong>of</strong> <strong>the</strong> western United States.<br />

STUDY AREA<br />

Both <strong>of</strong> <strong>the</strong>se studies were conducted in<br />

<strong>the</strong> Ca<strong>the</strong>rine Creek Wildlife Management<br />

Unit in nor<strong>the</strong>ast Oregon. The Ca<strong>the</strong>rine<br />

Creek study area is approximately 845 km 2<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

in size. Elevations range from 940 to 2,450<br />

m. This area is flanked on <strong>the</strong> west by range<br />

and agricultural lands <strong>of</strong> <strong>the</strong> Grande Ronde<br />

Valley and on <strong>the</strong> east by <strong>the</strong> Wallowa<br />

<strong>Mountain</strong>s within <strong>the</strong> Eagle Cap Wilderness<br />

Area. Most <strong>of</strong> <strong>the</strong> area (60%) is on <strong>the</strong><br />

Wallowa Whitman National Forest, with <strong>the</strong><br />

remaining being divided between Boise<br />

Cascade Corporation lands and o<strong>the</strong>r private<br />

ownership. Vegetation varies from<br />

subalpine coniferous forest to mixed conifer<br />

forest to rangeland and cropland. Road<br />

density varies from medium-high density<br />

(1.4 km/km 2 ) to small road closure areas.<br />

Approximately 20% <strong>of</strong> <strong>the</strong> work from <strong>the</strong>se<br />

studies was conducted within a Boise<br />

Cascade road closure area that had received<br />

various levels <strong>of</strong> logging activity. The<br />

majority <strong>of</strong> this area is mid-elevation<br />

coniferous forest with various forms <strong>of</strong> rock<br />

structure including rimrocks and outcrops.<br />

METHODS<br />

We compared <strong>the</strong> primary findings <strong>of</strong><br />

habitat characteristics at diurnal bed sites in<br />

Akenson et al. (1996) and at kill and cache<br />

sites in Nowak (1999). The 2 studies were<br />

compared qualitatively and <strong>the</strong> similarities<br />

and differences were described and<br />

discussed in an ecological context. The<br />

methods utilized in <strong>the</strong> 2 studies are briefly<br />

described below.<br />

Akenson et al. (1996) used various<br />

methods <strong>of</strong> locating and identifying<br />

mountain lion diurnal bed sites including<br />

snow tracking, radio telemetry, and a trained<br />

lion hound that located scent at bed sites.<br />

These methods were modified from<br />

Anderson (1990) for locating bobcat loafing<br />

sites in Colorado. A bed site was confirmed<br />

through visible evidence <strong>of</strong> ei<strong>the</strong>r soil or<br />

litter disturbance or tracks, and by alert<br />

reactions <strong>of</strong> a reliable dog. Beds were<br />

typically visible as a depression in snow or<br />

duff, or flattened grass. Once a bed site was<br />

identified, <strong>the</strong> surrounding area was<br />

searched for prey remains to determine


whe<strong>the</strong>r <strong>the</strong> bed was associated with a kill.<br />

The actual bed site became <strong>the</strong> center <strong>of</strong> a<br />

50-meter radius plot for collection <strong>of</strong> data to<br />

determine <strong>the</strong> physiographic and vegetative<br />

composition <strong>of</strong> <strong>the</strong> site. Habitat descriptions<br />

were aided by <strong>the</strong> handbook “Plant<br />

Associations <strong>of</strong> <strong>the</strong> Wallowa - Snake<br />

Province” (Johnson and Simon 1987).<br />

Akenson et al. (1996) evaluated 6<br />

primary habitat features at each plot site<br />

including rock structure, forest structure,<br />

canopy cover, shrub cover, plot visibility<br />

and overall security from human<br />

disturbance. This study emphasized <strong>the</strong><br />

structural influence <strong>of</strong> vegetation and<br />

topography on a mountain lion’s security<br />

from detection. O<strong>the</strong>r environmental<br />

influences such as distance to roads and<br />

abrupt habitat edges were also recorded.<br />

The distance to road measurement was<br />

recorded from <strong>the</strong> plot center to <strong>the</strong> nearest<br />

drivable road. A habitat edge typically<br />

marked a forest break or <strong>the</strong> beginning <strong>of</strong> a<br />

rock wall or large rock outcrop. For<br />

comparison, habitat data were also collected<br />

at randomly selected sites distributed<br />

throughout <strong>the</strong> study area. Random sites<br />

corresponded to <strong>the</strong> same square-mile<br />

section corner in 30 sections drawn from a<br />

pool <strong>of</strong> 185 possibilities, which all occurred<br />

in <strong>the</strong> known home ranges <strong>of</strong> <strong>the</strong> 5 subject<br />

mountain lions. All mountain lion age and<br />

sex classes were included. Habitat plots<br />

were categorized as summer (April 15 to<br />

September 1), winter (December 15 to<br />

March 15) or random, and data were<br />

summarized and compared using chisquared<br />

tests for differences between <strong>the</strong> 3<br />

plot types. Values were considered<br />

significant at α = 0.05.<br />

Nowak (1999) applied <strong>the</strong> term “cache<br />

site” to <strong>the</strong> location where a mountain lion<br />

kill was first found, whe<strong>the</strong>r or not <strong>the</strong> lion<br />

had moved it after making <strong>the</strong> kill. The<br />

exception to this was if <strong>the</strong> kill had<br />

obviously been moved from <strong>the</strong> original<br />

MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al. 113<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

cache site for subsequent feeding and <strong>the</strong><br />

original cache site could be identified. The<br />

term “kill site” referred to <strong>the</strong> location<br />

where <strong>the</strong> mountain lion actually killed its<br />

prey. The distinction between cache and kill<br />

site involved a combination <strong>of</strong> telemetry<br />

triangulation when <strong>the</strong> lion was present, and<br />

<strong>the</strong>n an investigation <strong>of</strong> <strong>the</strong> area after <strong>the</strong><br />

lion moved a safe distance away. As with<br />

o<strong>the</strong>r studies on lions, <strong>the</strong> majority <strong>of</strong><br />

information was obtained from locating<br />

radio instrumented animals on <strong>the</strong> ground<br />

(Anderson et al. 1992). Once <strong>the</strong> cache or<br />

kill site was determined, <strong>the</strong>n this site<br />

became <strong>the</strong> center <strong>of</strong> a 25-meter radius plot<br />

for collection <strong>of</strong> physiographic and<br />

vegetative data.<br />

Work closely followed Akenson et al.<br />

(1996) to facilitate comparisons between <strong>the</strong><br />

2 studies. Data were collected for 25 habitat<br />

variables to evaluate rock structure, forest<br />

structure, canopy cover, plot visibility and<br />

proximity to potential disturbance. This<br />

study likewise emphasized <strong>the</strong> influences <strong>of</strong><br />

vegetation and topography on mountain lion<br />

security but also on <strong>the</strong> security <strong>of</strong> kills,<br />

which may be left unattended for long<br />

periods <strong>of</strong> time. In this study, distance was<br />

recorded to both <strong>the</strong> nearest open, drivable<br />

road and to <strong>the</strong> nearest road <strong>of</strong> any kind,<br />

open or closed. As with Akenson et al., a<br />

habitat edge was typically a relatively abrupt<br />

change in stand composition and/or structure<br />

or topography. For comparison, habitat data<br />

were also collected at randomly selected<br />

sites distributed throughout <strong>the</strong> study area<br />

but within <strong>the</strong> subject lions’ home ranges.<br />

UTMs for random plots were generated by a<br />

computer random number generator<br />

(Micros<strong>of</strong>t Excel) using known study animal<br />

home ranges as limits to <strong>the</strong> generated<br />

coordinates. Habitat plots were categorized<br />

as cache, kill or random, and data were<br />

summarized and compared using forward,<br />

stepwise, logistic regression for differences<br />

between <strong>the</strong> 3 plot types. Values were


114 MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al.<br />

considered significant at α = 0.05. Only<br />

adult female mountain lions, with and<br />

without young, were included.<br />

RESULTS<br />

Akenson et al. (1996) recorded habitat<br />

characteristics at 61 diurnal bed sites, 32<br />

during winter and 29 during summer. Most<br />

(87%) <strong>of</strong> <strong>the</strong>se sites were not associated<br />

with kills. They collected <strong>the</strong> same habitat<br />

data at 30 randomly selected plots. Nowak<br />

(1999) collected habitat data at 79 cache<br />

sites, 19 kill sites and 101 randomly selected<br />

sites.<br />

Akenson et al. (1996) found significant<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

differences between diurnal bed sites and<br />

randomly selected sites in presence <strong>of</strong> rock<br />

structure, number <strong>of</strong> down logs in <strong>the</strong> plot,<br />

distance to habitat edge, sight distance (<strong>the</strong><br />

median distance at which a person could be<br />

seen from plot center at about lion height),<br />

understory density and management status<br />

(Table 1). Nowak (1999) found significant<br />

differences between cache sites and<br />

randomly selected sites in canopy cover,<br />

understory density, elevation, and<br />

management status. Significant differences<br />

between kill and random sites were in<br />

elevation, management status and plot<br />

visibility (<strong>the</strong> mean distance at which a<br />

Table 1. Habitat characteristics at mountain lion diurnal bed sites, summer and winter, at cache<br />

sites, and at randomly selected sites associated with each study (Akenson et al. 1996, Nowak 1999).<br />

Asterisks (*) indicate features significantly different (p


person could be seen from plot center at<br />

about lion height). Kill and cache sites<br />

differed only in canopy cover (Table 1).<br />

Large rock structure (forested rimrock)<br />

and down logs were present in significantly<br />

more diurnal bed site plots than expected but<br />

that was not <strong>the</strong> case for cache sites,<br />

although cache sites were slightly more<br />

likely to contain rock ledges than were <strong>the</strong><br />

random sites in that study. Canopy cover<br />

was significantly greater in cache sites than<br />

in ei<strong>the</strong>r kill or random sites but was not<br />

different between bed sites and random<br />

sites. Understory density was lower in<br />

cache sites but higher in summer diurnal bed<br />

sites. Akenson et al. (1996) found greater<br />

use <strong>of</strong> <strong>the</strong> old logged management type for<br />

diurnal beds in winter; Nowak (1999) found<br />

cache sites in old logged with similar<br />

frequency to random plots (Table 2). A<br />

relatively high percentage <strong>of</strong> cache sites<br />

were located in shelterwood but diurnal beds<br />

were in that management type with similar<br />

frequency to random plots. Cache sites were<br />

in <strong>the</strong> rangeland management type with less<br />

frequency than <strong>the</strong> random sites but bed<br />

sites were located in rangeland with about<br />

<strong>the</strong> same frequency as random sites.<br />

Nei<strong>the</strong>r study documented significant<br />

differences in <strong>the</strong> distance to <strong>the</strong> nearest<br />

open road although both winter bed sites and<br />

cache sites tended to be far<strong>the</strong>r from open<br />

roads than random sites. In Akenson et al.<br />

MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al. 115<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

(1996), summer diurnal beds were<br />

significantly closer to a habitat edge than<br />

were random sites. Although not<br />

statistically significant, Akenson et al.<br />

(1996) and Nowak (1999) found that winter<br />

diurnal beds and cache sites both tended to<br />

be closer to a habitat edge than <strong>the</strong> random<br />

sites. Both studies documented significantly<br />

lower plot visibility/sight distance in sample<br />

plots compared with random sites. Both<br />

studies also showed seasonal variation, in<br />

elevation with both bed sites and caches at<br />

lower elevation in winter than in summer.<br />

When 4 seasons were considered, Nowak<br />

found cache sites were at higher elevation in<br />

fall than in summer, spring, or winter.<br />

DISCUSSION<br />

Several authors have addressed <strong>the</strong><br />

question <strong>of</strong> mountain lion habitat use,<br />

conducted studies in some diverse<br />

environments, and concluded that a primary<br />

factor in habitat selection for this carnivore<br />

was <strong>the</strong> presence <strong>of</strong> vegetation and terrain<br />

cover to enhance <strong>the</strong> stalking <strong>of</strong> prey,<br />

usually deer or elk. Hornocker (1970) felt<br />

that lions in his Idaho study area selected<br />

habitat on <strong>the</strong> basis <strong>of</strong> prey density and<br />

terrain features that were advantages for<br />

hunting. Logan and Irwin (1985) also noted<br />

a high occurrence <strong>of</strong> lion caches within<br />

canyon vegetation, draws, and on steep<br />

ridges demonstrating <strong>the</strong> importance <strong>of</strong> both<br />

Table 2. Management status at mountain lion diurnal bed sites, summer and winter, at cache sites,<br />

and at randomly selected sites associated with each study (Akenson et al. 1996, Nowak 1999).<br />

Asterisks (*) indicate features significantly different (p


116 MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al.<br />

vegetative and terrain cover. Seidensticker<br />

et al. (1973) concluded that a “vegetation –<br />

topography/prey numbers – vulnerability<br />

complex” determined both lion home range<br />

size and population density. We agree that<br />

<strong>the</strong> need for cover while bedding, hunting,<br />

or guarding a cache site is ecologically<br />

important. Our findings indicate that forest<br />

management strategies contribute to both<br />

prey abundance and enhanced stalking cover<br />

for mountain lions (Table 2).<br />

Van Dyke et al. (1986) concluded that<br />

resident lions avoided portions <strong>of</strong> <strong>the</strong>ir home<br />

ranges with active logging activity, and<br />

found that transient lions were <strong>the</strong> primary<br />

users <strong>of</strong> areas with active timber harvest, or<br />

even newly logged areas. By contrast,<br />

Gagliuso (1991) found in southwestern<br />

Oregon that lions did not avoid timber<br />

harvest sites but ra<strong>the</strong>r were closer to <strong>the</strong>se<br />

activities than expected at random. We<br />

observed a similar attraction to new logging,<br />

which we believed was related to <strong>the</strong><br />

abundant “candy food” made newly<br />

available to deer and elk by logging that<br />

brought branches laden with lichen and<br />

mosses down to ground level. Once this<br />

resource was exhausted, deer and elk quit<br />

using <strong>the</strong>se sites, as did hunting lions. We<br />

concluded from track evidence made in<br />

snow during winter, or dust during summer,<br />

that lions were using newly logged areas at<br />

night. Nocturnal movement patterns, in<br />

association with sub-optimal habitat cover,<br />

was also documented by Beier (1995) in<br />

California and Van Dyke et al. (1986) in<br />

Utah where <strong>the</strong>y documented mountain lions<br />

using <strong>the</strong> most undisturbed habitats in <strong>the</strong>ir<br />

home ranges for diurnal localization. Our<br />

findings concur with <strong>the</strong>se authors. On a<br />

micro-habitat scale, our findings also show<br />

<strong>the</strong> importance <strong>of</strong> specific features, such as<br />

forested rimrock and downed logs for<br />

diurnal bed sites, understory density for<br />

hunting and stalking cover, and canopy<br />

cover for kill cache sites.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

The documentation <strong>of</strong> micro-habitat use<br />

is essential in understanding mountain lion<br />

daily adaptation to multiple environmental<br />

influences and disturbances. The use <strong>of</strong><br />

specific habitat types by lions is largely<br />

dependent on <strong>the</strong> activity <strong>of</strong> <strong>the</strong> individual.<br />

A cougar that is bedding for <strong>the</strong> day selects<br />

a location that <strong>of</strong>fers both concealment and<br />

nearby escape terrain, as indicated in our<br />

study by a strong selection for forested<br />

rimrock structure with a component <strong>of</strong><br />

downed logs. Whereas a lion that is hunting<br />

is going to use areas preferred by prey<br />

species that also afford stalking<br />

concealment, usually in <strong>the</strong> form <strong>of</strong> understory<br />

vegetation or o<strong>the</strong>r close to <strong>the</strong> ground<br />

structure. Then, once <strong>the</strong> kill has been<br />

made, <strong>the</strong>re is typically an effort made by<br />

<strong>the</strong> lion to cache <strong>the</strong> kill under a tree or<br />

brush, presumably to reduce detection by<br />

avian scavengers.<br />

MANAGEMENT IMPLICATIONS<br />

Our findings on mountain lion habitat<br />

use have implications to both wildlife and<br />

habitat managers. There are many complex<br />

variables influencing mountain lion habitat<br />

use in different regions and levels <strong>of</strong> human<br />

influence. Several factors influence <strong>the</strong> way<br />

in which lions use <strong>the</strong>ir environment, or<br />

conduct “land tenure” as described by John<br />

Seidensticker (1973). Obtaining food,<br />

establishing and defending territories,<br />

breeding, reproducing, and raising kittens to<br />

dispersal age all have a bearing on how<br />

mountain lions use a given landscape. In<br />

comparing findings from this study with<br />

o<strong>the</strong>r studies, it appears that factors vary<br />

from region to region. However, habitat use<br />

seems to be driven by three ecological<br />

needs: security, cover, and food.<br />

The mountain lions that we studied have<br />

co-existed with timber harvest for several<br />

lion generations. The literature suggests that<br />

lions will still use habitats that have been<br />

logged as long as <strong>the</strong> harvest areas are


Gagliuso 1991). Leaving strips <strong>of</strong> trees for<br />

buffers, in conjunction with small harvest<br />

units, creates an extensive habitat edge<br />

effect beneficial to mountain lions. O<strong>the</strong>r<br />

important features are vegetative cover<br />

around rock structure for bedding security,<br />

downed logs, and ample understory density<br />

to allow for successful stalking. All <strong>of</strong> <strong>the</strong><br />

diurnal bed-sites occurring in rimrock had<br />

ei<strong>the</strong>r brush or trees at <strong>the</strong> bed. We did not<br />

document bed-use in newly logged areas or<br />

in rock structure without some form <strong>of</strong><br />

vegetative cover. A timber management<br />

practice that leaves a forested buffer around<br />

rock structure is advantageous for mountain<br />

lion security. The size <strong>of</strong> <strong>the</strong> buffer would<br />

vary with vegetation type and density, but<br />

generally a 50-meter buffer would afford<br />

concealment for lions in our study area. We<br />

did not find a significant aversion to roads in<br />

<strong>the</strong> Ca<strong>the</strong>rine Creek study area, but our<br />

methods may not have effectively addressed<br />

this issue since most <strong>of</strong> our data was<br />

ga<strong>the</strong>red in or near a Boise Cascade<br />

Corporation road closure area. The two<br />

primary land managers, <strong>the</strong> US Forest<br />

Service and Boise Cascade Corporation,<br />

have implemented travel management plans<br />

that vastly reduce human disturbance<br />

through established road closure areas. In<br />

general, our findings are more similar to<br />

results produced in southwest Oregon by<br />

Gagliuso (1991) than those described by<br />

Van Dyke et al. (1986) in Arizona and Utah.<br />

We feel <strong>the</strong>se differences are due to<br />

mountain lions in Oregon having long-term<br />

exposure to logging, and <strong>the</strong> habitat having a<br />

quicker capability for regrowth with higher<br />

amounts <strong>of</strong> precipitation in two areas <strong>of</strong><br />

Oregon than <strong>the</strong> more arid Southwest.<br />

SUMMARY<br />

In conclusion, we have added more<br />

information to <strong>the</strong> pool <strong>of</strong> knowledge<br />

supporting <strong>the</strong> concept <strong>of</strong> mountain lions as<br />

an adaptable, yet vulnerable species. Logan<br />

MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al. 117<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

and Sweanor (2001) emphasize <strong>the</strong><br />

importance <strong>of</strong> gaining a better understanding<br />

<strong>of</strong> mountain lion habitat use through<br />

identifying critical habitats, landscape<br />

linkages, and by assessing how human<br />

development, resource extraction, and<br />

habitat modification can degrade or enhance<br />

<strong>the</strong>se habitats. We have demonstrated <strong>the</strong><br />

importance <strong>of</strong> small-scale physiographic<br />

features within <strong>the</strong> larger scale habitat<br />

complex. Scientific management <strong>of</strong><br />

mountain lions depends on both wildlife<br />

managers and land managers understanding<br />

this species’ requirements <strong>of</strong> security, cover,<br />

and food, and how obtaining <strong>the</strong>se<br />

ecological needs varies between regions and<br />

physiographic and climatological situations<br />

and conditions.<br />

ACKNOWLEDGEMENTS<br />

We are indebted to several agencies and<br />

individuals who made our studies<br />

successful. The study by Akenson et al. was<br />

for <strong>the</strong> Oregon Department <strong>of</strong> Fish and<br />

Wildlife. Funding was mostly from a grant<br />

from <strong>the</strong> Federal Aid in Fish and Wildlife<br />

Restoration Act. The generous contributions<br />

<strong>of</strong> time, knowledge, and hound services by<br />

Ted Craddock, Gale Culver, Loren Brown<br />

and field assistant Paul Alexander were<br />

invaluable. Financial support for <strong>the</strong> study<br />

by Nowak came from Washington State<br />

University, <strong>the</strong> National Wildlife Research<br />

Center <strong>of</strong> <strong>the</strong> U.S.D.A. Animal and Plant<br />

Health Inspection Service, and <strong>the</strong> Oregon<br />

Department <strong>of</strong> Fish and Wildlife. Nowak is<br />

greatly indebted for assistance from her field<br />

crew <strong>of</strong>: Craig Whitman, Gail Collins, Brett<br />

Lyndaker, Jeff Olmstead, Tracy Taylor, and<br />

volunteers: Renan Bagley, Doug Wolf,<br />

Mark Berrest, Kate Richardson, Keith<br />

Wehner, Eric Macy, and Mark Squire. The<br />

authors wish to thank Kerry Murphy and<br />

Mark Penninger for <strong>the</strong>ir constructive<br />

comments on an earlier version <strong>of</strong> <strong>the</strong><br />

manuscript.


118 MOUNTAIN LION BED, CACHE AND KILL SITES · Akenson et al.<br />

LITERATURE CITED<br />

ACKERMAN, B.B. 1982. Cougar predation<br />

and ecological energetics in sou<strong>the</strong>rn<br />

Utah. Thesis, Utah State University,<br />

Logan, Utah, USA.<br />

ANDERSON, A.E., D.C. BOWDEN, AND D.M.<br />

KATNER. 1992. The Puma on <strong>the</strong><br />

Uncompahgre Plateau. Colorado<br />

Division <strong>of</strong> Wildlife. Tech. Pub. No. 40.<br />

ANDERSON, E.M. 1990. Bobcat diurnal<br />

loafing sites in sou<strong>the</strong>astern Colorado.<br />

Journal <strong>of</strong> Wildlife Management.<br />

54:600-602.<br />

AKENSON, J.J., M.G. HENJUM, AND T.J.<br />

CRADDOCK. 1996. Diurnal bedding<br />

habitat <strong>of</strong> mountain lions in nor<strong>the</strong>ast<br />

Oregon. Abstract in Fifth <strong>Mountain</strong><br />

<strong>Lion</strong> <strong>Workshop</strong>, 27 February-1 March,<br />

1996, San Diego, California.<br />

BEIER, P. 1995. Dispersal <strong>of</strong> juvenile<br />

cougars in fragmented habitat. Journal<br />

<strong>of</strong> Wildlife Management. 59:228-237.<br />

GAGLIUSO, R.A. 1991. Habitat alteration<br />

and human disturbance: <strong>the</strong>ir impact on<br />

cougar habitat utilization in southwest<br />

Oregon. Thesis, Oregon State<br />

University, Corvallis, Oregon, USA.<br />

HORNOCKER, M.G. 1970. An analysis <strong>of</strong><br />

mountain lion predation upon mule deer<br />

and elk in <strong>the</strong> Idaho Primitive Area.<br />

Wildlife Monograph No. 21: 1-39.<br />

_____. 1976. Cougars up close. National<br />

Wildlife. 14(6):42-47.<br />

JOHNSON, C.G. AND S.A. SIMON.1987. Plant<br />

Associations <strong>of</strong> <strong>the</strong> Wallowa-Snake<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Province. U.S. Forest Service handbook<br />

No. R-6 ECOL-TP-225B-86.<br />

LOGAN, K.A. AND L.L. IRWIN. 1985.<br />

<strong>Mountain</strong> lion habitats in <strong>the</strong> Bighorn<br />

<strong>Mountain</strong>s, Wyoming. Wildlife Society<br />

Bulletin 13:257-262.<br />

LOGAN, K.A. AND L.L. SWEANOR. 2001.<br />

Desert Puma, evolutionary ecology and<br />

conservation <strong>of</strong> an enduring carnivore.<br />

Island Press. Washington, D.C., USA.<br />

NOWAK, M.C. 1999. Predation rates and<br />

foraging ecology <strong>of</strong> adult female<br />

mountain lions in nor<strong>the</strong>astern Oregon.<br />

Thesis. Washington State University,<br />

Pullman, Washington, USA.<br />

SEIDENSTICKER, J.C. IV, M.G. HORNOCKER,<br />

W.V. WILES AND J.P. MESSICK. 1973.<br />

<strong>Mountain</strong> lion social organization in <strong>the</strong><br />

Idaho Primitive Area. Wildlife<br />

Monograph 35:1-60.<br />

SEIDENSTICKER, J.C. IV. 1973. <strong>Mountain</strong><br />

lion social organization in <strong>the</strong> Idaho<br />

Primitive Area. Dissertation, University<br />

<strong>of</strong> Idaho, Moscow, Idaho, USA.<br />

VAN DYKE, F.G., R.H. BROCK, H.G. SHAW,<br />

B.B. ACKERMAN, T.P. HEMKER AND<br />

F.G. LINDZEY. 1986. Reactions <strong>of</strong><br />

mountain lions to logging and human<br />

activity. Journal <strong>of</strong> Wildlife<br />

Management, 50:95-102.<br />

YOUNG, S.P. 1946. History, life habits,<br />

economic status, and control, Part 1.<br />

Pages 1-173 in S.P. Young and E.A.<br />

Goldman, eds. The puma, mysterious<br />

American cat. The American Wildlife<br />

Institute, Washington, D.C. USA.


IMPACT OF EDGE HABITAT ON HOME RANGE SIZE IN PUMAS<br />

JOHN W. LAUNDRÉ, Instituto de Ecología, A.C. Apartado Postal 632, 34100 Durango, Dgo.,<br />

México, email: launjohn@prodigy.net.mx<br />

LUCINA HERNÁNDEZ, Instituto de Ecología, A.C. Apartado Postal 632, 34100 Durango,<br />

Dgo., México, email: lucina@sequia.edu.mx<br />

Abstract: In <strong>the</strong> previous workshop in San Antonio, researchers from Wyoming reported that<br />

pumas from two areas with different amounts <strong>of</strong> fragmentation still had home range areas that<br />

contained equal amounts <strong>of</strong> periphery (= edge). In <strong>the</strong> same workshop, we reported that edge<br />

habitat was critical for successful hunting <strong>of</strong> deer by pumas. These two results indicate that <strong>the</strong><br />

amount <strong>of</strong> edge habitat in an area may be an important factor in determining home range size <strong>of</strong><br />

pumas. We tested this hypo<strong>the</strong>sis with data we have on home ranges <strong>of</strong> pumas in sou<strong>the</strong>rn<br />

Idaho/northwestern Utah. The study area is highly fragmented into forest patches and sagebrush<br />

open areas. We tested three predictions: 1) <strong>the</strong> amount <strong>of</strong> edge habitat in <strong>the</strong> home ranges <strong>of</strong><br />

pumas would be similar, regardless <strong>of</strong> <strong>the</strong> size <strong>of</strong> <strong>the</strong> home range, 2) <strong>the</strong> percent <strong>of</strong> edge would<br />

be negatively related to home range size, and 3) <strong>the</strong>re would be more edge habitat within home<br />

range boundaries than in general areas <strong>of</strong> similar size. We tested <strong>the</strong>se predictions by overlaying<br />

telemetry locations on habitat maps <strong>of</strong> <strong>the</strong> area, determining <strong>the</strong> home range boundaries with <strong>the</strong><br />

minimum convex polygon method and <strong>the</strong>n estimating <strong>the</strong> amount <strong>of</strong> forest edge (km 2 ) that<br />

occurred in each home range. The analysis was conducted with standard GIS s<strong>of</strong>tware and we<br />

had 20 pumas where <strong>the</strong> home range was adequately determined (> 30 relocations). Home<br />

range size varied from 38 to 120 km 2 . However, 14 (70%) <strong>of</strong> <strong>the</strong> home ranges were between 38<br />

to 105 km 2 . The amount <strong>of</strong> edge habitat within all <strong>the</strong> home ranges varied from 13 to 35 km 2 .<br />

Within <strong>the</strong> 14 smaller home ranges, <strong>the</strong> amount <strong>of</strong> edge varied from 13 to 20 km 2 . The percent<br />

<strong>of</strong> edge within home ranges was negatively correlated with home range size. The amount <strong>of</strong><br />

edge within <strong>the</strong> home range boundaries was significantly greater (F = 15.05, P < 0.001) than<br />

general areas <strong>of</strong> similar size. We concluded that <strong>the</strong> amount <strong>of</strong> edge within an area was<br />

influencing <strong>the</strong> size <strong>of</strong> home ranges. We proposed that pumas needed a certain minimum<br />

amount <strong>of</strong> edge (hunting habitat) to successfully hunt <strong>the</strong>ir prey and that <strong>the</strong> amount <strong>of</strong><br />

“catchable” prey was more important than just general prey abundance.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

119


120<br />

EFFECT OF ROADS ON HABITAT USE BY COUGARS<br />

DOROTHY M. FECSKE, Department <strong>of</strong> Wildlife and Fisheries Sciences, South Dakota State<br />

University, Brookings, SD 57007, USA, email: gdf@rapidnet.com<br />

JONATHAN A. JENKS, Department <strong>of</strong> Wildlife and Fisheries Sciences, South Dakota State<br />

University, Brookings, SD 57007, USA, email: Jonathan_Jenks@sdstate.edu<br />

FREDERICK G. LINDZEY, USGS Biological Resources Division, Wyoming Cooperative Fish<br />

and Wildlife Research Unit, University Station, Laramie, WY 82071, USA, email:<br />

Flindzey@uwyo.edu<br />

STEVEN L. GRIFFIN, South Dakota Department <strong>of</strong> Game, Fish and Parks, 3305 W. South<br />

Street, Rapid City, SD 57702, USA, email: Steve.Griffin@state.sd.us<br />

Abstract: We examined effect <strong>of</strong> roads on habitat use by cougars, Puma concolor, in <strong>the</strong> Black<br />

Hills, South Dakota. A total <strong>of</strong> 768 daytime locations <strong>of</strong> 12 radio-collared cougars were<br />

obtained during weekly flights (1999 - 2001) using aerial telemetry techniques. Locations were<br />

incorporated into a geographic information system (GIS) <strong>of</strong> roads (Class 1, 2, 3, and 4). We<br />

tested <strong>the</strong> null hypo<strong>the</strong>ses that cougars select habitat at random distances to roads and at random<br />

road densities and cougar use <strong>of</strong> habitat near roads did not differ with respect to road class, sex,<br />

age class, and habitat quality (based on a ranked cougar habitat-relation model). We examined<br />

use <strong>of</strong> habitat near roads for an adult female cougar fitted with a Global Position System (GPS)<br />

collar during crepuscular, diurnal, and nocturnal periods. Also, we identified road classes where<br />

cougar snow tracks were located and cougar/vehicle collisions occurred. During daylight hours,<br />

cougars avoided habitat near Class 3 roads (P < 0.001), <strong>the</strong> predominant road class in <strong>the</strong> Black<br />

Hills. However, on occasions where cougars were located near Class 3 roads, high quality<br />

habitat was selected. Cougars in <strong>the</strong> 5 to 6-year age class were located far<strong>the</strong>r from Class 1 roads<br />

than younger animals (P < 0.0001). Females in <strong>the</strong> 1 to 2-year age class were located closer to<br />

Class 1 and Class 2 roads than older females (P < 0.0001). Females in 5 to 6 and 7 to 8-year age<br />

classes were located closer to Class 4 roads (P = 0.0047) than younger females. Road densities<br />

(km road/km 2 ) in annual home ranges <strong>of</strong> male cougars did not differ (P = 0.5000) from road<br />

densities throughout <strong>the</strong> Black Hills study area but densities in annual ranges <strong>of</strong> females were<br />

greater (P = 0.0078) than those <strong>of</strong> <strong>the</strong> study area. Cougars in <strong>the</strong> Black Hills have adapted to a<br />

heavily roaded landscape but presence <strong>of</strong> roads is impacting cougar use <strong>of</strong> habitat and survival.<br />

We suggest use <strong>of</strong> habitat near Class 3 roads by cougars would increase if roads were closed or<br />

had limited access, and if thinned ponderosa pine stands adjacent to Class 3 and 4 roads were<br />

managed for understory vegetation.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


ECOLOGY OF SYMPATRIC PUMAS AND JAGUARS IN NORTHWESTERN<br />

MEXICO<br />

CARLOS A. LOPEZ GONZALEZ, Universidad Autonoma de Queretaro, Cerro de las<br />

Campanas s/n, Mexico, email: Cats4mex@aol.com<br />

SAMIA E. CARRILLO PERCASTEGUI, Nor<strong>the</strong>rn Jaguar Project, 2114 W. Grant Rd. #121,<br />

Tucson, AZ 85745, USA, email: Cats4mex@aol.com<br />

Abstract: Pumas (Puma concolor) are usually considered subordinate species where jaguars<br />

(Pan<strong>the</strong>ra onca) are present. Most <strong>of</strong> <strong>the</strong> current information on resource partitioning by <strong>the</strong>se<br />

two species comes from tropical sites. Out study area is located in <strong>the</strong> limits <strong>of</strong> <strong>the</strong> tropical realm<br />

and consequently could be characterized as puma habitat. Our objectives were to describe <strong>the</strong><br />

ecology <strong>of</strong> both large felids in an area located 135 s. <strong>of</strong> <strong>the</strong> United States border, in <strong>the</strong> state <strong>of</strong><br />

Sonora, Mexico. From July 1999 to December 2002 using a suite <strong>of</strong> methodologies (cameratraps,<br />

radio-telemetry, scat, track and prey surveys), we surveyed an area ≈1000 km². The study<br />

area is a matrix <strong>of</strong> oak-woodland, tropical thornscrub, and upper sonoran desert; ranging in<br />

elevation from 200 to 1200 m. The main economic activity within <strong>the</strong> region is ranching. We<br />

determined through radio-telemetry a density <strong>of</strong> 3 pumas/100 km², and through camera-trap<br />

surveys a density <strong>of</strong> 1.4±0.4 jaguars/100 km². Camera-trap capture rates are three times higher<br />

for pumas than jaguars. Both species are feeding on white-tailed deer and to a lesser extent on<br />

livestock. Pumas are a ca<strong>the</strong>meral species whereas jaguars are nocturnal-crepuscular. Jaguars are<br />

using oak woodlands more than expected by chance, where pumas are using habitats according<br />

to availability. The number <strong>of</strong> pumas present may be an artifact <strong>of</strong> less prosecution by cowboys<br />

(only 1 puma killed since 1999), where jaguars are constantly prosecuted as <strong>the</strong>y are perceived as<br />

liable <strong>of</strong> most livestock depredations (22 jaguars killed since 1999). During 2002 we began a<br />

program to help local ranchers on maintaining infrastructure, and apparent result has been less<br />

pressure on <strong>the</strong> jaguar population within <strong>the</strong> area.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

121


122<br />

COUGAR ECOLOGY AND COUGAR-WOLF INTERACTIONS IN YELLOWSTONE<br />

NATIONAL PARK: A GUILD APPROCH TO LARGE CARNIVORE CONSERVATION<br />

TONI K. RUTH, Associate Conservation Scientist, Wildlife Conservation Society, 2023 Stadium<br />

Dr. Suite 1A, Bozeman, MT 59030, USA, email: truth@montanadsl.net<br />

POLLY C. BUOTTE, Research Assistant, Wildlife Conservation Society, 2023 Stadium Dr.<br />

Suite 1A, Bozeman, MT 59030, USA, email: polly_thornton@hotmail.com<br />

HOWARD B. QUIGLEY, Beringia South, 3610 W. Broadwater, Suite 111, Bozeman, MT<br />

59715, USA<br />

MAURICE G. HORNOCKER, Senior Scientist, Wildlife Conservation Society, 2023 Stadium<br />

Dr. Suite 1A, Bozeman, MT 59030, USA<br />

Abstract: Successful restoration <strong>of</strong> large carnivores in <strong>the</strong> Nor<strong>the</strong>rn Rockies and <strong>the</strong> concomitant<br />

increase in carnivore abundance and distribution will challenge humans as human development<br />

increases throughout <strong>the</strong> West. Presently, <strong>the</strong>re is little understanding <strong>of</strong> how<br />

reintroduction/reestablishment <strong>of</strong> endangered large carnivores (wolves and grizzly bears) may<br />

affect <strong>the</strong> population characteristics, distribution, and behavior <strong>of</strong> o<strong>the</strong>r large carnivore<br />

populations, such as cougars. If restored wolves limit cougar populations in number or<br />

distribution, this limitation may have synergistic effects with current relaxation <strong>of</strong> cougar<br />

hunting regulations and rapid development. An added stress such as low prey availability (e.g.<br />

caused by hard winter or disease) could fur<strong>the</strong>r impact populations. Understanding competitive<br />

relationships between large carnivores and <strong>the</strong> role that habitat and prey availability play is<br />

paramount to predicting and preparing for changes in <strong>the</strong> Greater Yellowstone region. In order<br />

to assess population-level effects <strong>of</strong> wolf (Canis lupus) reestablishment on cougars (Puma<br />

concolor) in and near Yellowstone National Park (YNP), we initiated a Phase II study <strong>of</strong> YNP<br />

cougars in 1998. The study is designed to examine <strong>the</strong> characteristics <strong>of</strong> <strong>the</strong> cougar population<br />

including: sex and age structure, density, reproductive and survival rates, dispersal and<br />

recruitment events, rate <strong>of</strong> predation on prey, and spatial and temporal movements. These<br />

parameters will be compared with analogous estimates made prior to <strong>the</strong> wolf restoration event<br />

in 1995 (Phase I data, Murphy 1998) and similar parameters documented for <strong>the</strong> wolf population<br />

to assess competition and resource partitioning between <strong>the</strong> two species. During 1998-2002, 56<br />

cougars were captured in and adjacent to areas used by 35-88 wolves within 3-5 wolf packs on<br />

<strong>the</strong> Nor<strong>the</strong>rn Yellowstone Study Area, Montana and Wyoming. A sample <strong>of</strong> 3 to 10 radiocollared<br />

wolves was maintained within each wolf pack by <strong>the</strong> Yellowstone Wolf Restoration<br />

program. In this paper we summarize current research findings relative to cougar population<br />

changes pre- and post-wolf reintroduction, species interactions, and discuss future study<br />

direction.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


EVALUATION OF HABITAT FACTORS THAT AFFECT THE ABUNDANCE OF<br />

PUMAS IN THE CHIHUAHUAN DESERT<br />

JOEL LOREDO SALAZAR, Instituto de Ecología, A.C. Apdo. Postal 63. CP 91070 Jalapa, Ver.<br />

México, email: loredosj@ecologia.edu.mx<br />

LUCINA HERNÁNDEZ, Instituto de Ecología, A.C., Apartado Postal 632, 34100 Durango,<br />

Dgo., México, email: lucina@sequia.edu.mx<br />

JOHN W. LAUNDRÉ, Instituto de Ecología, A.C., Apartado Postal 632, 34100 Durango, Dgo.,<br />

México, email: launjohn@prodigy.net.mx<br />

Abstract: Pumas originally occupied all <strong>of</strong> Mexico but <strong>the</strong>ir current status is not well known.<br />

This is especially true in <strong>the</strong> Chihuahuan desert <strong>of</strong> Nor<strong>the</strong>rn Mexico. To manage this species in<br />

this area, it is important to have some estimation <strong>of</strong> <strong>the</strong>ir status. To evaluate <strong>the</strong> status <strong>of</strong> pumas<br />

in this area we need to first identify factors that may contribute to <strong>the</strong>ir rarity or abundance.<br />

Such factors can be placed into three separate but related groups: habitat quality, prey<br />

abundance, and human impacts. To evaluate <strong>the</strong>se various factors, we selected two mountain<br />

ranges in <strong>the</strong> nor<strong>the</strong>rn Chihuahuan desert where previous work indicated differences in <strong>the</strong><br />

relative abundance <strong>of</strong> pumas. The area <strong>of</strong> low puma abundance was El Cuervo near Aldama,<br />

Chihuahua and <strong>the</strong> area <strong>of</strong> high abundance was Sierra Rica in <strong>the</strong> Canyon de Santa Elena<br />

protected area. In <strong>the</strong> field we estimated habitat quality by measuring shrub density, cover, and<br />

height. We also estimated prey (wild and domestic) abundance by counting fecal groups along<br />

random transects. With <strong>the</strong> use <strong>of</strong> GIS technology we assessed human impacts by determining<br />

<strong>the</strong> number <strong>of</strong> roads, number and size <strong>of</strong> towns, and overall density <strong>of</strong> humans in a 20 km radius<br />

around each range. Our results indicate that habitat quality was similar between <strong>the</strong> two areas.<br />

However, wild and domestic prey was higher in Santa Elena and all measurements <strong>of</strong> human<br />

impact were higher in El Cuervo. We concluded that habitat quality was not a factor<br />

contributing to relative puma abundance. However, <strong>the</strong> increased presence <strong>of</strong> and access by<br />

humans in El Cuervo is <strong>the</strong> main contributing factor via illegal hunting <strong>of</strong> pumas and <strong>the</strong>ir prey.<br />

Future work will test this hypo<strong>the</strong>sis in o<strong>the</strong>r areas <strong>of</strong> <strong>the</strong> Chihuahuan desert. If this hypo<strong>the</strong>sis<br />

is supported, it indicates that conservation efforts <strong>of</strong> pumas in Nor<strong>the</strong>rn Mexico need center on<br />

environmental education ra<strong>the</strong>r than habitat protection/restoration.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

123


124<br />

ARE PUMAS OPPORTUNISTIC SCAVENGERS?<br />

JIM W. BAUER, University <strong>of</strong> California-Davis, Wildlife Health Center, Sou<strong>the</strong>rn California<br />

Puma Project Field Station, P.O. Box 1203, Julian, CA 92036, USA, email:<br />

jwbauer@uia.net<br />

KENNETH A. LOGAN, University <strong>of</strong> California-Davis, Wildlife Health Center, Sou<strong>the</strong>rn<br />

California Puma Project Field Station, P.O. Box 1114, Julian, CA 92036, USA, email:<br />

klogan2@mindspring.com<br />

LINDA L. SWEANOR, University <strong>of</strong> California-Davis, Wildlife Health Center, Sou<strong>the</strong>rn<br />

California Puma Project Field Station, P.O. Box 1114, Julian, CA 92036, USA, email:<br />

lsweanor@mindspring.com<br />

WALTER M. BOYCE, University <strong>of</strong> California-Davis, Wildlife Health Center, One Shields<br />

Avenue, Davis, CA 95616, USA, email: wmboyce@ucdavis.edu<br />

Abstract: We examined scavenging on mule deer (Odocoileus hemionus) carcasses by pumas<br />

(Puma concolor) in <strong>the</strong> Peninsular Ranges <strong>of</strong> Sou<strong>the</strong>rn California. Between 23 January 2001<br />

and 21 November 2002, a total <strong>of</strong> 42 deer carcasses from road kills, depredation permits, and<br />

euthanized deer were used to determine scavenging events. Seventeen <strong>of</strong> 42 deer carcasses<br />

(40.5%) were scavenged by 7 to 10 different pumas. Two <strong>of</strong> <strong>the</strong> scavenging pumas (males) were<br />

previously telemetered, while 4 pumas (3 male, 1 female) were captured and instrumented at <strong>the</strong><br />

scavenging site. Telemetered pumas ranged in age from 11 months to 9 years. Deer carcasses<br />

were found and scavenged by pumas within 1 to 14 days, when carcass conditions ranged from<br />

fresh to rotting and maggot infested. Pumas treated scavenged carcasses as <strong>the</strong>y would <strong>the</strong>ir own<br />

kills, dragging unte<strong>the</strong>red carcasses to preferred sites and caching, as well as depositing scats and<br />

making scrapes in <strong>the</strong> area. However, pumas did not always attempt to cache te<strong>the</strong>red carcasses.<br />

During <strong>the</strong> course <strong>of</strong> our fieldwork we also discovered that one telemetered puma was repeatedly<br />

visiting a domestic livestock graveyard and scavenging on surface-discarded horse and cattle<br />

carcasses. While pumas are known to be opportunistic predators, our results would suggest that<br />

<strong>the</strong>y are opportunistic scavengers as well. Due to pumas’ propensity to scavenge, it is likely that<br />

some perceived puma kills may in fact be scavenging events. Frequent monitoring and timely<br />

field investigation <strong>of</strong> mortality signals detected from telemetered prey species will help<br />

investigators identify those events. Scavenging behavior should be considered when evaluating<br />

or predicting <strong>the</strong> effects <strong>of</strong> puma predation on prey species.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


COUGAR-INDUCED INDIRECT EFFECTS: DOES THE RISK OF PREDATION<br />

INFLUENCE UNUGULATE FORAGING BEHAVIOR ON THE NATIONAL BISON<br />

RANGE?<br />

DAVID M. CHOATE, Ph.D. candidate, Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Notre<br />

Dame, 107 Galvin Life Science Center, Notre Dame, IN 46556, USA; and, Department<br />

<strong>of</strong> Forestry, Range & Wildlife, Utah State University, Logan, UT 84322, USA, email:<br />

dchoate@nd.edu<br />

GARY E. BELOVSKY, Pr<strong>of</strong>essor, Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Notre<br />

Dame, 107 Galvin Life Science Center, Notre Dame, IN 46556, USA, email:<br />

Gary.E.Belovsky.1@nd.edu<br />

MICHAEL L. WOLFE, Pr<strong>of</strong>essor, Department <strong>of</strong> Forestry, Range & Wildlife, Utah State<br />

University, Logan, UT 84322, USA, email: mlwolfe@cc.usu.edu<br />

Abstract: Ecologists have long debated whe<strong>the</strong>r predators (“top-down”) or nutrients/food<br />

(“bottom-up”) limit prey populations. Evidence supporting <strong>the</strong> importance <strong>of</strong> predation is<br />

frequently based on <strong>the</strong> number <strong>of</strong> prey killed by predators – a direct effect. By examining only<br />

this direct effect many predation studies fail to consider behavioral changes arising from <strong>the</strong> risk<br />

<strong>of</strong> predation - indirect effects. Fur<strong>the</strong>rmore, <strong>the</strong>se behavioral indirect effects can be more<br />

important than <strong>the</strong> direct effect <strong>of</strong> predator-caused mortality, influencing both top-down and<br />

bottom-up processes. In this study we capitalize on a “natural experiment” on a suite <strong>of</strong> large<br />

mammalian herbivores, in a system (National Bison Range, MT) where <strong>the</strong> behavior and<br />

population dynamics <strong>of</strong> ungulate prey species (whitetail deer, Odocoileus virginianus; mule deer,<br />

O. hemionus; elk, Cervus elaphus) can be compared before and after an increase in risk <strong>of</strong><br />

predation by cougar (Puma concolor). We present preliminary data demonstrating that cougars<br />

can influence several aspects <strong>of</strong> prey behavior. With an increase in predation risk, mule deer and<br />

elk total daily activity time has declined by 35.9% and 31.8% (P


126<br />

COUGAR PREDATION ON PREY IN YELLOWSTONE NATIONAL PARK: A<br />

PRELIMINARY COMPARISON PRE- AND POST-WOLF REESTABLISHMENT<br />

TONI K. RUTH, Wildlife Conservation Society, 2023 Stadium Dr. Suite 1A, Bozeman, MT<br />

59030, USA, email: truth@montanadsl.net<br />

POLLY C. BUOTTE, Wildlife Conservation Society, 2023 Stadium Dr. Suite 1A, Bozeman, MT<br />

59030, USA, email: polly_thornton@hotmail.com<br />

KERRY M. MURPHY, Yellowstone Center for Resources, P.O. Box 168, Yellowstone National<br />

Park, Mammoth, WY 89210, USA, email: kerry_murphy@nps.gov<br />

MAURICE G. HORNOCKER, Wildlife Conservation Society, 2023 Stadium Dr. Suite 1A,<br />

Bozeman, MT 59030, USA<br />

Abstract: On Yellowstone National Park’s Nor<strong>the</strong>rn Range cougars and wolves rely on<br />

economically important prey species, particularly elk. Understanding how <strong>the</strong>se large carnivores<br />

partition prey resources and <strong>the</strong>ir combined affect on prey is important for management and<br />

conservation <strong>of</strong> cougars, wolves, and ungulate species. As part <strong>of</strong> a cougar-wolf interactions<br />

study, we quantified predation rates and prey selection by cougars on Yellowstone’s nor<strong>the</strong>rn<br />

range prior to (Phase I) and post wolf (Phase II) reestablishment. During Phase II, cougars spent<br />

an average <strong>of</strong> 3.7 days at kills and 4.4 days between each kill. The mean annual rate <strong>of</strong> cougar<br />

predation in Phase I was 9.4 (SD = 4.0; 95% CI = 7.8 to 11.0) days per ungulate kill, and<br />

10.9(SD = 8.5; 95% CI = 6.7 to 15.1) days per ungulate kill in Phase II. Rate <strong>of</strong> predation varied<br />

by cougar social class. When converted to biomass killed per day, cougars averaged 12.2 kg per<br />

day during Phase I and 12.9 kg per day during Phase II. We documented a total <strong>of</strong> 306 and 256<br />

positive and probable cougar kills during Phase I and Phase II, respectively. During Phase II,<br />

70% (n = 179) <strong>of</strong> cougar kills were elk, 17% (n = 43) were mule deer and 13% (n = 34) were<br />

o<strong>the</strong>r prey. During both Phase I and II more than 50% <strong>of</strong> cougar kills were elk calves, with cow<br />

elk making up <strong>the</strong> next largest category. During Phase I, cougar predation was nei<strong>the</strong>r a major<br />

source <strong>of</strong> mortality nor a significant factor limiting <strong>the</strong> numbers or growth rates <strong>of</strong> elk and mule<br />

deer populations in nor<strong>the</strong>rn Yellowstone. Cougars present on <strong>the</strong> study area killed 2-3% <strong>of</strong> <strong>the</strong><br />

elk and 3-5% <strong>of</strong> <strong>the</strong> mule deer estimated to be available during 5 years spanning <strong>the</strong> Phase I<br />

study. Simultaneous to our Phase II study, <strong>the</strong> Yellowstone Wolf Project quantifies wolf<br />

predation rates and prey selection. Cougars killed proportionally more elk calves and fewer bull<br />

elk than wolves between 1998 and 2002. We are continuing our data collection and analyses and<br />

plan to: 1) compare cougar and wolf per capita rate <strong>of</strong> predation, 2) contrast femur marrow fat<br />

content <strong>of</strong> cougar and wolf kills, by season killed and prey age, and 3) compare yearly <strong>of</strong>f-take<br />

<strong>of</strong> elk and mule deer by cougars and wolves. Cougar per capita predation rate averaged across<br />

social classes was 0.06 kills/cougar/day. When kittens were included with maternal females, that<br />

group had <strong>the</strong> lowest predation rate <strong>of</strong> 0.01. Without including kittens, maternal females<br />

averaged 0.15. Subadult males had an equally high rate <strong>of</strong> 0.15. Wolf predation ranged from<br />

0.03 to 0.078 kills per wolf per day (Smith et al., In Press).<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


FOUR DECADES OF COUGAR-UNGULATE RELATIONSHIPS IN THE CENTRAL<br />

IDAHO WILDERNESS<br />

HOLLY A. AKENSON, University <strong>of</strong> Idaho, Taylor Ranch Field Station, HC 83 Box 8070,<br />

Cascade, ID 83611, USA, e-mail: tayranch@direcpc.com<br />

JAMES J. AKENSON, University <strong>of</strong> Idaho, Taylor Ranch Field Station, HC 83 Box 8070,<br />

Cascade, ID 83611, USA; e-mail: tayranch@direcpc.com<br />

HOWARD B. QUIGLEY, Beringia South, 2023 Stadium Drive, Suite 1A, Bozeman, MT<br />

59715, USA<br />

MAURICE G. HORNOCKER, Wildlife Conservation Society, 2023 Stadium Drive, Suite 1A,<br />

Bozeman, MT 59715, USA<br />

Abstract: Research conducted on cougars (Puma concolor) in <strong>the</strong> Big Creek drainage in each <strong>of</strong><br />

<strong>the</strong> last four decades has enhanced <strong>the</strong> understanding <strong>of</strong> <strong>the</strong> dynamic nature <strong>of</strong> cougar – ungulate<br />

relationships. In 1964, Maurice Hornocker initiated his benchmark research on this cougar<br />

population and assessed <strong>the</strong> role <strong>of</strong> cougar predation in regulating ungulate populations. Each<br />

study that followed has had different objectives, yet, combined <strong>the</strong>se projects provide a rare<br />

continuum <strong>of</strong> ecological information on <strong>the</strong> dynamics <strong>of</strong> cougar – prey relationships. This<br />

cougar population has been influenced by significant environmental changes over <strong>the</strong> last 40<br />

years. The ungulate prey base has fluctuated, but generally elk numbers have increased and deer<br />

have decreased. Total ungulate biomass was similar in <strong>the</strong> 1960’s and 1980’s, but was 12%<br />

lower in <strong>the</strong> study just completed. The dynamics <strong>of</strong> carnivore competition, both inter-specific<br />

and intra-specific, has changed since introduced wolves recolonized <strong>the</strong> drainage in <strong>the</strong> 1990s. A<br />

large-scale forest fire 2 years ago drastically altered winter and summer ranges and affected<br />

predator – prey relationships. We compared cougar population size, structure, reproduction, and<br />

mortality factors; prey selection during 3 time periods; and evaluated pre and post-fire data in <strong>the</strong><br />

recent study. The estimated resident cougar population was 9 adults during <strong>the</strong> first 2 studies in<br />

<strong>the</strong> 1960’s and early 1970’s. The resident population grew to an estimated 13 adults in <strong>the</strong> mid-<br />

1980’s, but dropped to 10 individuals by 2000, and down to 6 resident cougars by 2002. The<br />

population increase during <strong>the</strong> 1980’s was in <strong>the</strong> adult female segment and it corresponded with<br />

an increasing elk population. The current low population is a result <strong>of</strong> a decreasing elk<br />

population, ungulate displacement from fire, increased hunter harvest <strong>of</strong> cougars, increased<br />

intraspecific strife, and competition with wolves for <strong>the</strong> same prey base. Cougars selected for<br />

elk ra<strong>the</strong>r than mule deer during <strong>the</strong> first study, but killed elk in proportion to <strong>the</strong>ir relative<br />

abundance during <strong>the</strong> study in <strong>the</strong> 1980’s and recent study (2000). Historical perspectives from<br />

pioneer diaries indicate similar cougar population numbers. In 1888 a bounty hunter removed 12<br />

cougars from <strong>the</strong> drainage, <strong>the</strong>n ten years later a different cougar hunter noted trapping and<br />

poisoning 12 individuals on Big Creek. Archeological evidence, old newspaper articles and<br />

diaries, and early agency field notes are all integrated into this discussion <strong>of</strong> long-term predator -<br />

prey relationships. The lengthy record <strong>of</strong> information on predator and prey populations in <strong>the</strong><br />

Big Creek drainage arguably makes this cougar population <strong>the</strong> best understood in North<br />

America.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

127


128<br />

COUGAR TOTAL PREDATION RESPONSE TO DIFFERING PREY DENSITIES: A<br />

PROPOSED EXPERIMENT TO TEST THE APPARENT COMPETITION<br />

HYPOTHESIS<br />

HUGH ROBINSON, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99614-6410,<br />

USA, email: hsrobins@wsunix.wsu.edu<br />

ROBERT WIELGUS, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99614-6410,<br />

USA, email: wielgus@wsu.edu<br />

HILARY CRUICKSHANK, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, PO Box 646410, Pullman, WA 99614-<br />

6410, USA, email: hcruicks@mail.wsu.edu<br />

CATHERINE LAMBERT, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99614-6410,<br />

USA, email: lambertcath@wsu.edu<br />

Abstract: Mule deer populations throughout <strong>the</strong> west are declining whereas white-tailed deer<br />

populations are increasing. We compared abundance, fetal rate, recruitment rate, and causespecific<br />

adult (≥1 yr. old) mortality rates <strong>of</strong> sympatric mule and white-tailed deer in south-central<br />

British Columbia to assess <strong>the</strong> population growth <strong>of</strong> each species. White-tailed deer were three<br />

times as abundant (908±152) as mule deer (336±122) (± 1SE). Fetal rates <strong>of</strong> white-tailed deer<br />

(1.83) were similar to mule deer (1.78) (t = 0.15, df = 13, P = 0.44) as was recruitment <strong>of</strong> whitetailed<br />

deer (56 fawns:100 does) and mule deer (38 fawns:100 does) (χ 2 = 0.91, df = 1, P=0.34).<br />

Annual adult white-tailed deer survival (SWT = 0.81) was significantly higher (z = 1.32, df = 1, P<br />

= 0.09) than mule deer survival (SMD = 0.72). The main source <strong>of</strong> mortality in both populations<br />

was cougar predation. The lower survival rate <strong>of</strong> mule deer could be directly linked to a higher<br />

predation rate (0.17) compared to white-tailed deer (0.09) (z = 1.57, df = 1, P = 0.06). The finite<br />

growth rate (λ) <strong>of</strong> mule deer was 0.88 and 1.02 for white-tailed deer. We suggest that <strong>the</strong><br />

disparate survival and predation rates are caused by apparent competition between <strong>the</strong> two deer<br />

species, facilitated through a shared predator; cougar. The apparent competition hypo<strong>the</strong>sis<br />

predicts that as alternate prey (white-tailed deer) densities increase, so do densities <strong>of</strong> predators,<br />

resulting in increased incidental predation on sympatric native prey (mule deer). Apparent<br />

competition can result in population declines and even extirpation <strong>of</strong> native prey in some cases.<br />

Such a phenomenon may account for declines <strong>of</strong> mule deer throughout <strong>the</strong> arid and semi-arid<br />

West where irrigation agriculture is practiced. We are in year two <strong>of</strong> a proposed five-year study.<br />

We will test <strong>the</strong> apparent competition hypo<strong>the</strong>sis by conducting a controlled, replicated “press”<br />

experiment in 2 treatment and 2 control areas in North-eastern Washington by reducing densities<br />

<strong>of</strong> white-tailed deer and observing any changes in cougar predation on mule deer. Washington<br />

Fish and Wildlife personnel using annual aerial surveys and/or o<strong>the</strong>r trend indices will monitor<br />

deer densities. Predation rates and population growth rates <strong>of</strong> deer will be determined using radio<br />

telemetry. Changes in cougar functional (kills/unit time), aggregative (cougars/unit area),<br />

numerical (<strong>of</strong>fspring/cougar), and total (predation rate) responses on deer will also be monitored<br />

using radio telemetry. Results will be used to determine <strong>the</strong> effect <strong>of</strong> increased white-tailed<br />

densities on cougar predation <strong>of</strong> mule deer.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


CHARACTERISTICS OF COUGAR HARVEST WITH AND WITHOUT THE USE OF<br />

DOGS<br />

DONALD A. MARTORELLO, Washington Department <strong>of</strong> Fish and Wildlife, 600 Capitol Way<br />

North, Olympia, WA 98501-1091, USA, email: martodam@dfw.wa.gov<br />

RICHARD A. BEAUSOLEIL, Washington Department <strong>of</strong> Fish and Wildlife, 3515 State<br />

Highway 97A, Wenatchee, WA 98801, USA, email: beausrab@dfw.wa.gov<br />

Abstract: Prior to 1996, dogs were used to harvest <strong>the</strong> majority (99%) <strong>of</strong> cougar during recreational hunting seasons<br />

in Washington State. However, in 1996 Voter Initiative 655 banned <strong>the</strong> use <strong>of</strong> dogs to aid in <strong>the</strong> harvest <strong>of</strong> cougar.<br />

As a result, harvest methods shifted to spot and stalk, predator calling, and incidental encounters between deer and<br />

elk hunters and cougar. We examined <strong>the</strong> sex and age structure <strong>of</strong> harvested cougar and compared <strong>the</strong>se data<br />

between seasons with (selective harvest) and without <strong>the</strong> use <strong>of</strong> dogs (non-selective harvest). We detected a<br />

significant increase in percent female cougars in <strong>the</strong> total harvest, from 42% to 59% during selective versus nonselective<br />

seasons (T = -7.85, P < 0.0001). We also found that non-selective harvest seasons had significantly more<br />

2<br />

2<br />

juvenile male ( χ = 98.1790, d.f. = 10, P < 0.0001) and female ( χ = 66.5116, d.f. = 10, P < 0.0001) cougars<br />

compared to selective seasons. We <strong>the</strong>n used program RISKMAN to evaluate <strong>the</strong> potential impacts to population<br />

growth (finite rate <strong>of</strong> increase) from changes we observed in harvest vulnerability <strong>of</strong> specific sex and age classes.<br />

Our sensitivity analysis suggests that changes in female adult and cub survival are <strong>the</strong> most influential parameters to<br />

population growth and <strong>the</strong> increased harvest <strong>of</strong> female cougars in non-selective harvest methods decreased <strong>the</strong> finite<br />

rate <strong>of</strong> increase by about 0.01–0.02. Harvest methods that increase <strong>the</strong> relative harvest vulnerability <strong>of</strong> <strong>the</strong>se cohorts<br />

have a greater potential for impacting population growth. In Washington State, <strong>the</strong> current level <strong>of</strong> cougar harvest<br />

and increased vulnerability <strong>of</strong> females and juvenile cougar have likely increased <strong>the</strong> risk <strong>of</strong> impacting population<br />

growth.<br />

129<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

Key Words: Voter Initiative 655, selective harvest, non-selective harvest, Washington, cougar, Puma concolor,<br />

RISKMAN<br />

INTRODUCTION<br />

In Washington State, cougar (Puma<br />

concolor) hunting regulations and harvest<br />

levels have changed dramatically during <strong>the</strong><br />

last decade (Table 1). Pivotal to <strong>the</strong>se<br />

changes was Voter Initiative 655 (I-655),<br />

which banned <strong>the</strong> use <strong>of</strong> dogs for hunting<br />

cougar in Washington in 1996. Prior to I-<br />

655, hunters used dogs to take about 99% <strong>of</strong><br />

<strong>the</strong> harvested cougar. Recognizing that<br />

using dogs was more effective than spotand-stalk<br />

or predator calling methods to<br />

harvest cougar, Washington Department <strong>of</strong><br />

Fish and Wildlife (WDFW) instituted<br />

several changes in an effort to maintain<br />

harvest levels similar to pre-initiative<br />

seasons; including leng<strong>the</strong>ning <strong>the</strong> season<br />

from about 86 to 227 days, increasing <strong>the</strong><br />

annual bag limit from 1 to 2 cougars/hunter,<br />

and decreasing <strong>the</strong> cost <strong>of</strong> a cougar transport<br />

tag from $24 to $10.<br />

Both prior to and since I-655, wildlife<br />

managers in Washington used harvest<br />

characteristics to assess <strong>the</strong> status <strong>of</strong> <strong>the</strong><br />

living cougar population. Managers<br />

monitored trends in total harvest, harvest<br />

success, percent females in <strong>the</strong> harvest, and<br />

median ages <strong>of</strong> harvested males and<br />

females. Using harvest information alone to<br />

assess a living population is fraught with<br />

problems. The value <strong>of</strong> harvest information<br />

becomes even more suspect when <strong>the</strong><br />

harvest methods change, because <strong>the</strong><br />

relationship between trend data and <strong>the</strong>


130 CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil<br />

Table 1. Recreational cougar hunting seasons in Washington, 1990-2001.<br />

Year Season Dates Days Harvest Restrictions Selectivity<br />

1990-1991 Nov. 21 – Jan. 15 57 102 Permit only season Dogs allowed<br />

1991-1992 Nov. 27 – Jan. 15 50 120 Permit only season Dogs allowed<br />

1992-1993 Oct. 17 – Jan. 31 107 140 Permit only season Dogs allowed<br />

1993-1994 Oct. 16 – Jan. 31 108 121 Permit only season Dogs allowed<br />

1994-1995 Oct. 15 – Jan. 31 109 177 Permit only season Dogs allowed<br />

1995-1996 Oct. 14 – Jan. 31 110 283 Permit only season Dogs allowed<br />

1996-1997 Oct. 12 – Mar. 15 155 178 General season No dogs allowed<br />

1997-1998 Aug. 1 – Mar. 15 227 132 General season No dogs allowed<br />

1998-1999 Aug. 1 – Mar. 15 227 184 General season No dogs allowed<br />

1999-2000 Aug. 1 – Mar. 15 227 273 General season No dogs allowed<br />

2000-2001 Aug. 1 – Mar. 15 227 208 General season No dogs allowed<br />

living population also may have changed.<br />

As a result, trend information based on<br />

harvest data may be <strong>of</strong> limited value in<br />

Washington because <strong>of</strong> <strong>the</strong> restriction on<br />

using dogs to hunt cougar.<br />

Little information is known about how<br />

restrictions on using dogs to hunt cougar<br />

impacts harvest, trends based on harvest<br />

data, or cougar populations. To that end,<br />

our objectives are to: 1) determine if <strong>the</strong>re<br />

are differences in cougar harvest<br />

characteristics (i.e., total harvest, percent<br />

female cougar in <strong>the</strong> harvest, and age<br />

structure <strong>of</strong> harvest cougar) during seasons<br />

with and without <strong>the</strong> use <strong>of</strong> dogs, and 2)<br />

identify potential impacts to <strong>the</strong> cougar<br />

population given changes in season structure<br />

and harvest levels in Washington between<br />

1990-2001.<br />

METHODS<br />

We obtained cougar harvest data from<br />

1990 to 2001 through a mandatory harvest<br />

reporting system implemented by WDFW.<br />

Successful hunters were required to report<br />

<strong>the</strong>ir harvested cougar and present <strong>the</strong> hide<br />

and skull to WDFW, where Agency staff<br />

collected a tooth sample, and documented<br />

sex, kill location, kill date, and kill type<br />

(depredation, recreational, public safety<br />

cougar removal, or o<strong>the</strong>r). We determined<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

ages for 79% <strong>of</strong> <strong>the</strong> harvested cougar using<br />

cementum annuli analysis.<br />

We classified harvest data from 1990-<br />

1995 as “selective” and from 1996-2001 as<br />

“non-selective” to compare harvest<br />

characteristics between periods when dogs<br />

were legal to years when <strong>the</strong>y were not,<br />

respectively. We used a T-test to determine<br />

if cougar harvest (recreational only) and<br />

percent females in <strong>the</strong> harvest (recreational<br />

only) differed between selective versus nonselective<br />

periods. To compare harvest age<br />

structures between selective and nonselective<br />

periods, we generated a mean age<br />

structure for each sex and each period by<br />

pooling cougar management units (CMUs)<br />

(Washington is divided into 9 administrative<br />

CMUs) and years. We <strong>the</strong>n used a chisquare<br />

test to determine if mean age<br />

structures for each sex differed between<br />

selective versus non-selective periods. We<br />

considered all tests significant at α ≤ 0.<br />

05 .<br />

We used program RISKMAN to assess<br />

potential impacts to <strong>the</strong> cougar population<br />

from changes in harvest methods and rates<br />

(Taylor et al. 2002). We incorporated<br />

parameter estimates from Logan and<br />

Sweanor (2001) and Spencer et al. (2001)<br />

(Table 2). For initial population size we<br />

multiplied <strong>the</strong> amount <strong>of</strong> suitable cougar<br />

habitat in Washington (88,497 km 2 ) by <strong>the</strong>


CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil 131<br />

Table 2. Parameter inputs for program RISKMAN simulations in Washington.<br />

Parameter Value Source<br />

Recruitment Probability <strong>of</strong> 1 cub 0.00 Logan and Sweanor 2001<br />

Probability <strong>of</strong> 2 cubs 0.25 Logan and Sweanor 2001<br />

Probability <strong>of</strong> 3 cubs 0.49 Logan and Sweanor 2001<br />

Probability <strong>of</strong> 4 cubs 0.26 Logan and Sweanor 2001<br />

Mean litter size 3.01 Logan and Sweanor 2001<br />

Proportion <strong>of</strong> females with litters 0.80 Logan and Sweanor 2001<br />

Proportion <strong>of</strong> males at birth 0.50 Logan and Sweanor 2001<br />

Survival Male cubs (age 0) 0.67 Logan and Sweanor 2001<br />

Female cubs (age 0) 0.67 Logan and Sweanor 2001<br />

Male yearling (age 1-2) 0.64 Logan and Sweanor 2001<br />

Female yearlings (age 1-2) 0.88 Spencer et al. 2001<br />

Male adults (age 3-12) 0.91 Logan and Sweanor 2001<br />

Female adults (age 3-12) 0.82 Logan and Sweanor 2001<br />

Litter survival rate 0.93 Logan and Sweanor 2001<br />

Population size 4159 WDFW a , unpublished data<br />

a<br />

Washington Department <strong>of</strong> Fish and Wildlife<br />

highest cougar density reported in <strong>the</strong><br />

literature (4.7 cougars/100 km 2 ) (Ross and<br />

Jalkotzy 1992). Each <strong>of</strong> <strong>the</strong> parameters<br />

values, including density, are at <strong>the</strong> high end<br />

in terms <strong>of</strong> <strong>the</strong> range reported in <strong>the</strong><br />

literature. We chose to model a population<br />

with high productivity, survival, and density<br />

so our analysis would reflect how harvest<br />

strategies and rates might impact even <strong>the</strong><br />

most robust cougar population. Therefore,<br />

potential impacts to a more realistic<br />

population would be at least as great, if not<br />

more, than those to our modeled population.<br />

We used <strong>the</strong> age structure from cougars<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

harvested between 1997-1999 to develop <strong>the</strong><br />

input age distribution for <strong>the</strong> model. We<br />

selected <strong>the</strong>se years because <strong>the</strong>y probably<br />

better reflect <strong>the</strong> standing population than<br />

years when dogs were legal or years when<br />

potential impacts may begin to be apparent.<br />

From <strong>the</strong> average age structure we generated<br />

a stable age distribution and used this as <strong>the</strong><br />

age distribution in <strong>the</strong> model.<br />

To assess <strong>the</strong> impacts <strong>of</strong> selective versus<br />

non-selective harvest methods, we<br />

conducted two simulations; one where <strong>the</strong><br />

strata specific harvest vulnerabilities were<br />

set to levels observed during <strong>the</strong> selective<br />

Table 3. Vulnerability inputs for selective and non-selective model simulations in Washington.<br />

Relative vulnerability<br />

Parameter Age class Selective Non-selective<br />

Cubs (age 0) 0.000 1.0<br />

Males<br />

Females without <strong>of</strong>fspring<br />

Juveniles (age 1-2) 0.118 1.0<br />

Adults (age 3-12) 0.464 1.0<br />

Cubs (age 0) 0.001 1.0<br />

Juveniles (age 1-2) 0.109 1.0<br />

Adults (age 3-12) 0.277 1.0<br />

Females with <strong>of</strong>fspring Adults (age 3-12) 0.031 1.0


132 CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil<br />

period (1990-1995) and one where <strong>the</strong><br />

vulnerabilities were constant across strata,<br />

<strong>the</strong>reby mimicking a non-selective method<br />

(1996-2001) (Table 3). We repeated each<br />

simulation for harvest rates ranging from 0%<br />

(no harvest) to 14% <strong>of</strong> <strong>the</strong> censused<br />

population. All simulations were<br />

deterministic and <strong>the</strong> realized finite rate <strong>of</strong><br />

increase with no harvest was 1.062.<br />

RESULTS<br />

Mean harvest increased from 157 during<br />

<strong>the</strong> selective period to 199 in <strong>the</strong> nonselective<br />

period; however <strong>the</strong> increase was<br />

not statistically significant (T = -1.26, P =<br />

0.2368). We detected a significant increase<br />

in percent females in <strong>the</strong> harvest (T = -7.85,<br />

P < 0.0001). Percent females in <strong>the</strong> harvest<br />

increased from 42 to 59% during selective<br />

versus non-selective periods, respectively<br />

(Figure 1). We detected a significant<br />

difference in male age distributions for<br />

2<br />

selective versus non-selective periods ( χ =<br />

98.1790, d.f. = 10, P < 0.0001). On a<br />

statewide level, <strong>the</strong>re were a greater<br />

proportion <strong>of</strong> younger males in <strong>the</strong> harvest<br />

during <strong>the</strong> non-selective period. Although<br />

less pronounced, <strong>the</strong>re also were a lower<br />

proportion <strong>of</strong> adult males in <strong>the</strong> non-<br />

Recreational harvest<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

1990<br />

1991<br />

1992<br />

1993<br />

1994<br />

1995<br />

1996<br />

1997<br />

1998<br />

1999<br />

2000<br />

2001<br />

Year<br />

Recreational harvest % Female in harvest<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Figure 1. Recreational cougar harvest and<br />

percent female in <strong>the</strong> harvest during selective<br />

(1990-1995) and non-selective (1996-2001)<br />

years, Washington, 1990-2001.<br />

% Female in harvest<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

selective period (Figure 2). We detected a<br />

significant difference in female age<br />

distributions for selective versus non-<br />

2<br />

selective periods ( χ = 66.5116, d.f. = 10, P<br />

< 0.0001). On a statewide level, <strong>the</strong>re also<br />

were a greater proportion <strong>of</strong> younger<br />

females in <strong>the</strong> harvest during <strong>the</strong> nonselective<br />

period (Figure 2).<br />

Using model simulations, we found that<br />

changing from a selective harvest to a nonselective<br />

harvest (with all o<strong>the</strong>r parameters<br />

held constant) decreased <strong>the</strong> population’s<br />

finite rate <strong>of</strong> increase by about 0.01–0.02<br />

annually (Figure 3). This is roughly<br />

equivalent to about a 1.5% increase in<br />

harvest rate (proportion <strong>of</strong> population<br />

harvested).<br />

Frequency<br />

Frequency<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

A.<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 16 17<br />

Age<br />

B.<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19<br />

Age<br />

1990-1995 1996-2001<br />

Figure 2. Age structures <strong>of</strong> harvested male<br />

(A) and female (B) cougar during selective<br />

(1990-1995) and non-selective (1996-2001)<br />

periods, Washington, 1990-2001.


Finite rate <strong>of</strong> increase<br />

1.08<br />

1.06<br />

1.04<br />

1.02<br />

1.00<br />

0.98<br />

0.96<br />

0.94<br />

0.92<br />

0.90<br />

Non-selective<br />

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14<br />

Harvest rate<br />

Figure 3. Relationship between harvest rate<br />

and finite rate <strong>of</strong> increase for selective and<br />

non-selective harvest strategies.<br />

DISCUSSION<br />

In Washington, when dogs were legal<br />

for hunting cougar, seasons primarily<br />

occurred from October to January. In<br />

contrast, when dogs were banned, season<br />

length increased and, more importantly,<br />

overlapped with deer and elk seasons. The<br />

reduced cougar tag and overlapping seasons<br />

made purchasing a cougar tag more<br />

attractive for deer or elk hunters, and<br />

licensed cougar hunters increased from less<br />

than 1,000 annually prior to I-655 to about<br />

58,000 post I-655. This in turn created a<br />

situation where <strong>the</strong> majority <strong>of</strong> <strong>the</strong> harvest<br />

was by deer and elk hunters that took a<br />

cougar incidentally during <strong>the</strong>ir deer or elk<br />

CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil 133<br />

Selective<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

hunt. We believe season timing, tag cost,<br />

and <strong>the</strong> large number <strong>of</strong> deer and elk hunters<br />

resulted in post I-655 harvest levels that<br />

were similar to pre I-655 levels.<br />

The proportional increase <strong>of</strong> females and<br />

juveniles in <strong>the</strong> harvest during years with<br />

and without dogs probably was correlated<br />

with <strong>the</strong> proportions <strong>of</strong> <strong>the</strong>se cohorts in <strong>the</strong><br />

standing population. That is, individual sex<br />

and age classes <strong>of</strong> cougar were probably<br />

taken relative to <strong>the</strong>ir availability.<br />

Deviations between sex and age specific<br />

proportions in <strong>the</strong> population versus harvest<br />

probably were influenced by cougar and<br />

hunter distributions, cougar behavioral<br />

patterns and home range sizes, and prey<br />

distribution.<br />

Our model simulations suggest that<br />

population growth is sensitive to female<br />

survival. The importance <strong>of</strong> female survival<br />

for population growth has been well<br />

documented (Clark 1999). A simple<br />

sensitivity analysis done by halving each<br />

parameter one-by-one and documented <strong>the</strong><br />

percent decline in lambda fur<strong>the</strong>r illustrates<br />

that female survival is <strong>the</strong> most influential<br />

parameter on lambda (Table 4). It <strong>the</strong>refore<br />

seems reasonable that changing from<br />

selective to non-selective harvest methods<br />

can impact a population’s growth rate. Our<br />

model, although crude, suggests that <strong>the</strong><br />

Table 4. Percent decrease in λ when cougar parameter values are divided by two. Realized<br />

λ =1.062 with original parameter values (see Table 2 for parameter estimate sources) (adapted<br />

from Clark 1999).<br />

Actual Reduced<br />

Percent decrease<br />

Parameter<br />

estimate estimate λ<br />

in λ<br />

Female survival 0.82 0.41 0.73 31.3<br />

Female juvenile survival 0.88 0.44 0.84 20.9<br />

Cub survival 0.67 0.34 0.94 11.5<br />

Litter size 3.01 1.50 0.94 11.5<br />

Litter survival 0.93 0.47 0.96 9.6<br />

Prop. <strong>of</strong> females with litters 0.80 0.40 0.97 8.7<br />

Male adult survival 0.91 0.46 1.06 0.0<br />

Male juvenile survival 0.64 0.32 1.06 0.0


134 CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil<br />

impact may be in <strong>the</strong> range <strong>of</strong> increasing<br />

lambda by 0.01–0.02. In more familiar<br />

terms, this is analogous to a 1.5% increase in<br />

harvest rate. At first glance this appears<br />

small, but in our example population a 1.5%<br />

increase in harvest rate equals a 38%<br />

increase in observed harvest level. So<br />

changing to a non-selective harvest method<br />

was biologically equivalent to increasing <strong>the</strong><br />

harvest by about 38%.<br />

Of course, our model and <strong>the</strong><br />

corresponding impacts to population growth<br />

assume that harvest is constant between<br />

selective and non-selective harvest methods.<br />

This is probably not a reasonable<br />

assumption unless season adjustments are<br />

made to mitigate <strong>the</strong> inefficiencies <strong>of</strong> boot<br />

hunting (e.g., spot and stalk, predator<br />

calling, and incidental take) for killing<br />

cougar. Without season changes, similar to<br />

those we discussed earlier, harvest would<br />

probably decline significantly without <strong>the</strong><br />

use <strong>of</strong> dogs. The population growth rate<br />

would likely increase accordingly until<br />

density dependence occurred.<br />

We also assumed <strong>the</strong>re was no<br />

immigration or emigration occurring. Logan<br />

and Sweanor (2001) found that immigration<br />

could have as large <strong>of</strong> an impact to<br />

population growth as reproduction. This<br />

suggests that immigration could potentially<br />

counter decreases in lambda that resulted<br />

from increased female harvest. The effect <strong>of</strong><br />

immigration acting in this manor decreases<br />

as <strong>the</strong> perimeter-to-area ratio <strong>of</strong> <strong>the</strong><br />

population boundary decreases.<br />

MANAGEMENT IMPLICATIONS<br />

Several factors probably influence<br />

whe<strong>the</strong>r changing from a selective to nonselective<br />

harvest method will cause a cougar<br />

population to decline. In addition to<br />

parameter estimates and standard errors,<br />

knowing <strong>the</strong> age structure and growth rate<br />

are essential for predicting how an actual<br />

population might respond to changes in<br />

harvest vulnerability (Caughley 1977).<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Unfortunately, collecting biological data on<br />

cougar population dynamics is difficult and<br />

costly because cougars are extremely<br />

secretive, difficult to count, have large home<br />

ranges, and occur at relatively low densities.<br />

As such, wildlife managers <strong>of</strong>ten use harvest<br />

information as a surrogate to data on <strong>the</strong><br />

living population. The risk <strong>of</strong> using harvest<br />

information is more acceptable when <strong>the</strong><br />

majority <strong>of</strong> <strong>the</strong> harvest is males, because<br />

male survival has a relatively small impact<br />

on population growth. In contrast, <strong>the</strong> risk<br />

<strong>of</strong> using harvest data to guide management<br />

decisions is higher when <strong>the</strong> majority <strong>of</strong> <strong>the</strong><br />

harvest is females, because female survival<br />

has a greater impact on population growth.<br />

ACKNOWLEDGMENTS<br />

We thank J. Beecham and K. Logan for<br />

reviewing earlier versions <strong>of</strong> our model<br />

simulations. This study received funding<br />

from Federal Aid in Wildlife Restoration,<br />

Project W-97-R.<br />

LITERATURE CITED<br />

CAUGHLEY, G. 1977. Analysis <strong>of</strong> vertebrate<br />

populations. John Wiley and Sons, New<br />

York, New York, USA.<br />

CLARK, J.D. 1999. Black bear population<br />

dynamics in <strong>the</strong> Sou<strong>the</strong>ast: some new<br />

perspectives on some old problems.<br />

Eastern Black Bear <strong>Workshop</strong><br />

<strong>Proceedings</strong> 15:97-115.<br />

LOGAN, K.A. AND L.L. SWEANOR. 2001.<br />

Desert puma: evolutionary ecology and<br />

conservation <strong>of</strong> an enduring carnivore.<br />

Island Press. Washington D. C., USA.<br />

ROSS, P.I., AND M.G. JALKOTZY. 1992.<br />

Characteristics <strong>of</strong> a hunted population <strong>of</strong><br />

cougars in southwestern Alberta.<br />

Journal <strong>of</strong> Wildlife Management 56:417-<br />

426.<br />

SPENCER, R.D., D.J. PIERCE, G.A.<br />

SCHIRATO, K.R. DIXON, AND C.B.<br />

RICHARDS. 2001. <strong>Mountain</strong> lion home<br />

range, dispersal, mortality and survival<br />

in <strong>the</strong> western Cascades <strong>Mountain</strong>s <strong>of</strong>


Washington. Final Report. Washington<br />

Department <strong>of</strong> Fish and Wildlife,<br />

Olympia, Washington, USA.<br />

CHARACTERISTICS OF COUGAR HARVEST · Martorello and Beausoleil 135<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

TAYLOR, M., M. OBBARD, B. POND, M.<br />

KUC, D. ABRAHAM. 2002. Riskman:<br />

user manual. The Queens Printer for<br />

Ontario, Ontario, Canada.


136<br />

DEFINING AND DELINEATING DE FACTO REFUGIA: A PRELIMIARY ANALYSIS<br />

OF THE SPATIAL DISTRIBUTION OF COUGAR HARVEST IN UTAH AND<br />

IMPLICATIONS FOR CONSERVATION<br />

DAVID C. STONER, Utah State University, Dept. <strong>of</strong> Forest, Range, and Wildlife Sciences,<br />

5230 Old Main Hill, Logan, UT 84322-5230, USA, email: dstoner@cc.usu.edu<br />

MICHAEL L. WOLFE, Utah State University, Dept. <strong>of</strong> Forest, Range, and Wildlife Sciences,<br />

5230 Old Main Hill, Logan, UT 84322-5230, USA, email: mlwolfe@cnr.usu.edu<br />

Abstract: Cougars (Puma concolor) in Utah are managed at two scales, <strong>of</strong>ten with differing<br />

objectives. The statewide population is managed for persistence and sustainable hunting<br />

opportunities, while at <strong>the</strong> finer scale <strong>of</strong> an individual management unit, a sub-population may be<br />

managed to accomplish density reductions, depending on local priorities. However, cougars<br />

have large and variable spatial requirements, and management unit boundaries may not coincide<br />

with actual demes. Compounding <strong>the</strong>se constraints, <strong>the</strong> lack <strong>of</strong> cost-effective and reliable<br />

enumeration techniques increases <strong>the</strong> risk <strong>of</strong> inadvertent over-harvest. Current research suggests<br />

that for species difficult to enumerate, greater emphasis be placed on metapopulation-scale<br />

management in order to minimize <strong>the</strong> effects <strong>of</strong> uncertainties with respect to demography and<br />

dispersal behavior. Because harvest is <strong>the</strong> primary variable that managers can manipulate and<br />

measure, it is important to understand how recruitment patterns in minimally exploited<br />

populations may influence <strong>the</strong> persistence and recovery <strong>of</strong> heavily exploited populations. In this<br />

paper we discuss some <strong>of</strong> <strong>the</strong> factors that account for <strong>the</strong> spatial distribution <strong>of</strong> harvest and how<br />

this information can be used to develop management strategies in <strong>the</strong> absence <strong>of</strong> census data.<br />

We used 6 years <strong>of</strong> radio-telemetry data from a lightly exploited population in <strong>the</strong> Oquirrh<br />

<strong>Mountain</strong>s <strong>of</strong> north-central Utah to quantify <strong>the</strong> effect <strong>of</strong> a small sanctuary (480 km²) on cougar<br />

survivorship, fecundity, and dispersal. We <strong>the</strong>n mapped <strong>the</strong> locations <strong>of</strong> cougars harvested<br />

across <strong>the</strong> state from 1996-2001, and attempted to: (1) identify <strong>the</strong> factors that influence <strong>the</strong>se<br />

patterns, and (2) determine <strong>the</strong> size and distribution <strong>of</strong> potential harvest sinks and de facto<br />

refugia in <strong>the</strong> state. Finally, we identified habitat patches on <strong>the</strong> periphery <strong>of</strong> <strong>the</strong> state that<br />

straddle management jurisdictions, representing areas <strong>of</strong> possible inter-state cooperation. We<br />

recommend that managers consider a metapopulation perspective and attempt to distribute<br />

harvest pressure in a spatially and ecologically relevant manner. In <strong>the</strong> absence <strong>of</strong> large (2400<br />

km²), contiguous refugia, small sanctuaries adjacent to areas <strong>of</strong> high exploitation may be mapped<br />

and utilized as a deterrent against potential over-exploitation.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


MONITORING CHANGES IN COUGAR SEX/AGE STRUCTURE WITH CHANGES IN<br />

ABUNDANCE AS AN INDEX TO POPULATION TREND<br />

CHUCK R. ANDERSON, JR., Wyoming Cooperative Fish and Wildlife Research Unit, Box<br />

3166, University Station, Laramie, WY 82071, USA, email: cander@uwyo.edu<br />

FRED G. LINDZEY, Wyoming Cooperative Fish and Wildlife Research Unit, Box 3166,<br />

University Station, Laramie, WY 82071 USA, email: flindzey@uwyo.edu<br />

Abstract: Cougar (Puma concolor) management has traditionally been plagued by <strong>the</strong> inability to<br />

identify population trends for adequate assessment <strong>of</strong> management strategies. Monitoring<br />

changes in harvest sex-age structure as an index to population trends appears useful in tracking<br />

black bear populations and may be applicable to cougar management, especially given <strong>the</strong>ir strict<br />

social structure and territorial behavior. We documented changes in cougar harvest structure<br />

(sex-age) through experimental population reduction and recovery to better understand <strong>the</strong><br />

relationship between sex-age composition and population trend in exploited populations. The<br />

cougar population in <strong>the</strong> Snowy Range, sou<strong>the</strong>ast Wyoming, declined from 58 (90% CI = 36 to<br />

81) in <strong>the</strong> fall <strong>of</strong> 1998 to 20 (90% CI = 14 to 26) independent cougars (>1 year old) by <strong>the</strong> spring<br />

<strong>of</strong> 2000 following 2 years <strong>of</strong> increased exploitation (mean exploitation rate = 43%) and increased<br />

to 46 (90% CI = 33 to 60) by <strong>the</strong> spring <strong>of</strong> 2003 following 3 years <strong>of</strong> reduced harvest levels<br />

(mean exploitation rate = 18%). Pre-treatment harvest composition was 63% subadults (1.0-2.5<br />

years old), 24% adult males, and 14% adult females (2 seasons; n = 22). A reduction in subadult<br />

harvest, an initial increase followed by a reduction in adult male harvest, and a steady increase in<br />

adult female harvest was consistent with hypo<strong>the</strong>sized harvest vulnerability for a declining<br />

population. Harvest composition was similar at high and low densities with light harvest, but <strong>the</strong><br />

proportion <strong>of</strong> male subadults increased at low density as adult males removed during <strong>the</strong><br />

treatment period (high harvest) were replaced. Examining cougar sex ratios (m:f) alone appears<br />

to be <strong>of</strong> limited utility for identifying population change. Including age class, however, provides<br />

a useful metric in monitoring cougar population trend. We feel this approach could be applied to<br />

adaptively manage cougar populations where adequate sex and age data are collected from<br />

harvested animals.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

137


138<br />

MANAGEMENT OF COUGARS (PUMA CONCOLOR) IN THE WESTERN UNITED<br />

STATES<br />

DEANNA DAWN, San Jose State University, Biology Department<br />

MICHEAL KUTILEK, San Jose State University, Biology Department<br />

RICK HOPKINS, Live Oak Associates, Inc.<br />

SELEHKA ANAND, San Jose State University, Biology Department<br />

STEVE TORRES, California Department <strong>of</strong> Fish and Game<br />

Abstract: In <strong>the</strong> U.S., cougar (Puma concolor) populations still exist in 13 western states. While<br />

sport hunting <strong>of</strong> cougars remains a management goal for 10 <strong>of</strong> <strong>the</strong>se states, <strong>the</strong>re is little<br />

information on how different hunting harvest strategies affect <strong>the</strong>ir biology. Both <strong>the</strong> rate <strong>of</strong><br />

harvest and <strong>the</strong> percentage <strong>of</strong> females in <strong>the</strong> harvest affect population stability. Therefore, <strong>the</strong><br />

purpose <strong>of</strong> this study was to examine <strong>the</strong> effect <strong>of</strong> different harvest strategies on <strong>the</strong> harvest rate<br />

and <strong>the</strong> percentage <strong>of</strong> females in <strong>the</strong> harvest. Annual hunting harvest records were requested<br />

from all 10 states and were summarized into a database for analysis. Harvest strategies that<br />

included female sub-quotas were associated with <strong>the</strong> lowest percentage <strong>of</strong> females removed,<br />

however <strong>the</strong>y also had some <strong>of</strong> <strong>the</strong> highest annual rates <strong>of</strong> harvest. These results suggest that, for<br />

some states, management strategies used in regulating sport hunting may <strong>of</strong>fer little protection<br />

against over-harvesting <strong>the</strong> population.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


DYNAMICS AND VIABILITY OF A COUGAR POPULATION IN THE PACIFIC<br />

NORTHWEST<br />

CATHERINE LAMBERT, Large Carnivore Conservation Laboratory, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, Pullman, WA 99164-6410, USA,<br />

email: cathlambert@mac.com<br />

ROBERT B. WIELGUS, Large Carnivore Conservation Laboratory, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, Pullman, WA 99164-6410, USA,<br />

email: wielgus@wsu.edu<br />

HUGH S. ROBINSON, Large Carnivore Conservation Laboratory, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, Pullman, WA 99164-6410, USA,<br />

email: hsrobins@wsunix.wsu.edu<br />

DONALD D. KATNIK, Large Carnivore Conservation Laboratory, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, Pullman, WA 99164-6410, USA<br />

HILARY CRUICKSHANK, Large Carnivore Conservation Laboratory, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, Pullman, WA 99164-6410, USA,<br />

email: hcruicks@mail.wsu.edu<br />

ROSS CLARKE, Columbia Basin Fish & Wildlife compensation program, 103-133 Victoria St.,<br />

Nelson, BC V1L 4K3, Canada<br />

Abstract: Cougar (Puma concolor) populations are believed to be at high density and increasing<br />

throughout western North America, especially in <strong>the</strong> Pacific Northwest, as evidenced by<br />

increasing cougars/humans encounters. Harvest rates have increased as a result. To test this<br />

hypo<strong>the</strong>sis, we determined <strong>the</strong> density, fecundity, survival, and growth rate <strong>of</strong> a cougar<br />

population in nor<strong>the</strong>astern Washington, northwestern Idaho, and sou<strong>the</strong>rn British Columbia.<br />

From 1998 to 2003, 52 cougars were captured, radio-collared, and monitored. We recorded<br />

fecundity through den site investigation and snow tracking, and mortality by weekly telemetry.<br />

Survival rates were estimated for kittens (0-1 yr), yearlings (1-2 yr), and adult (2-12 yr) males<br />

and females. Average overall density was 1.09 cougars/100km 2 or 0.46 adults/100km 2 . We<br />

estimated litter size at 2.53 kittens, birth interval at 18 months, proportion <strong>of</strong> reproductively<br />

successful females at 0.75, and age <strong>of</strong> first reproduction at 30 months, for a maternity rate <strong>of</strong><br />

0.63 male or female kitten/year/adult female. Average survival rate for all radio-collared<br />

cougars was 0.59, 0.77 for adult females, 0.44 for adult males, 0.37 for yearlings, and 0.57 for<br />

kittens. Hunting accounted for 92% <strong>of</strong> <strong>the</strong> mortalities <strong>of</strong> radio-collared cougars. Age- and sexspecific<br />

survival and fecundity were entered into a stochastic two-sex matrix model. We used<br />

computer simulations to determine <strong>the</strong> population stochastic growth rate and to assess its<br />

viability over 25 years. The annual stochastic growth rate <strong>of</strong> this population was λ = 0.80<br />

(95%CI = 0.11). Starting with a total initial abundance <strong>of</strong> 357, <strong>the</strong> median times to fall below a<br />

demographic collapse (N = 30 adults) and extirpation (N = 0) were 8.5 and 25.9 years. Our<br />

findings suggest that, contrary to popular belief, cougars in <strong>the</strong> Pacific Northwest are currently at<br />

low to moderate densities and are declining. Alternative hypo<strong>the</strong>ses may account for <strong>the</strong><br />

increased conflicts between cougars and humans in this area.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

139


140<br />

PROJECT CAT (COUGARS AND TEACHING): INTEGRATING SCIENCE, SCHOOLS<br />

AND COMMUNITY IN DEVELOPMENT PLANNING<br />

GARY M. KOEHLER, Wildlife Research Scientist, Washington Department <strong>of</strong> Fish and<br />

Wildlife, 600 Capitol Way N., Olympia, WA 98501 USA, email: koehlgmk@dfw.wa.gov<br />

EVELYN NELSON, Superintendent, Cle Elum-Roslyn School District, 2690 SR 903, Cle Elum,<br />

WA 98922, USA, email: Nelsone@cleelum.wednet.edu<br />

Abstract: Complaint reports <strong>of</strong> cougars (Puma concolor) venturing into urban areas, killing<br />

livestock and pets, and threatening humans have increased to more than 1,000 reports filed<br />

annually in Washington; where in <strong>the</strong> past 5 years cougars have mauled 2 children. Increased<br />

reports are coupled with human population increases <strong>of</strong> over 1 million in <strong>the</strong> past decade and an<br />

annual loss to development <strong>of</strong> over 28,000 ha <strong>of</strong> land. The rural Cle Elum-Roslyn community is<br />

experiencing similar growth and development with >5,000 ha <strong>of</strong> development and over 1,400<br />

new homes planned, but presently with few complaints <strong>of</strong> cougars. The winter <strong>of</strong> 2002-2003<br />

marked <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> 2 nd year <strong>of</strong> an 8-year scientific investigation on cougar and ecology<br />

by <strong>the</strong> Washington Department <strong>of</strong> Fish and Wildlife and data collection and analysis by teachers<br />

and students at Cle Elum-Roslyn School District. To date we have captured and marked with<br />

GPS collars 4 adult and 2 subadult male and 4 adult female cougars. GPS transmitter collars<br />

collect GPS coordinates at 4-hour intervals throughout <strong>the</strong> year. This data is plotted onto GIS to<br />

assess proximity to human residence, planned development, recreational centers, and to assess<br />

predation events and habitat use patterns. This investigation is used to engage students in an<br />

experiential learning activity whose focus is application <strong>of</strong> technology and learning about <strong>the</strong>ir<br />

ecological and social community. Students in kindergarten to senior high help collect and<br />

analyze data. Junior-Senior students in Advanced Placement Biology assist with cougar capture<br />

and marking efforts and correlate location data with GIS habitat, topographic, and human<br />

residence parameters. They will use DNA isolated from cougar scats to determine species and<br />

gender <strong>of</strong> animals depositing scats while 8 th grade students analyze scats for contents to correlate<br />

food habits with gender <strong>of</strong> cougars. Elementary students learn plant identification for plotting<br />

habitat types and learn animal track identification for reporting locations <strong>of</strong> carnivores and<br />

ungulate prey species near <strong>the</strong>ir residence. Students count ungulates along bus routes for longterm<br />

monitoring <strong>of</strong> prey distribution in relation to seasons and development. Students and<br />

community member conduct tests <strong>of</strong> GPS collars to assess influences <strong>of</strong> vegetative and<br />

physiographic conditions on satellite acquisition rate and accuracy. Students are assessed on<br />

<strong>the</strong>ir abilities to collect qualitative and quantitative data. Community members help collect data<br />

and help train students in outdoor and data collection skills. Central Washington University<br />

incorporates Project CAT objectives into training teachers. Information on ungulate habitat and<br />

cougar travel corridors is shared with community planners to incorporate into planning processes<br />

to minimize human-cougar interactions.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


MONITORING MOUNTAIN LIONS IN THE TUCSON MOUNTAIN DISTRICT OF<br />

SAGUARO NATIONAL PARK, ARIZONA, USING NONINVASIVE TECHNIQUES<br />

LISA HAYNES, Wild Cat and Carnivore Studies, 133 W. 2 nd St., Tucson, AZ 85705, USA,<br />

email: lynxrufus@earthlink.net<br />

DON SWANN, Saguaro National Park, 3693 S. Old Spanish Trail, Tucson, AZ 85730, USA,<br />

email: Don_Swann@nps.gov<br />

MELANIE CULVER, Arizona Cooperative Fish and Wildlife Research Unit, 104 Biosciences<br />

East, University <strong>of</strong> Arizona, Tucson, AZ 85721, USA, email: culver@ag.arizona.edu<br />

Abstract: This presentation will summarize a three-year effort to noninvasively monitor<br />

mountain lions in <strong>the</strong> Tucson <strong>Mountain</strong> District <strong>of</strong> Saguaro National Park, Arizona from 2001<br />

through <strong>the</strong> spring <strong>of</strong> 2003. Park managers are concerned about this mountain lion population,<br />

because <strong>the</strong> Tucson <strong>Mountain</strong>s are becoming surrounded by human development. The continued<br />

existence <strong>of</strong> <strong>the</strong> population is threatened due to habitat loss, potential inbreeding, and disrupted<br />

demographics. Noninvasive methods used to monitor mountain lions included track surveys,<br />

infrared-triggered cameras, and molecular genetic analysis <strong>of</strong> hair (collected from hair snares)<br />

and scat (feces). In <strong>the</strong> first two years, 2001 and 2002, we documented a total <strong>of</strong> 19 sets <strong>of</strong><br />

mountain lion tracks found during 2 winter surveys <strong>of</strong> 30 transects and during seasonal surveys<br />

<strong>of</strong> 4 transects. From track data we determined that mountain lions are consistently detected in<br />

most areas <strong>of</strong> <strong>the</strong> park; however, preliminary evidence suggests a paucity <strong>of</strong> adult males. One <strong>of</strong><br />

<strong>the</strong> most important aspects <strong>of</strong> this project is <strong>the</strong> opportunity for educating volunteers from <strong>the</strong><br />

general public who participate in <strong>the</strong> track surveys. In 2002, 31 <strong>of</strong> 35 hair snares distributed<br />

throughout <strong>the</strong> park had hair deposited on <strong>the</strong>m. Planned genetic analysis will determine<br />

efficacy <strong>of</strong> using hair snares and scat to obtain biological information on mountain lions. We<br />

will present additional data from <strong>the</strong> 2003 track survey and we will provide recommendations for<br />

future monitoring and research efforts.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

141


142<br />

ESTIMATING COUGAR ABUNDANCE USING PROBABILITY SAMPLING: AN<br />

EVALUATION OF TRANSECT VERSUS BLOCK DESIGN<br />

CHUCK R. ANDERSON, JR., Wyoming Cooperative Fish and Wildlife Research Unit, Box<br />

3166, University Station, Laramie, WY 82071. USA, email: cander@uwyo.edu<br />

FRED G. LINDZEY, Wyoming Cooperative Fish and Wildlife Research Unit, Box 3166,<br />

University Station, Laramie, WY 82071, USA, email: flindzey@uwyo.edu<br />

NATE P. NIBBELINK. Wyoming Geographic Information Science Center, University <strong>of</strong><br />

Wyoming, Box 4008, University Station, Laramie, WY 82071, USA, email:<br />

nathan@uwyo.edu<br />

Abstract: We used GPS data records <strong>of</strong> cougar track sets (n = 5-6 locations/night) to evaluate<br />

accuracy and precision <strong>of</strong> Transect Probability Sampling (TPS) and Block Probability Sampling<br />

(BPS) from spatial simulations <strong>of</strong> varying cougar densities, sampling efforts, and number <strong>of</strong><br />

track set nights. GPS data records yielded 446 1-night track sets and 225 2-night track sets from<br />

12 cougars (2 adult males, 6 adult females, and 4 subadults) for simulations. Accuracy and<br />

precision <strong>of</strong> TPS and BPS estimates generally improved with increased cougar density, sampling<br />

effort, and number <strong>of</strong> track set nights, but TPS estimates were vulnerable to extremely short<br />

track sets (e.g., cougars at kill sites) and BPS estimates were exceedingly imprecise. To address<br />

<strong>the</strong>se problems, we adjusted TPS estimates based on <strong>the</strong> proportion <strong>of</strong> cougar track sets<br />

estimated to be at kill sites and used bootstrap techniques to estimate 90% confidence intervals<br />

(CIs) around BPS estimates. TPS estimates adjusted for cougars at kill sites typically improved<br />

accuracy, precision, and estimator reliability (CIs approaching 90% coverage). Bootstrapping<br />

greatly reduced variance around BPS estimates but exaggerated precision (i.e., CI coverage<br />

typically below 90%), likely due to low cougar detection rates. Comparisons <strong>of</strong> adjusted TPS<br />

and BPS estimates suggested higher cougar detection rates and improved accuracy from TPS<br />

surveys, with more reliable CI coverage. TPS simulations suggested reliable cougar population<br />

estimates could be obtained from high-effort surveys (~2 km transect spacing) regardless <strong>of</strong><br />

cougar density or number <strong>of</strong> track set nights, or from medium-effort surveys (~3 km transect<br />

spacing) <strong>of</strong> medium-high density populations (2.3-3.5 independent cougars/100 km 2 ) sampling<br />

2-night track sets. Ninety-percent CIs suggested population changes <strong>of</strong> 27-30% could be<br />

detected using high effort surveys <strong>of</strong> 1-night track sets, 20-24% from medium effort surveys <strong>of</strong><br />

2-night track sets, and 15-18% from high effort surveys <strong>of</strong> 2-night track sets. Because <strong>of</strong> <strong>the</strong><br />

time and expense required to conduct high effort TPS surveys, we propose sampling cougar track<br />

sets without intense tracking efforts and applying perpendicular track lengths we measured to<br />

estimate cougar population parameters.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


EVALUATING MOUNTAIN LION MONITORING TECHNIQUES IN THE GARNET<br />

MOUNTAINS OF WEST CENTRAL MONTANA<br />

RICH DeSIMONE, Montana Fish, Wildlife & Parks, 1420 East Sixth Avenue, Helena, MT<br />

59620, USA, email: rdesimone@state.mt.us<br />

Abstract: Research began in 1998 to document characteristics <strong>of</strong> a hunted mountain lion<br />

population and develop survey techniques to detect trends in lion abundance. Efforts to capture<br />

and radio-collar mountain lions have focused on <strong>the</strong> 850 km 2 eastern half <strong>of</strong> <strong>the</strong> Garnet<br />

<strong>Mountain</strong>s where lion hunting was suspended from 2000 to 2002, allowing <strong>the</strong> lion population to<br />

increase. <strong>Lion</strong> hunting will resume, reducing <strong>the</strong> number <strong>of</strong> lions in <strong>the</strong> study area. Fluctuations<br />

in a known lion population will provide <strong>the</strong> opportunity to determine <strong>the</strong> sensitivity <strong>of</strong> population<br />

indicators to changes in lion abundance. <strong>Mountain</strong> lion population trend indicators being<br />

evaluated include lion track survey-routes and statewide telephone surveys <strong>of</strong> houndsmen and<br />

deer hunters. Eleven lion track snow-survey-routes totaling approximately 105 km were<br />

established in 2000 throughout <strong>the</strong> study area to determine <strong>the</strong> relationship between lion track<br />

density and <strong>the</strong> actual density <strong>of</strong> lions. Track densities ranged from 0 to 2 per 10 km.<br />

Preliminary results indicate that <strong>the</strong> densities <strong>of</strong> lion tracks recorded in different portions <strong>of</strong> <strong>the</strong><br />

study area correlate with <strong>the</strong> densities <strong>of</strong> lion home ranges. A statewide telephone survey <strong>of</strong><br />

houndsmen began in 2001 with approximately 300 houndsmen interviewed annually.<br />

Houndsmen took fewer days <strong>of</strong> hunting to tree lions (3 days) and encounter a lion family group<br />

(8 days) in northwest Montana, while in eastern Montana houndsmen took 10-35 days to tree a<br />

lion and 35-45 days to encounter a family group. Starting in 2001, <strong>the</strong> statewide telephone<br />

survey <strong>of</strong> deer hunters included asking hunters if <strong>the</strong>y observed lions. The percentage <strong>of</strong> deer<br />

hunters observing lions ranged from 4% in northwest Montana to less than 1% in eastern<br />

Montana. Seventy lions have been captured and radio-collared. Eleven <strong>of</strong> 26 radioed kittens died<br />

during <strong>the</strong>ir first year <strong>of</strong> life. Malnutrition due to orphaning was <strong>the</strong> most common cause <strong>of</strong><br />

death. Hunters harvested adult and subadult lions at a high rate. In portions <strong>of</strong> <strong>the</strong> study area<br />

where hunting was allowed, hunters harvested an average <strong>of</strong> 63% <strong>of</strong> <strong>the</strong> radioed lions annually<br />

from 1998 to 2001. Overall, 36 <strong>of</strong> 38 radioed lion deaths were human related.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

143


144<br />

PRESENCE AND MOVEMENTS OF LACTATING AND MATERNAL FEMALE<br />

COUGARS: IMPLICATIONS FOR STATE HUNTING REGULATIONS<br />

TONI K. RUTH, Wildlife Conservation Society, 2023 Stadium Dr. Suite 1A, Bozeman, MT<br />

59030, USA, email: truth@montanadsl.net<br />

KERRY M. MURPHY, Yellowstone Center for Resources, P.O. Box 168, Yellowstone National<br />

Park, Mammoth, WY 89210, USA, email: kerry_murphy@nps.gov<br />

POLLY C. BUOTTE, Wildlife Conservation Society, 2023 Stadium Dr. Suite 1A, Bozeman, MT<br />

59030, USA, email: polly_thornton@hotmail.com<br />

Abstract: Established in <strong>the</strong> early 1970’s, <strong>the</strong> regulation <strong>of</strong> cougar harvest through hunting<br />

seasons and quotas contributed to increases in cougar abundance and distribution in most<br />

western states during <strong>the</strong> past 30 years. Today, 10 <strong>of</strong> 12 western U.S. states regulate cougar<br />

harvest through hunting seasons and quotas, which vary by state, year, and in season length and<br />

quota numbers. California prohibits cougar hunting and Texas allows unlimited hunting and<br />

<strong>the</strong>refore lacks regulations. Three <strong>of</strong> 11 states that allow cougar hunting regulate <strong>the</strong> harvest <strong>of</strong><br />

female lions through subquotas. Two states that do not regulate female harvest do not allow <strong>the</strong><br />

use <strong>of</strong> hounds for hunting cougars. Hunting regulations in 9 <strong>of</strong> 11 states prohibit killing spotted<br />

kittens and females with spotted kittens. Only one state, Montana, requires hunters to backtrack<br />

lactating females that have been killed in order to locate dependent young. While regulations<br />

prohibiting <strong>the</strong> take <strong>of</strong> maternal females and affording protection to nonmaternal females through<br />

subquotas should remain in place, little information has been provided to hunters, guides and<br />

outfitters, or state managers on <strong>the</strong> proportion and movements <strong>of</strong> lactating and maternal females<br />

which may be encountered during hunting seasons. We examined reproductive data for a<br />

moderately hunted cougar population (1988-1992) and a primarily non-hunted population (1998-<br />

2002) on <strong>the</strong> Nor<strong>the</strong>rn Range <strong>of</strong> Yellowstone National Park. Proportion <strong>of</strong> females with<br />

dependent (pre-dispersal) <strong>of</strong>fspring was calculated across winters. We summarized breeding,<br />

denning, and lactation chronology for maternal females from both study periods. Peak breeding<br />

occurred in March through May and denning followed approximately 3 months later, peaking in<br />

June through August. Given a 4-month lactation period, proportion <strong>of</strong> lactating females (n = 19<br />

known date births or births estimated to


MYSTERY, MYTH AND LEGEND: THE POLITICS OF COUGAR MANAGEMENT IN<br />

THE NEW MILLENNIUM<br />

RICK A. HOPKINS, Live Oak Associates, Inc., 6830 Via Del Oro, Suite 805, San Jose, CA<br />

95119, USA, email: rhopkins@loainc.com<br />

Abstract: The cougar as America’s cat is a large, ghost like predator that usually hunts game as<br />

large as or larger than it is. As Teddy Roosevelt noted in <strong>the</strong> late 1800’s, “No American beast<br />

has been <strong>the</strong> subject <strong>of</strong> so much loose writing or <strong>of</strong> such wild fables as <strong>the</strong> cougar”. More than<br />

100 years and dozens <strong>of</strong> scientific studies later, we are no better <strong>of</strong>f in unraveling <strong>the</strong> “loose<br />

writing” and in some cases, <strong>the</strong> management objectives for this predator, than we were at <strong>the</strong><br />

beginning <strong>of</strong> <strong>the</strong> twenty century. The life style <strong>of</strong> <strong>the</strong> cat has long resulted in polarized attitudes<br />

toward <strong>the</strong> development <strong>of</strong> policies for its management throughout North America. The general<br />

“truth” that has evolved with wildlife managers regarding increasing cougar numbers throughout<br />

<strong>the</strong> West over <strong>the</strong> last 2 to 3 decades is believed to be born from 30 years <strong>of</strong> research. While<br />

numbers <strong>of</strong> cougars may have increased in portions <strong>of</strong> <strong>the</strong>ir range over <strong>the</strong> last 2-3 decades, <strong>the</strong><br />

general perception that cougars are more abundant in <strong>the</strong> western U.S. is based not on empirical<br />

data, but one based more on oral traditions passed on from one wildlife pr<strong>of</strong>essional to ano<strong>the</strong>r.<br />

Cougar management in <strong>the</strong> last two decades focused on <strong>the</strong> impact <strong>of</strong> prey populations and<br />

depredation <strong>of</strong> livestock more than on direct encounters with humans. However, an increase in<br />

human attacks in <strong>the</strong> 1990’s has not only heightened public awareness <strong>of</strong> cougars, but appears to<br />

have explicitly shifted management in some western states to focus more on “controlling” <strong>the</strong><br />

species. Lost in this debate and <strong>the</strong> objectives <strong>of</strong> <strong>the</strong> management <strong>of</strong> <strong>the</strong> species is <strong>the</strong><br />

preservation <strong>of</strong> those elements that will truly lead to its conservation. California represents an<br />

interesting living lab, as cougars have not been sport hunted in this state since 1972; it also<br />

supports one <strong>of</strong> <strong>the</strong> largest cougar populations in <strong>the</strong> U.S. if not <strong>the</strong> largest, and clearly supports<br />

<strong>the</strong> largest human population. The lessons learned in California can serve as a model for<br />

continued efforts to focus management objectives on <strong>the</strong> conservation <strong>of</strong> <strong>the</strong> species and not<br />

solely on equating management with harvest as is so <strong>of</strong>ten done.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

145


146<br />

RECONCILING SCIENCE AND POLITICS IN PUMA MANAGEMENT IN THE<br />

WEST: NEW MEXICO AS A TEMPLATE<br />

KENNETH A. LOGAN, Carnivore Researcher, Colorado Division <strong>of</strong> Wildlife, 2300 South<br />

Townsend Avenue, Montrose, CO 81401, USA, email: Ken.Logan@state.co.us<br />

LINDA L. SWEANOR, Scientist, Wildlife Health Center, University <strong>of</strong> California, TB128, Old<br />

Davis Road, Davis, CA 95616, USA, email: lsweanor@mindspring.com<br />

MAURICE G. HORNOCKER, Senior Scientist, Wildlife Conservation Society, Box 929,<br />

Bellevue, ID 83313, USA<br />

Abstract: The puma is <strong>the</strong> only large obligate carnivore thriving today in self-sustaining populations<br />

distributed across western North America. As such, <strong>the</strong> puma contributes to ecosystem integrity because<br />

<strong>the</strong> puma: 1) strongly influences energy flow and nutrient cycling; 2) is a strong natural selective force on<br />

prey animals; 3) modulates prey population dynamics; 4) indirectly affects herbivory on plant<br />

communities; 5) indirectly influences competition among herbivores; and 6) competes with o<strong>the</strong>r<br />

carnivores. Fur<strong>the</strong>rmore, because persisting puma populations depend on expansive, connected wild<br />

landscapes with thriving prey populations, <strong>the</strong> puma is also a potential focal species for designing nature<br />

reserve networks. Wildlife managers have <strong>the</strong> responsibility <strong>of</strong> weighing <strong>the</strong> natural value <strong>of</strong> <strong>the</strong> puma<br />

with <strong>the</strong> diverse needs <strong>of</strong> people. Yet, <strong>the</strong>ir tools for scientific puma management are crude mainly<br />

because pumas are very cryptic and exist in very low population densities. People in New Mexico<br />

identified 10 puma management issues: 1) pumas kill livestock and threaten rancher’s livelihoods; 2)<br />

pumas kill deer that could be taken by hunters; 3) pumas threaten conservation <strong>of</strong> endangered populations<br />

<strong>of</strong> mountain sheep; 4) some pumas threaten public safety; 5) sustainable puma hunting is desirable; 6)<br />

puma hunting should focus on taking males and protecting females and cubs; 7) hunting pumas with dogs<br />

is undesirable; 8) puma hunting is undesirable; 9) increased human development threatens puma<br />

conservation; 10) diverse interests make puma management difficult. Unknowns and uncertainties<br />

specific to puma management included: 1) number <strong>of</strong> pumas in populations; 2) population trends; 3)<br />

population growth rates; 4) population responses to management prescriptions; 5) effects <strong>of</strong> hunter<br />

selection; 6) density distributions; 7) age and sex structure <strong>of</strong> populations; 8) reproductive rates; 9) agespecific<br />

survival rates; 10) immigration and emigration rates; 11) validity <strong>of</strong> puma population simulation<br />

models. These unknowns and uncertainties along with <strong>the</strong> broad diversity <strong>of</strong> human values toward <strong>the</strong><br />

puma make management very difficult and challenge <strong>the</strong> pr<strong>of</strong>essional integrity <strong>of</strong> agencies. In New<br />

Mexico, we developed a robust, biologically sound, adaptive puma management structure that considers<br />

<strong>the</strong> role <strong>of</strong> <strong>the</strong> puma in ecosystems, <strong>the</strong> needs <strong>of</strong> people, and <strong>the</strong> unknowns and uncertainties in puma<br />

management. We called this structure Zone Management. Zone Management uses zones with lethal<br />

control, sport-hunting, and refuges. Control zones allow experimental puma control in focal areas to<br />

protect private property, human safety, endangered species, or game animals. Hunting zones allow sporthunting<br />

opportunity sustained by quotas on <strong>the</strong> number <strong>of</strong> pumas that can be killed, with emphasis on<br />

protecting females and cubs. Refuge zones (i.e., no hunting zones) are >3,000 sq. km and act as biological<br />

savings accounts that assist wildlife managers by countering mistakes made in <strong>the</strong> control and hunt zones,<br />

allowing natural selection to occur in puma populations, and providing numeric and genetic augmentation<br />

<strong>of</strong> human impacted zones via puma dispersal from refuges and immigration into human exploited zones.<br />

The zone management structure uses <strong>the</strong> source-sink metapopulation paradigm we developed for pumas<br />

in New Mexico.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


COMMUNITY-BASED CONSERVATION OF MOUNTAIN LIONS<br />

LYNN MICHELLE CULLENS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95812, USA, email: cullens@mountainlion.org<br />

CHRISTOPHER M. PAPOUCHIS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95812, USA, email: cpapouchis@mountainlion.org<br />

Abstract: The western United States is experiencing a rapid growth in human population, with a<br />

commensurate loss and fragmentation <strong>of</strong> wildlife habitat. As <strong>the</strong> only top predator with viable<br />

populations throughout <strong>the</strong> West, <strong>the</strong> long-term conservation <strong>of</strong> this species is vital for<br />

maintaining <strong>the</strong> health and integrity <strong>of</strong> <strong>the</strong> region’s native ecosystems. Unfortunately,<br />

populations <strong>of</strong> mountain lions in areas subjected to intensive human development or activity may<br />

become threatened unless conservation efforts are implemented and realized. For example, in<br />

several regions <strong>of</strong> California, including <strong>the</strong> West slope <strong>of</strong> California’s Sierra Nevada <strong>Mountain</strong>s,<br />

<strong>the</strong> number <strong>of</strong> mountain lions killed as <strong>the</strong> result <strong>of</strong> conflicts with domestic animals (e.g., goats<br />

and pets) has increased dramatically over <strong>the</strong> past decade. According to California Dept. <strong>of</strong> Fish<br />

and Game, mountain lions may be extirpated from this area within 40 years due to habitat loss<br />

and excessive human caused mortality. Accordingly, a substantial policy shift is needed. In o<strong>the</strong>r<br />

areas, such as Sou<strong>the</strong>rn California, few lions remain in fragmented wildlands, and different<br />

conservation strategies are required to reach urban and suburban residents who live on <strong>the</strong> edge<br />

<strong>of</strong> wildlife areas. We review efforts to conserve mountain lions at <strong>the</strong> community level and<br />

elaborate new approaches that stress science to establish a factual basis for dialogue, community<br />

involvement to identify shared goals, and developing partnerships with diverse organizations and<br />

pr<strong>of</strong>essions to broaden conservation efforts.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

147


148<br />

PUMA MANAGEMENT IN WESTERN NORTH AMERICA: A 100-YEAR<br />

RETROSPECTIVE<br />

STEVEN TORRES, California Department <strong>of</strong> Fish and Game, 1416 Ninth Street, Room 1342A,<br />

Sacramento, CA 95814, USA, email: storres@dfg.ca.gov<br />

HEATHER KEOUGH, Utah State University, Logan, UT 84322, USA<br />

DEANNA DAWN, South Dakota State University, Brookings, SD 57007, USA<br />

Abstract: Puma (Puma concolor) populations have had a diverse and long history <strong>of</strong><br />

management in western North America. For <strong>the</strong> most part <strong>of</strong> <strong>the</strong> last century, pumas were a<br />

bountied predator. By <strong>the</strong> early 1970s, <strong>the</strong>y had transitioned to game mammal status. In <strong>the</strong><br />

period since bounties ended, most states and provinces have reported increased puma activity<br />

that has been simultaneous with increased human populations and land conversion. We will<br />

present an analysis <strong>of</strong> <strong>the</strong> political and biological effects influencing puma populations during<br />

this period to provide perspective on <strong>the</strong> potential effects <strong>of</strong> bounty removals as <strong>the</strong>y may relate<br />

to hypo<strong>the</strong>sized increased populations in <strong>the</strong> latter part <strong>of</strong> <strong>the</strong> last century. This presentation will<br />

also explore <strong>the</strong> changing philosophy <strong>of</strong> predator management and <strong>the</strong> importance <strong>of</strong> maintaining<br />

predator-prey systems and redefining puma management to include <strong>the</strong>ir beneficial role in<br />

defining large blocks <strong>of</strong> habitat and movement corridors.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


USING COUGARS TO DESIGN A WILDERNESS NETWORK IN CALIFORNIA’S<br />

SOUTH COAST ECOREGION<br />

PAUL BEIER, Nor<strong>the</strong>rn Arizona University, Flagstaff, AZ 86011, USA, email:<br />

Paul.Beier@nau.edu<br />

KRISTEEN PENROD, South Coast Wildlands Project, PO Box 2493, Monrovia, CA 91016,<br />

USA, email: Kristeen@scwildlands.org<br />

Abstract: The groundbreaking “Missing Linkages” report published in fall 2001<br />

(www.scwildlands.org) identified over 200 linkages needed to prevent isolation <strong>of</strong> wildlands in<br />

California. South Coast Wildlands Project (SCWP) immediately spearheaded an effort to<br />

prioritize, protect and (where necessary) restore linkages in <strong>the</strong> South Coast Ecoregion. SCWP<br />

first assessed <strong>the</strong> ecoregion’s 69 linkages with respect to biological irreplaceability (size and<br />

quality <strong>of</strong> core areas served by a linkage were important criteria) and vulnerability (to<br />

urbanization and roads). This process identified 15 linkages as top priorities. We are now in <strong>the</strong><br />

process <strong>of</strong> conducting a series <strong>of</strong> action workshops for each linkage. At <strong>the</strong> first workshop, local<br />

biologists, government agencies and conservation NGO representatives developed lists <strong>of</strong> focal<br />

species and ecological processes that a linkage is intended to serve. Thus, although carnivores<br />

helped to initially identify important linkage areas, we are designing each linkage to serve<br />

broader biodiversity goals. Our personnel are researching <strong>the</strong> needs <strong>of</strong> <strong>the</strong> focal species,<br />

obtaining high-resolution photographs and parcel maps, and conducting field visits. One or more<br />

linkage designs will be presented at a second workshop, where participants will volunteer for<br />

various tasks (e.g., procuring easements, acquiring land, changing zoning, restoring habitat, or<br />

mitigating transportation projects) to preserve and enhance <strong>the</strong> linkage. By partnering with<br />

agencies and NGOs from <strong>the</strong> start ra<strong>the</strong>r than developing a plan on our own and asking o<strong>the</strong>rs to<br />

unite under us, our effort has attracted funding and cooperation from diverse sources and is<br />

making rapid progress.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

149


150<br />

MOUNTAIN LIONS AND BIGHORN SHEEP: FACING THE CHALLENGES<br />

CHRISTOPHER M. PAPOUCHIS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95812, USA, email: cpapouchis@mountainlion.org<br />

JOHN D. WEHAUSEN, U.C. White <strong>Mountain</strong> Research Station, 3000 E. Line Street, Bishop,<br />

CA 93514, USA<br />

Abstract: <strong>Mountain</strong> lions (Puma concolor) and bighorn sheep (Ovis canadensis) have coevolved<br />

as predator and prey. Bighorn sheep populations have declined over <strong>the</strong> past several<br />

centuries across <strong>the</strong>ir range due to a variety <strong>of</strong> anthropogenic causes, and have been <strong>the</strong> subject<br />

<strong>of</strong> extensive translocation efforts in recent years in an effort to reestablish populations in historic<br />

habitat. In <strong>the</strong> past several decades mountain lion predation has also been implicated in <strong>the</strong><br />

decline <strong>of</strong> several populations <strong>of</strong> endangered bighorn sheep, including <strong>the</strong> federally listed<br />

Peninsular (O. c. cremnobates) and Sierra Nevada populations in California and state listed<br />

desert bighorn sheep (O.c. mexicana) populations in New Mexico. We hypo<strong>the</strong>size that this<br />

recent phenomenon has resulted from land and wildlife management practices that have affected<br />

both species. Restoring <strong>the</strong> natural relationship between mountain lions and bighorn sheep<br />

presents both biological and ethical challenges. We discuss <strong>the</strong> current status <strong>of</strong> management<br />

efforts, review several hypo<strong>the</strong>sis <strong>of</strong> why predation has become a limiting factor for bighorn<br />

sheep recovery, and discuss current and potential management options.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


FACTORS AFFECTING DISPERSAL IN YOUNG MALE PUMAS<br />

JOHN W. LAUNDRÉ, Instituto de Ecologia, A.C, Km 5 a Carr. de Mazatlán s/n, C.P 34100,<br />

Durango, Dgo México, email: launjohn@prodigy.net.mx<br />

LUCINA HERNÁNDEZ, Instituto de Ecologia, A.C, Km 5 a Carr. de Mazatlán s/n, C.P 34100,<br />

Durango, Dgo México, email: lucina@fauna.edu.mx<br />

Abstract: Numerous studies have demonstrated that nearly all young male pumas disperse from <strong>the</strong>ir natal home<br />

range area while most females are philopatric. There are 2 hypo<strong>the</strong>ses for <strong>the</strong> driving force behind dispersal in<br />

young male pumas. The first is <strong>the</strong> competition (aggression avoidance) model where dispersal is because <strong>of</strong><br />

competition between young males and <strong>the</strong>ir fa<strong>the</strong>rs/incoming transient males for mates and resources. The second is<br />

<strong>the</strong> inbreeding model where young males are thought to disperse to avoid inbreeding with <strong>the</strong>ir mo<strong>the</strong>rs/sisters.<br />

Under each model, specific predictions can be made to test <strong>the</strong>ir validity. Under <strong>the</strong> competition model, we predict<br />

that <strong>the</strong>re should be physical conflicts between sons and <strong>the</strong>ir fa<strong>the</strong>rs, including infanticide and that competitive<br />

ability should increase with age and thus, <strong>the</strong> longer time and fur<strong>the</strong>r distance <strong>the</strong>y are from <strong>the</strong>ir home area. Thus,<br />

dispersal distances should reflect <strong>the</strong>ir competitive ability, i.e. few males will establish territories close to <strong>the</strong>ir natal<br />

home range. Under <strong>the</strong> inbreeding avoidance model, we predict fewer physical conflicts (young males leave on<br />

<strong>the</strong>ir own), no inbreeding between fa<strong>the</strong>rs and daughters, which is genetically equivalent to sons mating with <strong>the</strong>ir<br />

mo<strong>the</strong>rs, and no males should establish <strong>the</strong>ir territories adjacent (1-2 home range diameters) to <strong>the</strong>ir natal home<br />

range. We tested <strong>the</strong>se predictions with dispersal data from our study in sou<strong>the</strong>rn Idaho/northwestern Utah and<br />

published data. We refuted <strong>the</strong> inbreeding model because resident males do fight and kill <strong>the</strong>ir male <strong>of</strong>fspring,<br />

resident males do mate with <strong>the</strong>ir daughters, and <strong>the</strong>re is a high percent <strong>of</strong> males that establish <strong>the</strong>ir territories within<br />

2-4 HRDs <strong>of</strong> <strong>the</strong>ir natal home ranges. Our data supported <strong>the</strong> competition model with an increase in frequency <strong>of</strong><br />

dispersal distances at >2 diameters. We conclude that young males are forced out <strong>of</strong> <strong>the</strong>ir natal home range by <strong>the</strong>ir<br />

fa<strong>the</strong>rs or incoming males who, by default, will be older and stronger. We propose that <strong>the</strong>y continue to disperse<br />

until <strong>the</strong>y gain enough weight and experience to successfully takeover a territory.<br />

Key Words: male pumas, dispersal, inbreeding, competition, Idaho, Utah<br />

Numerous studies have demonstrated<br />

that nearly all young male pumas disperse<br />

from <strong>the</strong>ir home range area while most<br />

females are philopatric (Ross and<br />

Jalkotzy1992, Sweanor et al. 2000). There<br />

are 2 hypo<strong>the</strong>ses for <strong>the</strong> driving force behind<br />

dispersal in young male pumas that have<br />

been developed considering <strong>the</strong> various<br />

costs and benefits to dispersing vs<br />

philopatric behaviors (Shields 1987). The<br />

first is <strong>the</strong> competition model where<br />

dispersal is motivated by interference<br />

competition and aggression between young<br />

males and <strong>the</strong>ir fa<strong>the</strong>rs/incoming transient<br />

males for mates and resources (Dobson<br />

1982, Moore and Ali 1984). The second is<br />

<strong>the</strong> inbreeding model where young males are<br />

thought to disperse to avoid inbreeding<br />

151<br />

<strong>Proceedings</strong> <strong>of</strong> <strong>the</strong> <strong>Seventh</strong> <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong><br />

primarily with <strong>the</strong>ir mo<strong>the</strong>rs/sisters (Wolfe<br />

1994). There have been extensive reviews<br />

and criticisms <strong>of</strong> both hypo<strong>the</strong>ses and it<br />

continues to be a hotly debated topic (See<br />

Shields 1987 for a review). Relative to<br />

pumas, discussions <strong>of</strong> dispersal have<br />

primarily centered around dispersal ages and<br />

distances but little on <strong>the</strong> driving force<br />

behind dispersal, especially for young<br />

males. Sweanor (1990) reported on<br />

dispersal in male pumas in New Mexico and<br />

concluded that <strong>the</strong> likely driving force was<br />

competition among males. However, later,<br />

Logan and Sweanor (2001) concluded that<br />

male pumas were dispersing primarily to<br />

avoid inbreeding. Thus at this time it is<br />

uncertain which hypo<strong>the</strong>sis might best<br />

explain dispersal in young male pumas.


152 DISPERSAL IN MALE PUMAS · Laundré and Hernández<br />

Under each model, specific predictions<br />

can be made to test <strong>the</strong>ir application to male<br />

pumas. Under <strong>the</strong> competition model, we<br />

predict that <strong>the</strong>re should be physical<br />

conflicts between adult males/kittens<br />

(infanticide), adult males/juveniles,<br />

including fa<strong>the</strong>r/son, and adult male/adult<br />

male. Older, more experience males should<br />

win <strong>the</strong>se conflicts and competitive ability<br />

<strong>of</strong> dispersers should increase with age and<br />

thus, <strong>the</strong> longer time and fur<strong>the</strong>r distance<br />

<strong>the</strong>y are from <strong>the</strong>ir home area (Sweanor<br />

1990). Thus, ano<strong>the</strong>r prediction for this<br />

model is that dispersal distances should<br />

reflect <strong>the</strong>ir competitive ability, i.e. few<br />

males will establish territories close to <strong>the</strong>ir<br />

natal home range. Under <strong>the</strong> inbreeding<br />

avoidance model, we predict less physical<br />

conflicts (young males leave on <strong>the</strong>ir own),<br />

no inbreeding between fa<strong>the</strong>rs and<br />

daughters, which is genetically equivalent to<br />

sons mating with <strong>the</strong>ir mo<strong>the</strong>rs, and no<br />

males establishing territories within a<br />

minimum <strong>of</strong> 2 home range diameters <strong>of</strong> <strong>the</strong>ir<br />

natal home range (Sweanor 1990). We<br />

attempted to test <strong>the</strong>se predictions with data<br />

from our study in sou<strong>the</strong>rn<br />

Idaho/northwestern Utah and published data<br />

from a variety <strong>of</strong> o<strong>the</strong>r studies <strong>of</strong> pumas.<br />

We recognize that it is <strong>of</strong>ten difficult to<br />

obtain good dispersal data on a species such<br />

as <strong>the</strong> puma but combining <strong>the</strong> data from <strong>the</strong><br />

various studies that exist should provide us<br />

<strong>the</strong> best opportunity to examine possible<br />

causes <strong>of</strong> dispersal in male pumas.<br />

METHODS<br />

To test <strong>the</strong> predictions made, we used a<br />

combination <strong>of</strong> data from our long-term<br />

study <strong>of</strong> pumas and published data on o<strong>the</strong>r<br />

puma studies. In some <strong>of</strong> <strong>the</strong>se studies,<br />

pumas were protected from hunting while in<br />

<strong>the</strong> o<strong>the</strong>rs <strong>the</strong>y were exposed to sport<br />

harvest. However, in <strong>the</strong> case <strong>of</strong> <strong>the</strong><br />

protected populations, <strong>the</strong> protection<br />

extended only to <strong>the</strong> limits <strong>of</strong> <strong>the</strong> study<br />

areas, which ranged in size from ≈ 500 to<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

2,000 km 2 or a maximum diameter <strong>of</strong> 50<br />

km. Because in all <strong>the</strong> studies, most young<br />

male pumas dispersed to outside <strong>of</strong> <strong>the</strong><br />

designated study areas, we consider that<br />

pumas were dispersing under unprotected<br />

conditions. Consequently, we did not<br />

subdivide <strong>the</strong> studies into protected and<br />

unprotected populations.<br />

Our study <strong>of</strong> a harvested puma<br />

population was conducted in sou<strong>the</strong>astern<br />

Idaho and northwestern Utah. The study<br />

area was approximately 2,000 km 2 and<br />

consisted <strong>of</strong> 5 small mountain ranges<br />

(approximately 1,000 km 2 total area)<br />

separated by valleys where human activity<br />

predominated. Over 16 years, we conducted<br />

intensive capture efforts each winter and<br />

with <strong>the</strong> help <strong>of</strong> trained dogs, we captured<br />

and marked approximately 150 pumas. We<br />

documented male-male conflicts based upon<br />

our intensive field efforts. We obtained<br />

dispersal data <strong>of</strong> young radio collared male<br />

pumas primarily from hunter returns. We<br />

measured straight-line dispersal distance<br />

from <strong>the</strong> center <strong>of</strong> <strong>the</strong> natal home range to<br />

<strong>the</strong> point <strong>of</strong> harvest. In addition to our data,<br />

we searched <strong>the</strong> published literature for<br />

accounts <strong>of</strong> male-male conflicts, incidences<br />

<strong>of</strong> inbreeding, and estimates <strong>of</strong> dispersal<br />

distances. For male conflicts, we considered<br />

3 levels <strong>of</strong> conflict, adult male/kitten<br />

(infanticide), adult male/subadult male, and<br />

adult male/adult male. Incidences <strong>of</strong><br />

inbreeding were obtained via field<br />

observations and, in some cases, genetic<br />

testing.<br />

For dispersal, we expressed male<br />

dispersal distances from natal home ranges<br />

in multiples <strong>of</strong> <strong>the</strong> average male home range<br />

diameter (HRD) (Waser 1985). We<br />

calculated HRDs for each study assuming a<br />

circular home range (Sweanor et al. 2000).<br />

We <strong>the</strong>n estimated <strong>the</strong> cross study HRD<br />

average and <strong>the</strong>n divided dispersal distances<br />

<strong>of</strong> each study by <strong>the</strong> cross study mean HRD.<br />

Although we assumed <strong>the</strong> mean <strong>of</strong> <strong>the</strong>


different HRD estimates would be a better<br />

estimate <strong>of</strong> <strong>the</strong> true HRD for male pumas<br />

overall, we recognize that <strong>the</strong> variation<br />

among studies might represent true regional<br />

differences and not just random variation<br />

around <strong>the</strong> population mean. Thus, we also<br />

expressed dispersal distances <strong>of</strong> each study<br />

in multiples <strong>of</strong> <strong>the</strong> individual study estimates<br />

<strong>of</strong> HRD.<br />

We plotted <strong>the</strong> frequency distributions <strong>of</strong><br />

dispersal distances expressed as multiples <strong>of</strong><br />

HRDs for both individual study estimates<br />

and for <strong>the</strong> across study estimate <strong>of</strong> HRDs.<br />

To test if competition drives male dispersal,<br />

we compared <strong>the</strong> observed percent<br />

frequency distribution <strong>of</strong> dispersal distances<br />

(x) with that expected (f(x)) calculated from<br />

Equation 1 (Miller and Carroll 1989).<br />

Equation 1:<br />

x<br />

F(x) = p(x) ∏ [1-p(x-i)]<br />

i = 1<br />

In this equation, f(x) is <strong>the</strong> probability an<br />

animal will establish a territory at x<br />

dispersal distance units, in this case HRDs,<br />

from <strong>the</strong> natal home area. The value p(x) is<br />

<strong>the</strong> probability that an animal will obtain a<br />

territory at x dispersal distances and varies<br />

from p(0), <strong>the</strong> probability <strong>the</strong> animal will not<br />

leave its natal area, to a maximum value,<br />

usually represented by <strong>the</strong> average mortality<br />

rate (Waser 1985). The quantity, 1-p(x-i), is<br />

<strong>the</strong> probability <strong>of</strong> continuing through<br />

previous habitats. We selected <strong>the</strong> model <strong>of</strong><br />

Miller and Carroll (1989) over o<strong>the</strong>r<br />

geometric models such as that <strong>of</strong> Waser<br />

(1985 and 1987) because Equation 1<br />

incorporates changes over distance in <strong>the</strong><br />

probability <strong>of</strong> an animal securing a territory.<br />

We considered <strong>the</strong> model <strong>of</strong> Miller and<br />

Carrol (1989) more realistic because based<br />

on all <strong>the</strong> reported data (Logan and Sweanor<br />

2001), <strong>the</strong> probability <strong>of</strong> a young male<br />

DISPERSAL IN MALE PUMAS · Laundré and Hernández 153<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

setting up his territory within his natal area<br />

(p(0)) is zero.<br />

We estimated 2 expected frequency<br />

distributions calculated from Eq. 1. The<br />

first was an estimation <strong>of</strong> what we would<br />

expect under <strong>the</strong> competition model. We<br />

assumed that once young males start<br />

dispersing, <strong>the</strong>ir competitive ability should<br />

increase with more time, and thus, distance<br />

<strong>the</strong>y disperse until <strong>the</strong>y are equally<br />

competitive for any open territories (Logan<br />

and Sweanor 2001). Logan and Sweanor<br />

(2001) reported that dispersing males in<br />

<strong>the</strong>ir study rarely move constantly but move<br />

from one transient home range to ano<strong>the</strong>r.<br />

However, <strong>the</strong> amount <strong>of</strong> time and distance<br />

<strong>the</strong>y traveled between <strong>the</strong> natal and<br />

independent home range varied extensively.<br />

Of 4 dispersing males Logan and Sweanor<br />

(2001) monitored, <strong>the</strong> dispersal time<br />

averaged 211 + 192 days and straight-line<br />

dispersal distances averaged 143.2 + 26.5<br />

km. Based on <strong>the</strong>se estimates, male pumas<br />

will move on average 0.7 km per day. They<br />

disperse at approximately 16-18 months<br />

(Anderson et al. 1992, Ross and Jalkotzy<br />

1992, Hemker et al. 1984, Beier 1995,<br />

Maehr et al. 1991, Logan and Sweanor<br />

2001) and in 5 months would move<br />

approximately 100 km which, based on<br />

results presented later, represents<br />

approximately 5 HRDs. At 21-24 months<br />

<strong>the</strong>y are near <strong>the</strong>ir full body mass (Laundré<br />

and Hernández 2002) and we assumed to be<br />

competitive for territories. Thus <strong>the</strong><br />

competition model incorporated a linear<br />

change in p(x) from p(0) = 0 to <strong>the</strong> average<br />

mortality rate over <strong>the</strong> first 5 HRDs (x = 1<br />

to 5) and <strong>the</strong>n after that, it was held constant<br />

at <strong>the</strong> average mortality rate.<br />

For <strong>the</strong> inbreeding model, we set p(x) =<br />

0.0 for <strong>the</strong> first few HRDs (depending on <strong>the</strong><br />

average dispersal distance <strong>of</strong> females) and<br />

after that, set it to <strong>the</strong> average mortality rate;<br />

young male pumas should move beyond <strong>the</strong><br />

areas where <strong>the</strong>re is a high probability <strong>of</strong>


154 DISPERSAL IN MALE PUMAS · Laundré and Hernández<br />

mating with sisters or mo<strong>the</strong>rs that have<br />

shifted <strong>the</strong>ir home range areas. This<br />

minimal distance should be represented by<br />

<strong>the</strong> average dispersal distance <strong>of</strong> females<br />

(sisters) from <strong>the</strong>ir natal home range.<br />

RESULTS<br />

Prediction #1: Male conflicts<br />

Relative to levels <strong>of</strong> adult male/kitten<br />

conflict or infanticide, we found ample<br />

evidence for infanticide in our study and in<br />

<strong>the</strong> literature. We recorded 4 cases <strong>of</strong> adult<br />

male pumas killing kittens (9 kittens total) in<br />

our study area. Robinette et al. (1961) cited<br />

reports from Utah <strong>of</strong> young pumas being<br />

killed by adult males on 2 occasions.<br />

Hornocker (1970) reported 1 instance <strong>of</strong> an<br />

adult male killing 2 kittens in his study area<br />

in central Idaho. Lindzey et al. (1989)<br />

reported 4 kittens being killed by pumas in<br />

sou<strong>the</strong>rn Utah. However, <strong>the</strong>y were not sure<br />

<strong>of</strong> <strong>the</strong> sex <strong>of</strong> <strong>the</strong> responsible animals.<br />

Spreadbury et al. (1996) documented 2 male<br />

kittens being killed by a transient male in<br />

British Columbia. Logan and Sweanor<br />

(2001) reported 12 kittens being killed by<br />

adult males in <strong>the</strong>ir study in New Mexico.<br />

With respect to adult male/subadult male<br />

conflicts, Lindzey et al. (1989) reported a<br />

transient male being cannibalized by a<br />

resident male. Murphy (1998) reported 2<br />

incidences <strong>of</strong> territorial males killing<br />

yearling males in his study in Yellowstone<br />

National Park. Logan and Sweanor (2001)<br />

reported 4 mortalities <strong>of</strong> younger (14.3-21<br />

months) males being killed by older males,<br />

including 1 instance <strong>of</strong> a fa<strong>the</strong>r killing his<br />

son.<br />

Although Hornocker (1970) and<br />

Seidensticker et al. (1973) did not report<br />

conflicts among adult male pumas, various<br />

o<strong>the</strong>r investigators reported it as ra<strong>the</strong>r<br />

common. In our area, adult males we<br />

captured <strong>of</strong>ten had facial scars and torn ears,<br />

presumably from conflicts with o<strong>the</strong>r adult<br />

males or, perhaps, females. Sitton and<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Weaver (1977) reported 2 adult pumas being<br />

killed by o<strong>the</strong>r adult pumas as a result <strong>of</strong><br />

fighting. However, <strong>the</strong> sex <strong>of</strong> <strong>the</strong><br />

combatants was unknown. McBride (1976)<br />

reported 2 incidences <strong>of</strong> adult males with<br />

extensive facial injuries and 1 instance <strong>of</strong> an<br />

adult male being killed by ano<strong>the</strong>r. Murphy<br />

(1998) reported an adult immigrant male<br />

being killed by <strong>the</strong> resident male. Logan<br />

and Sweanor (2001) found 5 adult males<br />

killed by o<strong>the</strong>r males. They also reported<br />

that 4 <strong>of</strong> <strong>the</strong> 5 winners outweighed <strong>the</strong><br />

losers, 2 <strong>of</strong> <strong>the</strong> 5 losers were killed by older<br />

animals, and 2 <strong>of</strong> <strong>the</strong> o<strong>the</strong>r 3 were 8 and 12<br />

years old, well past <strong>the</strong>ir prime.<br />

Prediction # 2: occurrence <strong>of</strong> inbreeding<br />

Concerning <strong>the</strong> presence <strong>of</strong> inbreeding<br />

within puma populations, it is well known<br />

that <strong>the</strong> Florida pan<strong>the</strong>r is highly inbred,<br />

including first order matings (Roelke et al.<br />

1993, Barone et al. 1994 and o<strong>the</strong>rs). This<br />

indicates that <strong>the</strong>re is no innate “aversion” to<br />

inbreeding. Logan and Sweanor (2001)<br />

reported an average residency <strong>of</strong> males <strong>of</strong> 6<br />

years (maximum = 8 yrs) in <strong>the</strong>ir study in<br />

New Mexico. If we assume an average <strong>of</strong> 2<br />

years per litter (Logan and Sweanor 2001),<br />

males were resident on average for at least 3<br />

generations. Given that females are<br />

primarily philopatric (Logan and Sweanor<br />

2001), and <strong>the</strong>re is no aversion to inbreeding<br />

with <strong>the</strong>ir fa<strong>the</strong>rs, we predict that with this<br />

length <strong>of</strong> residency a fa<strong>the</strong>r would have <strong>the</strong><br />

opportunity to mate twice with his daughters<br />

and once with granddaughters. If <strong>the</strong>re is an<br />

innate aversion to inbreeding, such matings<br />

should not occur. Logan and Sweanor<br />

(2001) report that 2 females mated with <strong>the</strong>ir<br />

fa<strong>the</strong>rs, 1 <strong>of</strong> <strong>the</strong>m mated with her fa<strong>the</strong>r<br />

twice, supporting <strong>the</strong> prediction based on no<br />

aversion to inbreeding.<br />

Prediction # 3: Dispersal distances<br />

For dispersal distances <strong>of</strong> males, we<br />

limited our analysis to animals that were<br />

known to have set up territories or were


DISPERSAL IN MALE PUMAS · Laundré and Hernández 155<br />

Table 1. Average dispersal distances (km), home range sizes (HRS; km 2 ), and home range<br />

diameters (HRD; km) <strong>of</strong> male pumas from various studies. The home range sizes are based on<br />

resident males in <strong>the</strong> study areas. Dispersal distances are only <strong>of</strong> individuals who were known to<br />

establish territories after dispersing or were at least > 24 months old before being killed.<br />

Source n Distance H.R. Size HRD<br />

Lindzey et al. (1989 & 1994) 4 142.2 731.0 30.5<br />

Logan et al. (1986) 3 175.7 320.0 20.2<br />

Spreadbury et al. (1996) 4 74.5 132.3 13.0<br />

Hornocker (1960) &<br />

Seidensticker et al. (1973)<br />

5 102.4 379.5 22.0<br />

Ashman et al. (1983) 6 48.9 355.2 21.3<br />

Logan and Sweanor (2001) 13 128.2 187.1 15.4<br />

This study 13 190.6 302.5 19.6<br />

Averages 123.2 343.9 20.3<br />

at least >24 months old at <strong>the</strong> time <strong>the</strong>y were<br />

killed (Sweanor et al. 2000). We found<br />

dispersal distances for 48 male pumas from<br />

7 different studies, including 13 estimates<br />

from our study area (Table 1). We did not<br />

use <strong>the</strong> dispersal distances reported by Beier<br />

(1995) because his study area, although<br />

large (2070 km 2 ), was completely<br />

surrounded by urban development, which<br />

limited dispersal opportunities. Dispersal<br />

distances ranged from 31 km (Spreadbury et<br />

al. 1996) to a maximum <strong>of</strong> 550 km (our<br />

study). Estimates <strong>of</strong> resident home range<br />

sizes <strong>of</strong> males in <strong>the</strong>se studies varied<br />

between 132.3 to 731 km 2 (Table 1). The<br />

mean <strong>of</strong> <strong>the</strong>se averages was 343.9 km 2 with<br />

a HRD <strong>of</strong> 20.3 km. Dispersal distances<br />

varied from 0.9 to 26 HRDs with a mean <strong>of</strong><br />

6.0 HRDs (Fig. 1). The majority (≈ 50 %)<br />

<strong>of</strong> dispersal distances were between 2 and 4<br />

HRDs with a second peak (25 %) between<br />

8-9 HRD’s. There was a noticeable<br />

reduction <strong>of</strong> dispersal settlements between 5<br />

and 7 HRDs (Fig. 1).<br />

Based on <strong>the</strong> individual estimates <strong>of</strong><br />

home range size, <strong>the</strong> frequency <strong>of</strong> HRDs<br />

changed in some categories (Fig. 1).<br />

However, <strong>the</strong> pattern was similar to <strong>the</strong><br />

distribution based on <strong>the</strong> overall average<br />

HRD, e.g. <strong>the</strong> majority <strong>of</strong> <strong>the</strong> dispersal<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

distances were between 2 and 4 HRDs. The<br />

one difference was <strong>the</strong> appearance <strong>of</strong> more<br />

settlements in <strong>the</strong> 5-7 HRD range (Fig. 1)<br />

We found estimates <strong>of</strong> male mortality<br />

rates from 4 o<strong>the</strong>r studies (Ashman et al.<br />

1983, Ross and Jalkotzy 1992, Anderson et<br />

al. 1992, Logan and Sweanor 2001) plus our<br />

Figure 1. Frequency distribution <strong>of</strong><br />

dispersal distances, expressed as multiples<br />

<strong>of</strong> <strong>the</strong> cross study average home range<br />

diameter (HRD) and as multiples <strong>of</strong> <strong>the</strong><br />

individual study HRDs. The two curves<br />

represent <strong>the</strong> expected frequency<br />

distribution for <strong>the</strong> competition and <strong>the</strong><br />

inbreeding models.


156 DISPERSAL IN MALE PUMAS · Laundré and Hernández<br />

estimate and <strong>the</strong>y ranged from 9.0 % to 31.0<br />

% with an average mortality rate <strong>of</strong> 20.5%.<br />

Based on movement distances and times<br />

presented in Logan and Sweanor (2001), we<br />

assumed <strong>the</strong> competitive ability <strong>of</strong> a young<br />

male puma would be at its maximum by 5<br />

HRDs. Thus, in Eq. 1, for <strong>the</strong> competition<br />

model, we let p(x) vary from 0.04 to 0.18<br />

over <strong>the</strong> first 4 HRDs (x = 1 to 4) and <strong>the</strong>n<br />

after that, let it be a constant at 0.205, <strong>the</strong><br />

average mortality rate. The resulting curve<br />

underestimated <strong>the</strong> number <strong>of</strong> settlements in<br />

<strong>the</strong> second and third HRD but <strong>the</strong> pattern fit<br />

<strong>the</strong> data relatively well in <strong>the</strong> first 4 HRDs.<br />

In contrast it predicted more settlements in<br />

<strong>the</strong> 5-7 HRD range and less in <strong>the</strong> 8-9 range<br />

than what we found (Fig. 1).<br />

For <strong>the</strong> inbreeding model, we found in<br />

our study and in that <strong>of</strong> Logan and Sweanor<br />

(2001), approximately 42.0 % <strong>of</strong> <strong>the</strong> young<br />

females dispersed an average <strong>of</strong> 2 HRDs<br />

(range = 1 to 8 HRDs, n = 14). Based on<br />

this, a dispersing male would almost be as<br />

likely to mate with a sister within <strong>the</strong> first 2<br />

HRDs as in his natal home range. Also,<br />

considering that resident females can easily<br />

shift <strong>the</strong>ir home range area by at least 1<br />

HRD during <strong>the</strong>ir reproductive years, we<br />

concluded that to avoid inbreeding, a male<br />

should disperse at least a minimum <strong>of</strong> 3<br />

HRDs before seeking a territory.<br />

Consequently, we set p(x) = 0.0 for x = 1 to<br />

2 and after that, set it to <strong>the</strong> average<br />

mortality rate <strong>of</strong> 20.5 %. As 25.0 % <strong>of</strong> <strong>the</strong><br />

dispersing males settled within <strong>the</strong> first 2<br />

HRDs (Fig. 1), this model initially did not fit<br />

<strong>the</strong> data. It also predicted a higher<br />

occurrence <strong>of</strong> settlements in <strong>the</strong> third and<br />

fourth HRDs than what we found. From <strong>the</strong><br />

fifth HRD this model patterned after <strong>the</strong><br />

competition model.<br />

DISCUSSION<br />

Regarding prediction # 1, we found<br />

ample evidence <strong>the</strong>re are high levels <strong>of</strong><br />

conflict between adult territorial males and<br />

all age classes <strong>of</strong> non-territorial males<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

(kittens to adults, including between fa<strong>the</strong>rs<br />

and sons). These conflicts at times appear to<br />

be over food resources but primarily are<br />

over breeding resources. Also in almost all<br />

cases, <strong>the</strong> older, heavier, more experienced<br />

male wins <strong>the</strong> conflict. These data support<br />

<strong>the</strong> competition hypo<strong>the</strong>sis for dispersal<br />

because if young males do not disperse <strong>the</strong>y<br />

will be killed by <strong>the</strong>ir fa<strong>the</strong>rs or nearby<br />

resident males, all <strong>of</strong> whom are bigger and<br />

more experienced.<br />

For prediction # 2, although it can be<br />

argued that <strong>the</strong> Florida pumas inbreed<br />

because it is a closed population, it still<br />

refutes <strong>the</strong> myth <strong>of</strong> a natural aversion to<br />

inbreeding. The innate aversion to close<br />

inbreeding is <strong>the</strong> driving force behind <strong>the</strong><br />

inbreeding avoidance model and, for <strong>the</strong><br />

model to work, should be evident under all<br />

conditions. It is a little too anthropomorphic<br />

to expect young male pumas to avoid<br />

inbreeding under one situation and not<br />

ano<strong>the</strong>r, especially since <strong>the</strong>y don’t know if<br />

<strong>the</strong>y are in Florida or Montana, i.e. <strong>the</strong>y can<br />

only gain information about <strong>the</strong> number and<br />

distribution <strong>of</strong> potential mating opportunities<br />

by dispersing. Relative to <strong>the</strong> New Mexico<br />

data, based on our calculations <strong>of</strong> residency,<br />

we predict we should find cases <strong>of</strong> fa<strong>the</strong>rs<br />

not only mating with <strong>the</strong>ir daughters but<br />

could do so at least twice during <strong>the</strong>ir<br />

tenure. The data from Logan and Sweanor<br />

(2001) supported this prediction, thus<br />

refuting <strong>the</strong> inbreeding hypo<strong>the</strong>sis. As<br />

fa<strong>the</strong>rs mating with daughters is genetically<br />

equivalent to sons mating with <strong>the</strong>ir<br />

mo<strong>the</strong>rs, we see no reason why a son would<br />

refuse to mate with his mo<strong>the</strong>r but <strong>the</strong>n later<br />

mate with his daughter. It might be argued<br />

that a larger fa<strong>the</strong>r could dominate over a<br />

smaller unreceptive daughter. However, we<br />

reject this argument because by <strong>the</strong> time<br />

young males reach dispersal age, <strong>the</strong>y weigh<br />

equal to or more than <strong>the</strong>ir mo<strong>the</strong>rs<br />

(Laundré and Hernández 2002) and upon<br />

sexual maturity (21-27 months; Logan and


Sweanor 2001) could easily dominate over<br />

<strong>the</strong>ir smaller mo<strong>the</strong>r or sisters. As males do<br />

not seem reluctant to mate with close<br />

relatives, i.e. daughters, why don’t <strong>the</strong>y wait<br />

around to mate with <strong>the</strong>ir sisters or mo<strong>the</strong>rs?<br />

We conclude that <strong>the</strong> only logical answer to<br />

this question is that <strong>the</strong> high level <strong>of</strong> conflict<br />

with <strong>the</strong>ir fa<strong>the</strong>rs or, by default, even more<br />

competitive older bigger new male arrivals<br />

(who have displaced <strong>the</strong>ir fa<strong>the</strong>rs), force<br />

young males to disperse out <strong>of</strong> <strong>the</strong>ir natal<br />

area at an age (13-15 months <strong>of</strong> age) before<br />

<strong>the</strong>y are sexually mature (Sweanor 1990).<br />

Regarding prediction #3, <strong>the</strong> observed<br />

frequency distribution <strong>of</strong> HRDs was best fit<br />

by <strong>the</strong> competition model with a p(0) = 0<br />

and an increasing p(x) until a distance <strong>of</strong> 5<br />

HRDs. This model would reflect <strong>the</strong> effect<br />

<strong>of</strong> an increasing competitive ability <strong>of</strong> <strong>the</strong><br />

young males as <strong>the</strong>y disperse. In contrast, it<br />

is obvious that <strong>the</strong> data do not fit <strong>the</strong> most<br />

critical part <strong>of</strong> <strong>the</strong> inbreeding avoidance<br />

model: males should not settle within 2<br />

HRDs <strong>of</strong> <strong>the</strong>ir natal home range.<br />

Additionally, considering that even 5 HRDs<br />

may not be sufficient to reduce <strong>the</strong> chance <strong>of</strong><br />

inbreeding with sisters (Shields 1987,<br />

Sweanor 1990), <strong>the</strong> high percent (>50%) <strong>of</strong><br />

settlements within <strong>the</strong> first 4 HRDs argues<br />

strongly against inbreeding avoidance being<br />

<strong>the</strong> driving force behind male dispersal.<br />

The decreasing number <strong>of</strong> settlements<br />

between 5 and 7 HRDs likely reflects <strong>the</strong><br />

fragmented nature <strong>of</strong> <strong>the</strong> 2 studies that<br />

contributed <strong>the</strong> majority <strong>of</strong> <strong>the</strong> data (54.2 %;<br />

Logan and Sweanor 2001 and this study). In<br />

both studies, <strong>the</strong> principal mountain range(s)<br />

<strong>of</strong> <strong>the</strong> study areas are separated from nearby<br />

larger mountain ranges (places young pumas<br />

are most likely to find vacant territories) by<br />

valleys 75 to 100 + km wide. Thus, <strong>the</strong>re is<br />

a reduced likelihood <strong>of</strong> a male puma finding<br />

a territory in <strong>the</strong> 5-7 HRD range.<br />

Although nei<strong>the</strong>r model predicted <strong>the</strong><br />

higher number <strong>of</strong> settlements we found at >9<br />

and up to 20 HRDs, this finding also does<br />

DISPERSAL IN MALE PUMAS · Laundré and Hernández 157<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

not support <strong>the</strong> inbreeding avoidance<br />

hypo<strong>the</strong>sis. If a young male was dispersing<br />

primarily to avoid inbreeding, we would not<br />

predict such long distance dispersals where<br />

animals passed up what would appear to be<br />

ample suitable habitat in <strong>the</strong>ir dispersal<br />

movements. The only explanation for such<br />

long distance dispersals is <strong>the</strong>y had to<br />

continue to travel because <strong>the</strong>y were outcompeted<br />

by resident males or transients<br />

that had traveled fur<strong>the</strong>r than <strong>the</strong>mselves.<br />

Eventually, <strong>the</strong>y would gain weight and<br />

experience enough to successfully compete<br />

for a territory (Sweanor 1990).<br />

Sweanor (1990) originally suggested that<br />

male residency times, <strong>the</strong> polygynous<br />

mating system, and male dispersal distances<br />

in pumas argued against inbreeding<br />

avoidance being <strong>the</strong> main driving force in<br />

dispersal <strong>of</strong> young male pumas. However,<br />

after presenting fur<strong>the</strong>r evidence <strong>of</strong><br />

aggression among male pumas and<br />

inbreeding between fa<strong>the</strong>rs and daughters,<br />

Logan and Sweanor (2001) concluded that<br />

inbreeding avoidance was likely <strong>the</strong> main<br />

driving force behind male puma dispersal.<br />

Their argument centered on <strong>the</strong> assumption<br />

that young males should only disperse far<br />

enough to “avoid” competition. We contend<br />

that with 15 - 18 % <strong>of</strong> puma populations<br />

being primarily transient males (Hemker et<br />

al. 1984, Ross and Jalkotzy 1992,<br />

Spreadbury et al. 1996, Laundré and Clark<br />

2003), no dispersing male ever has <strong>the</strong><br />

luxury <strong>of</strong> avoiding competition for a<br />

territory. They concurred with <strong>the</strong> presence<br />

<strong>of</strong> <strong>the</strong>se transients and observed that new<br />

males were constantly coming into <strong>the</strong>ir<br />

population and concluded from this that<br />

competition must be “tolerable”. They<br />

fur<strong>the</strong>r argued that under <strong>the</strong> competition<br />

hypo<strong>the</strong>sis, however, most <strong>of</strong> <strong>the</strong> male<br />

recruits in <strong>the</strong>ir growing population should<br />

have been <strong>of</strong>fspring <strong>of</strong> <strong>the</strong> area. Because in<br />

<strong>the</strong>ir study most male recruits came from<br />

outside, <strong>the</strong>y argued that this refuted <strong>the</strong>


158 DISPERSAL IN MALE PUMAS · Laundré and Hernández<br />

competition hypo<strong>the</strong>sis. It is unclear as to<br />

what <strong>the</strong>y meant by “tolerable competition”<br />

but at any level <strong>of</strong> competition within its<br />

natal home range, a young male will be at a<br />

competitive disadvantage to his fa<strong>the</strong>r or any<br />

new male immigrant; <strong>the</strong>y will, by default,<br />

all be older and larger. Thus, under <strong>the</strong><br />

competition hypo<strong>the</strong>sis, we would expect no<br />

young males to compete successfully with<br />

<strong>the</strong>ir fa<strong>the</strong>rs or any older immigrant male<br />

that arrives to <strong>the</strong>ir home area.<br />

Consequently, we would predict that young<br />

males would all have to disperse, which is<br />

what all studies have found. Their chances<br />

<strong>of</strong> competing for a home range 1 HRD from<br />

<strong>the</strong>ir natal area is still low but greater than<br />

zero because <strong>the</strong>y are now slightly older and<br />

bigger and we would predict that some, not<br />

many, might establish territories that close<br />

to <strong>the</strong>ir natal home range. This is supported<br />

by <strong>the</strong> data presented, 2 out <strong>of</strong> 48 dispersing<br />

males set up territories within <strong>the</strong> first HRD.<br />

The fur<strong>the</strong>r <strong>the</strong>y disperse in time and space,<br />

<strong>the</strong> more competitive <strong>the</strong>y should be and we<br />

would predict that <strong>the</strong>y would start winning<br />

<strong>the</strong> competition for territories. Thus ra<strong>the</strong>r<br />

than refuting <strong>the</strong> competition model, <strong>the</strong> fact<br />

that Logan and Sweanor (2001) found only<br />

6 males in 10 years born in <strong>the</strong> area setting<br />

up territories within <strong>the</strong> study area starting at<br />

≈ 2.5 HRDs, actually supports and would<br />

have been predicted by <strong>the</strong> competition<br />

hypo<strong>the</strong>sis. Consequently, we contend that<br />

all <strong>the</strong> data presented by Logan and Sweanor<br />

(2001) concur with Sweanor’s (1990)<br />

original conclusion to reject <strong>the</strong> inbreeding<br />

hypo<strong>the</strong>sis and provide strong support for<br />

<strong>the</strong> competition model.<br />

In conclusion, after our analysis, we also<br />

concur with Sweanor’s (1990) original<br />

assessment and rejected <strong>the</strong> inbreeding<br />

model because resident males do fight and<br />

kill <strong>the</strong>ir male <strong>of</strong>fspring as well as younger<br />

transients, resident males do mate with <strong>the</strong>ir<br />

daughters, and dispersal distances were<br />

found to best fit that predicted by <strong>the</strong><br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

competition model. We conclude that young<br />

males are forced out <strong>of</strong> <strong>the</strong>ir natal home<br />

range by <strong>the</strong>ir fa<strong>the</strong>rs or incoming males<br />

who, by default, will be older and stronger.<br />

We propose that <strong>the</strong>y continue to disperse<br />

until <strong>the</strong>y gain weight and experience<br />

enough to successfully takeover a territory.<br />

What we propose is that <strong>the</strong> avoidance <strong>of</strong><br />

inbreeding is very likely not <strong>the</strong> driving<br />

force behind dispersal in young male pumas<br />

but ra<strong>the</strong>r, likely is a serendipitous<br />

consequence <strong>of</strong> dispersal driven by<br />

competitive interactions.<br />

ACKNOWLEDGEMENTS<br />

The data from our study was ga<strong>the</strong>red<br />

with <strong>the</strong> financial and logistic support <strong>of</strong> <strong>the</strong><br />

following foundations and agencies:<br />

ALSAM Foundation, Boone and Crockett<br />

Club, Earthwatch Institute, Idaho State<br />

University, National Rifle Association, The<br />

Eppley Foundation, U.S. Bureau <strong>of</strong> Land<br />

Management, The Nor<strong>the</strong>rn Rockies<br />

Conservation Cooperative, Idaho<br />

Department <strong>of</strong> Fish and Game, Mazamas,<br />

The Merril G. and Emita E. Hasting<br />

Foundation, Patagonia, Inc., SEACON <strong>of</strong><br />

<strong>the</strong> Chicago Zoological Society, The<br />

William H. and Mattie Wattis Harris<br />

Foundation, and Utah Division <strong>of</strong> Wildlife.<br />

We would like to thank <strong>the</strong> many<br />

Earthwatch volunteers without whose help<br />

this work would not have been<br />

accomplished. We thank J. Linnell and M.<br />

Culver for <strong>the</strong>ir helpful comments on this<br />

manuscript. We also thank <strong>the</strong> various<br />

researchers whose published data we used in<br />

our analysis. It is through research efforts<br />

like <strong>the</strong>irs that we are able to advance our<br />

understanding <strong>of</strong> puma ecology and<br />

behavior. Lastly, we especially thank Kevin<br />

Allred and Ken Jafek. It is only through<br />

<strong>the</strong>ir tireless enthusiasm and willing use <strong>of</strong><br />

<strong>the</strong>ir tracking dogs that our study was<br />

possible.


LITERATURE CITED<br />

ANDERSON, A.E., D.C. BOWDEN, AND D.M.<br />

KATTNER. 1992. The puma on <strong>the</strong><br />

Uncompahgre plateau, Colorado.<br />

Colorado Division <strong>of</strong> Wildlife Technical<br />

Publication No. 40.<br />

ASHMAN, D.J., G.C. CHRISTENSEN, M.L.<br />

HESS, G.K. TSUKAMOTO AND M.S.<br />

WICHERSHAM. 1983. The mountain lion<br />

in Nevada. Nevada Department <strong>of</strong><br />

Wildlife Report W-48-15, Reno,<br />

Nevada, USA.<br />

BARONE, M.A., M.E. ROELKE, J. HOWARD,<br />

J.L. BROWN, A.E. ANDERSON, AND D.E.<br />

WILDT. 1994. Reproductive<br />

characteristics <strong>of</strong> male Florida pan<strong>the</strong>rs:<br />

Comparative studies from Florida,<br />

Texas, Colorado, Latin America, and<br />

North American Zoos. Journal <strong>of</strong><br />

Mammalogy 75:150-162.<br />

BEIER, P. 1995. Dispersal <strong>of</strong> juvenile<br />

cougars in fragmented habitat. Journal<br />

<strong>of</strong> Wildlife Management 59:228-237.<br />

DOBSON, F.S. 1982. Competition for mates<br />

and predominant juvenile male dispersal<br />

in mammals. Animal Behavior 30:1183-<br />

1192.<br />

HEMKER, T.P., F.G. LINDZEY, AND B.B.<br />

ACKERMAN. 1984. Population<br />

characteristics and movement patterns <strong>of</strong><br />

cougars in sou<strong>the</strong>rn Utah. Journal <strong>of</strong><br />

Wildlife Management 48:1275-1284.<br />

HORNOCKER, M.G. 1970. An analysis <strong>of</strong><br />

mountain lion predation upon mule deer<br />

and elk in <strong>the</strong> Idaho Primitive area.<br />

Wildlife Monographs # 21, The Wildlife<br />

Society.<br />

LAUNDRÉ, J.W. AND L. HERNÁNDEZ. 2002.<br />

Growth curve models and age estimation<br />

<strong>of</strong> young cougars in <strong>the</strong> nor<strong>the</strong>rn Great<br />

Basin. Journal <strong>of</strong> Wildlife Management<br />

66:849-858.<br />

LAUNDRÉ, J.W. AND T.W. CLARK. 2003.<br />

Managing puma hunting in <strong>the</strong> western<br />

United States: through a metapopulation<br />

DISPERSAL IN MALE PUMAS · Laundré and Hernández 159<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

approach. Animal Conservation 6:159-<br />

170.<br />

LINDZEY, F.G., B.B. ACKERMAN, D.<br />

BARNHURST, T. BECKER, T.P. HEMKER,<br />

S.P. LAING, C. MECHAM, AND W.D.<br />

VANSICKLE. 1989. Boulder-Escalante<br />

cougar project. Final Report Utah<br />

Division <strong>of</strong> Wildlife Resources, Salt<br />

Lake City, Utah. USA.<br />

LINDZEY, F.G., W.D. VAN SICKEL, B.B.<br />

ACHERMAN, D. BARNHURST, T.P.<br />

HEMKER, AND S.P. LAING. 1994.<br />

Cougar population dynamics in sou<strong>the</strong>rn<br />

Utah. Journal <strong>of</strong> Wildlife Management<br />

58:619-624.<br />

LOGAN, K.A., L.L. IRWIN, AND R. SKINNER.<br />

1986. Characteristics <strong>of</strong> a hunted<br />

mountain lion population in Wyoming.<br />

Journal <strong>of</strong> Wildlife Management 50:648-<br />

654.<br />

LOGAN, K.A. AND L.L. SWEANOR. 2001.<br />

Desert Puma Evolutionary Ecology and<br />

conservation <strong>of</strong> an Enduring Carnivore.<br />

Island Press, Washington D.C, USA.<br />

MAEHR, D., E.D. LAND, AND J.C. ROOF.<br />

1991. Social ecology <strong>of</strong> Florida<br />

pan<strong>the</strong>rs. National Geographic Research<br />

and Exploration 7:414-431.<br />

MCBRIDE, R.T. 1976. The status and<br />

ecology <strong>of</strong> <strong>the</strong> mountain lion (Felis<br />

concolor) <strong>of</strong> <strong>the</strong> Texas-Mexico border.<br />

Thesis, Sul Ross State University,<br />

Alpine, Texas, USA.<br />

MILLER, G.L. AND B.W. CARROLL. 1989.<br />

Modeling vertebrate dispersal distances:<br />

alternatives to <strong>the</strong> geometric distribution.<br />

Ecology 70:977-986.<br />

MOORE, J. AND R. ALI. 1984. Are dispersal<br />

and inbreeding avoidance related?<br />

Animal Behaviour 32:94-112.<br />

MURPHY, K.M. 1998. The ecology <strong>of</strong> <strong>the</strong><br />

cougar (Puma concolor) in <strong>the</strong> nor<strong>the</strong>rn<br />

Yellowstone ecosystem: Interactions<br />

with prey, bears, and humans.<br />

Dissertation. University <strong>of</strong> Idaho,<br />

Moscow, Idaho, USA.


160 DISPERSAL IN MALE PUMAS · Laundré and Hernández<br />

ROBINETTE, W.L., J.S. GASHWILER, AND<br />

O.W. MORRIS. 1961. Notes on cougar<br />

productivity and life history. Journal <strong>of</strong><br />

Mammalogy 42:204-217.<br />

ROELKE, M.E., J.S. MARTENSON, AND S.J.<br />

O’BRIEN. 1993. The consequences <strong>of</strong><br />

demographic reduction and genetic<br />

depletion in <strong>the</strong> endangered Florida<br />

pan<strong>the</strong>r. Current Biology 3:340-350.<br />

ROSS, P.I. AND M.G. JALKOTZY. 1992.<br />

Characteristics <strong>of</strong> a hunted population <strong>of</strong><br />

cougars in southwestern Alberta.<br />

Journal <strong>of</strong> Wildlife Management 56:417-<br />

426.<br />

SEIDENSTICKER, J.C. IV, M.G. HORNOCKER,<br />

W.V. WILES, AND J.P. MESSICK. 1973.<br />

<strong>Mountain</strong> lion social organization in <strong>the</strong><br />

Idaho Primitive area. Wildlife<br />

Monographs # 35. The Wildlife Society.<br />

SHIELDS, W.M. 1987. Dispersal and mating<br />

systems: investigating <strong>the</strong>ir causal<br />

connections. Pages 3-24 in B.D.<br />

Chepko-Sade and S. T. Halpin editors.<br />

Mammalian Dispersal Patterns <strong>the</strong><br />

Effects <strong>of</strong> Social Structure on Population<br />

Genetics. University <strong>of</strong> Chicago Press,<br />

Chicago, Illinois, USA.<br />

SITTON, L.W. AND R.A. WEAVER. 1977.<br />

California mountain lion investigations<br />

with recommendations for management.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Report to <strong>the</strong> California Department <strong>of</strong><br />

Fish and Game. Sacramento, California,<br />

USA.<br />

SPREADBURY, B.R., K. MUSIL, J. MUSIL, C.<br />

KAISNER, AND J. KOVAK. 1996. Cougar<br />

population characteristics in<br />

sou<strong>the</strong>astern British Columbia. Journal<br />

<strong>of</strong> Wildlife Management. 60:962-969.<br />

SWEANOR, L.L. 1990. <strong>Mountain</strong> lion social<br />

organization in a desert environment.<br />

Thesis, University <strong>of</strong> Idaho, Moscow,<br />

Idaho, USA.<br />

SWEANOR, L.L., K.A. LOGAN, AND M.G.<br />

HORNOCKER. 2000. Cougar dispersal<br />

patterns, metapopulation dynamics and<br />

conservation. Conservation Biology<br />

14:798-808Waser, P. M. 1985. Does<br />

competition drive dispersal? Ecology<br />

66:1170-1175.<br />

WASER, P.M. 1987. A model predicting<br />

dispersal distance distributions. Pages<br />

251-256 in B.D. Chepko-Sade and S. T.<br />

Halpin editors Mammalian Dispersal<br />

Patterns <strong>the</strong> Effects <strong>of</strong> Social Structure<br />

on Population Genetics. University <strong>of</strong><br />

Chicago Press, Chicago, Illinois, USA.<br />

WOLFE, J.O. 1994. More on juvenile<br />

dispersal in mammals. Oikos 71:349-<br />

352.


COUGAR EXPLOITATION LEVELS AND LANDSCAPE CONFIGURATION:<br />

IMPLICATIONS FOR DEMOGRAPHIC STRUCTURE AND METAPOPULATION<br />

DYNAMICS<br />

DAVID C. STONER, Utah State University, Dept. <strong>of</strong> Forest, Range, and Wildlife Sciences,<br />

5230 Old Main Hill, Logan, UT, 84322-5230, USA, email: dstoner@cc.usu.edu<br />

MICHAEL L. WOLFE, Utah State University, Dept. <strong>of</strong> Forest, Range, and Wildlife Sciences,<br />

5230 Old Main Hill, Logan, UT, 84322-5230, USA, email: mlwolfe@cnr.usu.edu<br />

Abstract: Currently eleven states and two Canadian provinces utilize sport hunting as <strong>the</strong><br />

primary mechanism for managing cougar (Puma concolor) populations. However <strong>the</strong> impacts <strong>of</strong><br />

sustained harvest on demographic structure and population persistence are not well understood.<br />

Additionally, <strong>the</strong> range <strong>of</strong> non-biological factors influencing <strong>the</strong> rate <strong>of</strong> population recovery has<br />

not been thoroughly examined. We have been monitoring <strong>the</strong> cougar populations on Monroe<br />

<strong>Mountain</strong> in south-central Utah, and in <strong>the</strong> Oquirrh <strong>Mountain</strong>s <strong>of</strong> north-central Utah since 1996<br />

and 1997, respectively. The critical management distinction between <strong>the</strong>se sites is <strong>the</strong> degree <strong>of</strong><br />

exploitation. The Monroe population is subjected to heavy annual hunting pressure and is<br />

characterized demographically by a younger age distribution, low survivorship, low fecundity,<br />

and declining density. In contrast, <strong>the</strong> population inhabiting <strong>the</strong> nor<strong>the</strong>astern slope <strong>of</strong> <strong>the</strong><br />

Oquirrhs is subjected to little or no hunting pressure and exhibits an older age distribution,<br />

relatively high survivorship and fecundity, a stable density, and a high emigration rate. Due in<br />

part to <strong>the</strong>se differences, <strong>the</strong> Oquirrh and Monroe populations appear to exhibit source and sink<br />

dynamics within <strong>the</strong> regional metapopulation. Therefore <strong>the</strong> temporal scale <strong>of</strong> population<br />

recovery may depend on <strong>the</strong> interaction between <strong>the</strong> dominant harvest regime and <strong>the</strong> degree <strong>of</strong><br />

landscape connectivity with neighboring patches. Aside from harvest, <strong>the</strong> interaction between<br />

patch configuration and anthropogenic fragmentation may be highly influential in <strong>the</strong> long-term<br />

prognosis for <strong>the</strong>se populations. We discuss <strong>the</strong> implications <strong>of</strong> <strong>the</strong>se demographic distinctions<br />

in light <strong>of</strong> enumeration uncertainties, habitat fragmentation, and landscape structure.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

161


162<br />

ASSESSING GPS RADIOTELEMETRY RELIABILITY IN COUGAR HABITAT<br />

TRISH GRISWOLD<br />

JAMES BRIGGS<br />

GARY KOEHLER<br />

STUDENTS AT CLE ELUM-ROSLYN SCHOOL DISTRICT, Cle Elum, Washington<br />

Abstract: Studies evaluating <strong>the</strong> effectiveness <strong>of</strong> GPS radiotelemetry have shown that <strong>the</strong><br />

positional accuracy and rate <strong>of</strong> GPS fixes declines with increased forest canopy coverage<br />

(D’Eon, Serrouya, Smith, and Kochanny, 2002. Wildlife Society Bulletin, 30(2):430-439). Since<br />

GPS collars are being used to mark and monitor cougars (Puma concolor; Koehler and Nelson,<br />

7 th <strong>Mountain</strong> <strong>Lion</strong> <strong>Workshop</strong>), students, faculty, and volunteers at <strong>the</strong> Cle Elum-Roslyn Middle<br />

School, Washington, tested GPS location accuracy as part <strong>of</strong> Project CAT (Cougars and<br />

Teaching). We fitted domestic dogs (Canis familiaris) with <strong>the</strong> same GPS collars used to mark<br />

cougars and locational accuracy was measured in areas <strong>of</strong> known cougar habitat. GPS fixes were<br />

recorded and compared with UTM coordinates obtained from hand-held GPS receivers and 7.5minute<br />

topographic maps. Environmental factors, vegetation types, and physiographic<br />

parameters were recorded. It was felt that <strong>the</strong> dogs would closely approximate cougar movement<br />

patterns and give an index <strong>of</strong> reliability <strong>of</strong> GPS fixes for free-ranging cougars. While previous<br />

studies have addressed <strong>the</strong> reliability <strong>of</strong> GPS collar fixes, none have tested reliability <strong>of</strong> data<br />

collected in <strong>the</strong> rapidly suburbanizing ponderosa pine (Pinus ponderosa) and Douglas fir<br />

(Pseudosuga menziesii) forests <strong>of</strong> <strong>the</strong> eastern Cascade <strong>Mountain</strong>s. This project gives <strong>the</strong> middle<br />

school students an opportunity to participate in <strong>the</strong> school-wide educational effort <strong>of</strong> cougar<br />

ecology. Students proposed and tested hypo<strong>the</strong>ses and analyzed <strong>the</strong> data.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


USING GPS COLLARS TO DETERMINE COUGAR KILL RATES, ESTIMATE HOME<br />

RANGES, AND EXAMINE COUGAR-COUGAR INTERACTIONS<br />

POLLY C. BUOTTE, Research Assistant and GIS Specialist, Wildlife Conservation Society,<br />

2023 Stadium Dr. Suite 1A, Bozeman, MT 59030, USA, email:<br />

polly_thornton@hotmail.com<br />

TONI K. RUTH, Associate Conservation Scientist, Wildlife Conservation Society, 2023 Stadium<br />

Dr. Suite 1A, Bozeman, MT 59030, USA, email: truth@montanadsl.net<br />

Abstract: Single-species approaches to large carnivore conservation limits our understanding <strong>of</strong><br />

carnivore assemblages and interactions at a community level and obtaining data on wide ranging,<br />

secretive species such as cougars and wolves can be particularly challenging. Beginning in<br />

2001, we collaborated with <strong>the</strong> Yellowstone Wolf Project and Interagency Grizzly Bear Team in<br />

applying GPS technology to examine patterns <strong>of</strong> resource use among cougars, wolves, and<br />

grizzly bears. In this paper, we address three topics relative to future analysis <strong>of</strong> GPS data on all<br />

three carnivores: 1) <strong>the</strong> efficacy <strong>of</strong> finding cougar-killed elk and deer carcasses through GPS<br />

locations, 2) differences in home range estimation from GPS versus VHF locations with<br />

implications to analysis <strong>of</strong> species overlap; and 3) interaction between two cougars with<br />

implications to addressing spatial-temporal interactions between cougars, wolves, and bears. We<br />

deployed store-on-board GPS collars (GPS Generation II, Telonics, Inc.) on two adult male<br />

cougars (M137 and M127) during <strong>the</strong> winter <strong>of</strong> 2001. Male M137’s collar acquired 612 GPS<br />

locations between Feb 11 and June 13, with a successful fix rate <strong>of</strong> 59.9%. Male M127’s collar<br />

acquired 370 GPS locations between Feb 27 and May 1, with a successful fix rate <strong>of</strong> 73.4%.<br />

Each collar was programmed to attempt a GPS fix every third hour, or eight times per day. We<br />

identified clusters <strong>of</strong> locations by calculating distance moved between consecutive GPS locations<br />

and by selecting groups <strong>of</strong> locations within 200 meters <strong>of</strong> each o<strong>the</strong>r. Identified clusters were<br />

located and searched in <strong>the</strong> field utilizing a hand-held GPS. For cougar M127, we additionally<br />

documented kill rate via intensive daily ground-based VHF telemetry sampling between March 5<br />

and April 10. Both ground and GPS methods yielded four kills during that time span. To<br />

examine differences in identification <strong>of</strong> home ranges we calculated home ranges using fixed<br />

kernel analysis. Male M137’s GPS data included a disjunct area <strong>of</strong> approximately 15 km 2 that<br />

was not identified from VHF locations. Preliminary analysis <strong>of</strong> interaction indicated two times<br />

when both cougars were at <strong>the</strong> same location, after which <strong>the</strong> subordinate male M127 moved<br />

away from <strong>the</strong> dominant male M137. During winter 2003, we deployed 5 Televilt Simplex GPS<br />

collars on cougars. The collars allow for remote downloads <strong>of</strong> data and are programmed to<br />

acquire locations simultaneous to locations <strong>of</strong> GPS collared wolves. Our goals during <strong>the</strong> next<br />

two years are to: 1) develop correction factors for both ground-based and GPS collected kill rates<br />

and, in collaboration with <strong>the</strong> Yellowstone Wolf Project and <strong>the</strong> Interagency Grizzly Bear Team<br />

to, 2) quantify spatial-temporal interactions between cougars, wolves and bears via subsequent<br />

moves analyses and utilize <strong>the</strong>se data to develop a predictive model <strong>of</strong> carnivore movement and<br />

landscape use.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

163


164<br />

FUNCTIONAL RESPONSE OF COUGARS AND PREY AVAILABILITY IN<br />

NORTHEASTERN WASHINGTON<br />

HILARY S. CRUICKSHANK, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural<br />

Resource Sciences, Washington State University, PO Box 646410, Pullman, WA 99164,<br />

USA, email: hcruicks@mail.wsu.edu<br />

HUGH S. ROBINSON, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99164, USA,<br />

email: hsrobins@wsunix.wsu.edu<br />

CATHERINE LAMBERT, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99164, USA,<br />

email: lambertcath@hotmail.com<br />

ROBERT B. WIELGUS, Large Carnivore Conservation Lab, Department <strong>of</strong> Natural Resource<br />

Sciences, Washington State University, PO Box 646410, Pullman, WA 99164, USA,<br />

email: wielgus@mail.wsu.edu<br />

Abstract: Within <strong>the</strong> last ten years, a major change in <strong>the</strong> population structure <strong>of</strong> deer in western<br />

North America has taken place. Mule deer populations are sharply declining, while white-tailed<br />

deer populations are increasing. Researchers have suggested that cougar predation is a possible<br />

reason for <strong>the</strong> decline. The purpose <strong>of</strong> this project is to investigate cougar predation in a<br />

community where substantial populations <strong>of</strong> white-tailed deer, mule deer, and cougars overlap.<br />

We are testing two alternative hypo<strong>the</strong>ses <strong>of</strong> cougar prey selection. H1, or <strong>the</strong> apparent selection<br />

hypo<strong>the</strong>sis, states that equal selection by cougars for white-tailed deer and mule deer, but a<br />

higher reproductive rate by white-tailed deer is causing a decline in <strong>the</strong> mule deer population.<br />

H2 proposes that higher selection by cougars for mule deer is causing a decline in <strong>the</strong> population.<br />

Preliminary results suggest H2. The effect <strong>of</strong> predation on prey is determined by two factors: 1)<br />

functional response, and 2) prey availability. Functional response <strong>of</strong> cougars is quantified by <strong>the</strong><br />

number <strong>of</strong> kills, per cougar, per unit time, and prey availability provides an estimate <strong>of</strong> <strong>the</strong><br />

number and distribution <strong>of</strong> each prey species. The combination <strong>of</strong> <strong>the</strong>se two factors may <strong>of</strong>fer a<br />

more complete understanding <strong>of</strong> cougar prey selection. This research is in support <strong>of</strong> a larger<br />

study, which will use <strong>the</strong> apparent competition <strong>the</strong>ory to examine alternative cougar management<br />

strategies.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


WHAT DOES TEN YEARS (1993-2002) OF MOUNTAIN LION OBSERVATION DATA<br />

REVEAL ABOUT MOUNTAIN LION–HUMAN INTERACTIONS WITHIN REDWOOD<br />

NATIONAL AND STATE PARKS?<br />

GREGORY W. HOLM, Wildlife Biologist, Redwood National Park, 219 Hilton Rd., Orick, CA<br />

95519, USA, email: Gregory_Holm@nps.gov<br />

Abstract: <strong>Mountain</strong> lions (Puma concolor) occur throughout Redwood National and State Parks<br />

(RNSP) and most o<strong>the</strong>r portions <strong>of</strong> northwest California. However, because <strong>the</strong>y are not <strong>of</strong>ten<br />

observed, RNSP biologists have always been interested in recording mountain lion observations<br />

within <strong>the</strong> park. Prior to 4 mountain lion attacks in California between 1992-94, 2 resulting in<br />

fatalities, RNSP mountain lion observations were not compiled in a timely, consistent, or easily<br />

accessible manner. Since 1993, RNSP biologists have attempted to document and verify all<br />

mountain lion observations using a standard reporting form and database. Three hundred and<br />

seven mountain lion observations have been recorded within RNSP from 1993-2002 (mean ≅ 31;<br />

range 19-53). Most were observed between May and October during daylight hours, and<br />

involved a single mountain lion. While most observations (54%) involved a mountain lion near<br />

a road while people were in a vehicle, <strong>the</strong> remaining observations (46%) occurred while people<br />

were on trails or at o<strong>the</strong>r park facilities. The ultimate response <strong>of</strong> most mountain lions (68%)<br />

encountered on trails was to avoid humans, yet twenty percent <strong>of</strong> trail encounters involved some<br />

level <strong>of</strong> curiosity by <strong>the</strong> mountain lion towards humans. Although no human attacks were<br />

reported, <strong>the</strong>re were 8 reports <strong>of</strong> aggressive behavior towards humans, 6 reports <strong>of</strong> following<br />

humans, and 1 report <strong>of</strong> a dog on a leash getting attacked. The observation data does not<br />

accurately reflect <strong>the</strong> actual distribution or timing (seasonally or daily) <strong>of</strong> mountain lion activity,<br />

and should be interpreted with caution due to inherent problems with observer experience, and<br />

report quality and verification. However, <strong>the</strong> information does allow managers to quickly<br />

identify when and where mountain lion-human interactions have occurred, and more effectively<br />

focus management actions to prevent or reduce future mountain lion-human interactions within<br />

RNSP.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

165


166<br />

DEPREDATION TRENDS IN CALIFORNIA<br />

SARAH REED, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA 95814, USA<br />

CHRISTOPHER M. PAPOUCHIS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95812, USA, email: cpapouchis@mountainlion.org<br />

LYNN MICHELLE CULLENS, <strong>Mountain</strong> <strong>Lion</strong> Foundation, PO Box 1896, Sacramento, CA<br />

95812, USA, email: cullens@mountainlion.org<br />

Abstract: Since 1972 more than 1,600 California mountain lions have been killed under<br />

depredation permits. The number <strong>of</strong> lions killed annually has increased, with a peak <strong>of</strong> 149 lions<br />

killed in 2000. Although some permits are issued for losses incurred by traditional,<br />

economically viable, open range livestock operations, incidents on ranchettes and "hobby farms"<br />

are increasing. We evaluate trends in depredation permitting, including analysis <strong>of</strong> depredation<br />

events geographically, by parcel size and size <strong>of</strong> herd, and relative to human population and<br />

development trends. We conclude with new approaches to mountain lion conservation that stress<br />

science to establish a factual basis for dialogue, community involvement, and developing<br />

partnerships with diverse stakeholders.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


THE DISTRIBUTION OF PERCEIVED ENCOUNTERS WITH NON-NATIVE CATS IN<br />

SOUTH AND WEST WALES, UK: RELATIONSHIP TO MODELED HABITAT<br />

SUITABILITY<br />

A.B. SMITH, Exotic Cat Group, Geography Department, University <strong>of</strong> Wales Swansea,<br />

Singleton Park, Swansea SA2 8PP, UK, email: exoticcatproject@swan.ac.uk<br />

F.A. STREET PERROTT, Exotic Cat Group, Geography Department, University <strong>of</strong> Wales<br />

Swansea, Singleton Park, Swansea SA2 8PP, UK<br />

T. HOOPER, Exotic Animals Register, 85 Risedale Road, Ashton Vale, Bristol BS3 2RB, UK<br />

Abstract: Reports <strong>of</strong> perceived encounters with exotic cats in <strong>the</strong> British countryside have<br />

greatly increased in recent years. The species described (notably melanistic leopards, pumas and<br />

lynxes) were widely bred in <strong>the</strong> UK prior to <strong>the</strong> 1976 Dangerous Wild Animals Act, and do not<br />

correspond to those most familiar to <strong>the</strong> general public (such as lions, tigers and cheetahs).<br />

Some <strong>of</strong> <strong>the</strong>m are still being illegally imported or reared for ‘canned hunts’. Following <strong>the</strong><br />

recent discovery <strong>of</strong> leopard tracks in West Wales, and calls for action by Members <strong>of</strong> <strong>the</strong> UK<br />

Parliament and <strong>the</strong> Welsh National Assembly, <strong>the</strong> Welsh Agriculture Department has <strong>of</strong>ficially<br />

begun to collect statistics on sightings and livestock kills. In this independent study, we have<br />

analysed a database <strong>of</strong> 170 georeferenced encounter reports obtained from <strong>the</strong> police, news<br />

media and members <strong>of</strong> <strong>the</strong> public. In <strong>the</strong> absence <strong>of</strong> confirmatory DNA, physical or photographic<br />

evidence, encounter reports require very careful screening for reliability, based on <strong>the</strong><br />

characteristics <strong>of</strong> <strong>the</strong> witness(es); <strong>the</strong> validity <strong>of</strong> <strong>the</strong> identification, taking into account perceived<br />

cat size and shape (morphotype) and behaviour; indicators <strong>of</strong> scale, distance and lighting<br />

conditions; and <strong>the</strong> suitability <strong>of</strong> <strong>the</strong> habitat. The distributions <strong>of</strong> potential habitats within South<br />

and West Wales have been modelled with a GIS using standard habitat characteristics, such as<br />

prey-species presence, disturbance levels, geomorphology and land-use data. The spatial pattern<br />

<strong>of</strong> encounters does not show <strong>the</strong> clustering that might be expected if <strong>the</strong>y represent a purely<br />

sociological phenomenon. Instead, <strong>the</strong> distributions <strong>of</strong> specific morphotypes appear to be closely<br />

related to <strong>the</strong> degree <strong>of</strong> habitat suitability, <strong>the</strong>reby streng<strong>the</strong>ning <strong>the</strong> case for <strong>the</strong> presence <strong>of</strong> nonnative<br />

cat species in <strong>the</strong> UK.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

167


168<br />

PUMA ACTIVITY AND MOVEMENTS IN A HUMAN-DOMINATED LANDSCAPE:<br />

CUYAMACA RANCHO STATE PARK AND ADJACENT LANDS IN SOUTHERN<br />

CALIFORNIA<br />

LINDA L. SWEANOR, Wildlife Health Center, University <strong>of</strong> California - Davis, Sou<strong>the</strong>rn<br />

California Puma Project Field Station: P.O. Box 1114, Julian, CA 92036, USA, email:<br />

lsweanor@mindspring.com<br />

KENNETH A. LOGAN, Wildlife Health Center, University <strong>of</strong> California - Davis, Sou<strong>the</strong>rn<br />

California Puma Project Field Station: P.O. Box 1114, Julian, CA 92036, USA, email:<br />

klogan2@mindspring.com<br />

JIM W. BAUER, Wildlife Health Center, University <strong>of</strong> California - Davis, Sou<strong>the</strong>rn California<br />

Puma Project Field Station, P. O. Box 1203, Julian, CA 92036, USA, email:<br />

jwbauer@uia.net<br />

WALTER M. BOYCE, Wildlife Health Center, University <strong>of</strong> California - Davis, One Shields<br />

Avenue, Davis, CA 95616, USA, email: wmboyce@ucdavis.edu<br />

Abstract: Although puma attacks are exceedingly rare, statistics indicate dangerous encounters<br />

between humans and pumas are on <strong>the</strong> rise. In California <strong>the</strong>re have been 7 verified puma attacks<br />

resulting in 2 human deaths during <strong>the</strong> last 10 years; 2 <strong>of</strong> those attacks and 1 death occurred in<br />

Cuyamaca Rancho State Park (CRSP). Because <strong>of</strong> <strong>the</strong> high number <strong>of</strong> reported puma sightings<br />

in CRSP each year (range = 18-50 from 1993-2002) and <strong>the</strong> large, increasing number <strong>of</strong> human<br />

visitors (over 500,000 people visited <strong>the</strong> 50-square-mile park in 2001), park authorities were<br />

concerned about <strong>the</strong> potential for fur<strong>the</strong>r dangerous puma-human encounters. A study was<br />

initiated in January 2001 to understand puma behavior relative to human activity, to help<br />

minimize conflicts between pumas and humans, and to assist <strong>the</strong> development <strong>of</strong> long-term<br />

conservation strategies for pumas in <strong>the</strong> CRSP area. Specific objectives <strong>of</strong> <strong>the</strong> project were to:<br />

determine <strong>the</strong> number and characteristics <strong>of</strong> pumas using CRSP; map puma home ranges and<br />

determine important puma habitats and <strong>the</strong>ir juxtaposition relative to human use areas; examine<br />

puma movements (e.g., daily, seasonal) relative to areas <strong>of</strong> human activity (e.g., trails, roads,<br />

campgrounds); examine puma diet to determine what prey species are most important as puma<br />

food and to what extent, if any, domestic animals contribute to <strong>the</strong>ir diet; and use <strong>the</strong> data to<br />

formulate management recommendations. To obtain information on puma home ranges,<br />

movements and behavior, and to find prey killed by pumas, independent and adult pumas are<br />

being captured and fitted with Televilt GPS collars. As <strong>of</strong> March 2003, 11 pumas (6 adult males,<br />

5 adult females) had been captured in and around CRSP and fitted with GPS collars. To date,<br />

collars have yielded over 6400 locations. Additionally, human use <strong>of</strong> trails is being measured<br />

seasonally by placing infrared counters (TrailMaster monitors) along 4 trail systems within<br />

CRSP. This paper reports on project progress to date.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


MODELING OFFSPRING SEX RATIOS AND GROWTH OF COUGARS<br />

DIANA M. GHIKAS, Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Calgary, 2500<br />

University Drive N.W., Calgary, AB T2N 1N4, Canada, email: dghikas@ucalgary.ca<br />

MARTIN JALKOTZY, Arc Wildlife Services Ltd., 3527-35th Avenue S.W., Calgary, AB T3E<br />

1A2, Canada, email: martin_jalkotzy@nucleus.com<br />

IAN ROSS, Arc Wildlife Services Ltd., 3527-35th Avenue S.W., Calgary, AB T3E 1A2, Canada<br />

RALPH SCHMIDT, Arc Wildlife Services Ltd., 3527-35th Avenue S.W., Calgary, AB T3E<br />

1A2, Canada<br />

SHANE A. RICHARDS, Department <strong>of</strong> Biological Sciences, University <strong>of</strong> Calgary, 2500<br />

University Drive N.W., Calgary, AB T2N 1N4, Canada<br />

Abstract: We fur<strong>the</strong>r examined data from <strong>the</strong> Sheep River cougar study conducted in southwest<br />

Alberta from 1981-94. We asked is <strong>the</strong>re evidence <strong>of</strong> an equal or constant <strong>of</strong>fspring sex ratio, or<br />

do sex ratios vary over time, as a function <strong>of</strong> <strong>the</strong> mo<strong>the</strong>r's age, geographic location or population<br />

size. Logan and Sweanor (2001) analysed <strong>of</strong>fspring sex ratios as a function <strong>of</strong> mo<strong>the</strong>r's age and<br />

found that sex ratios <strong>of</strong> first litters were significantly different from subsequent litters and 1:1.<br />

Logan and Sweanor (2001) suggested that <strong>of</strong>fspring sex ratios might be influenced by <strong>the</strong><br />

mo<strong>the</strong>r's physical state (i.e., young mo<strong>the</strong>rs produce less-costly females so energy can be<br />

allocated to growth). We investigated possible relationships between <strong>of</strong>fspring sex ratios and<br />

cougar growth, and whe<strong>the</strong>r growth varied by sex and geographic location. The study area was<br />

divided into east and west locations based coarsely on prey abundance and cougar mortality.<br />

Probabilistic models were formulated for <strong>the</strong> sex-ratio analysis. A deterministic model based on<br />

<strong>the</strong> flexible Richards curve (Maehr and Moore 1992) was used to predict mass growth. Model<br />

predictions that were <strong>the</strong> most parsimonious with <strong>the</strong> data were identified using corrected Akaike<br />

Information Criterion. Parameters were estimated using maximum log-likelihood. The most<br />

parsimonious model predicted that <strong>of</strong>fspring sex ratios vary yearly. Evidence was not strong for<br />

sex ratios varying as a linear function <strong>of</strong> mo<strong>the</strong>r's age. The growth model predicted that females<br />

attain 91-92 % <strong>of</strong> adult mass by 25-26 mos, indicating that growth is largely completed prior to<br />

first reproduction ( x =30.0 ± 1.8 mos SE, Ross and Jalkotzy 1992). In years with poor resource<br />

conditions, <strong>the</strong> mo<strong>the</strong>r's physical state may result in more female <strong>of</strong>fspring being reared than<br />

male. Predicted mean mass at age <strong>of</strong> independence for male and female <strong>of</strong>fspring was 48.8 kg<br />

and 34.1 kg, respectively, inferring males are more costly to rear. The growth model that varied<br />

by sex only had <strong>the</strong> highest weight <strong>of</strong> evidence; adding geographic location did not result in a<br />

more parsimonious model. All growth models were unable to accurately estimate birth mass,<br />

which was also found by Maehr and Moore (1992).<br />

LITERATURE CITED<br />

Logan, K. A., and L.L. Sweanor. 2001. Desert Puma: Evolutionary ecology and conservation <strong>of</strong> an enduring carnivore. Island<br />

Press, Washington.<br />

Maehr, D.S., and C.T. Moore. 1992. Models <strong>of</strong> mass growth for 3 North American cougar populations. J. Wildl. Manage. 56:<br />

700-707.<br />

Ross, P.I., and M.G. Jalkotzy. 1992. Characteristics <strong>of</strong> a hunted population <strong>of</strong> cougars in southwestern Alberta. J. Wildl.<br />

Manage. 56: 417-426.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

169


170<br />

MOUNTAIN LION SURVEY TECHNIQUES IN NORTHERN IDAHO: A THREE-FOLD<br />

APPROACH<br />

CRAIG G. WHITE, Idaho Department <strong>of</strong> Fish and Game, 1540 Warner Avenue, Lewiston, ID<br />

83501, USA, email: cwhite@idfg.state.id.us<br />

PETER ZAGER, Idaho Department <strong>of</strong> Fish and Game, 1540 Warner Avenue, Lewiston, ID<br />

83501, USA, email: pzager@idfg.stat.id.us<br />

LISETTE WAITS, Department <strong>of</strong> Fish and Wildlife Resources, University <strong>of</strong> Idaho, Moscow,<br />

ID 83843, USA, email: lwaits@uidaho.edu<br />

Abstract: Management <strong>of</strong> mountain lions (Puma concolor) in Idaho relies largely on harvest<br />

data. This type <strong>of</strong> data is limited in scope and relays little information to <strong>the</strong> manager regarding<br />

population trend or density. Intensive radio telemetry studies involving capture and recapture<br />

can provide an estimation <strong>of</strong> density but are expensive. Currently researchers are exploring<br />

techniques to index or estimate population size by identifying individuals by <strong>the</strong>ir DNA. We<br />

outline three different techniques to “capture” and “recapture” mountain lion hair and/or tissue<br />

for DNA analysis: biopsy darts, rub tree stations, and legally harvested lions. Techniques are<br />

being implemented on two study areas in north-central Idaho, <strong>the</strong> Lochsa/North Fork <strong>of</strong> <strong>the</strong><br />

Clearwater River and <strong>the</strong> South Fork <strong>of</strong> <strong>the</strong> Clearwater River. Efforts by both volunteer<br />

houndsmen and hired houndsmen over 1½ lion harvest seasons have resulted in ≥15 DNA<br />

samples from lion treed and released. Over <strong>the</strong> same period ≥15 DNA samples have been turned<br />

in from legally harvested lions. In 2002, we placed 51 rub stations and recorded 42 visits over<br />

<strong>the</strong> 3 sampling periods. Seventeen <strong>of</strong> <strong>the</strong> visits resulted in ≥1 hair. Preliminary results indicate<br />

that 1 visit was from a lion, 7 visits were by bear, and 7 visits were possibly a lion or bear, and 2<br />

visits by o<strong>the</strong>r species. Improvements to <strong>the</strong> techniques are ongoing. This study will allow us to<br />

identify individuals in <strong>the</strong> mountain lion populations within our study areas and thus obtain a<br />

minimum population size. Number <strong>of</strong> captures each year will serve as an index <strong>of</strong> population<br />

trend. We will also explore <strong>the</strong> use <strong>of</strong> capture-recapture modeling to estimate population sizes.<br />

Our approach attempts to limit resources expended in capturing and marking animals, while still<br />

providing an index and potential population estimate within our study areas.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


MOUNTAIN LIONS IN SOUTH DAKOTA: RESULTS OF A 2002 PUBLIC OPINION<br />

SURVEY<br />

LARRY M. GIGLIOTTI, Planning Coordinator/Human Dimensions Specialist, South Dakota<br />

Department <strong>of</strong> Game, Fish and Parks, 523 E. Capitol, Pierre, SD 57501, USA, email:<br />

larry.gigliotti@state.sd.us<br />

DOROTHY M. FECSKE, Department <strong>of</strong> Wildlife and Fisheries Sciences, Box 2140B, South<br />

Dakota State University, Brookings, SD 57007, USA, email: gdf@rapidnet.com<br />

JONATHAN A. JENKS, Department <strong>of</strong> Wildlife and Fisheries Sciences, Box 2140B, South<br />

Dakota State University, Brookings, SD 57007, USA, email:<br />

Jonathan_Jenks@sdstate.edu<br />

Abstract: <strong>Mountain</strong> lions (Puma concolor) are a state threatened species in South Dakota,<br />

although <strong>the</strong>re is an established breeding population in <strong>the</strong> Black Hills. The Department <strong>of</strong><br />

Game, Fish and Parks (GFP) is currently funding a multi-year research project through South<br />

Dakota State University to learn more about <strong>the</strong> status <strong>of</strong> mountain lions. The information will<br />

be used by GFP to develop a mountain lion management plan. Public opinion and understanding<br />

<strong>of</strong> mountain lions will be a critical component for developing and implementing a management<br />

plan. This public opinion survey was <strong>the</strong> first step in developing <strong>the</strong> social component (human<br />

dimensions) <strong>of</strong> <strong>the</strong> plan. The survey was conducted in <strong>the</strong> early spring <strong>of</strong> 2002. Of 1,783<br />

deliverable questionnaires mailed to South Dakota residents, 1,114 usable questionnaires were<br />

returned for a total return rate <strong>of</strong> 62.57%. A one-page survey <strong>of</strong> non-respondents also was<br />

conducted; <strong>of</strong> those, 103 (19.5%) were returned. Overall, <strong>the</strong> majority <strong>of</strong> respondents (>50%)<br />

believed that presence <strong>of</strong> mountain lions was an indication <strong>of</strong> a healthy environment, lions and<br />

hunters did not compete for deer, if people modified a few behaviors <strong>the</strong>y could coexist with<br />

lions, and lions should be able to exist wherever <strong>the</strong>y occurred in South Dakota. Survey results<br />

were used to develop an attitude model to provide a framework for understanding public opinion<br />

<strong>of</strong> mountain lions. The model was intuitive, but derived empirically using a cluster analysis<br />

procedure from respondents’ answers to 12 questions. The model represented a continuum <strong>of</strong><br />

attitudes ranging from strongly supportive <strong>of</strong> to strongly disliking mountain lions. Based on <strong>the</strong><br />

model, 22.7% <strong>of</strong> <strong>the</strong> respondents were strongly pro-lion, 33.7% slightly pro-lion, 11.3% neutral,<br />

22.5% slightly contra-lion, and 9.8% <strong>of</strong> <strong>the</strong> respondents strongly contra-lion. Cluster names<br />

were descriptive <strong>of</strong> <strong>the</strong> general attitudes held toward mountain lions in South Dakota, and<br />

responses provided to o<strong>the</strong>r questions in <strong>the</strong> survey were used to fur<strong>the</strong>r describe each clustergroup.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

171


172<br />

CRITICAL COUGAR CROSSING AND BAY AREA REGIONAL PLANNING<br />

MICHELE KORPOS, Live Oak Associates, Inc., 6830 Via Del Oro, Suite 205, San Jose, CA<br />

95119, USA, email: mkorpos@loainc.com<br />

Abstract: The goal <strong>of</strong> this presentation is to illustrate <strong>the</strong> need for habitat conservation and<br />

regional planning on behalf <strong>of</strong> cougars (Puma concolor), which play an integral role in <strong>the</strong> health<br />

<strong>of</strong> <strong>the</strong>ir ecosystems. As a keystone species, “fur<strong>the</strong>r degradation <strong>of</strong> [cougar] habitat connectivity<br />

will lead to cascading impacts down through successively lower trophic levels…” (Jigour 2000).<br />

County borders are human constructs with no ecological relevance. It can be assumed certain<br />

cougar home ranges overlap San Mateo, Santa Clara and Santa Cruz Counties (California), while<br />

o<strong>the</strong>rs may overlap Monterey, San Benito and Santa Cruz Counties. These counties contain <strong>the</strong><br />

Santa Cruz <strong>Mountain</strong>s, and <strong>the</strong> Gabilan and Diablo <strong>Mountain</strong> Ranges. The challenge is to<br />

maintain land connections between large patches <strong>of</strong> intact habitat through open communication<br />

among county agencies and through regional planning efforts.<br />

By developing land on a project-by-project basis, counties promote habitat fragmentation. Left<br />

unchecked, human development in and around <strong>the</strong> Santa Cruz <strong>Mountain</strong>s will continue to<br />

fragment cougar habitat, leading to geographic isolation, and <strong>the</strong> eventual demise <strong>of</strong> our local<br />

population. Maintaining large tracts <strong>of</strong> land and providing connections through less hospitable<br />

landscapes are critical to ensuring <strong>the</strong> future health <strong>of</strong> cougar populations and <strong>the</strong> wildlife that<br />

share <strong>the</strong>ir ecosystems.<br />

LITERATURE CITED<br />

Jigour, V. 2000. Correspondence to Rusty Areias, Director, Department <strong>of</strong> Parks and Recreation, dated 31 January 2000.<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Holly Akenson<br />

Taylor Ranch Field Station, Univ. <strong>of</strong> Idaho<br />

HC 83 Box 8070<br />

Cascade, ID 83611<br />

USA<br />

888-842-7547<br />

tayranch@direcpc.com<br />

Jim Akenson<br />

Taylor Ranch Field Station, Univ. <strong>of</strong> Idaho<br />

HC 83 Box 8070<br />

Cascade, ID 83611<br />

USA<br />

888-842-7547<br />

tayranch@direcpc.com<br />

Kevin Allred<br />

373 East 300 South<br />

Burley, ID 83318<br />

USA<br />

208-678-3681<br />

Chuck Anderson<br />

Wyoming Coop. Fish and Wildlife Research Unit<br />

Box 3166, University Station<br />

Laramie, WY 82071<br />

USA<br />

307-766-2091<br />

cander@uwyo.edu<br />

Jerry Apker<br />

Colorado Division <strong>of</strong> Wildlife<br />

0722 South Road 1 East<br />

Monte Vista, CO 81144<br />

USA<br />

719-587-6922<br />

jerry.apker@state.co.us<br />

Fred Armstrong<br />

Guadalupe <strong>Mountain</strong> National Park<br />

HC 60 Box 400<br />

Salt Flat, TX 79847<br />

USA<br />

915-828-3251<br />

Fred_Armstrong@nps.gov<br />

LIST OF PARTICIPANTS<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 173<br />

Matt Austin<br />

BC Ministry <strong>of</strong> Water, Land, and Air Protection<br />

PO Box 9374<br />

Victoria, BC V8V 1N4<br />

Canada<br />

250-387-9799<br />

Matt.Austin@gems7.gov.bc.ca<br />

Dave Avey<br />

Wyoming Game and Fish Department<br />

3030 Energy Lane, Suite 100<br />

Casper, WY 82604<br />

USA<br />

307-473-3420<br />

Dave.Avey@wgf.state.wy.us<br />

David Baron<br />

2338 18 th Street<br />

Boulder, CO 80304<br />

USA<br />

303-443-2341<br />

dbaron@aya.yale.edu<br />

Jim Bauer<br />

Univ. <strong>of</strong> California-Davis, Wildlife Health Center<br />

PO Box 1203<br />

Julian, CA 92036<br />

USA<br />

760-767-4331<br />

jwbauer@uia.net<br />

Rich Beausoleil<br />

Washington Department <strong>of</strong> Fish and Wildlife<br />

3860 Chelan Hwy<br />

Wenatchee, WA 98801<br />

USA<br />

509-679-3858<br />

beausrab@dfw.wa.gov<br />

Scott Becker<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-2688<br />

scott.becker@wgf.state.wy.us


174 LIST OF PARTICIPANTS<br />

Chris Beldon<br />

Florida Fish and Wildlife Conservation Commission<br />

4005 South Main Street<br />

Gainesville, FL 32601<br />

USA<br />

352-955-2230<br />

Chris.Belden@fwc.state.fl.us<br />

William Betty<br />

Eastern Puma Research Network<br />

49B Punch Bowl Trail<br />

West Kingston, RI 02892<br />

USA<br />

401-789-4026<br />

wjbetty@peoplepc.com<br />

Roman Biek<br />

Fish and Wildlife Biology Program<br />

University <strong>of</strong> Montana<br />

Missoula, MT 59812<br />

USA<br />

406-243-6193<br />

rbiek@selway.umt.edu<br />

Mario Biondini<br />

Department <strong>of</strong> Animal and Range Sciences<br />

North Dakota State University<br />

Fargo, ND 58105<br />

USA<br />

701-231-8208<br />

mario.biondini@ndsu.nodak.edu<br />

Dan Bjornlie<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-2688<br />

dan.bjornlie@wgf.state.wy.us<br />

H. Webb Blessley<br />

The Cougar Fund<br />

910 Baja<br />

Laguna Beach, CA 92651<br />

USA<br />

702-275-7820<br />

wblessley@aol.com<br />

Steven Bobzien<br />

East Bay Regional Park District<br />

4422 Rockwood Avenue<br />

Napa, CA 94558<br />

USA<br />

510-544-2347<br />

sbobzien@ebparks.org<br />

Roger Bredeh<strong>of</strong>t<br />

Wyoming Game and Fish Department<br />

1017 Reynolds<br />

Laramie, WY 82070<br />

USA<br />

307-745-4402<br />

rbredeho@wyoming.com<br />

Jim Briggs<br />

Cle Elum-Roslyn School District - Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Doug Brimeyer<br />

Wyoming Game and Fish Department<br />

PO Box 67<br />

Jackson, WY 83001<br />

USA<br />

307-733-2321<br />

Doug.Brimeyer@wgf.state.wy.us<br />

Lara Brongo<br />

Auburn University<br />

1925 Trexler Court<br />

Raleigh, NC 27606<br />

USA<br />

334-332-1124<br />

llbrongo@yahoo.com<br />

Howard Buffet<br />

Howard G. Buffet Foundation<br />

PO Box 4508<br />

Decatur, IL 62521<br />

USA<br />

217-429-3988<br />

hgbfoundation@yahoo.com<br />

Polly Buotte<br />

Wildlife Conservation Society<br />

Box 299<br />

Gardiner, MT 59030<br />

USA<br />

406-848-7683<br />

polly_thornton@hotmail.com<br />

Clint Cabanero<br />

South Coast Wildlands Project<br />

PO Box 2493<br />

Monrovia, CA 91016<br />

USA<br />

626-599-9585<br />

clint@scwildlands.org<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Steve Cain<br />

Grand Teton National Park<br />

PO Box 170<br />

Moose, WY 83012<br />

USA<br />

307-739-3485<br />

Steve_Cain@nps.gov<br />

Franz Camenzid<br />

Jackson Hole Conservation Alliance<br />

PO Box 2835<br />

Jackson, WY 83001<br />

USA<br />

307-733-9417<br />

Franz@jhalliance.org<br />

David Choate<br />

University <strong>of</strong> Notre Dame<br />

107 Galvin Life Science Center<br />

Notre Dame, IN 46556<br />

USA<br />

574-631-0949<br />

dchoate@nd.edu<br />

Stacy Courville<br />

CSKT – Wildlife Management Program<br />

PO Box 278<br />

Pablo, MT 59855<br />

USA<br />

406-883-2888<br />

stacyac@cskt.org<br />

Hilary Cruickshank<br />

Washington State University<br />

PO Box 651<br />

Rossland, BC V0G 1Y0<br />

Canada<br />

250-362-3310<br />

hcruicks@mail.wsu.edu<br />

Michelle Cullens<br />

<strong>Mountain</strong> <strong>Lion</strong> Foundation<br />

PO Box 1896<br />

Sacramento, CA 95812<br />

USA<br />

916-442-2666<br />

cullens@mountainlion.org<br />

Chris Daubin<br />

Wyoming Game and Fish Department<br />

1205 Mary Anne<br />

Riverton, WY 82501<br />

USA<br />

307-856-4982<br />

cdaubin@tcinc.net<br />

Jeff Davis<br />

Wildlife Services<br />

PO Box 131<br />

Olancha, CA 93549<br />

USA<br />

760-937-6788<br />

Deanna Dawn<br />

University <strong>of</strong> California – Davis<br />

20231 Blauer Drive<br />

Saratoga, CA 95070<br />

USA<br />

408-741-5156<br />

Rich DeSimone<br />

Montana Fish, Wildlife, and Parks<br />

1420 East 6 th Avenue<br />

Helena, MT 59620<br />

USA<br />

406-444-0358<br />

rdesimone@state.mt.us<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 175<br />

Leonard Two Eagle<br />

Rosebud Sioux Tribe Game, Fish, and Parks<br />

PO Box 291<br />

Parmelee, SD 57566<br />

USA<br />

605-747-2289<br />

ln2eagle@hotmail.com<br />

Dorothy Fecske<br />

South Dakota State University<br />

Box 2140B<br />

Brookings, SD 57007<br />

USA<br />

605-688-6121<br />

gdf@rapidnet.com<br />

Lee Fitzhugh<br />

University <strong>of</strong> California – Davis<br />

One Shields Drive<br />

Davis, CA 95616<br />

USA<br />

530-752-1496<br />

elfitzhugh@ucdavis.com<br />

Flora Fitzhugh<br />

913 Purdue Drive<br />

Woodland, CA 95695<br />

USA<br />

530-668-1138


176 LIST OF PARTICIPANTS<br />

Steve Fitzwater<br />

National Trappers Association<br />

PO Box 106<br />

Dubois, ID 83423<br />

USA<br />

208-745-6664<br />

dubois5@yahoo.com<br />

Gary Fralick<br />

Wyoming Game and Fish Department<br />

PO Box 1022<br />

Thayne, WY 83127<br />

USA<br />

307-883-2998<br />

gfralick@silverstar.com<br />

Steve Galentine<br />

U.S. Department <strong>of</strong> Agriculture<br />

29469 East Vine Avenue<br />

Escalon, CA 95320<br />

USA<br />

209-605-4934<br />

sgalentine@mindspring.com<br />

Emily Garding<br />

Grand Canyon National Park<br />

Box 129<br />

Grand Canyon, AZ 86023<br />

USA<br />

928-638-7648<br />

Emily_Garding@nps.gov<br />

Jacquie Gerads<br />

North Dakota Game and Fish Department<br />

100 North Bismarck Expressway<br />

Bismarck, ND 58501<br />

USA<br />

701-328-6613<br />

jgerads@state.nd.us<br />

Diana Ghikas<br />

University <strong>of</strong> Calgary<br />

2500 University Drive NW<br />

Calgary, AB T3E 1A2<br />

Canada<br />

403-299-2797<br />

dghikas@ucalgary.ca<br />

Glenn Gibbons<br />

Oglala Sioux Parks and Recreation Authority<br />

PO Box 570<br />

Kyle, SD 57752<br />

USA<br />

605-455-2584<br />

aranger@gwtc.net<br />

Larry Gilbertson<br />

Nevada Division <strong>of</strong> Wildlife<br />

#60 Youth Center Road<br />

Elko, NV 89801<br />

USA<br />

775-777-2302<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

Carlos Lopez Gonzalez<br />

Escuela de Biologia-Facultad de Ciencias Naturales<br />

Universidad Autonoma de Queretaro<br />

Apdo. Postal 184<br />

Queretaro, Queretaro 76010<br />

Mexico<br />

442-215-4777<br />

Cats4mex@aol.com<br />

Chelsea Gordon<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Kevin Grady<br />

Boulder County Open Space<br />

1933 Geer Canyon Drive<br />

Boulder, CO 80302<br />

USA<br />

720-406-9178<br />

Kgrady@co.boulder.co.us<br />

Steve Griffin<br />

South Dakota Department <strong>of</strong> Game, Fish, and Parks<br />

3305 West South Street<br />

Rapid City, SD 57702<br />

USA<br />

605-394-2391<br />

steve.griffin@state.sd.us<br />

Torey Griswold<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Trish Griswold<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA


Carolyn Grygiel<br />

Natural Resources Management<br />

North Dakota State University<br />

Fargo, ND 58105<br />

USA<br />

701-231-8180<br />

carolyn.grygiel@ndsu.nodak.edu<br />

Kerry Gyekis<br />

Writer/Private Consultant Forester<br />

RR 1, Box 213<br />

Morris, PA 16938<br />

USA<br />

570-723-8251<br />

gyekis@epix.net<br />

Keto Gyekis<br />

107 Woods Trail<br />

Delton, MI 49046<br />

269-623-3240<br />

keto@uplink.net<br />

Jon Hanna<br />

Arizona Game and Fish Department<br />

7200 E. University<br />

Mesa, AZ 85207<br />

USA<br />

480-981-9400<br />

Kevin Hansen<br />

Yosemite National Park<br />

PO Box 434<br />

Yosemite, CA 95389<br />

USA<br />

209-372-8870<br />

Jan V. Hart<br />

USGS – Southwest Biological Research Center<br />

PO Box 5614, NAU<br />

Flagstaff, AZ 86011<br />

USA<br />

928-556-7466<br />

Jan.Hart@nau.edu<br />

John Hart<br />

Hartwood Natural Resource Consultants<br />

1390 Curt Gowdy Drive<br />

Cheyenne, WY 82009<br />

USA<br />

307-778-3993<br />

huntwyoming@aol.com<br />

Dee Dee Hawk<br />

Wyoming Game and Fish Lab<br />

PO Box 3312, University Station<br />

Laramie, WY 82071<br />

USA<br />

307-766-6313<br />

Dhawk@uwyo.edu<br />

Lisa Haynes<br />

University <strong>of</strong> Arizona<br />

133 West 2 nd Street, Rear<br />

Tucson, AZ 85705<br />

USA<br />

520-320-1841<br />

lynxrufus@earthlink.net<br />

Lucina Hernández<br />

Instituto de Ecologia, A.C.<br />

Km 5 a Carr. de Mazatlán s/n<br />

Durango, Dgo C.P 34100<br />

México<br />

+52-618-812-1483<br />

lucina@fauna.edu.mx<br />

Greg Hiatt<br />

Wyoming Game and Fish Department<br />

PO Box 186<br />

Sinclair, WY 82334<br />

USA<br />

307-324-7927<br />

gshiatt@trib.com<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 177<br />

Ryan Hill<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Mary Hindelang<br />

Michigan Tech University<br />

40032 N. Lower Worham Road<br />

Chassell, MI 49916<br />

USA<br />

906-523-4014<br />

mlhindel@mtu.edu<br />

Dave Hoerath<br />

Boulder County Parks and Open Space<br />

5305 Spine Road, Unit B<br />

Boulder, CO 80306<br />

USA<br />

303-516-9364<br />

dhoerath@co.boulder.co.us


178 LIST OF PARTICIPANTS<br />

Mollie Hogan<br />

The Nature <strong>of</strong> Wildlworks<br />

PO Box 109<br />

Topanga, CA 90290<br />

USA<br />

310-455-0550<br />

wildworks1@aol.com<br />

Greg Holm<br />

Redwood National Park<br />

PO Box 7<br />

Orick, CA 95555<br />

USA<br />

707-464-6101<br />

gregory_holm@nps.gov<br />

Bernie Holz<br />

Wyoming Game and Fish Department<br />

PO Box 850<br />

Pinedale, WY 82941<br />

USA<br />

307-367-4353<br />

Bernie.Holz@wgf.state.wy.us<br />

Mike Hooker<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-2688<br />

michael.hooker@wgf.state.wy.us<br />

Rick Hopkins<br />

Live Oak Associates, Inc.<br />

6830 Via del Oro, Suite 205<br />

San Jose, CA 95119<br />

USA<br />

408-281-5885<br />

rhopkins@loainc.net<br />

Betsy Howell<br />

525 Benton<br />

Port Townsend, WA 98368<br />

USA<br />

360-379-0582<br />

betsyhowell@olympus.net<br />

Neil Hymas<br />

Wyoming Game and Fish Department<br />

Box 368<br />

Cokeville, WY 83114<br />

USA<br />

307-279-3466<br />

nhymas@allwest.net<br />

Rose Jaffe<br />

Montana Fish, Wildlife, and Parks<br />

PO Box 200701<br />

Helena, MT 59620-0701<br />

USA<br />

406-444-1276<br />

rjaffe@state.mt.us<br />

Lynn Jahnke<br />

Wyoming Game and Fish Department<br />

PO Box 6249<br />

Sheridan, WY 82801<br />

USA<br />

307-672-7418<br />

Lynn.Jahnke@wgf.state.wy.us<br />

Martin Jalkotzy<br />

Arc Wildlife Services Ltd.<br />

3527 – 35 Ave SW<br />

Calgary, AB T3E 1A2<br />

Canada<br />

403-240-3361<br />

martin_jalkotzy@nucleus.com<br />

Jesse Millen-Johnson<br />

Bates College<br />

Box 486<br />

Lewiston, ME 04240<br />

USA<br />

207-674-2927<br />

jmillenj@bates.edu<br />

Don Jones<br />

Zion National Park<br />

State Route 9<br />

Springdale, UT 84767<br />

USA<br />

435-772-0212<br />

don_jones@nps.gov<br />

Tom Keegan<br />

Idaho Department <strong>of</strong> Fish and Game<br />

PO Box 1336<br />

Salmon, ID 83467<br />

USA<br />

208-756-2271<br />

tkeegan@idfg.state.id.us<br />

Sean Kelly<br />

Utah Division <strong>of</strong> Wildlife Resources<br />

PO Box 606<br />

Cedar City, UT 84720<br />

USA<br />

435-691-5701<br />

seankelly@utah.gov<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Marc Kenyon<br />

University <strong>of</strong> California – Davis<br />

7548 Event Way<br />

Sacramento, CA 95842<br />

USA<br />

916-320-6654<br />

mwkenyon@ucdavis.edu<br />

Hea<strong>the</strong>r Keough<br />

Utah State University<br />

6991 S. Hwy 165<br />

Hyrum, UT 84319<br />

USA<br />

435-881-2856<br />

hea<strong>the</strong>rkeough@hotmail.com<br />

Brian Kertson<br />

Washington Coop. Fish and Wildlife Research Unit<br />

23427 NE 28 th PL<br />

Sammamish, WA 98074<br />

USA<br />

425-941-0278<br />

gulogulo@hotmail.com<br />

Sharon Kim<br />

Zion National Park<br />

State Route 9<br />

Springdale, UT 84767<br />

USA<br />

435-772-0212<br />

sharon_kim@nps.gov<br />

Mike Kintigh<br />

South Dakota Game, Fish, and Parks<br />

3305 W. South Street<br />

Rapid City, SD 57702<br />

USA<br />

605-394-2391<br />

Mike.Kintigh@state.sd.us<br />

Gary Koehler<br />

Washington Department <strong>of</strong> Fish and Wildlife<br />

PO Box 102<br />

Cle Elum, WA 98922<br />

USA<br />

509-260-0477<br />

gmkmmk@cleelum.com<br />

Andrea Kortello<br />

University <strong>of</strong> Idaho<br />

PO Box 4297<br />

Banff, AB T1L 1E7<br />

Canada<br />

403-762-5339<br />

kortello@yahoo.com<br />

Caroline Krumm<br />

CSU – Colorado Division <strong>of</strong> Wildlife<br />

317 W. Prospect<br />

Fort Collins, CO 80524<br />

USA<br />

970-215-3759<br />

ckrumm@yahoo.com<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 179<br />

Pamela Kyselka<br />

Navajo Nation – Department <strong>of</strong> Fish and Wildlife<br />

PO Box 1480<br />

Window Rock, AZ 86515<br />

USA<br />

928-871-6451<br />

p_kyselka@hotmail.com<br />

Carl Lackey<br />

Nevada Division <strong>of</strong> Wildlife<br />

1060 Mallory Wy.<br />

Carson City, NV 89701<br />

USA<br />

775-720-6130<br />

cdembears@aol.com<br />

Ca<strong>the</strong>rine Lambert<br />

Washington State University<br />

1712 NW Lamont St.<br />

Pullman, WA 99163<br />

USA<br />

509-335-4084<br />

lambertcath@wsu.edu<br />

Melanie Lambert<br />

The Summerlee Foundation<br />

716 W. Tejon, Suite 9<br />

Colorado Springs, CO 80903<br />

USA<br />

800-256-7515<br />

mal3@summerlee.org<br />

Darrell Land<br />

Florida Fish and Wildlife Conservation Commission<br />

566 Commercial Blvd.<br />

Naples, FL 34104-4709<br />

USA<br />

239-643-4220<br />

Darrell.Land@fwc.state.fl.us<br />

John Laundré<br />

Instituto de Ecologia, A.C.<br />

Km 5 a Carr. de Mazatlán s/n<br />

Durango, Dgo C.P 34100<br />

México<br />

+52-618-812-1483<br />

launjohn@prodigy.net.mx


180 LIST OF PARTICIPANTS<br />

Cheryl Le Drew<br />

Lotek Wireless Inc.<br />

115 Pony Drive<br />

Newmarket, ON L3Y 7B5<br />

Canada<br />

905-836-6680<br />

cledrew@lotek.com<br />

Fred Lindzey<br />

Wyoming Coop. Fish and Wildlife Research Unit<br />

Box 3166, University Station<br />

Laramie, WY 82071<br />

USA<br />

307-766-5415<br />

flindzey@uwyo.edu<br />

Kenneth Logan<br />

Colorado Division <strong>of</strong> Wildlife<br />

2300 South Townsend Avenue<br />

Montrose, CO 81401<br />

USA<br />

970-252-6013<br />

Ken.Logan@state.co.us<br />

Mark Lotz<br />

Florida Fish and Wildlife Conservation Commission<br />

566 Commercial Blvd.<br />

Naples, FL 34104-4709<br />

USA<br />

239-735-0773<br />

Mark.Lotz@fwc.state.fl.us<br />

Cara Blessley Lowe<br />

The Cougar Fund<br />

Box 122<br />

Jackson, WY 83001<br />

USA<br />

310-562-4021<br />

cara@mangelson.com<br />

Marissa Luchau<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Lisa Lyren<br />

US Geological Survey<br />

1147 E. 6 th Street<br />

Corona, CA 92879<br />

USA<br />

909-735-0773<br />

llyren@usgs.gov<br />

Tom Mangelson<br />

The Cougar Fund<br />

Box 122<br />

Jackson, WY 83001<br />

USA<br />

307-733-6179<br />

tom@mangelson.com<br />

Marcie Maras<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Beth Marker<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

markerb@cleelum.wednet.edu<br />

Gary Matson<br />

Matson’s Lab<br />

PO Box 308<br />

Milltown, MT 59851<br />

USA<br />

406-258-6286<br />

gjmatson@montana.com<br />

David Mattson<br />

USGS – Southwest Biological Research Center<br />

PO Box 5614, NAU<br />

Flagstaff, AZ 86011<br />

USA<br />

928-556-7466<br />

David.Mattson@nau.edu<br />

Roy McBride<br />

Livestock Protection Co.<br />

Box 178<br />

Ochopee, FL 34141<br />

USA<br />

239-695-2287<br />

Helen McGinnis<br />

Eastern Cougar Foundation<br />

PO Box 300<br />

Harman, WV 26270<br />

USA<br />

304-227-4166<br />

helenmcginnis@meer.net<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Craig McLaughlin<br />

Utah Division <strong>of</strong> Wildlife Resources<br />

1594 W. North Temple<br />

Salt Lake City, UT 84114<br />

USA<br />

801-538-4758<br />

craigmclaughlin@utah.gov<br />

Daryl Meints<br />

Idaho Department <strong>of</strong> Fish and Game<br />

1515 Lincoln Road<br />

Idaho Falls, ID 83401<br />

USA<br />

208-525-7290<br />

dmeints@idfg.state.id.us<br />

Stephanie Middlebrooks<br />

Rosebud Sioux Tribe<br />

Box 300<br />

Rosebud, SD 57570<br />

USA<br />

605-747-2289<br />

middlebrooks@post.com<br />

Michael Middleton<br />

Muckleshoot Indian Tribe<br />

39015 172 nd Ave SE<br />

Auburn, WA 98092<br />

USA<br />

360-802-2202<br />

mikem@mitwildlifeculture.org<br />

Jessica Montag<br />

University <strong>of</strong> Montana<br />

2224 West Sussex<br />

Missoula, MT 59812<br />

USA<br />

406-243-6611<br />

jmontag@selway.umt.edu<br />

Dave Moody<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-2688<br />

dave.moody@wgf.state.wy.us<br />

Shane Moore<br />

Box 2980<br />

Jackson, WY 83001<br />

USA<br />

307-733-8862<br />

moorefilms@wyoming.com<br />

Don Morgan<br />

Sou<strong>the</strong>rn Hills Animal Clinic<br />

Box 67<br />

Pringle, SD 57773<br />

USA<br />

605-673-3503<br />

Susan Morse<br />

Keeping Track Inc.<br />

Wolfrun 55A Bentley Lane<br />

Jericho, VT 05465<br />

USA<br />

Kerry Murphy<br />

Yellowstone National Park<br />

PO Box 168<br />

Yellowstone National Park, WY 82190<br />

USA<br />

307-344-2393<br />

kerry_murphy@nps.gov<br />

Steve Nadeau<br />

Idaho Department <strong>of</strong> Fish and Game<br />

600 South Walnut St., Box 25<br />

Boise, ID 83707<br />

USA<br />

208-334-2920<br />

snadeau@idfg.state.id.us<br />

Sharon Negri<br />

Wild Futures<br />

353 Wallace Way<br />

Bainbridge Island, WA 98110<br />

USA<br />

206-780-9718<br />

snegri@igc.org<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 181<br />

Evelyn Nelson<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Mark Nelson<br />

Wyoming Game and Fish Department<br />

500 Western Hills Blvd.<br />

Cheyenne, WY 82009<br />

USA<br />

307-638-8354<br />

mnelson@wyoming.com


182 LIST OF PARTICIPANTS<br />

Ryan Nelson<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Jesse Newby<br />

Wildlife Conservation Society<br />

Box 299<br />

Gardiner, MT 59030<br />

USA<br />

406-848-7683<br />

newby_jr@yahoo.com<br />

M. Cathy Nowack<br />

Cat Tracks Wildlife Consulting<br />

PO Box 195<br />

Union, OR 97883<br />

USA<br />

541-562-1057<br />

mcnowak@eoni.com<br />

Jim Olson<br />

Wyoming Game and Fish Department<br />

1737 Hillcrest<br />

Evanston, WY 82930<br />

USA<br />

307-789-3285<br />

jwolson@wyoming.com<br />

Anne Orlando<br />

UC Davis – Agronomy and Range Science<br />

1 Shields Avenue<br />

Davis, CA 95616<br />

USA<br />

530-758-1204<br />

amorlando@ucdavis.edu<br />

Krishna Pacifici<br />

North Carolina State University<br />

1925 Trexler Court<br />

Raleigh, NC 27606<br />

USA<br />

919-233-1477<br />

kpacifici@yahoo.com<br />

Doug Padley<br />

Santa Clara Valley Water District<br />

PO Box 41306<br />

San Jose, CA 95160<br />

USA<br />

408-265-2607<br />

dougpadley@att.net<br />

Christopher Papouchis<br />

<strong>Mountain</strong> <strong>Lion</strong> Foundation<br />

PO Box 1896<br />

Sacramento, CA 95812<br />

USA<br />

916-442-2666<br />

cpapouchis@mountainlion.org<br />

Steve Pavlik<br />

4149 E. Waverly Street<br />

Tucson, AZ 85712<br />

USA<br />

520-327-0708<br />

Spavlik@aol.com<br />

James Pehringer<br />

USDA – Wildlife Services<br />

1302 Johnson<br />

Thermopolis, WY 82443<br />

USA<br />

307-272-3638<br />

Kristeen Penrod<br />

South Coast Wildlands Project<br />

PO Box 2493<br />

Monrovia, CA 91016<br />

USA<br />

626-599-9585<br />

Kristeen@scwildlands.org<br />

Howard Quigley<br />

Beringia South<br />

3610 Broadwater Street, Suite #111<br />

Bozeman, MT 59715<br />

USA<br />

406-556-2199<br />

hquigley@att.net<br />

Dick Ray<br />

Rocky <strong>Mountain</strong> Wildlife Park<br />

4821 A Hwy 84<br />

Pagosa Springs, CO 81147<br />

USA<br />

970-264-5546<br />

Tyler Riblett<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Wendy Reith<br />

Utah State University<br />

1143 Capitol Street<br />

Ogden, UT 84401<br />

USA<br />

435-797-4482<br />

wendyr@cc.usu.edu<br />

Seth Riley<br />

Santa Monica <strong>Mountain</strong>s National Recreation Area<br />

401 W. Hillcrest Drive<br />

Thousand Oaks, CA 91360<br />

USA<br />

805-370-2358<br />

seth_riley@nps.gov<br />

Wendy Keefover-Ring<br />

Sinapu<br />

4990 Pearl East Circle, Ste. 301<br />

Boulder, CO 80301<br />

USA<br />

303-447-8655<br />

wendy@sinapu.org<br />

Krissy Robertson<br />

The Cougar Fund<br />

Box 122<br />

Jackson, WY 83001<br />

USA<br />

307-733-0797<br />

krissy@images<strong>of</strong>naturestock.com<br />

Hugh Robinson<br />

Washington State Univ. – Dept. <strong>of</strong> Natural Resources<br />

PO Box 646410<br />

Pullman, WA 99164-6410<br />

USA<br />

250-362-2271<br />

hsrobins@wsunix.wsu.edu<br />

Kirk Robinson<br />

Western Wildlife Conservancy<br />

68 S. Main Street, 4 th Floor<br />

Salt Lake City, UT 84101<br />

USA<br />

801-575-7107<br />

predator@xmission.com<br />

Thiele Robinson<br />

PO Box 665<br />

Wilson, WY 83014<br />

USA<br />

307-734-1356<br />

trobinson@aol.com<br />

Toni Ruth<br />

Wildlife Conservation Society<br />

Box 299<br />

Gardiner, MT 59030<br />

USA<br />

406-848-7683<br />

truth@montanadsl.net<br />

Corey Rutledge<br />

Cougar Fund, Inc.<br />

PO Box 4068<br />

Cheyenne, WY 82003<br />

USA<br />

307-632-0554<br />

crutledge@lathropandrutledge.com<br />

Tom Ryder<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-7723<br />

Tom.Ryder@wgf.state.wy.us<br />

Tony Salandro<br />

5302 N. La Canada Drive<br />

Tucson, AZ 85704<br />

USA<br />

520-690-1794<br />

Tony_Salandro@hotmail.com<br />

Mike Sawaya<br />

Wildlife Conservation Society<br />

Box 299<br />

Gardiner, MT 59030<br />

USA<br />

406-848-7683<br />

mikesawaya@hotmail.com<br />

Ralph Schmidt<br />

ARC<br />

3108A – 14 St. NW<br />

Calgary, AB T3E 1A2<br />

Canada<br />

403-289-1164<br />

ElkRalph@aol.com<br />

Sharon Seneczko, DVM<br />

Sou<strong>the</strong>rn Hills Animal Clinic<br />

RR 1, Box 93F<br />

Custer, SD 57730<br />

USA<br />

605-673-4996<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 183


184 LIST OF PARTICIPANTS<br />

Harley Shaw<br />

The Juniper Institute<br />

PO Box 486<br />

Hillsboro, NM 88042<br />

USA<br />

505-895-5385<br />

hgshaw@zianet.com<br />

Benj Sinclair<br />

Teton Science School<br />

PO Box 7580<br />

Jackson, WY 83002<br />

USA<br />

307-733-2623<br />

bsinclair@wildlifeexpeditions.com<br />

Nick Smallwood<br />

Oglala Sioux Parks and Recreation Authority<br />

PO Box 570<br />

Kyle, SD 57752<br />

USA<br />

605-455-2584<br />

arranger@gwtc.net<br />

Alaric Smith<br />

Geography Dept., University <strong>of</strong> Wales Swansea<br />

Singleton Park, Swansea SA2 8PP<br />

United Kingdom<br />

+44(0)7770 542086<br />

exoticcatproject@swan.ac.uk<br />

Nick Smith<br />

Box 101<br />

Quemado, NM 87829<br />

USA<br />

505-773-4845<br />

Russell Sparkman<br />

Fusionpark Media Inc.<br />

PO Box 160<br />

Langley, WA 98260<br />

USA<br />

360-341-2020<br />

russell@owj.com<br />

Rocky Spencer<br />

Washington Department <strong>of</strong> Fish and Wildlife<br />

24916 255 th Pl. SE<br />

Ravendale, WA 98051<br />

USA<br />

206-799-3134<br />

John Steuber<br />

USDA – Wildlife Services<br />

2800 N. Lincoln Blvd.<br />

Oklahoma City, OK 73105<br />

USA<br />

405-521-4039<br />

john.e.steuber@usda.gov<br />

David Stoner<br />

Utah State University<br />

5230 Old Main Hill<br />

Logan, UT 84322-5230<br />

USA<br />

435-797-7125<br />

dstoner@cc.usu.edu<br />

Linda Sweanor<br />

Univ. <strong>of</strong> California-Davis, Wildlife Health Center<br />

1980 Stan Drive<br />

Montrose, CO 81401<br />

USA<br />

970-252-1928<br />

lsweanor@mindspring.com<br />

Scott Talbott<br />

Wyoming Game and Fish Department<br />

3030 Energy Lane, Suite 100<br />

Casper, WY 82604<br />

USA<br />

307-473-3400<br />

Scott.Talbott@wgf.state.wy.us<br />

Pete Taylor<br />

City <strong>of</strong> Boulder Open Space and Mtn. Parks Dept.<br />

PO Box 791<br />

Boulder, CO 80306<br />

USA<br />

303-413-7621<br />

Daniel Thompson<br />

South Dakota State University<br />

Box 2140B<br />

Brookings, SD 57007<br />

USA<br />

605-688-6121<br />

djthompson4@hotmail.com<br />

Ron Thompson<br />

Arizona Game and Fish Department<br />

Box 1588<br />

Pinetop, AZ 85935<br />

USA<br />

928-367-4342<br />

rthompson@gf.state.az.us<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP


Jay Tischendorf, DVM<br />

American Ecological Research Institute<br />

PO Box 1826<br />

Great Falls, MT 59403<br />

USA<br />

406-453-7233<br />

Tischendorfj@hotmail.com<br />

Bob Trebelcock<br />

Wyoming Game and Fish Department<br />

260 Buena Vista<br />

Lander, WY 82520<br />

USA<br />

307-332-2704<br />

bobt@wyoming.com<br />

Hank Uhden<br />

Wyoming Department <strong>of</strong> Agriculture<br />

2219 Carey<br />

Cheyenne, WY 82002-0100<br />

USA<br />

307-777-6574<br />

huhden@state.wy.us<br />

David Vales<br />

Muckleshoot Indian Tribe<br />

39015 172 nd Avenue SE<br />

Auburn, WA 98092<br />

USA<br />

360-802-2202<br />

davidv@mitwildlifeculture.org<br />

Brady Vandeburg<br />

Wyoming Game and Fish Department<br />

PO Box 286<br />

Kaycee, WY 82639<br />

USA<br />

307-738-2455<br />

bvande@kaycee.smalltown.net<br />

Winston Vickers<br />

Univ. <strong>of</strong> CA – Davis, Institute <strong>of</strong> Wildlife Studies<br />

125 Via Waziers<br />

Newport Beach, CA 92663<br />

USA<br />

949-929-8643<br />

Vickers@IWS.org<br />

Brian Wakeling<br />

Arizona Game and Fish Department<br />

2221 W. Greenway Road<br />

Phoenix, AZ 85023<br />

USA<br />

602-789-3385<br />

bwakeling@gf.state.az.us<br />

Bill Wall<br />

SCI Foundation<br />

501 2 nd Street NE<br />

Washington, DC 20002<br />

USA<br />

202-543-8733<br />

bwall@sci-dc.org<br />

Craig White<br />

Idaho Department <strong>of</strong> Fish and Game<br />

1540 Warner Avenue<br />

Lewiston, ID 83501<br />

USA<br />

208-799-5010<br />

cwhite@idfg.state.id.us<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP<br />

LIST OF PARTICIPANTS 185<br />

Kevin White<br />

Cle Elum-Roslyn School District – Project CAT<br />

2694 SR 903<br />

Cle Elum, WA 98922<br />

USA<br />

Rob Wielgus<br />

Washington State Univ. – Dept. <strong>of</strong> Natural Resources<br />

PO Box 646410<br />

Pullman, WA 99169-6410<br />

USA<br />

509-335-2796<br />

wielgus@wsu.edu<br />

Rick Winslow<br />

New Mexico Department <strong>of</strong> Fish and Game<br />

One Wildlife Way<br />

Santa Fe, NM 87507<br />

USA<br />

505-476-8046<br />

rwinslow@state.nm.us<br />

Greg Winston<br />

Box 505<br />

Wilson, WY 83014<br />

USA<br />

307-690-8161<br />

gwinston@spuynet.com<br />

Lisa Wolfe<br />

Colorado Division <strong>of</strong> Wildlife<br />

317 W. Prospect<br />

Fort Collins, CO 80526<br />

USA<br />

970-472-4312<br />

lisa.wolfe@state.co.us


186 LIST OF PARTICIPANTS<br />

Patricia Woodruff<br />

The Juniper Institute<br />

PO Box 486<br />

Hillsboro, NM 88042<br />

USA<br />

505-895-5385<br />

patita@zianet.com<br />

Russell Woolstenhulme<br />

Nevada Division <strong>of</strong> Wildlife<br />

1100 Valley Road<br />

Reno, NV 89512<br />

USA<br />

775-688-1992<br />

rwoolstenhulme@ndow.org<br />

Anthony Wright<br />

Utah Division <strong>of</strong> Wildlife Resources<br />

PO Box 495<br />

Price, UT 84501<br />

USA<br />

435-650-4016<br />

alwright61@hotmail.com<br />

Duggin Wroe<br />

Colorado Division <strong>of</strong> Wildlife<br />

29 <strong>Mountain</strong> Meadow Road<br />

Laramie, WY 82070<br />

USA<br />

307-760-8111<br />

dugginsw@yahoo.com<br />

Renan Yanish<br />

USGS<br />

PO Box 923<br />

East Helena, MT 59635-0923<br />

USA<br />

406-227-5140<br />

renan_yanish@yahoo.com<br />

Eric York<br />

Santa Monica <strong>Mountain</strong>s National Recreation Area<br />

401 W. Hillcrest Drive<br />

Thousand Oaks, CA 91360<br />

USA<br />

805-370-2363<br />

eric_york@nps.gov<br />

PROCEEDINGS OF THE SEVENTH MOUNTAIN LION WORKSHOP

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!