escaping the extinction vortex: identifying factors ... · johnson, heather e., ph.d., spring 2010...

243
University of Montana ScholarWorks at University of Montana eses, Dissertations, Professional Papers Graduate School 2010 Escaping the extinction vortex: identifying factors affecting population performance and recovery in endangered Sierra Nevada bighorn sheep Heather Elizabeth Johnson e University of Montana Follow this and additional works at: hp://scholarworks.umt.edu/etd is Dissertation is brought to you for free and open access by the Graduate School at ScholarWorks at University of Montana. It has been accepted for inclusion in eses, Dissertations, Professional Papers by an authorized administrator of ScholarWorks at University of Montana. For more information, please contact [email protected]. Recommended Citation Johnson, Heather Elizabeth, "Escaping the extinction vortex: identifying factors affecting population performance and recovery in endangered Sierra Nevada bighorn sheep" (2010). eses, Dissertations, Professional Papers. Paper 379.

Upload: others

Post on 29-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

University of MontanaScholarWorks at University of Montana

Theses, Dissertations, Professional Papers Graduate School

2010

Escaping the extinction vortex: identifying factorsaffecting population performance and recovery inendangered Sierra Nevada bighorn sheepHeather Elizabeth JohnsonThe University of Montana

Follow this and additional works at: http://scholarworks.umt.edu/etd

This Dissertation is brought to you for free and open access by the Graduate School at ScholarWorks at University of Montana. It has been accepted forinclusion in Theses, Dissertations, Professional Papers by an authorized administrator of ScholarWorks at University of Montana. For moreinformation, please contact [email protected].

Recommended CitationJohnson, Heather Elizabeth, "Escaping the extinction vortex: identifying factors affecting population performance and recovery inendangered Sierra Nevada bighorn sheep" (2010). Theses, Dissertations, Professional Papers. Paper 379.

Page 2: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

ESCAPING THE EXTINCTION VORTEX:

IDENTIFYING FACTORS AFFECTING POPULATION PERFORMANCE AND

RECOVERY IN ENDANGERED SIERRA NEVADA BIGHORN SHEEP

By

HEATHER ELIZABETH JOHNSON

B.S. University of California San Diego, La Jolla, California, 1999

M.S. University of Arizona, Tucson, Arizona, 2003

Dissertation

presented in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in Wildlife and Fisheries Biology

The University of Montana, Missoula, MT

May 2010

Approved by:

Perry Brown, Associate Provost for Graduate Education

Graduate School

L. Scott Mills, Chair

Wildlife Biology Program

Thomas R. Stephenson

California Department of Fish and Game

Mark Hebblewhite

Wildlife Biology Program

Daniel H. Pletscher

Wildlife Biology Program

Michael S. Mitchell

Wildlife Biology Program

Gordon Luikart

Division of Biological Sciences

Page 3: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

© COPYRIGHT

by

Heather Elizabeth Johnson

2010

All Rights Reserved

Page 4: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

iii

Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology

Escaping the extinction vortex: identifying factors affecting population performance and

recovery in endangered Sierra Nevada bighorn sheep

Chair: L. Scott Mills

An extinction vortex is one of the greatest threats to endangered species; when

demographic, environmental, and genetic stochasticity interact with each other and with

deterministic factors, such as habitat quality, to reinforce the demise of a small

population. To successfully escape an extinction vortex and enable species recovery, all

processes that affect endangered populations should be comprehensively assessed and

incorporated into conservation plans. For my dissertation, I worked in conjunction with

California Department of Fish and Game to develop a comprehensive research program

to guide recovery efforts for federally endangered Sierra Nevada bighorn sheep, the rarest

subspecies of mountain sheep in North America. I initiated a combination of

demographic, habitat and genetic analyses to identify the stochastic and deterministic

factors limiting the recovery of this subspecies, examine the relative and synergistic

impacts of these factors on the performance of Sierra Nevada bighorn sheep, and the

benefits of different management activities for stimulating recovery efforts. Just as the

extinction vortex predicts, I found that small populations of Sierra Nevada bighorn sheep

were driven by a number of stochastic and deterministic processes. Demographic, habitat,

climate, predation, and genetic factors operated singly and in concert to shape the overall

viability of this subspecies. The interaction of factors led to atypical demographic

patterns that deviated from theoretical expectations and increased extinction risk. To

alleviate extinction processes, I found that management strategies must be tailored to

population-specific dynamics, targeting those vital rates and ecological drivers which

have the greatest power to increase performance. Results from this study have elucidated

critical aspects of Sierra Nevada bighorn sheep ecology, provided a recovery strategy for

this subspecies, and supplied new quantitative tools for examining the dynamics of small

and endangered populations. Ultimately, this work offers an example of assessing

population viability, not in terms of probability of extinction, but in terms of quantifying

conservation measures that will alleviate extinction dynamics and achieve endangered

species recovery goals.

Page 5: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

iv

ACKNOWLEDGEMENTS

This work is a testament to the power of collaboration as it is the culmination of

the efforts of my graduate committee, the Sierra Nevada Bighorn Sheep Recovery

Program, the UM community, various funding sources, and my friends and family.

I am particularly grateful to my advisor, L. Scott Mills, for being an extraordinary

role model over the years. Scott’s intelligence, creativity, and passion and commitment to

conservation have continually challenged and motivated me. I am also deeply indebted to

my other major advisor, Tom Stephenson, for giving me the dissertation opportunity of

my dreams. Tom always endorsed my vision for Sierra Nevada bighorn sheep recovery,

and worked tirelessly to make my ideas a reality. I thank Mark Hebblewhite, my

surrogate advisor during the last months, as he was invaluable for helping me think

through complex ecological processes both conceptually and statistically. I am also

incredibly grateful for the insight, constructive feedback, and accessibility of my other

committee members, Dan Pletscher, Mike Mitchell, and Gordon Luikart.

This project was only possible with the support of the Sierra Nevada Bighorn

Sheep Recovery Program. Crew members had to hike, backpack, ski, snowshoe, climb,

ATV, snowmobile, and generally run themselves ragged chasing after elusive bighorn

sheep. Of the many people associated with the program I am especially thankful for the

efforts of Jonathan Fusaro, Kelsy Ellis, Lacey Greene, Dennis Jensen, Kathleen Knox,

Lora Konde, and Cody Schroeder. Mark Kiner not only brought ultra-athletism to our

program, but exceptional companionship. I am particularly grateful for my collaboration

with Dave German. His data management skills, enthusiasm for species recovery, and

desire for accountability in wildlife management have been vital to the program and to

Page 6: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

v

my education. Finally, this project could have happened without the support of John

Wehausen, Sierra bighorn sheep biologist extraordinaire. John was exceedingly generous

with his long-term data, time, and resources and has been a tremendous collaborator.

I also want to thank all the folks at the University of Montana that supported this

project. I am grateful to the Mills Lab for creating a supportive environment for learning

to think critically about scientific issues. Cindy Hartway, Elizabeth Crone, and Hugh

Robinson provided invaluable programming and statistical expertise. Jeanne Franz, Jodi

Todd, Jim Adams, and Reneea Gordon all helped administer contracts and grants, make

purchases, and hire a multitude of employees.

Several entities funded this work including the Sierra Nevada Bighorn Sheep

Recovery Program of California Department of Fish and Game, Canon National Parks

Science Scholars Program, the National Fish and Wildlife Foundation, Foundation for

North American Wild Sheep, and the Philanthropic Educational Organization. I sincerely

thank these groups for their financial commitment to endangered species conservation.

The past five years would not have possible without the incredible support of my

community of friends in both Missoula and Bishop. Thank you for welcoming me

seamlessly back into your lives after months of absence as I drifted between school-life

and field-life. I am particularly grateful to Scott Justham for his unfaltering smile and for

teaching me how to balance meaningful work and meaningful personal life.

I also want to thank my family - my dad, Ryan, Jan, Bill and Pam. Their

unconditional love and support through all of life’s challenges has been the foundation of

my sanity and spirit. This work is dedicated to the memory of my mother. Her strength,

integrity, and love of nature continue to inspire me.

Page 7: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

vi

TABLE OF CONTENTS

ABSTRACT ....................................................................................................................... iii

ACKNOWLEDGEMENTS ............................................................................................... iv

TABLE OF CONTENTS ................................................................................................... vi

LIST OF TABLES ............................................................................................................. xi

LIST OF FIGURES ...........................................................................................................xv

CHAPTER 1. Introduction and overview ............................................................................1

Background ......................................................................................................................1

Research objectives ..........................................................................................................4

General results ..................................................................................................................6

Management recommendations ......................................................................................11

Data collection and monitoring ..................................................................................11

Population-specific management recommendations ..................................................14

Planning future reintroductions ..................................................................................18

Research significance .................................................................................................20

Dissertation format .........................................................................................................21

Literature cited ...............................................................................................................22

CHAPTER 2. Population-specific vital rate contributions influence management of an

endangered ungulate ..........................................................................................................29

Abstract ..........................................................................................................................29

Introduction ....................................................................................................................30

Study area .......................................................................................................................34

Methods ..........................................................................................................................35

Vital rate parameter estimation ..................................................................................35

Asymptotic analyses ...................................................................................................39

Non-asymptotic life stage simulation analysis ...........................................................41

Results ............................................................................................................................43

Estimated vital rate parameters ..................................................................................43

Asymptotic analyses ...................................................................................................44

Non-asymptotic analyses ............................................................................................45

Discussion ......................................................................................................................46

Page 8: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

vii

Acknowledgements ........................................................................................................53

Literature cited ...............................................................................................................53

Appendix A. History and data collection of study populations .....................................73

Appendix B. Correlation matrices of population-specific vital rates .............................75

Appendix C. Population projections given correlations among vital rates ....................76

CHAPTER 3. Combining ground count, telemetry, and mark-resight data to infer

dynamics in an endangered species ...................................................................................78

Summary ........................................................................................................................78

Introduction ....................................................................................................................80

Materials and methods ...................................................................................................82

SNBS populations ......................................................................................................82

Data types ...................................................................................................................82

Model formulation and parameterization ...................................................................85

Results ............................................................................................................................92

Parameter estimates ....................................................................................................92

Covariate effects .........................................................................................................93

Discussion ......................................................................................................................94

Acknowledgements ......................................................................................................100

Literature cited .............................................................................................................100

Appendix D. Locations of study populations ...............................................................117

Appendix E. Pre-birth pulse survey modifications to demographic models ................118

Appendix F. WinBUGS code for post-birth pulse demographic models .....................120

Appendix G. Estimated annual vital rates and population growth rates ......................126

CHAPTER 4. From lambs to lambda: effects of inbreeding depression on the recovery of

endangered bighorn sheep................................................................................................130

Abstract ........................................................................................................................130

Introduction ..................................................................................................................131

Methods ........................................................................................................................134

Study populations .....................................................................................................134

Field sampling ..........................................................................................................135

Microsatellite analyses .............................................................................................135

Page 9: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

viii

Testing for inbreeding depression ............................................................................137

Quantifying population-level effects of inbreeding .................................................139

Results ..........................................................................................................................141

Genetic variation ......................................................................................................141

Detecting inbreeding depression ..............................................................................142

Effect of inbreeding depression on population dynamics ........................................143

Discussion ....................................................................................................................144

Acknowledgements ......................................................................................................148

Literature cited .............................................................................................................149

Appendix H. Identifying loci meeting neutral expectations ........................................166

Appendix I. Locus-specific genetic variation in microsatellite markers ......................172

Appendix J. Results from testing for locus-specific inbreeding depression effects .....177

CHAPTER 5. Apparent competition limits the recovery of an endangered ungulate:

quantifying the direct and indirect effects of predation ...................................................180

Abstract ........................................................................................................................180

Introduction ..................................................................................................................182

Methods ........................................................................................................................186

Study area and populations .......................................................................................186

Quantifying spatial overlap in bighorn and deer winter ranges ...............................187

Measuring direct effects of predation .......................................................................189

Measuring indirect effects of predation ....................................................................191

Results ..........................................................................................................................196

Spatial overlap of bighorn and deer winter ranges ...................................................196

Direct effects of predation ........................................................................................196

Indirect effects of predation ......................................................................................198

Discussion ....................................................................................................................199

Acknowledgements ......................................................................................................207

Literature cited .............................................................................................................207

Page 10: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

ix

LIST OF TABLES

CHAPTER 2

Table 2.1. Female Sierra Nevada bighorn sheep parameter estimates used in vital rate

analyses. Also provided for each vital rate are annual ranges, variances,

sensitivities (Sens), elasticities (Elast), and the proportion of variation in the

population growth rate explained (r2 of λ) given uncorrelated (Uncorr) and

correlated (Corr) vital rates. ……………………………………………….…… 61

Table 2.2. Median predicted sizes (NTmax) and stochastic growth rates (λs) of Sierra

Nevada bighorn sheep populations projected over 5 and 10 years. ……………. 64

Table 2.3. Predicted median size (NTmax) and growth rate (λs) of female Sierra Nevada

bighorn sheep populations given hypothetical management scenarios compared to

baseline predictions. ……………………………………………………………. 66

CHAPTER 3

Table 3.1. Number of years (n) that ground count, telemetry, and mark-resight data

were collected on the Warren, Wheeler, and Langley populations of Sierra

Nevada bighorn sheep……………………………………………….………… 105

Table 3.2. Annual adult female survival (ΦA), recruitment (R; for Wheeler) and

fecundity (F; for Warren and Langley) rates for Sierra Nevada bighorn sheep

populations. Estimates were obtained from ground count data only, telemetry data

only, and with the combined data model……………………………...………. 106

Table 3.3. Annual estimates of the number of adult females (NA) in the Wheeler and

Langley populations of Sierra Nevada bighorn sheep when estimated from ground

counts only, mark-resight (MR) only, and with the combined data model….... 108

Table 3.4. Posterior mean estimates, standard deviations, and 95% credible intervals

(95% CI) for regression coefficients from covariate models of adult female

survival and fecundity/recruitment rates from Sierra Nevada bighorn sheep

populations…………………………………………………………………….. 109

Page 11: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

x

CHAPTER 4

Table 4.1. Estimates of genetic variation for Sierra Nevada bighorn sheep populations

including observed heterozygosity (Ho), expected heterozygosity (He), and the

mean number of alleles (A)………………………………………………....… 156

Table 4.2. Pairwise FST values for Sierra Nevada bighorn sheep populations ……. 157

Table 4.3. Model selection metrics for non-genetic and genetic models of adult

survival and fecundity rates of Sierra Nevada bighorn sheep. ……………...… 158

Table 4.4. Parameter coefficients (± SE) of best non-genetic and genetic models for

adult survival and female fecundity. ………………………………………..… 160

Table 4.5. Means and variances of stochastic population growth rates for the Mono

Basin and Langley under current conditions (Baseline) and given the expected

effects of inbreeding depression (ID). Growth rates from simulated genetic

management at Mono Basin are also shown…………………………………... 161

Table 4.A.1. Comparison of population-level genetic diversity and differentiation

statistics for all polymorphic microsatellite markers (neutral and potentially

adaptive), regardless of meeting neutrality assumptions in Sierra Nevada bighorn

sheep populations. Individuals captured between 1999 and 2009 (n) were used to

calculate observed (Ho), expected heterozygosity (He), and the mean number of

alleles (A) for each population……..………………………...………………... 170

Table 4.A.2. Pairwise FST values of Sierra Nevada bighorn sheep populations using all

polymorphic markers, regardless of meeting the assumptions of neutrality….. 171

CHAPTER 5

Table 5.1. Mean annual probabilities of cause-specific mortality, survival, and overlap

with deer winter range for Sierra Nevada bighorn sheep populations………… 221

Table 5.2. Habitat model coefficient estimates from the top model of each population

of Sierra Nevada bighorn sheep ………………..…………………….……..… 222

Table 5.3. Comparison of models for winter habitat selection of Sierra Nevada

bighorn sheep populations with baseline parameters alone (slope, aspect, terrain

ruggedness, forest, and NDVI), with the addition of lion predation risk (R), and

with a risk by date interaction (R*D)…………………………………….….… 223

Page 12: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

xi

LIST OF FIGURES

CHAPTER 1

Figure 1.1. Diagram of the extinction vortex……………………………………….. 28

CHAPTER 2

Figure 2.1. Location of Mono Basin, Wheeler, and Langley Sierra Nevada bighorn

sheep populations, CA………………………………………………………….. 68

Figure 2.2. Number of adult females in the Wheeler, Langley and Mono Basin

populations of Sierra Nevada bighorn sheep, 1980-2007………………………. 69

Figure 2.3. Pre- and post-birth pulse matrix models used to simulate female Sierra

Nevada bighorn sheep population dynamics …………………………………... 70

Figure 2.4. Annual mean vital rates for the Mono Basin, Langley, Wheeler

populations of Sierra Nevada bighorn sheep, 1985-2007. …………………….. 71

Figure 2.5. Analytical elasticities and coefficients of determination (r2) for Sierra

Nevada bighorn sheep fecundity, yearling survival (yrl surv), and adult survival

(adt surv) rates in the Mono Basin and Langley populations. ………………… 72

CHAPTER 3

Figure 3.1. Sierra Nevada bighorn sheep demographic model for populations surveyed

post-birth pulse. Data inputs are represented by boxes and estimated parameters

are represented by circles. Solid arrows depict stochastic dependencies between

data and parameters and dashed arrows represent deterministic dependencies. The

model combines ground count (yCA, yCY, yCL), telemetry (yT), and mark-resight

(yMR) data to estimate numbers of adult females (NA), yearling females (NY) and

lambs (NL), adult female survival (ΦA), yearling female survival (ΦY), and

fecundity (F). The model also estimates detection probability for mark-resight

data (ρ) and variance of ground count data (σ2

y)…………………………….... 112

Figure 3.2. Estimates from the Wheeler population of Sierra Nevada bighorn sheep.

A) Annual recruitment rates (and SD) across all years of the study (1981-2009).

From 1981-2000 estimates were based only on ground count data and from 2001-

2009 estimates were based on the combined model (count and telemetry data). B)

Page 13: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

xii

Recruitment rate estimates when using only count data, only telemetry data, and

both data types combined. C) Standard deviations around recruitment estimates

when using only count data, only telemetry data, and both data types

combined…………………………………………………..…………………... 113

Figure 3.3. Estimated number of adult female Sierra Nevada bighorn sheep (and 95%

credible intervals) in the Warren, Wheeler, and Langley populations. Black circles

(●) signify years that demographic data were collected on the populations and

open squares (□) signify years that demographic data were not collected......... 114

Figure 3.4. Number of adult female Sierra Nevada bighorn sheep counted during

annual ground surveys and estimated from the Bayesian state-space demographic

model (with 95% credible intervals) in Warren from 1988 to 2008…………... 115

Figure 3.5. Predicted effects of winter snow depth, summer rainfall, and density on

adult female survival and fecundity/recruitment rates for populations of Sierra

Nevada bighorn sheep. ……………………………………………….……….. 116

CHAPTER 4

Figure 4.1. Location of Mono Basin, Wheeler, Baxter and Langley populations of

Sierra Nevada bighorn sheep, CA……………………………………………... 162

Figure 4.2. Number of adult females in the Wheeler, Langley, Baxter and Mono

Basin, 1998-2008. …………………………………………………………….. 163

Figure 4.3. Annual fecundity (± SE) for adult female Sierra Nevada bighorn sheep as a

function of multilocus heterozygosity. ……………………………………….. 164

Figure 4.4. Predicted size of the Mono Basin population over 10 generations if

heterozygosity (h) remains unchanged (h = 0.43; Baseline), if inbreeding

depression continues to reduce h and fecundity (Inbreeding Depression), if

average h was increased to 0.48, and if average h was increased to 0.59 …….. 165

CHAPTER 5

Figure 5.1. Mortalities of radio-collared Sierra Nevada bighorn sheep that have

occurred each month of the year from 2002 to 2010………………..………… 224

Figure 5.2. Location of Mono Basin, Wheeler, Baxter, and Langley populations of

Page 14: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

xiii

Sierra Nevada bighorn sheep. Green areas represent the winter minimum convex

polygons for bighorn sheep in each population and yellow areas delineate the

winter ranges of adjacent mule deer herds. ………………………………..….. 225

Figure 5.3. Locations of bighorn sheep mortalities in the Baxter population by

mountain lion predation. Green areas delineate the winter range of bighorn sheep

and yellow areas delineate the winter range of mule deer………………..…… 226

Figure 5.4. Probabilities of selection for elevation by Sierra Nevada bighorn sheep

populations……………………………………………………………..……… 227

Figure 5.5. Probability of selection for a) predation risk, b) slope, and c) terrain

ruggedness for populations of Sierra Nevada bighorn sheep wintering in close

proximity to deer herds……………………………………………...………… 228

Figure 5.6. Probability of selection for areas of low, medium, and high risk of lion

predation by Sierra Nevada bighorn sheep in the Baxter population over the

course of the winter (Dec-Apr)…………………..……………………………. 229

Page 15: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

1

CHAPTER 1

INTRODUCTION AND OVERVIEW

BACKGROUND

An increasing proportion of the world’s biodiversity is in danger of extinction, as

habitat loss, pollution, and invasive species rapidly alter biological conditions (Wilcove

et al. 1998). While endangered species conservation has been a key issue in natural

resource management, populations of many of these species remain precariously low. Of

the 1,370 species listed under the U.S. Endangered Species Act (ESA) only 21 have been

successfully recovered (www.fws.gov/endangerd). Male and Bean (2005) found that

among those species still listed, 48% were still declining, 42% were stable, and <10%

were improving. Given all the resources allocated to endangered species conservation,

why is recovery so difficult?

A growing body of evidence suggests that difficulties associated with species

recovery are often due to the “extinction vortex.” Gilpin and Soule (1986) were the first

to conceptualize this process, describing how demographic, environmental, and genetic

stochasticity could interact with each other and with deterministic factors, such as habitat

loss, to mutually reinforce and accelerate the extinction of small populations (Fig. 1.1). In

an analysis of 10 wildlife populations that have gone extinct, Fagan and Holmes (2006)

found that all exhibited dynamics indicative of an extinction vortex prior to collapse. A

classic example of a population that has experienced this phenomenon is the greater

prairie chicken (Tympanuchus cupido pinnatus) of Illinois. The prairie chicken

population was dramatically reduced due primarily to habitat loss, but despite efforts to

Page 16: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

2

improve habitat conditions the population continued to dwindle in size (Westemeier et al.

1998). The decline had significantly reduced genetic variation in the population, which in

turn, had led to inbreeding depression and impaired survival and reproductive rates. The

interaction of demographic, habitat, and genetic factors had initiated an extinction vortex

that could not be reversed through the management of habitat alone, even though this

caused the initial population crash. Similar extinction vortex dynamics have implicated in

the demise of several other endangered animals and plant populations (Saccheri et al.

1998, Madsen et al. 1999, Vergeer et al. 2003).

These studies have critical implications for wildlife conservation because they

suggest that the recovery of endangered species will often depend upon escaping an

extinction vortex. Wilcove et al. (1993) found that most populations are already

dangerously small when granted “endangered” status under the ESA, suggesting that

extinction dynamics are already in effect by the time recovery efforts are initiated. By

waiting until populations are critically low before initiating recovery efforts, we allow

stochastic factors to interfere and corrode the dynamics of populations (Brook et al. 2008,

Melbourne and Hastings 2008). By definition, stochastic factors elicit random and

unpredictable effects that complicate population processes and reduce the efficiency of

recovery actions. Thus, by allowing populations to exhibit extinction dynamics before

initiating aggressive recovery actions, we exacerbate the challenge of achieving

conservation goals. Indeed, this may explain why species recovery has been so difficult,

with <10% of the species listed under the ESA currently described as “improving” (Male

and Bean 2005), and only about 1.5% of those species successfully recovered.

Page 17: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

3

While the interaction of stochastic and deterministic factors introduces inherent

complexity into the dynamics of small populations, management programs often neglect

this complexity, focusing only on those factors responsible for initial population declines

rather than on those inhibiting recovery (Rabinowitz 1995, Asquith 2001, Brook et al.

2008). Such single-factor approaches have lead to costly errors when applied individually

to small populations, causing delays and failures in recovery (Armstrong and Ewen 2001,

Asquith 2001, Briskie and Mackintosh 2004). To achieve success in endangered species

conservation, wildlife professionals must recognize the role of extinction dynamics in

small populations and take a multifactor approach to investigate the relative and

synergistic effects of different factors. Indeed, this has been recognized as one of the

greatest needs in the field of conservation (Boyce 1992, Asquith 2001, Ouborg et al.

2006, Brook et al. 2008, Laurance and Useche 2009).

Sierra Nevada bighorn sheep (Ovis canadensis sierrae; hereafter SNBS) are the

rarest subspecies of mountain sheep in North America, with approximately 400

individuals (California Dept. Fish and Game, unpublished data). SNBS were listed under

the ESA in 1999, when surveys revealed that only 100 adults could be accounted for in

the wild, the lowest number ever recorded (U.S. Fish and Wildlife Service 2007). SNBS

are distributed in six small, isolated sub-populations that are highly vulnerable to

extinction dynamics due to both stochastic and deterministic factors. Being a highly

valued endemic subspecies of California, government agencies, non-profit organizations,

and research scientists have collected detailed information on SNBS since 1980. This

wealth of long-term data makes SNBS an excellent system for examining the dynamics

Page 18: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

4

of small populations and for developing new analytical tools to aid in the recovery of at-

risk species.

RESEARCH OJECTIVES

For my dissertation, I was given the opportunity by California Department of Fish

and Game to develop a comprehensive research program to guide recovery efforts for

Sierra Nevada bighorn sheep. I employed a multifactor approach to elucidate the

dynamics of SNBS populations and identify the most effective management strategies for

meeting recovery goals. Specifically my objectives were to:

Identify vital rates most critical to SNBS population performance (Chapter 2).

Because different vital rates (survival and reproductive parameters)

disproportionately affect the growth or decline of a population (Morris and Doak 2002,

Mills 2007), my first objective was to determine which rates were responsible for poor

SNBS performance, and those whose increase would most effectively stimulate recovery.

This is a critical first step in developing efficient recovery strategies as management

efforts which target the most influential vital rates will have the greatest potential to

redirect a population’s trajectory. To meet this objective I applied a suite of sensitivity

analyses to identify vital rates whose increase would most quickly accelerate subspecies

recovery.

Examine the stochastic and deterministic factors responsible for spatial and

temporal variation in key SNBS vital rates (Chapters 3, 4, and 5).

Theoretical expectations (Keller and Waller 2002, Lande et al. 2003) and

empirical evidence (Wehausen 1992, Wehausen 1996, U.S. Fish and Wildlife Service

2007) suggest that SNBS populations were highly vulnerable to demographic and

Page 19: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

5

environmental stochasticity, inbreeding depression due to genetic drift, variation in

habitat quality, and predation by mountain lions. To develop successful recovery plans

conservation practitioners need information on which stochastic and deterministic factors

drive the values of key vital rates. I initiated a combination of demographic, habitat,

predation, and genetic analyses to identify the ecological drivers limiting the recovery of

SNBS and to examine their relative and synergistic impacts on population performance.

Develop effective management strategies to increase SNBS population growth

rates and reach recovery goals (Chapters 2, 4, and 5).

The size, distribution, and connectivity of SNBS populations must be increased to

meet delisting requirements outlined in the Recovery Plan (U.S. Fish and Wildlife

Service 2007). Management options proposed for SNBS include reintroductions,

augmentations, habitat enhancement projects, predator management, and genetic

management, but the benefits of these different activities for stimulating recovery had yet

to be determined. I used information about key vital rates and their drivers to determine

the effectiveness of different management scenarios at increasing population growth rates

and escaping extinction dynamics.

Improve quantitative methods for evaluating demographic processes of

endangered populations (Chapters 2, 3, and 4).

Data on endangered populations are inherently piecemeal, having small sample

sizes, inconsistent data collection methods, intermittent data collection, and information

on only a subset of important parameters and covariates (Tear et al. 1995, Fieberg and

Ellner 2001, Morris et al. 2002). Traditional statistical approaches are limited in their

ability to analyze such data, and as a result, critical management and conservation

Page 20: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

6

decisions are routinely made with limited quantitative analysis. In assessing the

demographic drivers of SNBS dynamics, a major focus of this dissertation has been on

improving quantitative methods for small and endangered populations.

GENERAL RESULTS

Vital Rates Responsible for SNBS Dynamics – From the sensitivity analyses in

Chapter 2 I found that SNBS vital rates showed high spatial and temporal variation,

resulting in population-specific dynamics that did not fit general expectations from other

ungulates, particularly during periods of population decline. The dominant paradigm for

ungulates is that adult female survival is generally high with low variation, causing it to

contribute relatively little to changes in population growth rates. Juvenile survival,

however, tends to be low with high variation, making it the primary determinant of

population change (Gaillard et al. 1998; Gaillard et al. 2000; Raithel et al. 2007). In

contrast to this paradigm, I found that the growth rates of the Mono Basin and Wheeler

populations were driven by adult female survival, as rates were lower and more variable

than expected. Only Langley exhibited expected patterns, with fecundity explaining the

majority of the variation in population performance because adult survival was high and

relatively constant. Differences in the vital rate drivers of SNBS populations resulted

from population-specific stochastic and deterministic influences described below.

Such shifts in the means and variances of key vital rates may be largely

responsible for declining and endangered populations, a pattern that has not been

recognized by conservation practitioners. Given the lack of demographic data on many

populations, it has been suggested that important vital rates identified in other

populations of the same species, or those from similar species, be used to guide

Page 21: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

7

conservation efforts (Silvertown et al. 1996; Heppell 1998; Sæther and Bakke 2000).

These results illustrate the potential danger of this approach. For example, based on

studies of other ungulates, a reasonable assumption would be to focus SNBS

management on increasing juvenile survival, as this rate has been responsible for the

dynamics of other large herbivores. Data from SNBS suggest, however, that it is a

decrease in adult survival that is the primary driver of SNBS declines and should be the

focus of monitoring and management activities. As a result, inferences about the

importance of different vital rates from one species or population may not be applicable

to another, and could potentially misdirect critical resources if inappropriately employed

within conservation programs. Furthermore, endangered species recovery programs

should be responsive to deviations between observed vital rate values and those predicted

from classic life-history expectations. Such departures may be largely responsible for

population declines and serve as important targets for monitoring programs and

management actions.

Stochastic and Deterministic Drivers of Key Vital Rates – Chapters 3, 4, and 5

examine the influence of stochastic and deterministic factors on SNBS adult survival and

fecundity rates; those rates most important in determining performance. I found that

environmental stochasticity, indexed by winter severity and summer precipitation,

affected all SNBS populations, but to different degrees (Chapter 3). While weather had

strong effects on fecundity in Wheeler and Langley, it had negative effects on both

fecundity and adult survival in Mono Basin. In ungulates, environmental stochasticity

typically affects only the youngest stage classes (Gaillard et al. 2000), so its influence on

adult survival at Mono Basin is disconcerting and likely due to its extremely small size

Page 22: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

8

(Melbourne and Hastings 2008). As for other stochastic factors, I found evidence that

inbreeding depression has reduced SNBS fecundity, but had relatively little influence on

long-term population projections (Chapter 4). In evaluating deterministic factors, I found

evidence of positive density dependence (or an Allee effect; Courchamp et al. 1999) in

adult survival in Mono Basin, but negative density dependence in survival and fecundity

in Wheeler and Langley (Chapter 3). Mountain lion predation had no effect on Mono

Basin, but decreased adult survival in all other herds, particularly in Baxter (Chapter 5).

These analyses illustrate the diverse suite of factors driving the dynamics of

SNBS populations. While each stochastic and deterministic factor was significant in

affecting vital rates in at least one population, the relative influence of these factors was

unique to each herd. For example, population growth in Mono Basin was inhibited by

low adult survival, which was most influenced by environmental stochasticity and Allee

effects. Meanwhile, growth rates at Wheeler were also limited by adult survival, but due

to lion predation and density dependence. At Langley, the fastest growing herd, variation

in population growth was attributed to fecundity, with environmental and genetic

stochasticity having the greatest effects on that rate. Populations of SNBS are all small,

isolated, and in a relatively small geographic area, and yet, they are driven by entirely

different processes. As predicted by the extinction vortex, stochastic factors have

interceded in the dynamics of SNBS herds and must be accounted for in recovery

planning. Our results emphasize the importance of examining multiple factors in

assessing the ecological drivers of endangered species and that dynamics may often be

specific to individual populations.

Page 23: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

9

Developing Effective Management Strategies –In Chapters 2 and 4, I model the

predicted outcomes of different management scenarios for SNBS, given the distinct

dynamics of each herd. Just as dynamics were specific to each population I found that the

most effective management strategies were also specific to each population. Predator

management appeared to be most effective for boosting performance in Wheeler and

Baxter (Chapters 2 and 5), while an augmentation would be most effective increasing the

Mono Basin population (Chapter 2). Given the current high growth rate of the Langley

herd, management actions are not predicted to have an appreciable effect in the short-

term, and thus recovery efforts could be better invested elsewhere. Although evidence of

inbreeding depression was detected in all SNBS populations, simulations of genetic

management did not appear to have a significant effect over the next few decades

(Chapter 4). That said, because genetic variation was lower in SNBS than any other

bighorn sheep population, management strategies should work to maintain genetic

variation by restoring connectivity and gene flow among existing herds so as to maximize

the adaptive potential of the subspecies. This suite of analyses demonstrates that multiple

management actions will be required to increase growth rates and alleviate extinction

dynamics in SNBS. Based on detailed information of the demographic drivers of each

population, the SNBS Recovery Program can now implement recovery strategies tailored

specifically to each herd.

Improving Quantitative Methods for Endangered Populations – Statistical

analyses of endangered species are often limited by inadequate data and inappropriate

assumptions. In Chapters 2, 3, and 4, I confront these limitations by improving current

demographic and genetic quantitative methods. In Chapter 2 I develop a simulation-

Page 24: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

10

based, non-asymptotic “life-stage analysis” method (LSA; see Mills 2007) for examining

transient dynamics of small populations. Most vital rate sensitivity analyses rely on

asymptotic matrix properties which are inappropriate for small populations

(Bierzychudek 1999; Clutton-Brock and Coulson 2002; Fefferman and Reed 2006). The

non-asymptotic approach I applied to SNBS vital rates obtained more accurate short-term

predictions for endangered populations and provided non-intuitive results of relevance to

managers.

Given the wide application of telemetry in wildlife studies, combining telemetry

data with other data types has tremendous potential for enhancing demographic

parameter estimation. In Chapter 3, I extend the use of Bayesian state-space models to

combine all available data types on SNBS (minimum count, mark-resight, and telemetry

data) to estimate key demographic parameters (Brooks et al. 2004, Schaub et al. 2007).

Models combining disparate data types increased accuracy and precision in parameter

estimates, fit covariates to vital rates driving population performance, and standardized

the error structure across different data types. This analysis is the first example of

integrating telemetry and ground count data within the state-space approach.

Finally, in Chapter 4, I assess neutral and candidate adaptive genetic variation in

SNBS. Typically, only neutral genetic variation is employed in studies of wildlife

population genetics, with the assumption that it serves as an index of adaptive variation

and fitness. Candidate adaptive markers - those expected to be closely tied to individual

fitness (Luikart et al 2003) - have recently become available for some wildlife species.

This study is one of the first to compare genetic variation of neutral and candidate

markers in an endangered species. Inferences from each marker type was highly

Page 25: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

11

correlated in SNBS, and contrary to expectation, fitness traits were not more closely

related to adaptive markers than neutral ones.

MANAGEMENT RECOMMENDATIONS

A key objective of my dissertation was to provide critical information for future

management and recovery of SNBS. To date, the Sierra Nevada Bighorn Sheep Recovery

Plan (U.S. Fish and Wildlife Service 2007) has been the guiding document for recovery

activities for the subspecies. While the plan lists several actions expected to promote

recovery, these actions are not clearly prioritized at either the level of the subspecies or

for individual populations. As a result, managers require further direction about which

actions would be most beneficial for SNBS recovery, and the order in which suggested

actions should be taken. By identifying the vital rates driving population performance and

their ecological determinants, I have identified several key management

recommendations to direct future recovery efforts. I outline the management

recommendations produced from this dissertation as they pertain to 1) monitoring and

data collection priorities, 2) population-specific recovery actions, and 3) strategies for

future reintroductions and translocations.

Data Collection and Monitoring

Monitor adult female survival ─ Detailed in Chapter 1, I found that depressed adult

female survival rates, and increased variation in that rate, are largely responsible for

declining population trends in SNBS. Because adult female survival is the vital rate

with the greatest elasticity, reductions in that rate can precipitate a population decline

faster than a reduction in any other rate. For stable and increasing populations, mean

adult survival rates should generally be ≥90% in bovid populations (Gaillard et al.

Page 26: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

12

2000). By carefully monitoring adult female survival the Recovery Program will be

alerted to populations in need of swift management intervention.

Couple telemetry and ground count data to effectively track demographic parameters

and estimate population growth rates ─ A combination of ground count, telemetry,

and mark-resight data have been collected on SNBS populations. Chapter 2 describes

how these data types can be used individually or in combination to estimate different

demographic parameters and assess the factors influencing population trends. Based

on the size of individual SNBS populations, I found that different data types were

more effective at estimating key parameters. For example, in small populations (i.e.

Mono Basin) annual ground counts were most effective at estimating survival and

reproductive rates. In large populations, however, it was more efficient to use

telemetry for estimating adult survival (as long as >30% of the females were radio-

collared) and either telemetry or counts to estimate reproduction. As a result, data

collection strategies should be specified for each SNBS population based on size and

accessibility. When there are multiple existing data types that provide information on

the same demographic rates, these data should be combined to yield more precise and

accurate estimates, and better track population trends.

Evaluate genetic diversity every 6-7 years (every SNBS generation) ─ Allele

frequencies in populations change on the time-scale of generations, and as a result, it

will be important to quantify any losses in genetic diversity on that basis. Given

current levels of genetic variation in SNBS, relative to other wild bighorn sheep

populations, it appears likely that this subspecies has lost roughly one-third to one-

half of its variation (Chapter 4). As a result, SNBS are expected to have reduced

Page 27: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

13

evolutionary potential and be vulnerable to novel environmental stressors. Future

monitoring should track any additional losses of genetic variation due to drift, and

work to ameliorate those losses through targeted translocations (described below).

Identify the spatial distribution of mule deer during winter months ─ Overlap in the

winter ranges of SNBS and mule deer precipitate predation effects by mountain lions

on endangered bighorn sheep populations. This occurs through apparent competition,

when lions opportunistically take SNBS, a secondary prey species, when they are

found in close proximity to deer, their primary prey source. In Chapter 5, I found that

the spatial overlap between the two ungulate species has resulted in direct and

indirect predation effects on SNBS, increasing mortality rates and eliciting strong

anti-predator behavior. To assess the potential for apparent competition to limit SNBS

recovery in both occupied and unoccupied habitat, the distribution, density, and

population trend of mule deer should be routinely monitored. Any major changes to

the distribution of deer, particularly expansions of their range, will be important to

track as such changes could elicit key demographic consequences for recovering

SNBS populations.

Determine cause-specific mortality factors of collared individuals ─ Knowing the

causes of SNBS mortality is critical for implementing timely management actions.

For example, we found that lion predation was responsible for most mortality in the

Baxter and Wheeler populations, while unknown factors were responsible for most

mortality in Mono Basin and Langley. When specific mortality factors (i.e. lion

predation) are strongly linked to appropriate management actions (i.e. lion removal),

cause-specific mortality data can direct swift recovery actions. Based on collared

Page 28: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

14

individuals, deaths attributed to known cause had a median investigation time of 7.5

days, while those deaths that were attributed to an unknown cause had a median

investigation time of 30 days. Logistic regression revealed that mortality sites that

were investigated within 7 days of death had an 84% probability of being attributed to

a known cause (95% confidence interval from 77-90%). To adequately detect major

predation, disease, or other mortality factors affecting SNBS populations to direct

appropriate management actions, survival of collared individuals should be monitored

on a weekly basis.

Quantify the demographic effects of different management actions ─ Management

actions should be prioritized based on their ability to increase the size and distribution

of SNBS populations. While several potential actions are outlined in the Recovery

Plan, it is unclear how many of them are directly linked to specific vital rate

parameters or population performance. Managers should focus on those actions

demonstrated (through empirical data or modeling) to have the greatest impact on

vital rates that drive population growth (Chapter 2). When modeled management

outcomes are implemented on the ground, careful monitoring of the demographic

effects of those actions will need to occur. This information should be used to

improve future modeling efforts and management decisions, within the iterative

process of adaptive management (Williams 2001, Lyons et al. 2008).

Population-Specific Management Recommendations

The major management actions that can be used to boost population growth rates in

existing herds of SNBS are augmentations, genetic management, predator management,

prescribed fire, and disease prevention. For each of the major populations evaluated in

Page 29: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

15

this study, I outline specific management recommendations as they pertain to these

actions, with a couple of exceptions. Detailed analyses of the effects of prescribed fire are

described in Greene et al. (In Preparation); I only briefly touch on their potential benefit

here. Also, because this dissertation did not investigate disease factors of SNBS, I do not

make management recommendations relevant to disease prevention (but see Clifford et

al. 2009, Cahn et al. In Preparation).

Mono Basin – This population should be augmented with additional SNBS, as there

was evidence of positive density dependence, or Allee effects, in adult female survival

rates (Chapter 3). By increasing overall population size, adult female survival rates are

expected to increase, having a significant effect on population growth (Chapter 2).

Additionally, an augmentation may also slightly increase genetic variation in the Mono

Basin. While a minor increase in genetic variation was not predicted to have near-term

population-level effects (Chapter 4), heterosis (unaccounted for in our models; Tallmon

et al. 2004) may exacerbate the genetic benefit of an augmentation. Additionally, an

increase in population size may alter habitat use patterns of this herd. Collared SNBS in

Mono Basin did not use low elevation winter range, even though some of this habitat was

available. Instead, SNBS remained at high elevations throughout the winter, and both

adult survival and fecundity rates were negatively associated with severe snow

conditions. An augmentation may elicit changes in habitat use patterns, as SNBS from

other populations are accustomed to using lower elevation winter range. Given that there

are no deer herds that winter adjacent to the Mono Basin population (Chapter 5), and

thus, no current resident mountain lions, this population may also substantially benefit

from a prescribed fire. A fire could be used to increase the availability of non-forested,

Page 30: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

16

low-elevation habitat, allowing SNBS to descend to lower elevations and escape harsh

winter weather conditions.

Wheeler – Analyses suggest that the growth rate of the Wheeler population will

benefit most from mountain lion management (Chapter 2 and Chapter 5). Over the past

10 years at least 5% of adult SNBS in this population have been preyed upon by lions.

This is a highly conservative estimate as the cause of almost 40% of collared SNBS

mortalities were unknown, and many are suspected to have succumbed to lion predation.

Although this herd has been increasing in size (Chapter 2), lion management could be

used in the short-term to maximize population growth rates and generate additional

source stock for future reintroductions. Additionally, I found evidence of negative density

dependence in vital rate values, a trend that may be due to either food-based or predator-

based carrying capacity. By initiating predator management, these factors could be teased

apart. If vital rates continue to show signs of negative density dependence post-predator

management, this population should be prioritized as a source herd for reintroductions

and augmentations.

Baxter – Predator management should be initiated to increase the growth rate of the

Baxter population. Cause-specific mortality analyses of collared SNBS conservatively

revealed that a mean of 12% of the adults were annually eliminated due to lion predation

(Chapter 5). Additionally, the annual survival rate for Baxter was 80%, the lowest of any

SNBS population and significantly lower than those of other bovid populations (Gaillard

et al. 2000). Such high predation rates on adult SNBS are not sustainable. The winter

range of SNBS in Baxter has greater overlap with deer than any other SNBS herd,

leading to high lion predation via apparent competition (Chapter 5). When SNBS and

Page 31: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

17

deer have overlapping seasonal habitat, managers must expect to have to manage lion

predation to increase small and endangered secondary prey populations.

Langley – The Langley population was estimated to be growing at 18%/year during

the time period evaluated in this study (Chapter 2). Given the high growth rate of the

herd, I found that management activities would best be invested in other populations,

allowing Langley to grow independently of management. The high growth rate of this

herd, coupled with evidence of density dependence in fecundity rates (Chapter 3),

suggests that Langley is an ideal source for translocation stock. Individuals could be

removed for augmentations or reintroductions, pending that the risk of such removals is

found to be insignificant based on quantitative projection models. An important factor to

recognize about SNBS in Langley is their use of habitat during winter months. The

Recovery Plan currently stresses the importance of low elevation habitat use as a

prerequisite for recovering populations, however, SNBS in Langley have had limited use

of such areas (Chapter 5) yet still maintain high demographic rates (Chapters 2 and 3).

Maximizing the use of low elevation habitat (<2,750 m) may not be a requirement for

SNBS populations to increase in size and distribution, particularly when low elevation

habitat is associated with high mortality by mountain lions.

All populations – All SNBS have low genetic variation, as exhibited by their

heterozygosity and allelic diversity values (Chapter 4). Such limited individual genetic

variation was associated with reduced fecundity rates, a clear sign of inbreeding

depression. Although, I did not find that inbreeding depression affected short-term

population dynamics, maintaining genetic variation should still be a priority as this will

maximize the evolutionary potential of the subspecies. Given that SNBS populations are

Page 32: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

18

currently isolated, genetic variation could be maintained by increasing connectivity

among populations and by translocating individuals between populations.

Planning Future Reintroductions

To de-list SNBS from the Endangered Species Act, the Recovery Program must

increase the size, number, and distribution of SNBS populations (U.S. Fish and Wildlife

Service 2007). Meeting these requirements will necessitate reintroducing new

populations into unoccupied habitat. Results from this dissertation yield key

recommendations for such reintroductions.

Identify Appropriate Source Stock ─ Stochastic population viability models should be

used to estimate the risk of removing individuals from one population to start new

populations. To adequately model the relative risk for source populations, managers

must explicitly identify the probability of extinction they are willing to accept for the

source herd before conducting removals. In addition, the probability of persistence for

reintroduced populations should also be modeled, given the probable number of

animals that will be translocated and their expected vital rates (Morris and Doak

2002). By modeling projected outcomes in both the source and reintroduced

populations, managers can weigh the costs and benefits of such actions in a

transparent framework. Currently, Wheeler and Langley are exhibiting patterns of

negative density dependence (Chapter 3). Managers should determine whether these

patterns result from food-based or predator-based carrying capacities. If density

dependence is food-based, these populations should be targeted as reintroduction

source stock.

Page 33: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

19

Assess the potential for apparent competition to limit population performance ─ Prior

to reintroducing SNBS into unoccupied habitat, managers should determine the local

distribution of deer and lions. Low elevation winter ranges that are adjacent to, or

overlap with, deer range will also be inhabited by mountain lions, the main predator

of SNBS. Our habitat models found that SNBS selected for, rather than avoided, areas

of high lion density, a behavior which was associated with higher lion-caused

mortality. As such, reintroduction sites should either be in areas where low elevation

winter range is de-coupled from deer and lion populations or areas where high

elevation habitat can successfully sustain bighorn sheep throughout the winter (such

as high elevation winter range at Langley). Such places can serve as refugia for

bighorn sheep recovery, minimizing the need for continual predator management.

Evaluate local weather conditions ─ I found that severe winter weather in areas of

Mono Basin used by SNBS decreased adult survival and fecundity rates, while winter

weather did not have a negative effect on any other population. Mono Basin, the

northernmost SNBS herd, generally experiences the harshest winter conditions.

Weather stations within Mono Basin reported snow depths almost double those

reported for the other herds, indicating that large-scale precipitation patterns along the

Eastern Sierra may influence SNBS demographic rates. As a result, weather patterns

should be considered when prioritizing potential reintroduction sites.

Maximize genetic variation ─ Source stock for reintroductions should be carefully

selected to maximize the genetic variation of new populations. Ideally, animals

should be selected from multiple herds to increase mean heterozygosity, as existing

herds have experienced genetic drift for several generations.

Page 34: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

20

Monitor vital rates of new populations ─ Once a new population has been

reintroduced the Recovery Team should carefully monitor survival and reproductive

rates. By taking a targeted vital rate monitoring approach, managers will be able to 1)

determine the growth rate of the new population, 2) identify specific parameters that

may limit performance, and 3) direct management activities towards vital rates that

are most influential in driving population persistence.

RESEARCH SIGNIFICANCE

As anthropomorphic factors continue to reduce wildlife populations,

understanding the processes that govern the fate of small populations is becoming

increasingly urgent. Up to 30% of the species on earth are predicted to be threatened with

extinction within the next century due to climate change (IPCC 2007). As a result, the

challenge of preserving biodiversity will continue to dominate the field of natural

resource management well into the future. To succeed in the recovery of endangered

species managers must account for the powerful influence of stochasticity in the

dynamics of small populations and acknowledge that factors responsible for initial

population declines may not be the same ones inhibiting recovery. The demonstrated

ability of the extinction vortex to exacerbate the decline of small populations implies two

fundamental truths. First, practitioners should quantify the relative contribution of

stochastic and deterministic factors to develop more effective management strategies; and

second, that ESA listing and recovery should be initiated prior to the onset of extinction

dynamics when populations are larger and driven primarily by deterministic processes.

The following chapters on SNBS provide a scientific framework for assessing the

existence of extinction processes and alleviating them through directed management. To

Page 35: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

21

assess extinction dynamics managers must identify vital rates responsible for poor

population performance and take a multifactor approach to examine the factors that drive

the values of those rates. Just as the extinction vortex predicts, I found that small

populations of SNBS were driven by a number of stochastic and deterministic processes.

Demographic, habitat, climate, predation, and genetic factors were working singly and in

concert to shape the overall viability of this subspecies. The interaction of factors led to

atypical demographic patterns in SNBS that deviated from theoretical expectations and

increased extinction risk. To alleviate extinction processes, I found that management

strategies must be tailored to population-specific dynamics, targeting those vital rates and

ecological drivers which have the greatest power to increase performance. Results from

this study have elucidated critical aspects of SNBS ecology, provided a recovery strategy

for Sierra Nevada bighorn sheep, and supplied new quantitative tools for examining the

dynamics of small populations. Ultimately, this work offers an example of assessing

population viability, not in terms of probability of extinction, but in terms of quantifying

conservation measures that will alleviate extinction dynamics and achieve recovery goals.

DISSERTATION FORMAT

The following chapters were formatted for individual publication in specific peer-

reviewed scientific journals. Chapter 2 is currently “In Press” in Ecological Applications

(Johnson et al. 2010) and Chapter 3 is “In Press” in Journal of Applied Ecology (Johnson

et al. In Press). Because I worked in conjunction with several collaborators to meet the

objectives of this study, co-authors are listed at the start of each chapter and I shift from

the singular “I” to the collective “we” throughout the remainder of the dissertation.

Page 36: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

22

LITERATURE CITED

Armstrong, D.P., and J.G. Ewen. 2001. Assessing the value of follow-up translocations: a

case study using New Zealand robins. Biological Conservation 101:239-247.

Asquith, N.M. 2001. Misdirections in conservation biology. Conservation Biology

15:345-352.

Bierzychudek, P. 1999. Looking backwards: assessing the projections of a transition

matrix model: Ecological Applications 9:1278-1287.

Boyce, M.S. 1992. Population viability analysis. Annual Review of Ecology and

Systematics 23:481-506.

Briskie, J.V., and M. Machintosh. 2004. Hatching failure increases with severity of

population bottlenecks in birds. Proceedings of the National Academy of Sciences

101:558-561.

Brook, B.W, N.S. Sodhi, and C.J.A. Bradshaw. 2008. Synergies among extinction

drivers under global change. Trends in Ecology and Evolution 23:453-460.

Brooks, S.P., R. King, and B.J.T. Morgan. 2004. A Bayesian approach to combining

animal abundance and demographic data. Animal Biodiversity and Conservation

27:515-529.

Cahn, M.L., M.M. Conner, O.J. Schmitz, M.W. Miller, T.R. Stephenson, H.E. Johnson,

and J.D. Wehausen. Disease, population viability, and recovery of endangered

Sierra Nevada bighorn sheep. In Preparation.

Clifford, D.L., B.A. Schumaker, T.R. Stephenson, V.C. Bleich, M.L. Cahn, B.J.

Page 37: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

23

Gonzales, W.M. Boyce, and J.A.K. Mazet. 2009. Assessing disease risk at the

wildlife-livestock interface: a study of Sierra Nevada bighorn sheep. Biological

Conservation 142:2559-2568.

Clutton-Brock, T.H., and T. Coulson. 2002. Comparative ungulate dynamics: the devil is

in the detail: Philosophical Transactions of the Royal Society of London Series B

357:1285-1298.

Courchamp, F., T. Clutton-Brock, and B. Grenfell. 1999. Inverse density dependence and

the Allee effect. Trends in Ecology and Evolution 14:405-410.

Fagan, W.F., and E.E. Holmes. 2006. Quantifying the extinction vortex. Ecology Letters

9:51-60.

Fieberg, J., and S.P. Ellner. 2001. Stochastic matrix models for conservation and

management: a comparative review of methods. Ecology Letters 4:233-266.

Fefferman, N.H., and J.M. Reed. 2006. A vital rate sensitivity analysis for nonstable age

distributions and short-term planning. Journal of Wildlife Management 70:649-

656.

Gaillard, J.-M., M. Festa-Bianchet, and N.G. Yoccoz. 1998. Population dynamics of large

herbivores: variable recruitment with constant adult survival. Trends in Ecology

and Evolution 13:58-63.

Gaillard, J.-M., M. Festa-Bianchet, N.G. Yoccoz, A. Loison, and C. Toïgo. 2000.

Temporal variation in fitness components and population dynamics of large

herbivores. Annual Review of Ecology and Systematics 31:367-393.

Gilpin, M.E., and M.E. Soule. 1986. Minimum viable populations: process of species

Page 38: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

24

extinctions. Pages 19-34 in M. E. Soulé, ed. Conservation biology: the science of

scarcity and diversity. Sinauer Associates, Inc., Sunderland, Massachusetts.

Greene L., T. Stephenson, and M. Hebblewhite. Short-term effects of wildfire on the

winter range of Sierra Nevada bighorn sheep. In Preparation.

Heppell, S.S. 1998. Application of life-history theory and population model analysis to

turtle conservation. Copeia 2:367-375.

IPCC. 2007. Synthesis report: summary for policy makers. Contributions of Working

Groups I, II, and III to the Fourth Assessment Report of the Intergovernmental

Panel on Climate Change. Cambridge University Press, Cambridge.

Johnson, H. E., L. S. Mills, J. Wehausen, and T. R. Stephenson. 2010. Population-

specific vital rate contributions influence management of an endangered ungulate.

Ecological Applications 20:1753-1765.

Johnson, H. E., L. S. Mills, J. Wehausen, and T. R. Stephenson. In Press. Combining

ground count, telemetry, and mark-resight data to infer population dynamics in an

endangered species. Journal of Applied Ecology.

Keller, L.F., and D.M. Waller. 2002. Inbreeding effects in wild populations. Trends in

Ecology and Evolution 17:230-241.

Lande, R., S. Engen, and B. Saether. 2003. Stochastic population dynamics in ecology

and conservation. Oxford University Press, New York, New York.

Laurance, W.F., and D.C. Useche. 2009. Environmental synergisms and extinctions of

tropical species. Conservation Biology 23:1427-1437.

Luikart G., P.R. England, D. Tallmon, S. Jordan, and P. Taberlet. 2003. The power and

Page 39: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

25

promise of population genomics: from genotyping to genome typing. Nature

Reviews Genetics 4:981-994.

Lyons, J.E., M.C. Runge, H.P. Laskowski, and W.L. Kendall. 2008. Monitoring in the

context of structured decision-making and adaptive management. Journal of

Wildlife Management 72:1683-1692.

Madsen, T., R. Shine, M. Olsson, and H. Wittzell. 1999. Restoration of an inbred adder

population. Nature 402:34-35.

Male, T.D., and M.J. Bean. 2005. Measuring progress in US endangered species

conservation. Ecology Letters 8:986-992.

Melbourne, B.A., and A. Hastings. 2008. Extinction risk depends strongly on factors

contributing to stochasticity. Nature 454:100-103.

Mills, L.S. 2007. Conservation of wildlife populations: demography, genetics and

management. Blackwell Publishing, Malden, Massachusetts.

Morris, W.F., and D.F. Doak. 2002. Quantitative conservation biology: theory and

practice of population viability. Sinauer Associates, Sunderland, Massachussetts.

Morris, W.F., P.L. Bloch, B.R. Hudgens, L.C. Moyle, and J.R. Stinchcombe. 2002.

Population viability analysis in endangered species recovery plans: past use and

future improvements. Ecological Applications 12:708-712.

Ouborg, N.J., P. Vergeer, and C. Mix. 2006. The rough edges of the conservation

genetics paradigm for plants. Journal of Ecology 94:1233-1248.

Rabinowitz, A. 1995. Helping a species go extinct: the Sumatran rhino in Borneo.

Conservation Biology 9:482-488.

Raithel, J.D., M.J. Kauffman, and D.H. Pletscher. 2007. Impact of spatial and temporal

Page 40: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

26

variation in calf survival on the growth of elk populations: Journal of Wildlife

Management 71:795-803.

Saccheri, I., M. Kuussaari, M. Kankare, P. Vikman, W. Fortelius, and I. Hanski. 1998.

Inbreeding and extinction in a butterfly metapopulation. Nature 392:491-494.

Sæther, B.E., and O. Bakke. 2000. Avian life history variation and contribution of

demographic traits to the population growth rate. Ecology 81:642-653.

Schaub, M., O. Gimenez, A. Sierro, and R. Arlettaz. 2007. Use of integrated modeling

to enhance estimates of population dynamics obtained from limited data.

Conservation Biology 21:945-955.

Silvertown, J., M. Franco, and E. Menges. 1996. Interpretation of elasticity matrices as an

aid to the management of plant populations for conservation. Conservation

Biology 10:591-597.

Tallmon, D.A., G. Luikart, and R.S. Waples. 2004. The alluring simplicity and complex

reality of genetic rescue. Trends in Ecology and Evolution 19:489-496.

Tear, T.H., J.M. Scott, P.H. Hayward, and B. Griffith. 1995. Recovery plans and the

Endangered Species Act: are criticisms supported by data? Conservation Biology

9:182-195.

U.S. Fish and Wildlife Service. 2007. Recovery Plan for the Sierra Nevada bighorn

sheep. Sacramento, California.

Vergeer, P., R. Rengelink, A. Copal, and N.J. Ouborg. 2003. The interacting effects of

genetic variation, habitat quality, and population size on performance of Succisa

pratensis. Journal of Ecology 91:18-26.

Wehausen, J.D. 1992. The role of precipitation and temperature in the winter range diet

Page 41: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

27

quality of mountain sheep of the Mount Baxter herd, Sierra Nevada. Proceedings

of the Biennial Symposium of the Northern Wild Sheep and Goat Council 8:279-

292.

Wehausen, J.D. 1996. Effects of mountain lion predation on bighorn sheep in the Sierra

Nevada and Granite Mountains of California. Wildlife Society Bulletin 24:471-

479.

Westemeier, R.L, J.D. Brawn, S.A. Simpson, T.L. Esker, R.W. Jansen, J.W. Walk, E.L.

Kershner, J.L. Bouzat, and K.N. Paige. 1998. Tracking the long-term decline and

recovery of an isolated population. Science 282:1695-1698.

Wilcove, D.S., M. McMillan, K.C. Winston. 1993. What exactly is an endangered

species? An analysis of the U.S. Endangered Species list: 1985-1991.

Conservation Biology 7:87-93.

Wilcove, D.S., D. Rothstein, J. Dubow, A. Phillips, and E. Losos. 1998. Quantifying

threats to imperiled species in the United States. Bioscience 48:607-615.

Williams, B. K. 2001. Uncertainty, learning, and the optimal management of wildlife.

Environmental and Ecological Statistics 8:269–288.

Page 42: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

28

Figure 1.1. Diagram of the extinction vortex, modified from Mills (2007).

Page 43: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

29

CHAPTER 2

POPULATION-SPECIFIC VITAL RATE CONTRIBUTIONS

INFLUENCE MANAGEMENT OF AN ENDANGERED UNGULATE

HEATHER E. JOHNSON, University of Montana, Wildlife Biology Program, College

of Forestry and Conservation, Missoula, MT 59812, USA

L. SCOTT MILLS, University of Montana, Wildlife Biology Program, College of

Forestry and Conservation, Missoula, MT 59812, USA

THOMAS R. STEPHENSON, Sierra Nevada Bighorn Sheep Recovery Program,

California Department of Fish and Game, 407 West Line Street, Bishop, CA

93514, USA

JOHN D. WEHAUSEN, White Mountain Research Station, University of California,

3000 East Line Street, Bishop, CA 93514, USA

ABSTRACT

To develop effective management strategies for the recovery of threatened and

endangered species, it is critical to identify those vital rates (survival and reproductive

parameters) responsible for poor population performance and those whose increase will

most efficiently change a population’s trajectory. In actual application, however,

approaches identifying key vital rates are often limited by inadequate demographic data,

by unrealistic assumptions of asymptotic population dynamics and of equal, infinitesimal

changes in mean vital rates. We evaluated the consequences of these limitations in an

analysis of vital rates most important in the dynamics of federally endangered Sierra

Page 44: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

30

Nevada bighorn sheep (Ovis canadensis sierrae). Based on data collected from 1980-

2007, we estimated vital rates in three isolated populations, accounting for sampling

error, variance, and co-variance. We used analytical sensitivity analysis, life-stage

simulation analysis, and a novel non-asymptotic simulation approach to (a) identify vital

rates that should be targeted for subspecies recovery; (b) assess vital rate patterns of

endangered bighorn sheep relative to other ungulate populations; (c) evaluate the

performance of asymptotic versus non-asymptotic models for meeting short-term

management objectives; and (d) simulate management scenarios for boosting bighorn

sheep population growth rates. We found wide spatial and temporal variation in bighorn

sheep vital rates, causing rates to vary in their importance to different populations. As a

result, Sierra Nevada bighorn sheep exhibited population-specific dynamics that did not

follow theoretical expectations or those observed in other ungulates. Our study suggests

that vital rate inferences from large, increasing or healthy populations may not be

applicable to those that are small, declining or endangered. We also found that while

asymptotic approaches were generally applicable to bighorn sheep conservation planning,

our non-asymptotic population models yielded unexpected results of importance to

managers. Finally, extreme differences in the dynamics of individual bighorn sheep

populations imply that effective management strategies for endangered species recovery

may often need to be population-specific.

INTRODUCTION

If deterministic or stochastic factors trigger successive decreases in key vital rates,

such as stage-specific survival and reproductive parameters, a population will decline,

potentially to extinction. To develop effective management strategies for the recovery of

Page 45: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

31

threatened and endangered species, it is critical to identify those vital rates responsible for

poor population performance and those whose increase will most efficiently change a

population’s trajectory (Morris & Doak 2002; Mills 2007). While the disproportionate

impact of different vital rates on population growth is well recognized in basic and

applied ecology (Crouse et al. 1987; Heppell et al. 1996; Caswell 2001; Gaillard et al.

2001), it is still often overlooked in endangered species recovery programs. In many

cases, detailed demographic data are unavailable, but even when they exist the

application of vital rate analyses in conservation planning is often not prioritized. As a

result, well-intended conservation programs have misdirected their efforts towards

increasing survival or reproductive parameters relatively inconsequential to population

recovery efforts (Heppell et al. 1996).

Given the lack of demographic data on many small and endangered populations, it

has been suggested that important vital rates identified in other populations of the same

species, or those from similar species, be used to guide conservation efforts. The logic

being that demographic trends among species with analogous life-history traits should be

comparable and thus, information on the importance of vital rates from well-studied

populations should be applicable to those for which there is little information (Silvertown

et al. 1996; Heppell 1998; Sæther & Bakke 2000). While the application of life-history

expectations to the management of small or declining populations seems intuitive, its

relevance has not been well evaluated. In fact, some long-term studies examining the

dynamics of declining populations have reported that the most influential vital rates do

not follow life-history expectations (Schmidt et al. 2005; Owen-Smith & Mason 2005).

As a result, it remains unclear whether demographic trends in endangered or declining

Page 46: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

32

populations do indeed mirror those from large or healthy populations, and if inferences

among populations are appropriate.

When demographic data on endangered species are available, the most popular

tools for assessing the relative significance of different vital rates are analytical

sensitivity and elasticity analyses (de Kroon et al. 2000; Heppell et al. 2000; Morris &

Doak 2002). These matrix-based approaches identify vital rates whose equal and

infinitesimal changes have the greatest effect on population growth. In usual application

these metrics rely on asymptotic properties of population matrices, assuming populations

have constant mean vital rates, have converged to stable-stage-distribution (SSD), and are

large enough to be unaffected by demographic stochasticity (although stochastic

sensitivities and elasticities can be calculated, see Tuljapurkar et al. 2003; Morris & Doak

2005).

Assumptions inherent in traditional analytical analyses - asymptotic properties

and equal, infinitesimal changes in vital rates - are limiting for most conservation

applications. First, many small populations deviate from SSD and are subject to high

demographic stochasticity (Bierzychudek 1999; Clutton-Brock & Coulson 2002;

Fefferman and Reed 2006), causing asymptotic approaches to potentially misguide

critical management efforts, particularly over short time periods (Fox & Gurevitch 2000;

Merrill et al. 2003; Yearsley 2004; Koons et al. 2005; Koons et al. 2006). The second

assumption of equal, infinitesimal changes in mean vital rates ignores the amount of

variation that realistically occurs in those rates (Mills et al. 1999; Wisdom et al. 2000;

Mills et al. 2001; Norris & McCulloch 2003). For example, in ungulate populations

Gaillard et al. (1998) concluded that adult female survival consistently had the highest

Page 47: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

33

elasticity (and thereby had the greatest infinitesimal effect on population growth), but had

inherently low variability, allowing little room for management to have an appreciable

effect. Meanwhile, juvenile survival had low elasticity, but wide variation that was

primarily responsible for changes in population size, and thus, the key vital rate to target

for management purposes (see also Citta & Mills 1999; Gaillard et al. 2000; Wisdom et

al. 2000; Raithel et al. 2007).

Recognition that the contribution of a vital rate to population growth largely

depends on its actual range of variation has lead to alternative methods of sensitivity

analyses, including life-stage-simulation analysis (LSA; Wisdom & Mills 1997; Wisdom

et al. 2000). This approach readily incorporates variation (and covariation) among vital

rates and allows investigators to simulate the effects of different management scenarios

on population trajectories. Constrained by a lack of data on initial stage distribution and

population size, most applications of LSA have relied on asymptotic properties of matrix

models (Biek et al. 2002; Norris & McCulloch 2003; Hoekman et al. 2006; Raithel et al.

2007), however, this method could easily be extended to non-asymptotic projections with

a specified initial stage vector and projection interval (Mills & Lindberg 2002).

In identifying vital rates driving the dynamics of small or endangered populations,

the most useful sensitivity analyses should be those incorporating non-asymptotic

dynamics and actual vital rate variation (present either in nature or under management).

This would require data on vital rate means, variances and covariances, estimates of

initial population size and stage distribution, and a projection interval of significance to

managers. A unique data set allowed us to perform such an analysis on federally

endangered Sierra Nevada bighorn sheep (Ovis canadensis sierrae; SNBS) and evaluate

Page 48: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

34

the influence of asymptotic assumptions on inferences about different vital rates. SNBS

have the most restricted range and the fewest number of individuals of any subspecies of

bighorn sheep in North America (U.S. Fish & Wildlife Service 2007). While currently

there are populations of SNBS in five geographic areas, we focus only on three in this

paper: Mono Basin, Wheeler Ridge (Wheeler), and Mount Langley (Langley). These

herds are of particular interest because their mean vital rates, variances and covariances

can be estimated directly from annual survey data, and their population sizes and stage

distributions are known. Using detailed demographic data on SNBS, we applied

analytical sensitivity analysis, traditional (asymptotic) LSA, and a novel non-asymptotic

extension of LSA to (a) identify vital rates that should be targeted for subspecies

recovery, (b) assess vital rate patterns of endangered SNBS relative to other ungulate

populations, (c) compare the performance of asymptotic vs. non-asymptotic models for

meeting short-term SNBS management objectives, and (d) simulate management

scenarios for boosting SNBS population performance.

STUDY AREA

The Sierra Nevada mountain range forms the eastern backbone of California and

is approximately 650 km long and ranges from 75 to 125 km wide (Hill 1975). This range

is an uplifted fault block with a steep eastern slope that has been sculpted by Pleistocene

glaciers that created U-shaped canyons, steep cirque headwalls, and prominent peaks

(Hill 1975). Historical and current distributions of SNBS include only the southern half

of the Sierra Nevada, where these geologic processes have created the highest mountains

and the most alpine habitat. SNBS spend summers in the alpine along the crest of the

Sierra Nevada and winters either in the alpine or at lower elevations typically east of the

Page 49: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

35

crest, inhabiting elevations ranging from 1,525 to >4000 m (U.S. Fish & Wildlife Service

2007). Climate in the Sierra Nevada is characterized by relatively dry conditions in

summer (May-Sept), with most of the annual precipitation received as snow in winter

(Nov-Apr), varying considerably by year. There is a strong rain shadow effect in

precipitation east of the Sierra crest resulting in open, xeric vegetation communities. Low

elevations (1,500-2,500 m) are characterized by Great Basin sagebrush-bitterbrush scrub;

mid-elevations (2,500-3,300 m) by pinyon-juniper woodland, sub-alpine meadows, and

forests; and high elevations (>3,300 m) by sparse alpine vegetation including occasional

meadows. Virtually all SNBS habitat is public land, managed primarily by Yosemite and

Sequoia-Kings Canyon National Parks, and Inyo and Sierra National Forests.

METHODS

Vital Rate Parameter Estimation

We evaluated the three SNBS populations for which extensive demographic data

have been collected: Mono Basin, Wheeler Ridge (Wheeler), and Mount Langley

(Langley). These herds were reintroduced in the late 1970’s and 1980’s (Bleich et al.

1990), with Mono Basin the northernmost population, Langley the southernmost, and

Wheeler in the central part of the range (Fig. 2.1). Because SNBS is a highly valued

endemic subspecies of California, annual surveys have been routinely conducted by

California Department of Fish and Game (CDFG), the National Park Service, and

independent biologists. All populations are known to be geographically isolated so that

their dynamics are independent, and data from these herds encompass a wide range of

spatial and temporal demographic variability (Fig. 2.2; See Appendix A for detailed

population histories). After being reintroduced, the Wheeler and Langley populations

Page 50: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

36

remained relatively stationary until 1995, when they decreased slightly, and since then

have dramatically increased. Consecutive annual surveys began in 1995 for Wheeler, and

1996 for Langley, so analyses of these populations are relevant to the period when these

herds increased in size. Meanwhile, the Mono Basin population quickly grew following

its reintroduction in 1986, and has subsequently declined (Fig. 2.2). Because annual

surveys have been conducted since 1986, data from Mono Basin are analyzed across all

years, and separately for the increasing and decreasing periods. Causes of such disparate

population trends are not fully understood but suspected to be driven by differences in

predation, habitat quality, and use of low elevation winter ranges (U.S. Fish & Wildlife

Service 2007).

During annual population surveys each herd unit was systematically hiked and

scanned by experienced observers for bighorn sheep by sex and stage class. Field efforts

focused specifically on counts of females and lambs, as they represent the reproductive

segment of the population. The annual lambing period for SNBS occurs primarily from

mid-April through mid-June, and females give birth to one offspring/year (Wehausen

1996). Surveys of the Wheeler population were conducted in late March or early April

just before new lambs were born (pre-birth pulse), while surveys in Mono Basin and

Langley were conducted in summer, just after new lambs were born (post-birth pulse).

Three stage classes were observed during both surveys types; however, the timing

of surveys resulted in distinct differences in the data collected that translate to different

parameterizations of population projection matrices. During surveys at Wheeler (pre-birth

pulse) observers counted the number of adult females (≥2.7 yrs; NA), two-year-old

females (~1.7–1.9 yrs; NT) and yearlings (~0.7-0.9 yrs; NYS). Surveys at Mono Basin and

Page 51: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

37

Langley (post-birth pulse) counted the number of adult females (≥2.1 yrs; NA), yearling

females (~1.1-1.3 yrs; NYL), and newborn lambs (~0.1-0.3 yrs; NL). While yearlings are

present in both survey types, we refer to individuals in this stage class as “short”

yearlings (NYS) in pre-birth pulse surveys, and “long” yearlings (NYL) in the post-birth

pulse surveys to acknowledge age differences of these animals observed in the field. All

stage classes were uniquely identifiable by distinct horn and body size differences.

Although “two-year-olds” (designated in pre-birth pulse surveys) are not typically

classified in bighorn sheep studies, because these animals were not quite two-years-old

(being approximately 1.7-1.9 yrs) this stage class could still be identified as their horns

had not yet experienced a second season of growth. Annual surveys obtained minimum

count data for each stage class, but due to intensive monitoring, repeated field efforts, and

very small, observable populations (for example, numbers of adult females ranged from

approximately 5 to 35 in any population in any year), annual counts were highly

successful at being near-complete censuses.

We used counts conducted during consecutive years to estimate annual population

vital rates. Given the available data, different vital rates were calculated for pre- and post-

birth pulse surveys. For the Wheeler population (sampled pre-birth pulse) we estimated

adult female survival (SA), two-year-old female survival (ST), and recruitment (RA). We

calculated adult female survival in year t as NA(t)/(NA(t-1) + NT(t-1)). We calculated two-

year-old survival as NT(t)/NYS(t-1) and assumed equal survival among males and females

because short yearlings at Wheeler were not consistently identified by sex in the field.

Recruitment for year t was calculated as NYS♀(t)/NA(t-1). We assumed that two-year-olds

did not produce offspring, as ultrasonography on 8 yearlings (captured from 2003-2009)

Page 52: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

38

found only one to be pregnant (CDFG unpublished data). Because yearlings were not

consistently distinguished by sex, yet small numbers of yearlings/year were subject to

high demographic stochasticity, we did not assume a 50:50 sex ratio. Instead, we

attempted to correct for known numbers of yearling females/adult female by using the

number of two-year-old males and females in year t+1 to back-calculate minimum

numbers of yearlings by sex in year t. Where yearling survival was <1 (i.e. not all

yearlings survived to be two-year-olds and thus, there were yearlings counted in year t

not accounted for as two-year-olds in t+1), we assigned a 50:50 sex ratio to the remaining

animals of unknown gender.

For Mono Basin and Langley, populations surveyed just after the lambing period,

we estimated adult female survival (SA), yearling female survival (SY), and fecundity

(FA). We calculated adult female survival as NA(t)/(NA(t-1) + NYL(t-1)). Due to extremely

small population sizes in Mono Basin, calculations of adult female survival exceeded 1.0

in three years when one (in 1996 and 2002) or two (in 2001) additional females were

observed in year t than those known to be alive in the previous year t-1; survival in these

cases was truncated at 1.0. While field surveys were highly successful at being near-

complete census counts, these calculations demonstrate error in the data that we account

for later. Yearling survival was calculated as NYL(t)/NL(t-1), assuming equal survival

among males and females since newborn lambs were not identified by sex. Fecundity was

estimated as the number of female lambs/adult females or NL♀(t)/NA(t), assuming that

only adult females successfully reproduced. Again, given the influence of demographic

stochasticity inherent with small sample sizes we used available data on the sex of

Page 53: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

39

yearlings in year t+1 to correct for known numbers of female lambs in year t, and

assumed a 50:50 sex ratio for lambs of unknown gender.

Raw vital rate estimates included both process variance, the true biological

variation in a rate due to spatial and temporal factors (including environmental and

demographic stochasticity), and sampling variance, arising from inherent uncertainty in

parameter estimation (Link & Nichols 1994). Because we were only interested in the

influence of process variance in vital rate parameters on SNBS population performance

(Mills & Lindberg 2002), we used the method of Kendall (1998) to remove sampling

error from our binary vital rate data. We used the program Kendall.m in MatLab (Morris

& Doak 2002) to search over 1,000 combinations of means and variances for each rate to

estimate corrected population-specific vital rate parameters. For Mono Basin, in addition

to estimating vital rates for the entire study period (hereafter referred to as Mono

BasinALL), we estimated vital rate parameters for the period the population was increasing

(Mono BasinINCREASING; pre-1995), decreasing (Mono BasinDECREASING; post-1995), and

for recent population trends (Mono BasinRECENT; post-1998; Fig. 2.2).

Asymptotic Analyses

Because of differences in the timing of population surveys, we modeled Wheeler

using a pre-birth pulse stage-based matrix model and Mono Basin and Langley using a

post-birth pulse matrix model (Fig. 2.3). Both matrices were female-based, with a one

year projection interval derived from vital rates on the three observable stage classes. A

primary difference between the matrices is the recruitment term (RA) in the pre-birth

model (the number of lambs that are born and survive their first year/adult female) and

the fecundity term (FA) in the post-birth pulse model (the number of lambs born/adult

Page 54: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

40

female). The other main difference is that the survival of two-year-olds (ST) is included in

the pre-birth model (survival from yearling to two-years-old) while the survival of

yearlings (SY) is included in the post-birth pulse model (survival from newborn lamb to

yearling). Given that SNBS are long-lived (≤20 yrs), we consider adult female survival in

both matrices (≥2.1 yrs in the pre-birth model and ≥2.7 yrs in the post-birth model) to be

equivalent.

We evaluated demographic trends for Wheeler, Langley, and Mono Basin across

all years data were collected, and for Mono BasinINCREASING, Mono BasinDECREASING, and

Mono BasinRECENT. Using mean vital rate estimates we calculated the deterministic

asymptotic population growth rate (λ) for each population and time period. We also

calculated analytical sensitivity and elasticity values for vital rates of each population

scenario. We evaluated differences between asymptotic SSD and current SNBS stage

distributions (from 2007 surveys) using a χ2 test and Keyfitz’s Δ, a measure of the

Euclidean distance between actual and expected population vectors (Caswell 2001).

To determine life-history parameters having the greatest impact on SNBS

performance, we next performed a conventional LSA (Wisdom et al. 2000) to identify

vital rate “importance” in terms of the amount of variation in λ explained by variation in

each vital rate. Specifically we generated 1,000 matrices from distributions specifying the

means, variances, and covariances of demographic rates. We then regressed asymptotic λ

from each matrix against each vital rate to measure the relative value of different rates in

determining λ. Vital rate values for each time step were drawn from beta probability

distributions (bounded between 0 and 1) using mean and variance estimates. We

conducted analyses separately for uncorrelated and correlated vital rates. Correlated vital

Page 55: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

41

rates were based on the estimated covariance structure from population data (Appendix

B).

Non-Asymptotic LSA

We extended LSA to non-asymptotic projections using field surveys from 2007 to

specify the initial number of individuals in each stage class. Initial population vectors

describing the number of lambs, yearlings, and adults for each population were 6, 4, and

34 for Wheeler; 9, 11, and 38 for Langley; and 4, 3, and 11 for Mono Basin. Each matrix

was projected for periods of 5 and 10 years, time periods of management interest for the

SNBS Recovery Program and the U.S. Fish and Wildlife Service.

In each simulation, the population vector for each year was multiplied by a

randomly drawn matrix, where vital rate values were generated from beta distributions

given the means and variances specific to each population. Because SNBS populations

were small, we also included demographic stochasticity into simulations, as incorporated

by Mills and Smouse (1994) for survival and reproduction. Over the course of each

simulation we tracked the total change in population size (ΔN) over the projection

interval and stochastic lambda (λs) calculated as (NTmax/ N0)1/Tmax

. For each model we ran

1,000 replicates and calculated average ΔN and λs across replicated simulations. Using

this approach, we then evaluated a series of scenarios for each SNBS population to

predict performance given 1) baseline or non-manipulated vital rate values, 2)

proportional one-at-a-time increases in each individual vital rate, and 3) potential

management activities.

Page 56: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

42

The baseline scenario for each population used non-manipulated vital rate values

for 5 and 10 year projections. For Mono Basin, we used vital rate values post-1998

(Mono BasinRECENT) as they are representative of recent trends (Fig. 2.2).

Next, we simulated a one-at-a-time 5% proportional increase in each mean vital

rate value while maintaining estimated variances around those rates. We did this to

compare vital rate assessments from asymptotic analyses to those simulated from non-

asymptotic models and determine whether management recommendations would be

identical. As with baseline projections these scenarios were simulated for 5 and 10 years.

All vital rates were individually increased for each population except for adult female

survival at Langley, where high baseline survival (97.7%) prevented a biologically

meaningful increase.

Finally, we simulated the potential impact of two management activities that have

been proposed for SNBS conservation: predator control and augmentations. Management

scenarios were simulated for 5 years, a time period congruent with recovery effort

assessment. While we include these simulations as an example of how demographic

models can be used to evaluate potential management scenarios it is important to

acknowledge that other actions may be equally or more effective at boosting SNBS

performance (U.S. Fish & Wildlife Service 2007).

Mountain lions are the primary predators of SNBS and have been implicated in

impeding their recovery (Wehausen 1996, U.S. Fish & Wildlife Service 2007).

Additionally, other studies quantifying the effect of mountain lions on bighorn sheep

have found that predation can cause substantial reductions in survival and recruitment

rates (Ross et al. 1997; Hayes et al. 2000; Kamler et al. 2002; Rominger et al. 2004;

Page 57: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

43

Festa-Bianchet et al. 2006; Rominger & Goldstein 2008). While most studies monitor

predation rates on only adult bighorn sheep, younger stage classes may be subject to even

higher rates of lion predation (Ross et a. 1997; Festa-Bianchet et al. 2006). While lion

removal is expected to benefit SNBS, the precise effects on vital rates are unknown.

Based on predation rates from other studies and cause-specific mortality data from

SNBS, we conservatively modeled the effects of predator control on vital rates in two

ways, 1) a proportional 5% increase across all rates, and 2) a 5% proportional increase in

SA but a 10% increase in vital rates of younger stage classes.

In addition to predator control, we modeled the impact of an augmentation on the

performance of each SNBS population. CDFG has considered augmenting populations

with 5 -10 adult females to stimulate population growth, realistic numbers given limited

source stock for translocations. We modeled such increases by altering the initial

population vector to reflect potential augmentations, while leaving vital rate values

unchanged.

RESULTS

Estimated Vital Rate Parameters

Vital rate values showed strong spatial and temporal variation (Table 2.1; Fig.

2.4), with Langley having the highest mean vital rates, followed by Wheeler, and then

Mono Basin. After sampling variance was separated from process variance, yearling and

two-year-old survival were generally the most variable vital rates across all populations

and years. However, when analyses for Mono Basin were conducted for different trend

periods, adult survival was the most variable vital rate for Mono BasinDECREASING and

Mon BasinRECENT (Table 2.1). Contrary to typical patterns of ungulate dynamics, adult

Page 58: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

44

survival had greater process variation than recruitment at Wheeler, and than fecundity at

Mono BasinALL, Mono BasinDECREASING and Mono BasinRECENT.

Asymptotic Analyses

Based on average vital rates, asymptotic λ was 1.09 for Wheeler, 1.18 for

Langley, 0.99 for Mono BasinALL, 1.07 for Mono BasinINCREASING, 0.96 for Mono

BasinDECREASING, and 1.02 for Mono BasinRECENT. None of the observed stage

distributions from 2007 field surveys were significantly different from SSD (all

populations χ2 < 0.05, df = 2, p > 0.97; also Keyfitz’s Δ ≤ 0.10 for all herds). Consistent

with studies of other ungulates and long-lived species, adult female survival had the

highest analytical sensitivity and elasticity values across all populations and time periods

(Table 2.1).

LSA results showed that the proportion of variation in λ attributable to each vital

rate differed across SNBS populations (Table 2.1). Langley and Mono BasinINCREASING

exhibited classic patterns of ungulate dynamics where younger stage classes were

responsible for most of the variation in λ (Fig. 2.5; Gaillard et al. 1998; Gaillard et al.

2000; Raithel et al. 2007). In Langley, recruitment explained the highest percentage of

variation in λ (74%), while for Mono BasinINCREASING yearling survival explained most of

the variation (63%). Conversely, adult survival was most strongly associated with λ for

Wheeler, Mono BasinALL, Mono BasinDECREASING, and Mono BasinRECENT explaining

>82% of the variation in these growth rates (Table 2.1; Fig. 2.5). When we incorporated

correlations among vital rates into LSA (Appendix B), rankings of the relative

importance of different rates were qualitatively the same but there were differences in the

amount of variance in λ explained (Table 2.1).

Page 59: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

45

Non-Asymptotic Analyses

Because our populations were close to SSD (previous section), our non-

asymptotic LSA simulations incorporating field-estimated population sizes and initial

stage structures, as well as demographic stochasticity and short time periods, largely

agreed with those from asymptotic LSA results. Predictions from both showed that

increases in adult female survival would have the greatest recovery benefit for Wheeler

and Mono Basin, while increases in vital rates of the younger stage classes would be most

beneficial for Langley (Table 2.2). When vital rates were simulated to individually

increase by the same proportional amount, adult survival had the greatest effect on

projected median population sizes in Wheeler and Mono Basin. At Langley, however,

while asymptotic simulations clearly predicted that fecundity would be most beneficial

for population performance, non-asymptotic simulations demonstrated that an increase in

either fecundity or yearling survival would result in essentially equivalent gains in

population size (Table 2.2). Thus, given simulation results, recovery efforts would almost

equally benefit from increases in either stage class, a potentially critical observation, as

certain rates may be easier to manage than others. Correlations among vital rates had no

qualitative effect on results when included in population simulations (Appendix C).

Simulations of potential management actions on SNBS populations suggest that

effective conservation activities are largely population-specific (Table 2.3). Given the

two scenarios we modeled, for Wheeler it appears that predator control would be more

successful than augmentations at increasing population size in the short-term. For Mono

Basin, effects of predator control and augmentations were similar, although the one time

addition of 10 adult females was predicted to have the greatest effect on the population

Page 60: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

46

(Table 2.3), a result largely driven by its small size. For Langley, impacts of either

management action appear to be similar. Because baseline vital rate values are already

high, there is not predicted to be much gain in size over the next 5 years from either

scenario, with only a 7-20% proportional increase in numbers over baseline projections.

DISCUSSION

Vital rate analyses elucidated findings relevant both for the conservation of SNBS

and for the general application of these approaches to the management of declining and

endangered populations. First, we found that SNBS vital rate values showed high spatial

and temporal variation, resulting in population-specific dynamics that did not always fit

general expectations from other ungulates, particularly during the period of population

decline. We also found that while asymptotic approaches were generally applicable to

SNBS conservation planning, our non-asymptotic models yielded non-intuitive results

that could be important for managers. Finally, we found that due to extreme differences

in the dynamics of individual populations, effective management strategies for

endangered species recovery may often need to be population-specific.

Vital rate parameters showed dramatic spatial and temporal variation (Table 2.1;

Fig. 2.4), as SNBS populations have experienced increasing and decreasing trajectories

both within herds (Mono Basin) and recently among herds (i.e. Langley vs. Mono Basin).

Also, populations with seemingly synchronous trajectories, such as Langley and Wheeler,

were shown to be driven by entirely different vital rates (fecundity at Langley and adult

survival at Wheeler). These differences suggest substantial variation in the spatial and

temporal factors determining SNBS demographic processes, but there is uncertainty

about the specific factors driving this variation. Differences in low elevation winter range

Page 61: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

47

habitat-use are suspected to influence SNBS demographic rates (U.S. Fish and Wildlife

Service 2007), particularly for the Mono Basin. While predation rates are not known over

the entire period of this study, cause-specific mortality data collected over the past five

years suggests that predation pressure varies among herds (CDFG unpublished data);

Langley and Mono Basin generally experience low lion predation, while Wheeler

experiences moderate predation. Impacts of disease and genetic diversity may also

differentially influence SNBS demographic rates, but the effects of these factors are

currently unknown.

The dominant paradigm for ungulates is that adult female survival has the highest

elasticity, but its low variation causes it to contribute relatively little to changes in the

population growth rate compared to juvenile survival, which has low elasticity but high

variation, making it the primary determinant of population change (Gaillard et al. 1998;

Gaillard et al. 2000; Gaillard & Yoccoz 2003; Raithel et al. 2007). In contrast to this

paradigm, we found that while elasticity results were consistent across all SNBS herds

and followed classic expectations, vital rates explaining the most variation in population

growth differed among herds and contradicted theoretical expectations. For example, in

Wheeler, Mono BasinALL, Mono BasinDECREASING, and Mono BasinRECENT variation was

higher in adult survival than recruitment or fecundity, contributing to the pattern that

adult survival explained the highest proportion of variation in population growth. Only

growth rates for Langley and Mono BasinINCREASING followed general ungulate life-

history expectations, driven by changes in fecundity and yearling survival, respectively.

To date, few ungulate studies have observed such variation in the importance of different

vital rates within or among populations (Albon et al. 2000; Coulson et al. 2005, Nilsen et

Page 62: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

48

al. 2009), and consequently, the implications of such variation for conservation and

management purposes has been likely overlooked.

Such shifts in the means and variances of key vital rates may be largely

responsible for declining and endangered populations. Owen-Smith and Mason (2005)

found that decreases in adult survival were responsible for African ungulate populations

that transitioned from stable trajectories to declining ones. That this pattern was contrary

to other studies of ungulate dynamics was attributed to the fact that most investigations

have been conducted in temperate zones, not tropical ones with a large suite of predators.

However, our temperate-region Mono Basin population provides similar evidence for

how changes in vital rate values may trigger a declining growth rate. Mono

BasinINCREASING was characterized by high adult survival (92%) with extremely low

process variability (0.0003), and a growth rate that was most closely associated with

survival of the widely varying yearling stage class. Mono BasinDECREASING, on the other

hand, was characterized by much lower mean adult survival (84%), with almost a 100

fold increase in process variation (0.0288), and a growth rate almost entirely determined

by adult survival. Pfister (1998) suggested that demographic rates were unlikely to be

both highly variable and have a large effect on the growth rate of a population. This

observation, however, may only be relevant to stable or increasing populations. In small,

declining or endangered populations it might be quite common for vital rates with the

greatest elasticity to also be highly variable and have a large impact on population change

(Wisdom et al. 2000, Schmidt et al. 2005, Nilsen et al. 2009). In fact, several studies on

long-lived species have associated population declines to decreases in adult survival, the

Page 63: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

49

rate expected to have the highest elasticity (Wehausen 1996, Flint et al. 2000, Rubin et al.

2002, Pistorius et al. 2004, Wittmer et al. 2005).

As anthropomorphic factors continue to reduce wildlife populations,

understanding the processes that govern the fate of small populations is becoming

increasingly urgent. Because data are often sparse for threatened and endangered species

it seems intuitive to apply results of vital rate analyses from healthy, well-studied species

or populations to those of conservation concern. Our results, however, illustrate the

potential danger of this approach. Based on studies of other ungulate populations, a

reasonable assumption would be to focus SNBS recovery efforts on increasing juvenile

survival, as this rate has been responsible for the dynamics of other healthy herds of large

herbivores. Data from SNBS suggest, however, that it is a decrease in adult survival that

is the primary driver of SNBS declines and that it should be the focus of monitoring and

management activities. Shifts in the means or variances of key vital rates, particularly as

they differ from life-history expectations may frequently result in endangered, small, or

declining populations. As a result, it may be necessary to conduct a detailed demographic

analysis of these populations to identify appropriate management targets.

Recent papers have stressed the importance of considering transient dynamics

with initial population vectors when making short-term predictions, as those based on

asymptotic properties can yield misleading results (Fox & Gurevitch 2000; Koons et al.

2005; Fefferman & Reed 2006; Caswell 2007). This is particularly important for

“slower” species, such as bighorn sheep, having longer life-spans and lower reproductive

potential (Koons et al. 2005). In spite of this, our non-asymptotic LSA approach,

incorporating demographic stochasticity, initial population sizes, and short management

Page 64: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

50

timeframes, conferred qualitatively similar results to asymptotic predictions. Close

agreement between the methods are likely because SNBS stage class distributions were

very similar to SSD, and thus short-term predictions were in close alignment with

asymptotic expectations. For populations that are far from SSD, however, our non-

asymptotic approach should yield more accurate short-term predictions, particularly for

populations of management and conservation concern. Such an approach would be

particularly valuable for making predictions about populations that have been recently

“bumped” from SSD, such as after a major perturbation or mortality event (i.e. a disease

episode) that differentially affects distinct life stages, or for newly reintroduced

populations having artificially skewed (and known) initial stage distributions.

While our non-asymptotic simulations agreed with asymptotic LSA results in

terms of identifying vital rates contributing most to the variation in population growth,

they also yielded some non-intuitive results relevant to management. In some cases, we

found that by targeting an entirely different rate than the one identified by asymptotic

LSA, gave essentially equivalent results over time periods of management interest.

Depending on the ability of management actions to manipulate individual vital rates, such

simulations could recognize equally viable recovery alternatives that would not be

apparent from asymptotic analyses alone. For example, asymptotic analyses found that

the growth rate at Langley was most strongly correlated with fecundity rates; however,

non-asymptotic models predicted that an increase in either fecundity or yearling survival

would yield virtually identical results in population performance over the time period of

management interest. While we recognize that the detailed demographic data needed to

conduct non-asymptotic analyses do not exist for most endangered species, we feel that

Page 65: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

51

when data are available it is important to incorporate such approaches into population

models to avoid simplifying assumptions. When such data are not available, however,

traditional LSA will still provide critical information for management.

To design successful conservation plans, managers must first know how different

actions will affect key vital rates and to what degree. The two management scenarios we

simulated, predator control (an increase in mean vital rate values) and augmentations (an

increase in number of adult females in the initial population vector), illustrated that

effective strategies appear to be largely population-specific. For example, from the two

scenarios we modeled it appears that predator control will be most effective for

stimulating the Wheeler population, while an augmentation may be most effective for a

short-term boost in performance at Mono Basin. Given the current growth rate of the

Langley herd, management actions are not predicted to have an appreciable impact, and

thus recovery efforts could be better invested elsewhere. While predator control and

augmentations are two management options currently being considered for SNBS

recovery, other options considered in the Recovery Plan include habitat enhancement,

genetic management, and disease prevention (U.S. Fish & Wildlife Service 2007).

Unfortunately we have too little information from field data or the literature to adequately

model the effects of those activities on SNBS vital rates.

Our models incorporated numerous factors that we assumed were important for

short-term SNBS dynamics such as environmental and demographic stochasticity,

correlations (positive and negative) among vital rates, realistic management timeframes,

and actual initial population vectors. We did not, however, include functional changes in

vital rate values with respect to population size, as would be expected with density

Page 66: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

52

dependence. Instead, we assumed that negative density dependence would not be an issue

for this endangered subspecies at the small population sizes and short time periods we

modeled (Beissinger & Westphal 1998). Additionally, while we simulated the numeric

response of augmentations on SNBS population dynamics, we did not account for

potential positive density dependence (Allee effects) on mean vital rate values. If either

negative or positive density dependence occurs in SNBS populations our predictions

about population change could be either over- or under-estimated, and any such process

variation would be falsely attributed to stochasticity. We also did not include a senescent

stage class in our demographic models, as these animals are not uniquely identifiable in

the field. In an analysis based on marked individuals, Nilsen et al. (2009) found that the

inclusion of a senescent stage class slightly decreased the contribution of adult survival to

population growth, but there were no qualitative differences. Because our estimates of

adult female survival and reproduction are based on near-complete census counts (which

include both prime-age and senescent individuals), we assume that the demographic

impacts of senescent animals are incorporated into our projection models. As with all

matrix model simulations, predicted results should be regarded on a relative, rather than

absolute, basis (Beissinger & Westphal 1998; Morris & Doak 2002; Reed et al. 2002).

In conclusion, we demonstrate that the relative contribution of different vital rates

to population growth may vary among populations of the same species, and within the

same geographic region, not following expectations from life history theory. As a result,

inferences about the importance of different rates from one species or population may not

be applicable to another, and could potentially misdirect critical resources if

inappropriately employed within conservation programs. Furthermore, endangered

Page 67: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

53

species recovery programs should be responsive to deviations between observed vital rate

values and those predicted from classic life-history expectations. Such departures may be

largely responsible for population declines and serve as important targets for monitoring

programs and management actions.

ACKNOWLEDGEMENTS

We would like to thank all the people involved in collecting demographic data on

SNBS including T. Blankinship, L. Bowermaster, L. Brown, M. Cahn, C. Chang, L.

Chow, K. Ellis, D. German, A. Feinberg, J. Fusaro, L. Greene, D. Jensen, M. Kiner, K.

Knox, P. Moore, R. Ramey II, C. Schroeder, T. Taylor, and S. Thompson. We are

grateful to D. German, M. Hebblewhite, M. Kauffman, M. Mitchell, D. Pletscher, J.

Millspaugh, J.-M. Gaillard and an anonymous reviewer for providing constructive

comments on this manuscript. We would also like to thank C. Hartway for her

programming expertise in MatLab and the California Department of Fish and Game for

supporting this work.

LITERATURE CITED

Albon, S.D., T.N. Coulson, D. Brown, F.E. Guinness, J.M. Pemberton, and T.H. Clutton-

Brock. 2000. Temporal changes in key factors and key age groups influencing the

population dynamics of female red deer: Journal of Animal Ecology 69:1099-

1110.

Beissinger, S., and M.I. Westphal. 1998. On the use of demographic models of

population viability in endangered species management: Journal of Wildlife

Management 62:821-841.

Biek, R., W.C. Funk, B.A. Maxell, and L.S. Mills. 2002. What is missing in amphibian

Page 68: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

54

decline research: insights from ecological sensitivity analysis: Conservation

Biology 16:728-734.

Bierzychudek, P. 1999. Looking backwards: assessing the projections of a transition

matrix model: Ecological Applications 9:1278-1287.

Bleich, V.C., J.D. Wehausen, K.R. Jones, and R.A. Weaver. 1990. Status of bighorn

sheep in California, 1989 and translocations from 1971 through 1989: Desert

Bighorn Council Transactions 34:24-26.

Caswell, H. 2001. Matrix population models: Construction, analysis and interpretation.

Second Edition. Sinauer Associates, Sunderland, Massachusetts.

Caswell, H. 2007. Sensitivity analysis of transient population dynamics: Ecology

Letters 10:1-15.

Citta, J.J., and L.S. Mills. 1999. What do demographic sensitivity analyses tell us about

controlling brown-headed cowbirds? Studies in Avian Biology 18:121-134.

Clutton-Brock, T.H., and T. Coulson. 2002. Comparative ungulate dynamics: the devil is

in the detail: Philosophical Transactions of the Royal Society of London Series B

357:1285-1298.

Coulson, T., J.-M. Gaillard, and M. Festa-Bianchet. 2005. Decomposing the variation in

population growth into contributions from multiple demographic rates: Journal of

Animal Ecology 74:789-801.

Crouse, D.T., L.B. Crowder, and H. Caswell. 1987. A stage-based population model for

loggerhead sea-turtles and implications for conservation: Ecology 68:1412-1423.

de Kroon, H., J.van Groenendael, and J. Ehrlén. 2000. Elasticities: A review of methods

and model limitations: Ecology 81:607-618.

Page 69: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

55

Fefferman, N.H., and J.M. Reed. 2006. A vital rate sensitivity analysis for nonstable age

distributions and short-term planning: Journal of Wildlife Management 70:649-

656.

Festa-Bianchet, M., T. Coulson, J-M. Gaillard, J.T. Hogg, and F. Pelletier. 2006.

Stochastic predation events and population persistence in bighorn sheep:

Proceedings of the Royal Society of London Series B 273:1537-1543.

Flint, P.L., M.R. Petersen, C.P. Dau, J.E. Hines, and J.D. Nichols. 2000. Annual survival

and site fidelity of Steller’s eiders molting along the Alaska Peninsula. Journal of

Wildlife Management 64:261-268.

Fox, G.A., and J. Gurevitch. 2000. Population numbers count: Tools for near-term

demographic analysis: American Naturalist 156:242-256.

Gaillard, J.-M., M. Festa-Bianchet, and N.G. Yoccoz. 2001. Not all sheep are equal:

Science 292:1499-1500.

Gaillard, J.-M. and N.G. Yoccoz. 2003. Temporal variation in survival of mammals: A

case of environmental canalization? Ecology 84:3294-3306.

Gaillard, J.-M., M. Festa-Bianchet, N.G. Yoccoz, A. Loison, and C. Toïgo. 2000.

Temporal variation in fitness components and population dynamics of large

herbivores: Annual Review of Ecology and Systematics 31:367-393.

Gaillard, J.-M., M. Festa-Bianchet, and N.G. Yoccoz. 1998. Population dynamics of large

herbivores: Variable recruitment with constant adult survival: Trends in Ecology

and Evolution 13:58-63.

Hayes, C.L., E.S. Rubin, M.C. Jorgensen, R.A. Botta, and W.M. Boyce. 2000. Mountain

Page 70: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

56

lion predation of bighorn sheep in the Peninsular ranges, California: Journal of

Wildlife Management 64:954-959.

Heppell, S.S. 1998. Application of life-history theory and population model analysis to

turtle conservation: Copeia 2:367-375.

Heppell, S.S., L.B. Crowder, and D.T. Crouse. 1996. Models to evaluate headstarting

as a management tool for long-lived turtles: Ecological Applications 6:556-565.

Heppell, S., C. Pfister, and H. de Kroon. 2000. Elasticity analysis in population

biology: Methods and applications: Ecology 81:605-606.

Hill, M. 1975. Geology of the Sierra Nevada. The University of California Press,

Berkeley, California.

Hoekman, S.T., T.S. Gabor, M.J. Petrie, R. Maher, H.R. Murkin, and M.S. Lindberg.

2006. Population dynamics of mallards breeding in agricultural environments in

eastern Canada: Journal of Wildlife Management 70:121-128.

Kamler, J.F., R.M. Lee, J.C. deVos, Jr., W.B. Ballard, and H.A. Whitlaw. 2002.

Survival and cougar predation of translocated bighorn sheep in Arizona: Journal

of Wildlife Management 66:1267-1272.

Kendall, B.E. 1998. Estimating the magnitude of environmental stochasticity in

survivorship data: Ecological Applications 8:184-193.

Koons, D.N., J.B. Grand, B. Zinner, and R.F. Rockwell. 2005. Transient population

dynamics: Relations to life history and initial population size: Ecological

Modelling 185:283-297.

Koons, D.N., R.F. Rockwell, and J.B. Grand. 2006. Population Momentum: Implications

for wildlife management: Journal of Wildlife Management 70:19-26.

Page 71: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

57

Link, W.A., and J.D. Nichols. 1994. On the importance of sampling variance to

investigations of temporal variation in animal population size: Oikos 69:539-544.

Merrill, J.A., E.G. Cooch, and P.D. Curtis. 2003. Time to reduction: Factors influencing

management efficacy in sterilizing overabundant white-tailed deer: Journal of

Wildlife Management 67:267-279.

Mills, L.S. 2007. Conservation of wildlife populations: demography, genetics and

management. Blackwell Publishing, Malden, Massachusetts.

Mills, L.S., D.F. Doak, and M.J. Wisdom. 1999. Reliability of conservation actions based

on elasticity analysis of matrix models: Conservation Biology 13:815-829.

Mills, L.S., D.F. Doak, and M.J. Wisdom. 2001. Elasticity analysis for conservation

decision making: Reply to Ehrlén et al.: Conservation Biology 15:281-283.

Mills, L.S., and M.S. Lindberg. 2002. Sensitivity analysis to evaluate the consequences of

conservation actions. Pages 338-366 in S.R. Beissinger and D.R. McCullough,

editors. Population Viability Analysis. University of Chicago Press, Chicago,

Illinois.

Mills, L.S., and P.E. Smouse. 1994. Demographic consequences of inbreeding in remnant

populations: American Naturalist 144:412-431.

Morris, W.F., and D.F. Doak. 2002. Quantitative conservation biology: Theory and

practice of population viability. Sinauer Associates, Sunderland, Massachussetts.

Morris, W.F., and D.F. Doak. 2005. How general are the determinants of the stochastic

population growth rate across nearby sites? Ecological Monographs 75:119-137.

Nilsen, E.B., J.-M. Gaillard, R. Andersen, J. Odden, D. Delorme, G. Van Laere, and

Page 72: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

58

J.D.C. Linnell. 2009. A slow life in hell or a fast life in heaven: demographic

analyses of contrasting roe deer populations. Journal of Animal Ecology 78:585-

594.

Norris, K., and N. McCulloch. 2003. Demographic models and the management of

endangered species: A case study of the critically endangered Seychelles magpie

robin: Journal of Applied Ecology 40:890-899.

Owen-Smith, N., and D.R. Mason. 2005. Comparative changes in adult vs. juvenile

survival affecting population trends of African ungulates: Journal of Animal

Ecology 74:762-773.

Pfister, C.A. 1998. Patterns of variance in stage-structured populations: Evolutionary

predictions and ecological implications: Proceedings of the National Academy of

Sciences of the United States of America 95:213-218.

Pistorius, P.A., M.N. Bester, M.N. Lewis, F.E. Taylor, C. Campagna, and S.P. Kirkman.

2004. Adult female survival, population trend, and the implications of early

primiparity in a capital breeder, the southern elephant seal (Mirounga leonina).

Journal of Zoology 263:107-119.

Raithel, J.D., M.J. Kauffman, and D.H. Pletscher. 2007. Impact of spatial and temporal

variation in calf survival on the growth of elk populations: Journal of Wildlife

Management 71:795-803.

Reed, J.M., L.S. Mills, J.B. Dunning, Jr., E.S. Menges, K.S. McKelvey, R. Frye, S.R.

Beissinger, M.C. Anstett, and P. Miller. 2002. Emerging Issues in population

viability analysis: Conservation Biology 16:7-19.

Rominger, E.M., and E.J. Goldstein. 2008. Evaluation of an 8-year mountain lion

Page 73: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

59

removal management action on endangered desert bighorn sheep recovery.

Wildlife Management Division Report. New Mexico Department of Game and

Fish, Santa Fe, New Mexico.

Rominger, E.M., H.A. Whitlaw, D.L. Weybright, W.C. Dunn, and W.B. Ballard. 2004.

The influence of mountain lion predation on bighorn sheep translocations: Journal

of Wildlife Management 68:993-999.

Ross, P.I., M.G. Jalkotzy, and M. Festa-Bianchet. 1997. Cougar predation on bighorn

sheep in southwestern Alberta during winter: Canadian Journal of Zoology

75:771-775.

Rubin, E.S., W.M. Boyce, and E.P. Caswell-Chen. 2002. Modeling demographic

processes in an endangered population of bighorn sheep. Journal of Wildlife

Management 66:796-810.

Sæther, B.E., and O. Bakke. 2000. Avian life history variation and contribution of

demographic traits to the population growth rate: Ecology 81:642-653.

Schmidt, B.R., R. Feldmann, and M. Schaub. 2005. Demographic processes underlying

population growth and decline in Salamandra salamadra: Conservation Biology

19:1149-1156.

Silvertown, J., M. Franco, and E. Menges. 1996. Interpretation of elasticity matrices as an

aid to the management of plant populations for conservation: Conservation

Biology 10:591-597.

Tuljapurkar, S., C.C. Horvitz, and J.B. Pascarella. 2003. The many growth rates and

elasticities of populations in random environments: American Naturalist 162:489-

502.

Page 74: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

60

U.S. Fish and Wildlife Service. 2007. Recovery Plan for the Sierra Nevada bighorn

sheep. Sacramento, California.

Wehausen, J.D. 1996. Effects of mountain lion predation on bighorn sheep in the Sierra

Nevada and Granite Mountains of California: Wildlife Society Bulletin 24:471-

479.

Wittmer, H.U., B.N. McLellan, D.R. Seip, J.A. Young, T.A. Kinley, G.S. Watts, and D.

Hamilton. 2005. Population dynamics of the endangered mountain ecotype of

woodland caribou (Rangifer tarandus caribou) in British Columbia, Canada.

Canadian Journal of Zoology 83:407-418.

Wisdom, M.J., L.S. Mills, and D.F. Doak. 2000. Life stage simulation analysis:

Estimating vital-rate effects on population growth for conservation: Ecology

81:628-641.

Wisdom, M.J., and L.S. Mills. 1997. Sensitivity analysis to guide population recovery:

Prairie chickens as an example: Journal of Wildlife Management 61:302-312.

Yearsley, J.M. 2004. Transient population dynamics and short-term sensitivity analysis of

matrix population models: Ecological Modelling 177: 245-258.

Page 75: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

61

Table 2.1. Female Sierra Nevada bighorn sheep parameter estimates used in vital rate analyses, including the number of years data

were collected (n), and vital rate means estimated directly from survey data (estimated) and corrected with a maximum likelihood

approach to account for sampling variance (corrected). Also provided for each vital rate are annual ranges, variances, sensitivities

(Sens), elasticities (Elast), and the proportion of variation in the population growth rate explained (r2 of λ) given uncorrelated (Uncorr)

and correlated (Corr) vital rates.

Subpopulation n

Estimated

Mean

Corrected

Mean Min Max

Total

Variance

Process

Variance

Sens Elast

r2 of λ

Uncorr

r2 of λ

Corr

WHEELER

All Years

recruitment 13 0.3225 0.3126 0.2268 0.4254 0.0216 0.0055 0.4242 0.1211 0.1447 0.2893

2-yr survival 13 0.7561 0.7295 0.4444 1.0000 0.0355 0.0138 0.1818 0.1211 0.0678 0.5872

adult survival 13 0.9168 0.9197 0.6923 1.0000 0.0097 0.0083 0.9019 0.7577 0.8243 0.8740

LANGLEY

All Years

fecundity 11 0.3409 0.3311 0.1670 0.5450 0.0354 0.0068 0.4495 0.1265 0.7408 0.7506

Page 76: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

62

yearling survival 9 0.8986 0.8722 0.5556 1.0000 0.0243 0.0115 0.1706 0.1265 0.1759 0.0870

adult survival 9 0.9735 0.9772 0.9000 1.0000 0.0022 0.0001 1.0516 0.8735 0.0579 0.0002

MONO BASIN

All Years

fecundity 22 0.3048 0.2934 0.0556 0.5625 0.0126 0.0003 0.3558 0.1054 0.0051 0.1190

yearling survival 21 0.6115 0.6006 0.1000 1.0000 0.0461 0.0339 0.1738 0.1054 0.0350 0.3607

adult survival 17 0.8625 0.8583 0.4286 1.0000 0.0276 0.0189 1.0325 0.8946 0.9459 0.9651

Increasing (Pre-1995)

fecundity 8 0.2930 0.2716 0.1857 0.4444 0.0054 0.0003 0.4275 0.1086 0.0383 0.5134

yearling survival 7 0.7223 0.7364 0.8182 1.0000 0.0343 0.0217 0.1577 0.1086 0.6298 0.7489

adult survival 5 0.9019 0.9207 0.8182 1.0000 0.0066 0.0003 1.0350 0.8914 0.3373 0.4791

Decreasing (Post-1995)

fecundity 14 0.3116 0.3172 0.1333 0.5625 0.0172 0.0003 0.3072 0.1013 0.0059 0.0945

yearling survival 14 0.5562 0.5055 0.1000 1.0000 0.0452 0.0192 0.1927 0.1013 0.0170 0.3617

adult survival 12 0.8460 0.8395 0.4286 1.0000 0.0367 0.0288 1.0295 0.8987 0.9764 0.9835

Page 77: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

63

Recent (Post-1998)

fecundity 9 0.3315 0.3360 0.2000 0.4000 0.0073 0.0003 0.3632 0.1200 0.0000 0.0712

yearling survival 8 0.6769 0.6740 0.5000 0.8000 0.0075 0.0003 0.1811 0.1200 0.0012 0.0158

adult survival 8 0.8647 0.8563 0.5556 1.0000 0.0242 0.0110 1.0450 0.8800 0.9964 0.9963

Page 78: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

64

Table 2.2. Median predicted sizes (NTmax) and stochastic growth rates (λs) of Sierra

Nevada bighorn sheep populations projected over 5 and 10 years. Initial population sizes

and stage distributions were parameterized from 2007 field surveys.

Population Scenario

Projected

Yrs

Median

NTmax

% ↑ from

baseline λs Var λs

Wheeler

Baseline 5 72 - 1.0963 0.0024

10 110 - 1.0938 0.0011

Increase RA by 5% 5 75 4.17 1.1081 0.0022

10 120 9.09 1.1026 0.0010

Increase SY by 5% 5 73 1.39 1.1049 0.0022

10 117 6.36 1.1004 0.0011

Increase SA by 5% 5 88 22.22 1.1384 0.0026

10 167 51.82 1.1340 0.0013

Langley

Baseline 5 130 - 1.1765 0.0005

10 295 - 1.1760 0.0002

Increase FA by 5% 5 135 3.85 1.1841 0.0005

10 312 5.76 1.1832 0.0003

Increase SY by 5% 5 136 4.62 1.1845 0.0005

10 313 6.10 1.1834 0.0002

Increase SA by 5% 5 N/A* N/A* N/A* N/A*

10 N/A* N/A* N/A* N/A*

Page 79: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

65

Mono Basin - Post 1998

Baseline 5 19 - 1.0006 0.0056

10 20 - 1.0046 0.0035

Increase RA by 5% 5 20 5.26 1.0113 0.0053

10 21 5.00 1.0085 0.0031

Increase SY by 5% 5 20 5.26 1.0100 0.0054

10 21 5.00 1.0099 0.0030

Increase SA by 5% 5 24 26.32 1.0486 0.0053

10 31 55.00 1.0461 0.0027

* An increase in adult female survival for Langley was not modeled since the baseline

mean value was already so high (97.7%).

Page 80: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

66

Table 2.3. Predicted median size (NTmax) and growth rate (λs) of female Sierra Nevada bighorn sheep populations given hypothetical

management scenarios. Each scenario prediction is compared to baseline predictions. Predator control was modeled by first simulating

a 5% increase in all vital rates for each population. For Wheeler increases in recruitment (RA), two-year-old female survival (ST), and

adult survival (SA) were simulated, while for Langley and Mono Basin increases in fecundity (FA), yearling survival (SY), and adult

survival (SA) were simulated. In a second predator control scenario, we modeled an increase in RA and ST or FA and SY by 10% and an

increase in SA by 5%. We also simulated population effects of a one-time augmentation of 5 or 10 adult females into each population.

All scenarios were projected for 5 years.

Population Potential Effect of Management Median NTmax λs Var λs % ↑ from baseline

Predator Control

Wheeler Increase RA, SY & SA by 5% 94 1.1540 0.0022 30.56

Increase RA & SY by 10%, SA by 5% 100 1.1669 0.0025 38.89

Langley Increase FA & SY by 5% 140 1.1921 0.0005 7.69

Increase FA & SY by 10% 149 1.2077 0.0006 14.62

Mono Basin Increase FA, SY & SA by 5% 25 1.0603 0.0053 31.58

Increase FA & SY by 10%, SA by 5% 27 1.0700 0.0055 42.11

Page 81: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

67

Augmentation

Wheeler Augment 5 adult females 80 1.0985 0.0022 11.11

Augment 10 adult females 90 1.1000 0.0024 25.00

Langley Augment 5 adult females 143 1.1784 0.0006 10.00

Augment 10 adult females 156 1.1800 0.0005 20.00

Mono Basin Augment 5 adult females 25 1.0067 0.0033 31.58

Augment 10 adult females 31 1.0335 0.0031 63.16

Page 82: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

68

Figure 2.1. Location of Mono Basin, Wheeler, and Langley Sierra Nevada bighorn sheep

populations, CA.

Page 83: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

69

Figure 2.2. Number of adult females in the Wheeler, Langley and Mono Basin

populations of Sierra Nevada bighorn sheep, 1980-2007.

Page 84: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

70

Figure 2.3. Pre- and post-birth pulse matrix models used to simulate female Sierra

Nevada bighorn sheep population dynamics. Vital Rates in the pre-birth pulse model are

recruitment (RA), two-yr-old female survival (ST), and adult female survival (SA). Vital

rates in the post-birth pulse model are fecundity (FA), yearling female survival (SY), and

adult female survival (SA).

Pre-Birth Pulse Matrix Model (Wheeler)

AA

T

A

SS

S

R

0

00

00

Post-Birth Pulse Matrix Model (Langley and Mono Basin)

AA

Y

AA

SS

S

SF

0

00

*00

Page 85: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

71

Figure 2.4. Annual mean vital rates for the Mono Basin, Langley, Wheeler populations of

Sierra Nevada bighorn sheep, 1985-2007. Due to the timing of field surveys at Wheeler,

adult female survival is the only vital rate comparable to the other populations.

Page 86: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

72

Figure 2.5. Analytical elasticities and coefficients of determination (r2) for Sierra Nevada

bighorn sheep fecundity, yearling survival (yrl surv), and adult survival (adt surv) rates in

the Mono Basin and Langley populations. Values are shown for years the Mono Basin

population was increasing versus decreasing.

Page 87: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

73

APPENDIX A. History and data collection of Sierra Nevada bighorn sheep populations.

Mono Basin was initially reintroduced with 11 adult females in 1986 and

augmented with an additional 8 females in 1988. Between 1991 and 1994 male and

female yearlings were not distinguished during surveys and, as a result, adult female

survival was not calculated between 1992 and 1995. We estimated adult survival for all

other years between 1987 and 2007, accounting for the augmentation in 1988. Yearling

survival was estimated from 1987-2007, and recruitment from 1986-2007. After being

reintroduced the population initially grew to 35 adult females in 1993 but declined to 5

adult females in 1999. Since 1999 it has stabilized with gradual population gains (Fig.

2.2).

Wheeler was reintroduced in 1980 with 4 adult females. An additional 8 females

were added to the population in 1981 and 3 more in 1987. In 2005, 5 adult females were

removed from the herd for translocations. All additions and removals were accounted for

in survival rate calculations. Annual surveys were conducted between 1980 and 2007

except for the following years; 1981-82, 1985-86, 1988-1991, and 1993-94. Adult female

survival, yearling survival and recruitment were estimated between all years of

consecutive surveys. Since 1995, when consecutive annual surveys have been routinely

conducted at Wheeler, the population has grown from 6 adult females to 38 (Fig. 2.2).

The Langley herd was reintroduced in 1980 with 6 adult females and 1 female

lamb. An augmentation in 1982 added an additional 5 adult females and 1 female lamb.

Consecutive population surveys were conducted between 1996 and 2007, with the

exception of 2005. As a result, adult and yearling survival estimates were available for

Page 88: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

74

1997-2004, and 2006-07, and recruitment data for all years except 2005. The Langley

herd has been increasing since 1996 (Fig. 2.2), from 6 adult females to >35.

Page 89: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

75

APPENDIX B. Correlation matrices for Sierra Nevada bighorn sheep vital rates.

Population

Recruitment/

Fecundity

Two-year-old/

Yearling Survival

Adult

Survival

Wheeler Ridge

Recruitment 1.0

Two-year-old Survival 0.4415 1.0

Adult Survival 0.2616 0.6955 1.0

Mount Langley

Fecundity 1.0

Yearling Survival -0.1266 1.0

Adult Survival -0.1723 -0.2633 1.0

Mono Basin All Yrs

Fecundity 1.0

Yearling Survival 0.2823 1.0

Adult Survival 0.2836 0.4449 1.0

Mono Basin Increasing Yrs

Fecundity 1.0

Yearling Survival 0.5796 1.0

Adult Survival 0.3395 0.2716 1.0

Mono Basin Decreasing Yrs

Fecundity 1.0

Yearling Survival 0.3471 1.0

Adult Survival 0.2542 0.4965 1.0

Page 90: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

76

APPENDIX C. Median predicted sizes (NTmax) and stochastic growth rates (λs) of Sierra

Nevada bighorn sheep populations over 5 and 10 years given correlated vital rates,

demographic stochasticity, and initial sizes and stage distributions from 2007 field

surveys. No increase in adult female survival for Langley since the mean value is 97.7%.

Population Scenario

Projected

Yrs

Median

NTmax

% ↑ from

baseline λs Var λs

Wheeler

Baseline 5 72 - 1.0924 0.0034

10 110 - 1.0910 0.0017

Increase RA by 5% 5 74 2.78 1.1047 0.0033

10 119 8.18 1.1007 0.0016

Increase ST by 5% 5 74 2.78 1.1020 0.0033

10 115 4.55 1.0982 0.0016

Increase SA by 5% 5 89 23.61 1.1380 0.0031

10 165 50.00 1.1337 0.0017

Langley

Baseline 5 131 - 1.1765 0.0005

10 293 - 1.1755 0.0002

Increase FA by 5% 5 138 5.34 1.1888 0.0005

10 325 10.92 1.1878 0.0002

Increase SY by 5% 5 136 3.82 1.1856 0.0005

10 316 7.85 1.1839 0.0002

Increase SA by 5% 5 N/A* N/A* N/A* N/A*

Page 91: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

77

10 N/A* N/A* N/A* N/A*

Mono Basin - Post

1998

Baseline 5 19 - 1.0037 0.0040

10 20 - 1.0036 0.0029

Increase FA by 5% 5 20 5.26 1.0116 0.0037

10 21 5.00 1.0087 0.0025

Increase SY by 5% 5 20 5.26 1.0083 0.0038

10 22 10.00 1.0133 0.0019

Increase SA by 5% 5 25 31.58 1.0475 0.0031

10 33 65.00 1.0514 0.0018

Page 92: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

78

CHAPTER 3

COMBINING COUNT, TELEMETRY, AND MARK-RESIGHT DATA TO

INFER POPULATION DYNAMICS IN AN ENDANGERED SPECIES

Heather E. Johnson, University of Montana, Wildlife Biology Program, College of

Forestry and Conservation, Missoula, MT 59812, USA.

L. Scott Mills, University of Montana, Wildlife Biology Program, College of Forestry

and Conservation, Missoula, MT 59812, USA.

John D. Wehausen, White Mountain Research Station, University of California, 3000 E.

Line Street, Bishop, CA 93514, USA.

Thomas R. Stephenson, Sierra Nevada Bighorn Sheep Recovery Program, California

Department of Fish and Game, 407 W. Line Street, Bishop, CA 93514, USA.

SUMMARY

1. To successfully manipulate populations for management and conservation

purposes, managers must be able to track changes in demographic rates and

determine the factors driving spatial and temporal variation in those rates. For

populations of management concern, however, data deficiencies frequently limit

the use of traditional statistical methods for such analyses. Long-term

demographic data are often piecemeal, having small sample sizes, inconsistent

methodologies, intermittent data, and information on only a subset of important

parameters and covariates.

Page 93: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

79

2. We evaluated the effectiveness of Bayesian state-space models for meeting these

data limitations in elucidating dynamics of federally endangered Sierra Nevada

bighorn sheep, Ovis canadensis sierrae. We combined ground count, telemetry,

and mark-resight data to: 1) estimate demographic parameters in 3 populations

(including stage-specific abundances and vital rates); and 2) determine whether

density, summer precipitation, or winter severity were driving variation in key

demographic rates.

3. Models combining all existing data types increased the precision and accuracy in

parameter estimates and fit covariates to vital rates driving population

performance. They also provided estimates for all years of interest (including

years in which field data were not collected) and standardized the error structure

across data types.

4. Demographic rates indicated that recovery efforts should focus on increasing

adult and yearling survival in the smallest bighorn sheep population. In evaluating

covariates we found evidence of negative density dependence in the larger herds,

but a trend of positive density dependence in the smallest herd suggesting that an

augmentation may be needed to boost performance. We also found that vital rates

in all populations were positively associated with summer precipitation, but that

winter severity only had a negative effect on the smallest herd, the herd most

strongly impacted by environmental stochasticity.

5. Synthesis and applications. For populations with piecemeal data, a problem

common to both endangered and harvested species, obtaining precise

demographic parameter estimates is one of the greatest challenges in detecting

Page 94: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

80

population trends, diagnosing the causes of decline, and directing management.

Data on Sierra Nevada bighorn sheep provide an example of the application of

Bayesian state-space models for combining all existing data to meet these

objectives and better inform important management and conservation decisions.

INTRODUCTION

To successfully manipulate populations for management and conservation

purposes, managers must be able to track changes in demographic parameters, identify

vital rates (survival and reproductive rates) having the greatest influence on population

growth, and determine the factors driving spatial and temporal variation in those key rates

(Franklin et al. 2000; Morris & Doak 2002; Bakker et al. 2009). Unfortunately, data

deficiencies prohibit these critical analyses for many populations of management interest

(Tear et al. 1995; Fieberg & Ellner 2001; Morris et al. 2002). Demographic data are often

piecemeal, having small sample sizes, inconsistent methodologies, intermittent data

collection, and information on only a subset of important parameters and covariates.

Traditional statistical approaches are limited in their ability to analyze such demographic

data, and as a result, critical management decisions are routinely made with limited

quantitative analysis.

Bayesian state-space models provide a powerful statistical tool for evaluating the

dynamics of populations with messy or incomplete datasets. These models can account

for multiple data types, small sample sizes, and missing data to estimate key demographic

parameters and simultaneously fit covariates to those parameters (Brooks, King &

Morgan 2004; Goodman 2004; Schaub et al. 2007; King et al. 2008). As such, they

effectively “integrate” all available demographic data into a single, comprehensive model

Page 95: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

81

that can describe the behaviour of a population while standardizing the error structure

across different data types (Besbeas et al. 2002; Brooks, King & Morgan 2004; Goodman

2004). Another benefit of these models is that they are highly mechanistic, explicitly

linking variation in population size to changes in stage-specific vital rates and covariate

values. While this method holds tremendous potential for combining a wide range of

demographic data types, its application to wildlife populations has been limited largely to

merging ground surveys with capture-recapture data (Besbeas et al. 2002; Brooks, King

& Morgan 2004; Schaub et al. 2007; Véran & Lebreton 2008).

We used the Bayesian state-space approach to evaluate the dynamics of federally

endangered Sierra Nevada bighorn sheep, Ovis canadensis sierra (SNBS), the rarest

subspecies of bighorn sheep in North America (U.S. Fish & Wildlife Service 2007).

Although data have been collected intermittently on this subspecies for >30 years,

limitations of the dataset have prohibited comprehensive demographic analyses; despite

the value of such information for directing recovery efforts. Data available on SNBS

include ground counts, telemetry based known-fate survival and reproductive success,

and mark-resight surveys. Our objectives were to evaluate the effectiveness of the

Bayesian state-space approach for combining SNBS data types to 1) estimate annual

population size and vital rate parameters and, 2) determine factors potentially driving

variation in key vital rates.

To meet our second objective we specifically evaluated the factors driving those

vital rates that are most important to this subspecies. A recent sensitivity analysis found

that most variation in SNBS growth rates was attributable to variation in adult female

survival and fecundity (Johnson et al. 2010). We evaluated variation in these “key” vital

Page 96: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

82

rates with respect to variation in population density, winter severity, and summer

precipitation. These covariates are likely to affect bighorn sheep in the Sierra Nevada and

are commonly associated with the dynamics of other ungulate populations (Portier et al.

1998; Coulson, Milner-Gulland & Clutton-Brock 2000; Gaillard et al. 2000; Jacobson et

al. 2004).

MATERIALS AND METHODS

SNBS Populations

Seven SNBS populations currently exist, but we focus only on the three for which

there are long-term demographic data: Warren, Wheeler, and Langley. These populations

were reintroduced between 1979 and 1986 (Bleich et al. 1990), with Warren the

northernmost population, Langley the southernmost, and Wheeler in the central part of

the range (Appendix D). These herds represent approximately 60% of the overall

subspecies population, and exhibit high spatial and temporal variation in population

trends, density, and environmental conditions. All populations are known to be

geographically isolated so that their dynamics are independent. Detailed information

about the history of the populations and the study area is described in Johnson et al.

(2010).

Data Types

Ground count data (yC) – Annual ground counts were performed by experienced

observers, who systematically hiked and scanned each herd area for bighorn sheep by sex

and stage class (lambs, yearlings and adults). Due to small population sizes and repeated

surveys, in many cases counts were successful at being complete, or near-complete,

censuses of numbers in each stage class. We used counts collected at Warren from 1988

Page 97: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

83

to 2008, at Wheeler from 1981 to 2009, and at Langley from 1987 to 2008 (Table 3.1).

Annual surveys at Warren and Langley occurred in July or August, shortly after new

lambs were born (post-birth pulse), while surveys at Wheeler occurred in March or April

just before new lambs were born (pre-birth pulse). The lambing period primarily occurs

from mid-April to mid-June with adult females giving birth to one offspring/year

(Wehausen 1980, 1996). Although three stage classes were observed during both pre- and

post-birth pulse surveys, the timing of surveys resulted in distinct differences in the field

data that translate into different parameterizations of our demographic models.

Throughout the Methods we describe data and models relevant to post-birth pulse

surveys, and provide details on the pre-birth pulse modifications in Appendix E. Post-

birth pulse surveys counted the number of adult females (≥2.2 yrs; yA), yearling females

(~1.2 yrs; yY), and newborn lambs (~0.2 yrs; yL).

Telemetry data (yT) - California Department of Fish and Game (CDFG) personnel

radio-collared adult female SNBS for information on individual survival and

reproduction. Radio-collars were deployed in each herd 1-2 times/year using a net-gun

fired from a helicopter. Collaring efforts began at Wheeler in 2001, Warren in 2002, and

Langley in 2004. Collared females were monitored twice/month by ground and aerial

telemetry for survival (systematic survival monitoring began at Wheeler in 2002 and at

Warren in 2003) and they were observed annually to determine whether or not they had a

lamb (reproductive monitoring began in Langley in 2005). While telemetry data should

yield precise parameter estimates, small sample sizes limit statistical power in many

years (Table 3.1).

Page 98: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

84

Mark-resight data (yMR) – Mark-resight data were collected in Wheeler and

Langley from 2006 onward, following McClintock et al. (2007; Table 3.1). During mark-

resight surveys herd areas were systematically searched (without telemetry) for all adult

females, and the identities of marked (collared) females and the numbers of unmarked

adult females were recorded. Surveys were conducted in a single day by multiple

observers such that sampling of marked animals was done without replacement. We

performed multiple (2-3) mark-resight surveys within a season to estimate adult female

population size. The only exception to this was that Langley was only surveyed once in

2007. We did not collect mark-resight data on the same days as ground counts to ensure

independence among data types.

Covariate data - We evaluated the effects of population density, winter severity,

and summer rainfall on SNBS survival and reproductive rates. The effect of density on

vital rates in year t was modelled as the number of adult and yearling females estimated

in year t-1 (described in State Process). We indexed winter severity by the monthly

average depth of snowpack from February-April (cm). Snow data were obtained from

weather stations operated by the California Department of Water Resources (CDEC;

http://cdec.water.ca.gov). We selected population-specific stations to reflect differences

in local conditions, with stations located within or adjacent to each herd area and situated

at 2,775 - 3,050 m, an average winter elevation for SNBS. For summer rainfall we

calculated mean monthly precipitation from June-August (cm), as rain during these

months is likely to be important for maintaining growth and nutrient quality of forage in

the arid eastern Sierra Nevada. We obtained precipitation data from NOAA weather

Page 99: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

85

stations (http://lwf.ncdc.noaa.gov/oa/climate) located in the towns closest to each herd

because CDEC stations have not tracked long-term precipitation patterns.

Model Formulation and Parameterization

General approach – The Bayesian state-space approach identifies state processes

(equations describing the dynamics of the system) and observation processes (equations

linking state processes to empirical field data; Besbeas et al. 2002; Brooks, King &

Morgan 2004; Buckland et al. 2004; Schaub et al. 2007). For SNBS, the state process

describes annual changes in the size of each SNBS stage class as a function of changes in

stage-specific vital rates, modelled with a series of likelihood functions. The observation

process then links our various data types to the population size and vital rate parameters

describing our system.

For populations surveyed post-birth pulse the observed stage classes were adult

females, yearling females, and lambs, and the vital rates describing changes in these

stages were annual adult female survival (ΦA), yearling female survival (ΦY), and

fecundity (F; the number of lambs born/number of adult females). For Wheeler, surveyed

during the pre-birth pulse, the observed stage classes were adult females, two-year-old

females, and yearlings, and the associated vital rates were adult female survival, two-

year-old survival and recruitment (the number of lambs that were born and survived their

first year/adult female; see Appendix E). A consequence of this difference was that we

were able to estimate fecundity for Warren and Langley, and recruitment for Wheeler.

Each of our data types, count (yC), telemetry (yT), and mark-resight data (yMR),

provide information on a subset of demographic parameters (Fig. 3.1). Annual ground

counts provide direct information on the numbers of animals in each stage class, and

Page 100: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

86

consecutive annual counts provide indirect information on survival rates. Meanwhile,

telemetry data can be used to estimate adult female survival and fecundity, and mark-

resight data can be used to estimate the population size of adult females. We modelled the

Warren, Wheeler, and Langley populations independently, parameterizing models with

demographic and covariate data specific to each herd.

State process- We used a binomial distribution to model the number of adult

females in year t as a function of the adult female survival rate from t-1 to t (ΦA(t-1)) and

the number of adult and yearling females in year t-1:

NA♀(t) ~ Binomial (ΦA(t-1), NA♀(t-1) + NY♀(t-1))

Because field data did not identify lambs by sex, we assumed a 50:50 sex ratio and

described the number of yearling females at time t as a function of their survival from t-1

to t (ΦY(t-1)) and half the total number of lambs in t-1:

NY♀(t) ~ Binomial (ΦY(t-1), 0.5*NL(t-1))

We assumed that yearlings did not produce offspring, as ultrasonography on 8 yearlings

(captured from 2003-2009) found only one to be pregnant (CDFG unpublished data). We

therefore modelled the number of lambs in year t as a function of the annual fecundity

rate (F(t)) and the number of adult females in year t-1 multiplied by their survival rate

(ΦA(t-1)):

NL(t) ~ Binomial (F(t), NA♀(t-1)* ΦA(t-1))

We could model this process using a binomial distribution because SNBS only give birth

to 1 offspring/year (Wehausen 1980).

Ground count likelihood functions- Given that ground counts of SNBS have been

complete or near-complete censuses in most years we assumed that counts of adult

Page 101: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

87

females (yCA♀), yearling females (yCY♀), and lambs (yCL) in year t were normally

distributed as a function of the true number of adult females (NA♀), yearling females

(NY♀), and lambs (NL) in year t:

yCA♀(t) ~ Normal (NA♀(t), σyA2)

yCY♀(t) ~ Normal (NY♀(t), σyY2)

yCL(t) ~ Normal (NL(t), σyL2)

where σy2

terms represent the variance associated with counts of each stage class, and the

normal distribution is truncated at 0. These equations describe how ground count data are

linked to the true, but unknown, number of animals in the population.

Telemetry likelihood functions- While we obtained indirect information on adult

female survival and fecundity from consecutive annual counts, we also obtained direct

information on these vital rates from collared adult females. Given known-fate telemetry

data (yTKnown-Fate), we used a parametric exponential model to estimate annual adult

female survival. This model only required estimation of a single parameter which could

be accommodated by limited telemetry data. This model assumes that the baseline hazard

rate (H0; the probability that death occurs in given interval) is constant and is expressed as

the negative log of the survival rate (Hosmer & Lemeshow 1999; Ibrahim, Chen & Sinha

2001). To approximate the baseline hazard rate we described the probability of mortality

(D(i,t)) for female i in year t as a function of the population-specific hazard rate in that

year:

D(i,t) ~ exponential (H0(t)| Number of days at risk(i,t))

We needed to account for different numbers of days-at-risk because animals were

collared following a staggered-entry design. For females collared at the start of the year

Page 102: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

88

the days-at-risk was 365, and for females collared after the start of the year the days-at-

risk was reduced to reflect monitoring time. This model provided an estimate of the daily

(instantaneous) hazard rate. Annual adult female survival was then calculated by taking

the exponent of negative daily hazard rate multiplied by 365:

ΦA(t) = e-(H0(t) * 365)

We calculated annual known-fate survival rates to match the timing of ground counts (i.e.

the timing of pre-birth pulse vs. post-birth pulse surveys).

To incorporate data on the reproductive success of collared females (yTLambing)

into fecundity estimates, we used a binomial distribution to describe the number of

collared females with a lamb in year t (l(t)) as a function of the annual fecundity rate (F(t))

and the number of collared females monitored in that year (c(t)):

l(t) ~ binomial (F(t) , c(t))

Mark-Resight Likelihood Functions- We included mark-resight data (yMR) into

estimates of annual adult female population size (NA) by using a modified Bayesian

binomial model (McClintock & Hoeting 2009). This model describes the probability of

sighting individual i in year t (x(i,t)) as a function of the annual detection probability (ρ(t))

and the number of surveys (sampling occasions) conducted in that survey season (k(t)):

x(i,t) ~ binomial (ρ(t) , k(t))

Assumptions are that the number of marked animals is known, sampling is without

replacement, and there is no individual heterogeneity in sighting probabilities. Our study

design satisfied the first two assumptions. While there may be some heterogeneity in

sighting probabilities, small numbers of marked animals limited our ability to fit more

complex models. We assumed that the total number of unmarked adult females observed

Page 103: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

89

across sampling occasions for a given year (UF(t)) was a binomial function of the annual

detection probability (ρ(t)) and Uk(t), or the number of sampling occasions that occurred in

year t multiplied by the total number of unmarked adult females observed during those

occasions (the total number of adult females minus the number of marked females (m(t))),

as described by the equations:

UF(t) ~ binomial (ρ(t) , Uk(t)) , and

Uk(t) = ( NA♀(t) - m(t)) * k(t)

While the mark-resight model was developed by McClintock & Hoeting (2009) for

populations where the number of marked animals was unknown, the model could be

simplified to assess the abundance of populations like SNBS where the number of marks

is known with certainty.

Combined model - Likelihoods from different data types were combined to form a

joint model, where parameter estimates were maximized across individual component

likelihoods (Besbeas et al. 2002; Brooks, King & Morgan 2004; Buckland et al. 2004;

Goodman 2004). A key assumption in pooling multiple data types into a joint likelihood

is sampling independence among those data types (not independence among sampled

animals). Because count, mark-resight, and telemetry data were collected independently

of one another their component likelihoods in our post-birth pulse model can be

represented as:

Ground count: L ( yC | NA, NY, NL, ΦA, ΦY, F )

Telemetry: L ( yT | ΦA, F )

Mark-Resight: L ( yMR | NA, ρ )

Page 104: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

90

The demographic parameters adult female survival (ΦA), fecundity (F), and adult female

population size (NA) are present in multiple independent component likelihoods so we

combined likelihoods from the three different data types to yield the joint function:

L ( yC, yT, yMR | NA, NY, NL, ΦA, ΦY, F, ρ )

In addition to estimating annual population numbers and vital rates, we calculated annual

population growth rates (λt) for each herd, a derived parameter. This was obtained by

dividing the number of adult and yearling females in year t by the number of adult and

yearling females in year t-1:

λt = (NA(t) + NY(t))/(NA(t-1) + NY(t-1))

We used the geometric mean of annual λt’s for each population to estimate the average

growth rate over the time each herd was monitored.

Fitting covariates to key vital rates – After estimating baseline population size

and vital rate parameters we fit the covariates density, winter snowpack, and summer

precipitation to adult survival and fecundity/recruitment rates, assessing covariate effects

for each vital rate independently in separate models. We modelled these factors for

periods of complete or near-complete consecutive annual surveys: 1988-2008 for Warren,

1995-2009 for Wheeler, and 1996-2008 for Langley. Because vital rates were constrained

between 0 and 1, we used a logit transformation to model covariates as a linear function

of adult female survival and fecundity:

logit (ΦA(t)) = β0 + β1(density(t)) + β2(snow depth(t)) + β3(precipitation(t)), and

logit (F(t)) = β0 + β1(density(t)) + β2(snow depth(t)) + β3(precipitation(t))

Model implementation – We made inferences about demographic parameters by

drawing samples from the joint posterior distribution using Markov Chain Monte Carlo

Page 105: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

91

techniques in WinBUGS (Lunn et al. 2000). We ran each model for 1,100,000 iterations,

discarding the first 100,000 iterations as “burn-in” and sampling 1 out of every 10

iterations thereafter to estimate posterior distributions (PD) for parameter values.

Convergence occurred within 5,000 iterations, as indicated by the Brooks-Rubin-Gelman

diagnostic (Brooks & Gelman 1998). We estimated adult female survival, fecundity, and

the total number of adult females (parameters with data from multiple sources) based on

each data type independently, and in combination, to evaluate the effectiveness of

pooling across data types. We also estimated regression coefficients for overall covariate

effects, annual detection probabilities for mark-resight surveys, variance terms for stage-

specific ground counts, and annual population growth rates (Fig. 3.1).

We used uninformative priors to specify demographic parameters. We used 10, 5,

and 5, as prior initial sizes of adult, yearling, and lamb stage classes in all populations for

the first year of the simulation (values that reflected the small sizes of these reintroduced

populations), using large variances of 104

(Brooks, King & Morgan 2004; Schaub et al.

2007). We used an inverse gamma prior for estimates of σy2 with distribution parameters

equal to 0.001 (Brooks, King & Morgan 2004). We assumed uniform prior distributions

for detection probabilities that ranged from 0 to 1, and beta distributions (mean and

variance parameters of 1) for vital rates modeled without covariate effects. For vital rates

estimated with covariates we assumed all priors on regression coefficients were normally

distributed with a mean of 0 and variance of 2. To demonstrate the combination of count,

telemetry and mark-resight data we provide the baseline WinBUGS code in Appendix F.

Page 106: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

92

RESULTS

Parameter estimates

The inclusion of telemetry data in demographic models consistently reduced

posterior standard deviations around estimates of adult female survival, recruitment, and

fecundity, from those obtained from ground count data alone (Table 3.2, Fig. 3.2). When

estimates from ground count and telemetry data were quite different, the values obtained

from the combined model were generally intermediate, and weighted more strongly

towards the data providing higher precision (Table 3.2). Similarly, mark-resight data

increased the precision in estimates of adult female population size, although the effect

was slight (Table 3.3). When female population size was estimated independently for

each data type, posterior standard deviations around NA were significantly larger for

mark-resight estimates than for ground count estimates (~2-3 times as large; Table 3.3).

As a result, the combined estimate was consistently weighted towards values obtained

from counts (the data type with the higher precision). With the exception of Langley in

2006 (only 8 marked females), mark-resight abundance estimates were also consistently

higher than those generated from count data.

The models estimated demographic parameters even in years when field data

were not collected, although the posterior standard deviations of those estimates were

substantially larger than years for which data existed (NA estimates shown in Fig. 3.3).

These estimates can be derived because only some values are logically possible given the

population structure in the previous and following years. Models also provided PD

credible intervals (CIs) for parameters obtained from ground counts, for which there had

been no previous estimates of error. In total, 85% of the ground counts of each stage class

Page 107: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

93

fell within the credible intervals of predicted abundance values (see example of estimates

versus count data for adult females at Warren in Fig. 3.4).

We report estimated numbers of adult females, mean vital rate values, and mean

population growth rates for each herd from baseline demographic models (year-specific

estimates are provided in Appendix G). The Warren population was estimated to have

grown from 14 adult females in 1988, to a maximum of 32 in 1993. This herd

subsequently declined to an estimated 6 adult females in 1999, and has remained

relatively static since then with 5 adults estimated in 2008. Over all the years in which

data were collected, the mean adult female survival was 0.83 (SD = 0.02), yearling

survival was 0.50 (SD = 0.10), fecundity was 0.62 (SD = 0.05), and the mean growth rate

was 0.97 (95% CI = 0.86 – 1.08). For Wheeler 10 adult females were estimated in 1981,

which declined to a low of 7 in 1994, and grew to an estimated 38 as of 2009. From 1981

to 2008, average adult female survival was 0.90 (SD = 0.02), two-year-old survival was

0.67 (SD = 0.07), recruitment was 0.55 (SD = 0.03), and the mean growth rate was 1.04

(95% CI = 1.00 – 1.08). The Langley herd was estimated at 18 adult females in 1987,

declining to a low of 10 in 1997, and subsequently increasing to 40. From 1987 to 2008

the average adult female survival was 0.91 (SD=0.02), yearling survival was 0.77 (SD =

0.06), fecundity was 0.63 (SD = 0.05), and the mean growth rate was 1.05 (95% CI =

1.00 to 1.11).

Covariate effects

The increasing populations, Wheeler and Langley, exhibited negative density

dependence in both vital rate parameters (95% CIs of regression coefficients did not

overlap zero with the exception of adult survival at Wheeler which had 86% PD<0; Table

Page 108: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

94

3.4). Meanwhile, there was a trend suggesting positive density dependence in survival

rates at Warren (regression coefficient had 90% PD>0; Table 3.4; Fig. 3.5). Generally

summer rainfall had a positive influence on survival and reproductive rates in all

populations although the effect was greatest on fecundity rates for bighorn sheep in

Warren (95% CI of regression coefficients did not overlap 0; Table 3.4; Fig 3.5).

Regression coefficients demonstrated that increases in snow depth were positively

associated with adult survival at Wheeler (>83% PD>0) and with reproduction at Langley

(95% CIs did not overlap zero). At Warren, however, both adult survival and fecundity

were strongly negatively associated with winter snow depth (95% CIs did not overlap

zero; Fig. 3.5).

DISCUSSION

The Bayesian state-space models developed here, combining ground count,

telemetry, and mark-resight data, allowed us to integrate all the available data to increase

accuracy and precision in parameter estimates and to fit covariates to vital rates driving

population performance. Imprecise parameter estimates are one of the greatest limitations

in detecting population trends, diagnosing causes of declines, and directing management

actions (Taylor & Gerrodette 1993; Gibbs, Droege & Eagle 1998). By integrating all

available data to better track the spatial and temporal dynamics of SNBS populations we

are able to prioritize populations for management intervention, identify vital rates that

should be increased, and direct future recovery strategies.

We found that combining SNBS data types significantly improved the precision

of demographic parameter estimates, as has been found in other studies combining data

types (Fig. 3.2; Besbeas et al. 2002; White & Lubow 2002; Brooks, King & Morgan

Page 109: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

95

2004; Schaub et al. 2007). Given the wide application of telemetry in wildlife research

and monitoring, the integration of telemetry data with other data types, such as ground

counts, has tremendous potential for enhancing demographic parameter estimation. While

the integration of count and telemetry data dramatically decreased the variation in adult

survival and fecundity, the combination of count and mark-resight data just slightly

decreased the variation in estimates of adult female population size (Table 3.3). This

occurred because the count data model had a constant relationship to population size

while the mark-resight model replaced this assumption with an estimate of detection

probability, thus increasing the variance around abundance estimates. Given that the

count data yielded a much more precise estimate of adult abundance, the estimates from

the combined data model were biased towards those from counts. We suspect that

heterogeneity in resighting probabilities among individuals (unaccounted for in the

current model) may have caused the mark-resight data to overestimate the numbers of

adult females, while count data may have underestimated them, particularly as

populations increased in size. By combining data types, estimates were more intermediate

in value and were likely to be more accurate (Table 3.3).

Given the piecemeal nature of the SNBS data, this approach was also beneficial

for standardizing the error structure across different data types. For the first time,

demographic rates of SNBS can be directly compared among populations, regardless of

the various data types used in different years. The models also estimated parameter

values for years when field data were not collected, filling in gaps in our data set, and

estimating precision around ground counts that had no previous measurement of error

(Fig. 3.3).

Page 110: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

96

With these improved demographic parameter estimates we have greater power to

prioritize populations of conservation concern and detect demographic rates indicative of

decline. For example, Wheeler and Langley have been increasing in size with long-term

growth rates of 1.04 and 1.05, respectively. Meanwhile, Warren has had a negative long-

term growth rate of 0.97 and is the clear management priority. While fecundity rates at

Warren were comparable to the other herds, adult and yearling survival rates were ~10%

and 20% lower, respectively; suggesting that recovery activities should focus on

increasing these rates (Johnson et al. 2010). Greater precision in parameter estimates can

also be used to more quickly identify key changes in population trajectories. Since 2000

growth rates at Langley have been >1.13, however, in 2008 this rate dropped

dramatically. Based on only the ground count data, the annual growth rate was estimated

at 0.88 with a 95% credible interval ranging from 0.66 to 1.08, yielding uncertainty about

the status of the population. Given the improved combined data model, the annual growth

rate was estimated at 0.86 with a credible interval from 0.76 to 0.96, signalling to

managers the likelihood of a definitive short-term decline.

The results of our models can also be used to improve population monitoring

efficiency (Goodman 2004). For example, it appears that telemetry-based vital rate

estimates have comparable precision to ground counts, but only in the large populations

and when >30% of the females are collared. As populations increase in size and near-

complete census counts are harder to obtain, information on adult survival, fecundity, and

adult female abundance could be entirely derived from telemetry data. While telemetry

data were highly informative for estimating demographic rates in large SNBS

populations, count data were more effective for elucidating rates in small herds like

Page 111: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

97

Warren, where small sample sizes caused telemetry estimates to be less precise (Table

3.2). In the future, simulation studies could be used to identify the value of different data

types (both those currently collected and novel ones) to estimate parameters under a wide

range of conditions to improve monitoring programmes given logistical and budgetary

constraints.

By including covariates into demographic models we found evidence of negative

density dependence in both adult female survival and recruitment/fecundity in the

increasing populations of Wheeler and Langley (Fig. 3.5), despite their relatively small

sizes. Negative density effects may arise from a combination of high site fidelity, limited

female dispersal, and discrete habitat patches in the Sierra Nevada; all factors that may

constrain bighorn sheep from expanding into unoccupied ranges. In Wheeler and Langley

reproductive rates were impacted at lower densities than adult survival (Fig. 3.5), a

pattern commonly observed in other ungulates, as younger stage classes tend to be

disproportionately influenced by negative density dependence (Gaillard, Festa-Bianchet

& Yoccoz 1998; Gaillard et al. 2000).

While vital rates were depressed by density in Wheeler and Langley, it appeared

to be positively associated with adult survival at Warren (Fig. 3.5). Studies of density in

ungulate populations have largely focused on negative effects, with few studies observing

Allee effects, or positive density dependence, in small populations (Treydte et al. 2001;

Matson, Goldizen & Jarman 2004; Wittmer, Sinclair & McLellan 2005). While the

mechanism driving an Allee effect at Warren is unknown, small SNBS herds may need to

be augmented to alleviate depressed survival and/or reproductive rates. The large herds,

Page 112: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

98

showing signs of negative density effects, could potentially be used as the source stock

for such augmentations.

Similar to density, climate factors differentially influenced the small Warren

population compared to the larger populations of Wheeler and Langley. At Warren,

winter snow depth negatively affected both adult female survival and fecundity, as has

been found in other ungulates (Jacobson et al. 2004, Gaillard et al. 2000). Contrary to

this pattern, however, snow depth had a minor positive effect on survival at Wheeler and

on reproduction at Langley. This disparity may be due to the relative amounts of snowfall

each herd receives, as the weather station at Warren reported snow depths almost double

those reported for other herds (Fig. 3.5). Additionally, low elevation winter range is

abundant in Wheeler and Langley, and SNBS in these herds routinely descend below

snow line. Meanwhile, a majority of the winter observations of SNBS at Warren have

been at high elevations on slopes blown-free from snow, where snow patterns are

expected to have a greater effect (CDFG unpublished data). Given that most precipitation

in the arid Sierra Nevada is received as snow in winter, the positive effects of snow depth

at Wheeler and Langley may reflect a longer growing season the following spring and

summer. Unlike snow, summer precipitation affected all populations in a similar way,

having a positive influence on both survival and reproduction. While this effect was

slight at Wheeler and Langley, rainfall dramatically increased vital rates in Warren. As

with snow, this may be a function of the range of rainfall values that occurred during the

study, which were greatest at Warren (Fig. 3.5).

Although weather factors typically have a greater influence on younger stage

classes than older ones (Gaillard et al. 2000), we did not see this pattern in SNBS.

Page 113: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

99

Weather covariates had weak effects on both adult survival and reproduction at Wheeler

and Langley, but elicited strong effects on the vital rates of both young and old stage

classes at Warren. The powerful influence of weather on adult survival in this herd is

disconcerting, as this is atypical of ungulates and may contribute to Allee effects. We are

uncertain whether the strong influence of environmental stochasticity in this population

reflects a difference in habitat quality or is simply a function of its small size and

demographic stochasticity.

Mountain lions, Puma concolor, are the main predator of SNBS (Wehausen 1996)

and lion predation is the primary known-cause of mortality (CDFG, unpublished data).

Because consistent long-term data on mountain lions in SNBS populations were

unavailable, this factor could not be included in our demographic models. Thus, observed

negative density dependence could be due to food-based or predator-based carrying

capacity; further work on the role of predation in SNBS dynamics will be required. Other

factors that may significantly influence SNBS vital rates, including disease, habitat use

patterns, and genetic diversity, were not included in our analysis but are suspected to play

a significant role in the dynamics of these populations.

To successfully manage and conserve populations we must be able to accurately

estimate key demographic parameters and identify the deterministic and stochastic factors

driving the variation in those rates. Using traditional statistical methods, such analyses

have been limited for populations with piecemeal datasets, common for both endangered

and harvested species. We found that Bayesian state-space models were a powerful tool

for integrating count, telemetry and mark-resight data available for SNBS, identifying the

information content of different data types, determining demographic trends, and

Page 114: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

100

elucidating the ecological processes driving dynamics (King et al. 2008; Véron &

Lebreton 2008). For SNBS, our model results can be used to prioritize populations of

conservation concern, better detect population declines, improve monitoring schemes,

and direct management strategies; all capabilities that will improve recovery success in

this subspecies.

ACKNOWLEDGEMENTS

We thank E. Crone and M. Hebblewhite for discussion and statistical modelling

support, B. McClintock for providing code for Bayesian mark-resight estimation, and D.

German for compiling demographic data. We are grateful to all the people who have

collected data on SNBS including T. Blankinship, L. Bowermaster, L. Brown, M. Cahn,

K. Chang, L. Chow, K. Ellis, D. German, A. Feinberg, J. Fusaro, L. Greene, D. Jensen,

M. Kiner, K. Knox, P. Moore, R. Ramey II, C. Schroeder, T. Taylor, and S. Thompson.

CDFG and the Canon National Parks Scholars Program provided funding for this work.

LITERATURE CITED

Bakker, V.J., Doak, D.F., Roemer, G.W., Garcelon, D.K., Coonan, T.J., Morrison, S.A.,

Lynch, C., Ralls, K. & Shaw, R. (2009) Incorporating ecological drivers and

uncertainty into a demographic population viability analysis for the island fox.

Ecological Monographs, 79, 77-108.

Besbeas, P., Freeman, S.N., Morgan, B.J.T. & Catchpole, E.A. (2002) Integrating

mark-recapture-recovery and census data to estimate animal abundance and

demographic parameters. Biometrics, 58, 540-547.

Bleich, V.C., Wehausen, J.D., Jones, K.R. & Weaver, R.A. (1990) Status of bighorn

Page 115: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

101

sheep in California, 1989 and translocations from 1971 through 1989. Desert

Bighorn Council Transactions, 34, 24-26.

Brooks, S.P., & Gelman, A. (1998) Alternative methods for monitoring convergence of

iterative simulations. Journal of Computational and Graphical Statistics, 7, 434-

455.

Brooks, S.P., King, R. & Morgan, B.J.T. (2004) A Bayesian approach to combining

animal abundance and demographic data. Animal Biodiversity and Conservation,

27, 515-529.

Buckland, S.T., Newman, K.B., Thomas, L. & Koesters, N.B. (2004) State-space models

for the dynamics of wild animal populations. Ecological Modelling, 171, 157-175.

Coulson, T.E., Milner-Gulland, E.J. & Clutton-Brock, T. (2000) The relative roles of

density and climate variation on population dynamics and fecundity rates in three

contrasting ungulate species. Proceedings of the Royal Society of London Series

B, 267, 1771-1779.

Fieberg, J., & Ellner, S.P. (2001) Stochastic matrix models for conservation and

management: a comparative review of methods. Ecology Letters, 4, 233-266.

Franklin, A.B., Anderson, D.R., Gutierrez, R.J., & Burnham, K.P. (2000) Climate,

habitat quality, and fitness in Northern Spotted Owl populations in northwestern

California. Ecological Monographs, 70, 539-590.

Gaillard, J.-M., Festa-Bianchet, M. & Yoccoz, N.G. (1998) Population dynamics of large

herbivores: variable recruitment with constant adult survival. Trends in Ecology

and Evolution, 13, 58-63.

Gaillard, J.-M., Festa-Bianchet, M., Yoccoz, N.G., Loison, A. & Toïgo, C. (2000)

Page 116: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

102

Temporal variation in fitness components and population dynamics of large

herbivores. Annual Review of Ecology and Systematics, 31, 367-393.

Gibbs, J.P., Droege, S. & Eagle, P. (1998) Monitoring populations of plants and animals.

Bioscience, 48, 935-940.

Goodman, D. (2004) Methods for joint inference from multiple data sources for

improved estimates of population size and survival rates. Marine Mammal

Science, 20, 401-423.

Hosmer, D.W.Jr., & Lemeshow, S. (1999) Applied survival analysis: Regression

modeling of time to event data. John Wiley & Sons, Inc., New York.

Ibrahim, J.G., Chen, M.-H., & Sinha, D. (2001) Bayesian survival analysis. Springer

Sciencies & Business Media, Inc., New York.

Jacobson, A.R., Provenzale, A., Von Hardenberg, A., Bassano, B. & Festa-Bianchet, M.

(2004) Climate forcing and density dependence in a mountain ungulate

population. Ecology, 85, 1598-1610.

Johnson, H.E., Mills, L.S., Stephenson, T.R. & Wehausen, J.D. (2010) Population-

specific vital rate contributions influence management of an endangered ungulate.

Ecological Applications, 20, 1753-1765.

King, R., Brooks, S.P., Mazzetta, C., Freeman, S.N. & Morgan, B.J.T. (2008) Identifying

and diagnosing population declines: a Bayesian assessment of lapwings in the

UK. Journal of the Royal Statistical Society Series C. Applied Statistics, 57, 609-

632.

Lunn, D.J., Thomas, A., Best, N. & Spiegelhalter, D. (2000) WinBUGS – a Bayesian

Page 117: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

103

modeling framework: concepts, structure, and extensibility. Statistics and

Computing, 10, 325-337.

Matson, T.K., Goldizen, A.W. & Jarman, P.J. (2004) Factors affecting the success of

translocations of the black-faced impala in Namibia. Biological Conservation,

116, 359-365.

McClintock, B.T. & Hoeting, J.A. (2009) Bayesian analysis of abundance for binomial

sighting data with unknown number of marked individuals. Environmental and

Ecological Statistics. DOI 10.1007/s10651-009-0109-0.

McClintock, B.T. & White, G.C. (2007) Bighorn sheep abundance following a

suspected pneumonia epidemic in Rocky Mountain National Park. Journal of

Wildlife Management, 71, 183-189.

Morris, W.F., Bloch, P.L., Hudgens, B.R., Moyle, L.C. & Stinchcombe, J.R. (2002)

Population viability analysis in endangered species recovery plans: past use and

future improvements. Ecological Applications, 12, 708-712.

Morris, W.F. & Doak, D.F. (2002) Quantitative conservation biology: Theory and

practice of population viability analysis. Sinauer, Sunderland, Massachusetts.

Portier, C., Festa-Bianchet, M., Gaillard, J.-M., Jorgenson, J.T. & Yoccoz, N.G. (1998)

Effects of density and weather on survival of bighorn sheep lambs (Ovis

Canadensis). Journal of Zoology, 245, 271-278.

Schaub, M., Gimenez, O., Sierro, A. & Arlettaz, R. (2007) Use of integrated modeling

to enhance estimates of population dynamics obtained from limited data.

Conservation Biology, 21, 945-955.

Taylor, B.L. & Gerrodette, T. (1993) The uses of statistical power in conservation

Page 118: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

104

biology: the vaquita and northern spotted owl. Conservation Biology, 7, 489-500.

Tear, T.H., Scott, J.M., Hayward, P.H. & Griffith, B. (1995) Recovery plans and the

Endangered Species Act: are criticisms supported by data? Conservation Biology,

9, 182-195.

Treydte, A.C., Williams, J.B., Bedin, E., Ostrowski, S., Seddon, P.J., Marschall, E.A.,

Waite, T.A. & Ismail, K. (2001) In search of the optimal management strategy for

Arabian oryx. Animal Conservation, 4, 239-249.

U.S. Fish and Wildlife Service (2007) Recovery Plan for the Sierra Nevada bighorn

sheep. Sacramento, California.

Véron, S. & Lebreton, J-D. (2008) The potential of integrated modeling in conservation

biology: a case study of the black-footed albatross (Phoebastria nigripes).

Canadian Journal of Statistics, 36, 85-98.

Wehausen, J.D. (1980) Sierra Nevada bighorn sheep: history and population ecology.

Ph.D. dissertation, University of Michigan.

Wehausen, J.D. (1996) Effects of mountain lion predation on bighorn sheep in the Sierra

Nevada and Granite Mountains of California. Wildlife Society Bulletin, 24, 471-

479.

White, G.C. & Lubow, B.C. (2002) Fitting population models to multiple sources of

observed data. Journal of Wildlife Management, 66, 300-309.

Wittmer, H.U., Sinclair, A.R.E. & McLellan, B.N. (2005) The role of predation in the

decline and extirpation of woodland caribou. Oecologia, 144, 257-267.

Page 119: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

105

Table 3.1. Number of years (n) that ground count, telemetry, and mark-resight data were collected on the Warren, Wheeler, and

Langley populations of Sierra Nevada bighorn sheep. The minimum and maximum numbers of adult females that were radio-collared

for telemetry and mark-resight surveys are given in parentheses.

Ground Count Telemetry - Adult Survival Telemetry – Fecundity Mark – Resight

Population Years Collected n Years Collected n Years Collected n Years Collected n

Warren 1988-1999,

2001-2008

20A 2003-2008 6 (1-6) 2002, 2005-

2008

4 (1-5) N/A N/A

Wheeler 1981, 1983, 1984,

1987, 1992, 1995-

2009

20 2002-2009 8 (7-21) 2001-2009 9 (5-17) 2006-2009 4 (13-18)

Langley 1987, 1990, 1996-

2008

14B 2004-2008 5 (3-17) 2005-2008 4 (7-15) 2006-2008 3 (8-17)

A No ground count for adult females in 1994 and for yearling females from 1991-1994.

B No ground count for adult females in 2005.

Page 120: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

106

Table 3.2. Annual adult female survival (ΦA), recruitment (R; for pre-birth pulse surveyed Wheeler) and fecundity (F; for post-birth

pulse surveyed Warren and Langley) rates for Sierra Nevada bighorn sheep populations. Estimates were obtained from ground count

data only, telemetry data only, and with the combined data model (count and telemetry data; SD in parentheses).

Ground Count Only Telemetry Only Combined Model

Population

& Year ΦA R/F ΦA R/F ΦA R/F

Wheeler

2001 0.82 (0.10) 0.95 (0.05) 0.86 (0.12) 0.71 (0.16) 0.88 (0.08) 0.91 (0.06)

2002 0.91 (0.08) 0.79 (0.11) 0.89 (0.10) 0.75 (0.14) 0.94 (0.05) 0.81 (0.09)

2003 0.85 (0.09) 0.70 (0.11) 0.83 (0.11) 0.67 (0.15) 0.87 (0.07) 0.69 (0.09)

2004 0.91 (0.07) 0.65 (0.10) 0.80 (0.10) 0.55 (0.14) 0.86 (0.06) 0.61 (0.08)

2005 0.92 (0.07) 0.79 (0.09) 0.93 (0.07) 0.62 (0.13) 0.95 (0.04) 0.73 (0.07)

2006 0.76 (0.09) 0.58 (0.09) 0.77 (0.10) 0.57 (0.13) 0.80 (0.06) 0.58 (0.08)

2007 0.92 (0.07) 0.34 (0.08) 0.85 (0.08) 0.59 (0.12) 0.89 (0.05) 0.41 (0.07)

2008 0.89 (0.07) 0.32 (0.08) 0.94 (0.05) 0.26 (0.10) 0.95 (0.04) 0.28 (0.06)

Page 121: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

107

2009 NA 0.46 (0.08) NA 0.53 (0.12) NA 0.47 (0.07)

Warren

2004 0.73 (0.16) 0.75 (0.17) 0.57 (0.27) NA 0.82 (0.13) 0.74 (0.18)A

2005 0.76 (0.13) 0.57 (0.20) 0.47 (0.23) 0.40 (0.20) 0.72 (0.13) 0.50 (0.16)

2006 0.76 (0.14) 0.72 (0.18) 0.71 (0.16) 0.57 (0.18) 0.81 (0.09) 0.65 (0.15)

2007 0.69 (0.13) 0.70 (0.18) 0.44 (0.17) 0.67 (0.18) 0.54 (0.13) 0.78 (0.12)

2008 NA 0.76 (0.17) NA 0.50 (0.22) NA 0.63 (0.19)

Langley

2003 0.91 (0.07) 0.79 (0.12) 0.70 (0.22) NA 0.91 (0.07) 0.79 (0.12) A

2004 0.94 (0.05) 0.63 (0.13) 0.87 (0.11) NA 0.95 (0.04) 0.63 (0.12) A

2005 0.86 (0.09) 0.90 (0.07) 0.81 (0.12) 0.89 (0.10) 0.86 (0.06) 0.93 (0.05)

2006 0.76 (0.11) 0.76 (0.11) 0.94 (0.06) 0.60 (0.15) 0.93 (0.05) 0.72 (0.08)

2007 0.79 (0.11) 0.67 (0.14) 0.69 (0.10) 0.53 (0.12) 0.76 (0.06) 0.55 (0.08)

2008 NA 0.32 (0.13) NA 0.43 (0.13) NA 0.33 (0.08)

AThe combined model value reflects only the ground count data because no telemetry data was available in that year.

Page 122: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

108

Table 3.3. Annual estimates of the number of adult females (NA) in the Wheeler and Langley populations of Sierra Nevada bighorn

sheep when estimated from ground count data only, mark-resight (MR) data only, and with the combined data model (count and mark-

resight data; SD in parentheses).

Population Year #Marked Females #MR Surveys MR NA Ground count NA Combined

Wheeler

2006 13 2 36 (5.96) 33 (2.02) 34 (1.62)

2007 18 1 42 (5.81) 33 (1.94) 35 (1.39)

2008 16 2 41 (6.71) 35 (1.82) 36 (1.50)

2009 16 1 44 (6.52) 35 (2.07) 38 (1.53)

Langley

2006 8 3 29 (3.5) 33 (2.97) 32 (2.13)

2007 17 1 50 (10.9) 34 (3.49) 40 (2.38)

2008 12 2 47 (6.7) 35 (3.75) 40 (2.15)

Page 123: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

109

Table 3.4. Posterior mean estimates, standard deviations, and 95% credible intervals (95% CI) for regression coefficients from

covariate models of adult female survival and fecundity/recruitment rates from Sierra Nevada bighorn sheep populations. The

proportion of the coefficient posterior distribution (PD) on either side of zero is also reported.

Vital Rate & Population Parameter Estimate SD Lower 95% CI Upper 95% CI %PD<0 %PD>0

Adult Survival

Warren Intercept 2.667 0.906 0.955 4.423 — —

Density* 0.030 0.026 -0.016 0.088 9.7 90.3

Snowpack** -0.011 0.004 -0.020 -0.002 99.4 0.6

Summer Rain* 0.564 0.439 -0.274 1.481 7.8 92.2

Wheeler Intercept 2.579 1.018 0.563 4.571 — —

Density* -0.025 0.024 -0.077 0.017 86.1 13.9

Snowpack* 0.008 0.008 -0.008 0.023 16.8 83.2

Summer Rain* 0.951 0.898 -0.759 2.771 14.2 85.8

Page 124: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

110

Langley Intercept 3.923 0.726 2.394 5.352 — —

Density** -0.061 0.015 -0.091 -0.026 99.9 0.1

Snowpack 0.003 0.004 -0.005 0.012 21.4 78.6

Summer Rain* 1.138 1.092 -0.962 3.320 14.9 85.1

Fecundity/Recruitment

Warren Intercept 1.734 0.849 0.047 3.40 — —

Density -0.009 0.024 -0.053 0.040 67.2 32.8

Snowpack** -0.006 0.003 -0.013 0.000 97.8 2.2

Summer Rain** 0.624 0.391 0.029 1.590 1.8 98.2

Wheeler Intercept 1.804 0.658 0.563 3.159 — —

Density** -0.059 0.023 -0.107 -0.057 99.9 0.1

Snowpack 0.002 0.005 -0.009 0.002 38.0 62.0

Summer Rain -0.300 0.668 -1.601 1.031 67.8 32.2

Page 125: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

111

Langley Intercept 2.454 0.935 0.707 4.376 — —

Density** -0.073 0.023 -0.119 -0.029 99.6 0.4

Snowpack** 0.009 0.004 0.001 0.018 1.1 98.9

Summer Rain* 2.180 1.252 -0.306 4.619 5.3 94.7

** Indicates covariate coefficients with confidence intervals non-overlapping zero and having >95% of their posterior probability

distributions > or < than zero.

*Indicates covariate coefficients that have >80% of their posterior probability distributions > or < than zero.

Page 126: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

112

Figure 3.1. Sierra Nevada bighorn sheep demographic model for populations surveyed

post-birth pulse. Data are represented by boxes while estimated parameters are

represented by circles. Solid arrows depict stochastic dependencies between data and

parameters and dashed arrows represent deterministic dependencies. The model combines

ground count (yCA, yCY, yCL), telemetry (yT), and mark-resight (yMR) data to estimate

numbers of adult females (NA), yearling females (NY) and lambs (NL), adult female

survival (ΦA), yearling female survival (ΦY), and fecundity (F). The model also estimates

detection probability for mark-resight surveys (ρ) and the variance of ground counts (σ2

y).

yCL

yT

Known-fate

ΦA♀ ΦY F

NA♀ NY♀ NL

σ2

yA σ2

yY σ2

yL

yMR

ρ

yT

Lambing Rate

yCY

yCA

Covariates: Density

Winter Snow Depth Summer Rainfall

Page 127: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

113

Figure 3.2. Estimates from the Wheeler population of Sierra Nevada bighorn sheep. A) Annual recruitment rates (and SD) across all

years of the study (1981-2009). From 1981-2000 estimates were based only on ground count data and from 2001-2009 estimates were

based on the combined model (count and telemetry data). B) Recruitment rate estimates when using only count data, only telemetry

data, and both data types combined. C) Standard deviations around recruitment estimates when using only count data, only telemetry

data, and both data types combined.

B

Year

2002 2004 2006 2008

Rec

ruit

men

t R

ate

0.2

0.4

0.6

0.8

1.0

Count Data

Telemetry Data

Count + Telemetry Data

C

Year

2002 2004 2006 2008

SD

of

Rec

ruit

men

t R

ate

0.00

0.05

0.10

0.15

0.20

Count Data

Telemetry Data

Count + Telemetry Data

A

Year

1980 1985 1990 1995 2000 2005 2010

Rec

ruit

men

t R

ate

0.0

0.2

0.4

0.6

0.8

1.0

Count Data

Count + Telemetry Data

Page 128: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

114

Figure 3.3. Estimated number of adult female Sierra Nevada bighorn sheep (and 95% credible intervals) in the Warren, Wheeler, and

Langley populations. Black circles (●) signify years that demographic data were collected on the populations and open squares (□)

signify years that no demographic data were collected.

Year

1980 1985 1990 1995 2000 2005 2010

0

10

20

30

40

Wheeler

Year

1990 1995 2000 2005

0

10

20

30

40

Langley

Year

1990 1995 2000 2005

Nu

mb

er o

f A

du

lt F

emale

s

0

5

10

15

20

25

30

35

40

Warren

Page 129: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

115

Figure 3.4. Number of adult female Sierra Nevada bighorn sheep counted during annual

ground surveys and estimated from the Bayesian state-space demographic model (with

95% credible intervals) in the Warren population from 1988 to 2008.

Year

1990 1995 2000 2005

Num

ber

of

Adult

Fem

ales

0

5

10

15

20

25

30

35

40

Estimate

Ground Count

Page 130: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

116

Figure 3.5. Predicted effects of winter snow depth, summer rainfall, and density on adult female survival and fecundity/recruitment

rates for populations of Sierra Nevada bighorn sheep. Predictions are only shown for populations with >80% of their regression

coefficient posterior distributions > or < 0. Predictions for weather covariates were only modeled for the range of values experienced

by each herd during the study period.

0.5

0.6

0.7

0.8

0.9

1

20 120 220

Ad

ult

Fem

ale

Surv

ival

Avg Winter Snow Depth (cm)

Warren**

Wheeler0.5

0.6

0.7

0.8

0.9

1

0 0.5 1 1.5 2 2.5 3

Avg Summer Rainfall (cm)

WarrenWheeler

0.5

0.6

0.7

0.8

0.9

1

0 10 20 30 40 50Density

WarrenWheeler

0

0.2

0.4

0.6

0.8

1

20 120 220

Fecu

nd

ity/

Rec

ruit

me

nt

Avg Winter Snow Depth (cm)

Warren**Langley**

0

0.2

0.4

0.6

0.8

1

0 0.5 1 1.5 2 2.5 3

Avg Summer Rainfall (cm)

Warren**Langley

0

0.2

0.4

0.6

0.8

1

0 10 20 30 40 50

Density

Wheeler**Langley**

Page 131: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

117

APPENDIX D. Location of Warren, Wheeler, and Langley Sierra Nevada bighorn sheep

populations, CA.

Page 132: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

118

APPENDIX E. Pre-birth pulse survey modifications to the Sierra Nevada bighorn sheep

demographic model.

Ground counts of SNBS in the Wheeler population were conducted prior to the

lambing period, following a pre-birth pulse survey design, as opposed to the post-birth

pulse surveys that occurred at Warren and Langley. As a result, there were some key

differences in the stage classes observed in the field, resulting in slight modifications to

the model structure and parameterization. At Wheeler observers counted the number of

adult females (≥2.7 yrs; yA♀), two-year-old females (~1.7-1.9 yrs; yT♀) and yearlings

(~0.7-0.9 yrs; yY), rather than the number of adult females, yearling females, and lambs

recorded in post-birth pulse surveys. Given that ground counts of SNBS have been

complete or near-complete censuses in most years we assumed that counts of each stage

class were normally distributed as a function of the true number of adult females (NA♀),

two-year-old females (NT♀), and yearlings (NY):

yCA♀(t) ~ Normal (NA♀(t), σyA2)

yCT♀(t) ~ Normal (NT♀(t), σyT2)

yCY(t) ~ Normal (NY(t), σyY2)

where σy2

terms represent the variance associated with counts of each stage class.

Given Wheeler pre-birth pulse survey data, we estimated the vital rates adult

female survival (SA), two-year-old female survival (ST), and recruitment (RA; the number

of lambs born in t-1 and survived to year t per adult female in t-1). We used binomial

distributions to model the number of individuals in each stage class as a function of these

stage-specific vital rates. The number of adult females in year t was modeled as a

Page 133: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

119

function of the adult female survival rate from t-1 to t (ΦA(t-1)) and the number of adult

and two-year-old females in t-1:

NA♀(t) ~ Binomial (ΦA(t-1), NA♀(t-1) + NT♀(t-1))

Because yearlings at Wheeler were not reliably identified by sex, we assumed a 50:50 sex

ratio and described the number of two-year-old females at time t as a function of their

survival from t-1 to t (ΦT(t-1)) and half the total number of yearlings in t-1:

NT♀(t) ~ Binomial (ΦT(t-1), 0.5*NY(t-1))

We assumed that two-year-olds did not produce offspring and thus modeled the number

of yearlings in year t as a function of the annual recruitment rate (R(t)) and the number of

adult females in year t-1:

NY(t) ~ Binomial (R(t), NA♀(t-1))

Due to modifications in the vital rates estimated from pre-birth pulse surveys, the

likelihood function describing reproductive data on marked females was adapted to

account for recruitment rather than fecundity. We used a binomial distribution to describe

the probability of a marked female recruiting a yearling in year t (Y(t)) as a function of the

annual recruitment rate (R(t)) and the number of marked females monitored in that year

(m(t)):

Y(t) ~ Binomial (R(t) , m(t))

Pre-birth pulse surveys also required us to modify our calculation of annual

population growth rates. For Wheeler these were obtained by dividing the number of

adult and two-year-old females in year t by the number of adult and two-year-old females

in t-1:

λt = (NA(t) + NT(t))/(NA(t-1) + NT(t-1)).

Page 134: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

120

APPENDIX F. WinBUGS code for Sierra Nevada bighorn sheep post-birth pulse

demographic model.

To demonstrate the application of ground count, telemetry, and mark-resight data

in a single Bayesian state-space model, we provide bighorn sheep data and starting values

for the Langley population from 2000 to 2008. Because the model is recursive, note that

values for the first years of the simulation will be slightly different than those reported in

the manuscript which used data from 1987 to 2008.

model{

######################## PRIORS ##############################

#Observation error prior for each observed stage class; adults, yearlings and lambs.

“Vary” #represents variance and “tauy” represents precision.

avary <- 1/atauy

atauy ~ dgamma(0.001, 0.001)

yvary <- 1/ytauy

ytauy ~ dgamma(0.001, 0.001)

lvary <- 1/ltauy

ltauy ~ dgamma(0.001, 0.001)

#Initial population priors (year 1) for numbers of adults, yearlings & lambs. The (0, )

truncates #the lower bound of the distribution at zero.

Na[1] ~ dnorm(10,0.0001) I(0, )

Ny[1] ~ dnorm(5,0.0001) I(0, )

Page 135: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

121

Nl[1] ~ dnorm(5,0.0001) I(0, )

#Priors for survival and fecundity probabilities.

for(t in 1:(T-1)){

phiaf[t] ~ dbeta(1,1)

phiy[t] ~ dbeta(1,1)

}

for(t in 2:T){

fecund[t] ~ dbeta(1,1)

}

#################### GROUND COUNTS ###############################

#System process for ground count data which describes numbers of animals in each stage

class #in year t as a function of the number of animals in t-1 and their survival rates.

for(t in 2:T){

meana[t] <- Na[t-1]+Ny[t-1]

meany[t] <- Nl[t-1]*0.5

meanl[t] <- Na[t-1]*phiaf[t-1]

Na[t] ~ dbin(phiaf[t-1], meana[t])

Ny[t] ~ dbin(phiy[t-1], meany[t])

Nl[t] ~ dbin(fecund[t], meanl[t])

#Observation process for the ground survey data, connecting ground counts (y data) to

the #numbers of animals in each stage class.

for(t in 1:T){

yca[t] ~ dnorm(Na[t], atauy)

Page 136: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

122

ycy[t] ~ dnorm(Ny[t], ytauy)

ycl[t] ~ dnorm(Nl[t], ltauy)

}

#################### TELEMETRY DATA ##############################

#Model estimating fecundity from telemetry data. l(t) is the number of marked females

#observed with a lamb in a given year and c(t) represents the number of marked females

#monitored in a given year.

for (t in 6:T){

l[t] ~ dbin(pl[t], c[t])

pl[t] <- fecund[t]

}

#Model estimating adult female survival from telemetry data. d[t,i] represents when

female i in #year t died (if applicable) as a function of the hazard rate in that year

(lambda[t]).

for(t in 4:(T-1)){

lambda[t] <- (-log(phiaf[t]))/365

for(i in 1:N){

d[t,i] ~ dweib(1,lambda[t])I(t.cen[t,i],)

}

}

#################### MARK RESIGHT DATA #############################

#Model for estimating the number of adult females from mark-resight data, modified

from #McClintock and Hoeting (2009). Mu[t] represents detection probability, s is the

Page 137: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

123

number of #marked females, k[t] is the number of sighting occasions, UF[t] is the number

of unmarked #females observed across those sighting occasions, and x[t,s] represents the

number of sightings #of individual s in year t.

for(t in 7:T){

mu[t]~dunif(0,1)

for(s in 1:S) {

x[t,s]~dbin(mu[t],k[t])

}

Uk[t]<-(Na[t]-m[t])*k[t]

UF[t]~dbin(mu[t],Uk[t])

}

#################### DERIVED PARAMETERS #########################

#Annual population growth rate of females (adults and yearlings).

for(t in 2:T){

growrate[t] <- (Na[t] + Ny[t])/(Na[t-1] + Ny[t-1])

}

} #END MODEL

####################### LANGLEY DATA ##############################

list(T=9, N=17, S=17,

yca=c(9, 11, 14, 20, 27, NA, 34, 34, 36), ycy=c(2, 4, 6, 7, 6, 6, 11, 10, 3), ycl=c(9, 10, 11,

13, 11, 25, 18, 17, 8),

l=c(NA, NA, NA, NA, NA, 7, 5, 8, 5), c=c(NA, NA, NA, NA, NA, 7, 8, 15, 12),

Page 138: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

124

m=c(NA, NA, NA, NA, NA, NA, 8, 17, 12), k = c(NA, NA, NA, NA, NA, NA, 3, 1, 2),

UF = c(NA, NA, NA, NA, NA, NA, 44, 15, 44),

d = structure(. Data = c(NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, 156, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

320, 216, NA, NA, NA, NA, NA, NA, NA, 228, NA, NA, 17, 188, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA),.Dim = c(9,17)),

t.cen = structure(. Data = c(0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0,

0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 166, 166, 166, 0, 0, 0, 0,

0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 365, 365, 365, 205, 205, 205, 205, 177, 0, 0, 0, 0, 0, 0, 0, 0, 0,

365, 365, 365, 365, 365, 0, 365, 365, 209, 0, 0, 0, 0, 0, 0, 0, 0, 365, 365, 365, 365, 365,

365, 365, 365, 324, 324, 245, 245, 245, 245, 245, 245, 245, 365, 365, 0, 0, 365, 365, 365,

365, 365, 365, 365, 0, 365, 365, 0, 0, 365, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0,

0),.Dim = c(9,17)),

x = structure(.Data=c(NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

Page 139: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

125

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA,

NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, NA, 2, 1, 1, 3, 2,

2, 2, 2, NA, NA, NA, NA, NA, NA, NA, NA, NA, 1, 1, 1, 0, 1, 1, 1, 1, 0, 0, 0, 0, 1, 1, 0,

0, 0, 2, 2, 2, 1, 2, 2, 2, 2, 1, 0, 0, 0, NA, NA, NA, NA, NA),.Dim=c(9,17)))

####################### STARTING VALUES ############################

#Starting Na, Ny, and Nl values were fixed to reflect priors for survival and fecundity

rates and #which were large enough to be used in the binomial mark-resight function.

Ideally, we would #have used count data for starting N values, but this was not possible

due to our small population #sizes and demographic stochasticity in lamb sex ratio

(complicated by our use of a 50:50 sex #ratio for lambs and the fact that binomial

survival and fecundity functions need plausible #starting values).

list(atauy=1, ytauy=1, ltauy=1, yca=c(NA, NA, NA, NA, NA, 31, NA, NA, NA),

Na=c(34, 34, 34, 34, 34, 34, 34, 34, 34), Ny=c(4, 4, 4, 4, 4, 4, 4, 4, 4), Nl=c(15, 15, 15,

15, 15, 15, 15, 15, 15), mu=c(NA, NA, NA, NA, NA, NA, 0.5, 0.5, 0.5), phiaf=c(0.9, 0.9,

0.9, 0.9, 0.9, 0.9, 0.9, 0.9), phiy=c(0.5, 0.5, 0.5, 0.5, 0.5, 0.5, 0.5, 0.5), fecund=c(NA, 0.5,

0.5, 0.5, 0.5, 0.5, 0.5, 0.5, 0.5)))

Page 140: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

126

APPENDIX G. Estimated annual vital rates and population growth rates (λ) from

Bayesian state-space models of Sierra Nevada bighorn sheep populations (standard

deviations in parentheses). Due to differences in the timing of annual surveys, vital rates

at Warren and Langley (post-birth pulse) were adult female survival (ΦA), yearling

survival (ΦY) and fecundity (F), while vital rates at Wheeler (pre-birth pulse) were adult

female survival (ΦA), two-year-old survival (ΦT) and recruitment (R).

Population Year ΦA ΦY/ ΦT F/R λ

Warren

1988 0.89 (0.08) 0.77 (0.18) -- --

1989 0.93 (0.05) 0.75 (0.17) 0.69 (0.15) 1.12 (0.06)

1990 0.91 (0.07) 0.79 (0.17) 0.83 (0.11) 1.18 (0.04)

1991 0.94 (0.05) 0.83 (0.15) 0.73 (0.13) 1.22 (0.06)

1992 0.96 (0.04) 0.50 (0.29) 0.77 (0.11) 1.20 (0.05)

1993 0.55 (0.19) 0.54 (0.29) 0.48 (0.11) 1.11 (0.09)

1994 0.65 (0.19) 0.14 (0.13) 0.59 (0.20) 0.65 (0.19)

1995 0.93 (0.06) 0.50 (0.24) 0.35 (0.17) 0.68 (0.18)

1996 0.88 (0.10) 0.59 (0.17) 0.85 (0.11) 1.03 (0.05)

1997 0.39 (0.12) 0.15 (0.13) 0.67 (0.16) 0.19 (0.08)

1998 0.73 (0.17) 0.30 (0.23) 0.50 (0.25) 0.39 (0.07)

1999 0.85 (0.13) 0.30 (0.23) 0.51 (0.25) 0.82 (0.13)

2000 0.92 (0.06) 0.36 (0.27) 0.40 (0.28) 0.98 (0.05)

2001 0.86 (0.11) 0.80 (0.16) 0.83 (0.14) 0.99 (0.05)

2002 0.93 (0.07) 0.74 (0.20) 0.67 (0.20) 1.53 (0.09)

Page 141: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

127

2003 0.66 (0.12) 0.43 (0.22) 0.81 (0.16) 1.21 (0.06)

2004 0.82 (0.13) 0.68 (0.21) 0.74 (0.19) 0.80 (0.11)

2005 0.72 (0.13) 0.74 (0.20) 0.50 (0.16) 1.12 (0.13)

2006 0.81 (0.09) 0.73 (0.20) 0.65 (0.15) 1.01 (0.13)

2007 0.54 (0.13) 0.80 (0.16) 0.78 (0.12) 1.04 (0.13)

2008 -- -- 0.63 (0.19) 0.86 (0.14)

Wheeler

1981 0.79 (0.15) 0.61 (0.27) -- --

1982 0.85 (0.12) 0.45 (0.26) 0.70 (0.16) 0.99 (0.15)

1983 0.82 (0.14) 0.70 (0.23) 0.58 (0.17) 1.04 (0.15)

1984 0.76 (0.17) 0.53 (0.29) 0.39 (0.14) 1.06 (0.15)

1985 0.77 (0.16) 0.55 (0.28) 0.55 (0.28) 0.88 (0.17)

1986 0.79 (0.16) 0.55 (0.27) 0.53 (0.27) 0.98 (0.22)

1987 0.75 (0.20) 0.57 (0.28) 0.71 (0.17) 0.97 (0.19)

1988 0.76 (0.19) 0.55 (0.29) 0.55 (0.29) 1.01 (0.23)

1989 0.74 (0.20) 0.56 (0.28) 0.56 (0.28) 0.95 (0.25)

1990 0.77 (0.18) 0.60 (0.27) 0.60 (0.28) 0.94 (0.26)

1991 0.82 (0.14) 0.55 (0.27) 0.54 (0.27) 1.02 (0.25)

1992 0.73 (0.19) 0.54 (0.29) 0.52 (0.20) 0.87 (0.18)

1993 0.75 (018) 0.56 (0.28) 0.56 (0.28) 0.98 (0.21)

1994 0.78 (0.16) 0.52 (0.28) 0.51 (0.28) 0.96 (0.25)

1995 0.86 (0.12) 0.58 (0.26) 0.67 (0.19) 1.16 (0.19)

1996 0.89 (0.10) 0.70 (0.23) 0.54 (0.17) 1.14 (0.15)

Page 142: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

128

1997 0.91 (0.08) 0.76 (0.20) 0.60 (0.16) 1.22 (0.08)

1998 0.94 (0.06) 0.80 (0.17) 0.51 (0.14) 1.21 (0.06)

1999 0.95 (0.05) 0.66 (0.23) 0.82 (0.11) 1.21 (0.02)

2000 0.84 (0.10) 0.81 (0.16) 0.93 (0.06) 1.18 (0.07)

2001 0.88 (0.08) 0.51 (0.19) 0.91 (0.06) 1.07 (0.09)

2002 0.94 (0.05) 0.71 (0.18) 0.81 (0.09) 1.23 (0.10)

2003 0.87 (0.07) 0.71 (0.18) 0.69 (0.09) 1.07 (0.07)

2004 0.86 (0.06) 0.53 (0.20) 0.61 (0.08) 1.02 (0.07)

2005 0.95 (0.04) 0.81 (0.14) 0.73 (0.07) 1.24 (0.07)

2006 0.80 (0.06) 0.48 (0.19) 0.58 (0.08) 0.91 (0.06)

2007 0.89 (0.05) 0.59 (0.22) 0.41 (0.07) 0.87 (0.06)

2008 0.95 (0.04) 0.45 (0.25) 0.28 (0.06) 1.00 (0.06)

2009 -- -- 0.47 (0.07) 1.01 (0.05)

Langley

1987 0.80 (0.14) 0.51 (0.27) -- --

1988 0.82 (0.14) 0.51 (0.28) 0.53 (0.28) 0.94 (0.15)

1989 0.83 (0.13) 0.52 (0.28) 0.53 (0.28) 0.97 (0.16)

1990 0.80 (0.16) 0.55 (0.28) 0.71 (0.17) 0.97 (0.16)

1991 0.80 (0.16) 0.52 (0.28) 0.53 (0.28) 0.99 (0.18)

1992 0.80 (0.15) 0.51 (0.28) 0.52 (0.28) 0.94 (0.17)

1993 0.80 (0.16) 0.52 (0.28) 0.54 (0.28) 0.94 (0.17)

1994 0.77 (0.18) 0.55 (0.28) 0.57 (0.27) 0.95 (0.18)

1995 0.81 (0.15) 0.34 (0.26) 0.35 (0.27) 0.96 (0.19)

Page 143: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

129

1996 0.88 (0.10) 0.59 (0.24) 0.39 (0.19) 0.88 (0.13)

1997 0.89 (0.10) 0.76 (0.18) 0.68 (0.16) 1.04 (0.08)

1998 0.92 (0.07) 0.40 (0.27) 0.21 (0.17) 1.24 (0.07)

1999 0.93 (0.06) 0.61 (0.21) 0.59 (0.18) 0.99 (0.03)

2000 0.96 (0.03) 0.75 (0.17) 0.72 (0.14) 1.14 (0.03)

2001 0.96 (0.03) 0.87 (0.11) 0.88 (0.09) 1.25 (0.03)

2002 0.88 (0.08) 0.89 (0.10) 0.91 (0.07) 1.27 (0.05)

2003 0.91 (0.07) 0.81 (0.14) 0.79 (0.12) 1.19 (0.08)

2004 0.95 (0.04) 0.83 (0.13) 0.63 (0.12) 1.15 (0.06)

2005 0.86 (0.06) 0.86 (0.10) 0.93 (0.05) 1.13 (0.06)

2006 0.93 (0.05) 0.90 (0.09) 0.72 (0.08) 1.16 (0.08)

2007 0.76 (0.06) 0.39 (0.15) 0.55 (0.08) 1.16 (0.06)

2008 -- -- 0.33 (0.08) 0.86 (0.05)

Page 144: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

130

CHAPTER 4

FROM LAMBS TO LAMBDA: EFFECTS OF INBREEDING

DEPRESSION ON RECOVERY OF ENDANGERED BIGHORN SHEEP

HEATHER E. JOHNSON, University of Montana, Wildlife Biology Program, College

of Forestry and Conservation, Missoula, MT 59812, USA

L. SCOTT MILLS, University of Montana, Wildlife Biology Program, College of

Forestry and Conservation, Missoula, MT 59812, USA

JOHN D. WEHAUSEN, White Mountain Research Station, University of California,

3000 East Line Street, Bishop, CA 93514, USA

THOMAS R. STEPHENSON, Sierra Nevada Bighorn Sheep Recovery Program,

California Department of Fish and Game, 407 West Line Street, Bishop, CA

93514, USA

GORDON LUIKART, University of Montana, Division of Biological Sciences,

Missoula, MT 59812, USA

ABSTRACT

Small populations can suffer from inbreeding depression as matings between

related individuals reduce fitness, decrease population performance, and increase

extinction risk. While evidence of inbreeding depression is commonly detected in fitness

components in animals, the consequences for population performance in endangered

species are rarely assessed. We examined genetic diversity and inbreeding depression in

federally endangered Sierra Nevada bighorn sheep (Ovis canadensis sierrae), the rarest

Page 145: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

131

subspecies of bighorn sheep in North America. Our objectives were to 1) characterize

neutral and potentially adaptive genetic variation in this subspecies; 2) test for evidence

of inbreeding depression in vital rates driving population performance; 3) evaluate

whether inbreeding depression might limit subspecies recovery; and 4) examine the

potential for genetic management to stimulate population growth. We found that genetic

variation of Sierra Nevada bighorn sheep populations was among the lowest reported for

any bighorn sheep populations (with 29 loci mean heterozygosity ranged from 0.33 to

0.39), and exhibited inbreeding depression in adult female fecundity. Despite this,

matrix-based population projection models demonstrated that the costs of inbreeding

were not predicted to appreciably inhibit recovery in the next 3 decades. Furthermore,

simulations of genetic rescue with bighorn sheep within the Sierra Nevada did not

indicate significant population-level effects within time periods of interest to managers.

Only simulations of the effects of augmenting endangered populations with genetic

diversity from other subspecies predicted dramatic increases in population performance, a

scenario that is not currently a management option. While management activities should

minimize future losses of genetic variation in this subspecies, genetic effects within these

populations - either negative (inbreeding depression) or positive (genetic rescue) – appear

unlikely to substantially compromise or stimulate short-term conservation efforts.

INTRODUCTION

Small populations can suffer from inbreeding depression as matings between

related individuals reduce key fitness traits (Keller and Waller 2002). Inbreeding

depression can result from the increased expression of deleterious recessive alleles or

from a decrease in heterozygosity at loci with heterozygote advantage (Charlesworth and

Page 146: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

132

Charlesworth 1987, Roff 2002). The specific costs of inbreeding depression are difficult

to predict, because populations exhibit different effects depending on their demographic

history, the genetic diversity of founders, the occurrence of purging, the severity of

environmental conditions, and chance (Lacy et al. 1996, Lacy and Ballou 1998, Bijlsma

et al. 2000, Lesbarrères et al. 2005). Nevertheless, inbreeding depression is capable of

decreasing population performance, reducing evolutionary potential, and increasing

extinction risk (Newman and Pilson 1997, Saccheri et al. 1998, Westemeier et al. 1998,

Hogg et al. 2006). As a result, genetic factors are a major consideration in the

conservation and recovery of threatened and endangered populations (Hedrick and

Kalinowski 2000, Keller and Waller 2002).

Evidence of inbreeding is commonly based on heterozygosity-fitness correlations

(HFCs; Da Silva et al. 2006, Acevedo-Whitehouse et al. 2006, Ortego et al. 2007,

Mainguy et al. 2009) and while the consequences of HFCs for population viability are

often emphasized, their effects are rarely assessed (Keller and Waller 2002). A major

limitation in linking studies of inbreeding depression to population performance is the use

of indirect fitness correlates, such as morphometric (e.g. body size) or physiological traits

(e.g. parasite loads). These traits may be weakly correlated to individual fitness

(Chapman et al. 2009) and do not easily scale-up to population-level assessment.

To properly examine the influence of inbreeding on populations, genetic

variability must be evaluated relative to direct fitness measures or vital rates (survival and

reproductive rates) that have the greatest impact on the growth rates of populations (Mills

and Smouse 1994). While HFCs are often related to vital rates (Coulson et al. 1999,

Lesbarrères et al. 2005, Ortego et al. 2007, Cohas et al. 2009, Mainguy et al. 2009), this

Page 147: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

133

inbreeding depression is rarely connected to population performance. Because different

vital rates disproportionately affect populations (Mills 2007), significant HFCs – even

those related to vital rates –may not affect population growth rates in any measurable

way. Connecting vital rate-specific inbreeding costs to population growth is critical for

conservation because it allows managers to assess whether genetic factors are limiting

population performance and the potential for genetic rescue to stimulate recovery efforts

(Tallmon et al. 2004).

We examined inbreeding depression in federally endangered Sierra Nevada

bighorn sheep (Ovis canadensis sierrae; hereafter “bighorn sheep”). This is the rarest

subspecies of bighorn sheep in North America, totaling approximately 400 individuals in

2009 (California Dept. Fish and Game, unpublished data). A microsatellite analysis

included in the recovery plan for the subspecies (U.S. Fish and Wildlife Service 2007)

found heterozygosity levels for Sierra Nevada bighorn sheep to be among the lowest

reported for any wild bighorn sheep population in the U.S., comparable to values from

Rocky Mountain bighorn sheep (O. c. canadensis) in the National Bison Range where

inbreeding depression was detected in multiple fitness traits (Hogg et al. 2006). Sierra

Nevada bighorn sheep populations have a history of being bottlenecked, demographically

isolated, and small, raising significant concerns with conservation practitioners over the

potential for genetic factors to limit recovery.

To identify whether inbreeding depression may inhibit recovery efforts and

precipitate the need for genetic management, we assessed genetic variation of bighorn

sheep with respect to adult survival and fecundity rates, the vital rates found to explain

>80% of the variation in the growth rates of bighorn sheep populations (Johnson et al.

Page 148: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

134

2010). We evaluated genetic variation at neutral markers – the workhorse of classical

population genetics and inbreeding theory (Wright 1951) – and potentially adaptive

markers expected to be more closely tied to individual fitness (Luikart et al 2003).

Specifically, our objectives were to 1) characterize neutral and potentially adaptive

genetic variation in Sierra Nevada bighorn sheep, 2) test for evidence of inbreeding

depression in bighorn sheep vital rates (adult survival and fecundity), 3) evaluate whether

inbreeding depression may limit subspecies recovery, and 4) examine the ability of

genetic management to alleviate inbreeding effects and stimulate population growth.

METHODS

Study Populations

Baxter-Sawmill (hereafter “Baxter”) was the only major population of bighorn

sheep remaining in the Sierra Nevada in the late 1970’s, estimated to have approximately

250 individuals (Wehausen 1980). That herd was subsequently used as source stock for

reintroducing three additional populations; Wheeler, Langley, and Mono Basin in 1979,

1980, and 1986, respectively (Bleich et al. 1990; Fig. 4.1). By 1998, field surveys

revealed that only 125 adult bighorn sheep could be accounted for across all populations,

the lowest number ever recorded (U.S. Fish and Wildlife Service 2007). As a result,

Sierra Nevada bighorn sheep received emergency listing as an endangered species by the

federal government in 1999. Since then Wheeler, Baxter and Langley have increased

considerably in size, while Mono Basin has remained small (≤11 adult females; Fig. 4.2).

Each population is geographically isolated so their dynamics are independent. Detailed

information on the populations and study area is described in Johnson et al. (2010).

Page 149: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

135

Field Sampling

California Dept. of Fish and Game captured bighorn sheep via helicopter net-gun

operations between 1999 and 2009. Each captured animal was sampled for blood for

DNA extraction by filling a 10cc EDTA tube from the jugular vein. Animals were also

uniquely marked with a radio-collar that emitted a mortality signal. Radio-collared sheep

were monitored twice/month for survival using ground and fixed-wing aerial telemetry,

and adult females were annually observed in July-August for reproductive information.

The lambing period for bighorn sheep in the Sierra Nevada occurs primarily from mid-

April through mid-June, with females giving birth to one offspring/year (Wehausen

1996). A radio-collared adult female was recorded as successfully reproducing if

observed nursing a lamb during summer population ground surveys.

From 1995-2007 annual minimum counts of each bighorn sheep population were

performed by experienced observers, who systematically hiked and scanned each herd

area for bighorn sheep by sex and stage class. Due to small population sizes and repeated

surveys, in many cases counts were successful at being complete, or near-complete,

censuses of numbers in each stage class. Consecutive ground counts were used to

estimate means and variances of all vital rates needed to parameterize bighorn sheep

population matrix models (Johnson et al. 2010).

Microsatellite Analyses

We extracted DNA from a total of 128 unique individuals, 26 bighorn sheep in

Baxter (21 females and 5 males), 21 in Langley (all females), 29 in Mono Basin (13

females and 16 males), and 52 in Wheeler (29 females and 23 males). Blood samples

Page 150: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

136

were centrifuged, and the buffy coat was removed for DNA extraction using the Qiagen

blood and tissue kit™.

We used polymerase chain reactions (PCRs) to amplify microsatellite markers,

using 2 replicates and an additional 2 replicates to rescore any discrepancies (Wehausen

et al. 2004). PCRs were run in 96 well polycarbonate plates and each plate included a

positive and negative control. PCR volume was 15 - 20 uL and included 1X PCR buffer

(Applied Biosystems™), 2.6-3.75 mM MgCl2 depending on locus, 160mM dNTP,

400ug/ml bovine serum albumin (New England Biolabs™), 24-320 nM each primer

depending on locus, 0.035/ul taq polymerase (Amplitaq GoldTM

), 8.67-26uL/ml DNA,

and 5% extra H2O to counteract drydown. PCR cycling was performed with a 96C heated

lid and 40 cycles of 95C for 30s, 51-62C depending on locus for 40s, and 72C for 30s

after an initial 7.5 min at 93C to activate the taq polymerase. Forward primers were

tagged with florescent dye labels and PCR products were electrophoresed on an ABI

PRISM 377 DNA sequencer using tamra 350 (Applied Biosystems) size standards with

different loci in adjacent lanes for 96 lane runs. Many loci were run in various PCR

multiplexes of 2 and occasionally 3 loci. The last runs included 2 multiplexes each with 7

loci using the Qiagen Multiplex MixTM

in 15 uL reactions. We manually scored

chromatograms with GeneScan 3.12 software (Applied Biosystems).

We genotyped 29 microsatellite loci known to be polymorphic in Ovis species; 18

loci were assumed neutral (not in or near genes) and 11 loci were located in potentially

adaptive genes (see Appendix D for details). The neutral loci were AE16, AE129,

BM4513, CP20, CP128, FCB11, FCB193, FCB266, FCB304, HH47, HH62, HH64,

JMP29, MAF33, MAF36, MAF48, MAF65, and MAF209. Candidate adaptive

Page 151: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

137

microsatellite markers were ADC, BL4, IFNG, KERA, MHCI, MMP9, OIFNG, OLA,

SOMA, TCRB, and TGLA387. From those loci that met neutral expectations (see

Appendix D), we calculated general measures of genetic variation. For each population

we calculated observed and expected heterozygosity, allelic richness, and genetic

differentiation (FST), specifically evaluating whether genetic variation was reduced in the

reintroduced herds relative to the source herd.

Testing for Inbreeding Depression

We used two metrics of individual genetic variation to test for inbreeding

depression, heterozygosity (h; the proportion of typed loci that are heterozygous for an

individual; Mitton 1993) and mean d2 (the average squared distance in repeat units

between two alleles at any typed locus for an individual; Coulson et al. 1998). We

evaluated these measures across all polymorphic loci that fit neutral expectations to

detect “multilocus” effects, and individually for each locus to detect “locus-specific”

effects (Da Silva et al. 2009).

We then examined whether individual genetic variation was associated with the

vital rates most important to bighorn sheep dynamics (Johnson et al. 2010): adult survival

and fecundity of adult females. For each vital rate, we first developed a minimal non-

genetic model, testing the explanatory power of key covariates known or hypothesized to

be important (Mainguy et al. 2009, Da Silva et al. 2009). We included population as a

categorical covariate, using Baxter as the reference class. We also evaluated age, and age2

to account for potential asymptotic or curvilinear effects on vital rates. If the addition of

age2 improved model fit over age alone, we retained the quadratic term in any model

including the main age effect. Models were evaluated using Akaike Information Criterion

Page 152: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

138

(Burnham and Anderson 2002) with the small sample size correction (AICc). Models

with AICc values ≤ 2 relative other models were considered to be a better fit to the data.

All statistical analyses were conducted in STATA 9.0 (StataCorp 2007).

We used Cox proportional hazards models to evaluate adult survival (Cox 1972,

Cleves et al. 2008), an approach that readily accounts for staggered entry of marked

animals and the evaluation of covariates. In addition to age, age2, and population as

covariates, we included sex and tested for a sex by age interaction. We used a study-

based time scale for analysis (Fieberg and DelGiudice 2009), and considered each animal

to be at risk from the time they were radio-collared until their collar was heard emitting a

mortality signal.

To model adult female fecundity we used logistic mixed effects models (Rabe-

Hesketh and Skrondal 2008). Because each radio-collared female was monitored for

annual fecundity between 1 and 7 years (depending on the length of time collared), we

used a random effect to account for individual differences in data duration.

Once a best non-genetic model was identified, we included genetic factors. Mono

Basin was suspected to be particularly vulnerable to inbreeding depression due to its

chronic small size (Fig. 4.2) and severe environmental conditions (Johnson et al. In

Press). To determine whether genetic factors were evident in only particular populations

we also considered an interaction between population and genetic factors. We used

deviance residuals, link tests, and χ2 statistics to examine model fit (McCullagh and

Nelder 1989, Cleves et al. 2008). For survival models, we also evaluated the proportional

hazards assumption by testing for an interaction between covariates and time, and by

checking Schoenfeld residuals (Cleves et al. 2008).

Page 153: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

139

We discerned whether significant vital rate-genetic variation models were driven

by genome-wide or locus-specific effects. If a locus-specific effect significantly

improved model fit we would re-calculate multilocus h and d2 without that locus and re-

test our multilocus vital rate model. If the removal of specific loci altered the relationship

of a vital rate to multilocus genetic variation we assumed our results were driven by local,

as opposed to genome-wide, effects.

Quantifying Population-Level Effects of Inbreeding

We predicted the demographic consequences of inbreeding depression by

incorporating vital rate inbreeding costs into matrix projection models. Specifically we

evaluated the effects of inbreeding on population recovery by 1) estimating the long-term

influence of inbreeding depression on population growth rates, given current levels of

heterozygosity and future expected losses of heterozygosity, and 2) simulating a

management-induced increase in heterozygosity (“genetic rescue”) and its predicted

effects on population growth.

To assess the effects of inbreeding depression on population performance, we first

determined the rate at which heterozygosity would be expected to be lost in bighorn

sheep populations. We used the program LDNe (Waples 2006) to estimate effective

population size (Ne), using 0.02 as the lowest allele frequency in the analysis. We

estimated expected loss of heterozygosity per generation (Wright 1951):

012

11 H

NH

e

where H0 is the initial multilocus heterozygosity of the population, and H1 is the

heterozygosity expected after one generation. We used a generation time of 6 years as

this was the average age of female bighorn sheep having lambs in our study. Using our

Page 154: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

140

vital rate models, we then estimated the percent decrease in vital rates expected to occur

(per generation) given the predicted loss in heterozygosity. We modeled the impact of

this effect for 1, 5, and 10 bighorn sheep generations. We simulated these effects for

Mono Basin and Langley because 1) detailed demographic data are available for both

populations, 2) these herds represent the minimum and maximum population growth rates

observed in Sierra Nevada bighorn sheep (Fig. 4.2), and 3) individual survival and

fecundity data from radio-collared sheep directly aligned with population parameters

estimated from consecutive annual ground surveys (Johnson et al. 2010).

We used vital rate means and variances estimated from population ground data to

parameterize stochastic matrix models (Johnson et al. 2010) and initialized population

vectors based on 2008 survey counts. Matrix stage classes were those easily observed

during annual ground surveys: adults, yearlings, and lambs. Vital rates estimated from

those stage classes were adult survival, yearling survival, and adult fecundity. We

calculated the mean stochastic population growth rate (λs) and expected median

population size (Nt) based on 1,000 replicate simulations using an exponential growth

model. We first projected populations assuming no inbreeding depression (“Baseline”

scenario). Next we decremented vital rates to incorporate the accumulation of measured

inbreeding costs.

To simulate genetic rescue or management effects (Tallmon et al. 2004) we

modeled population responses to increasing average heterozygosity. We only conducted

management simulations for Mono Basin because this was the population of highest

management concern and the only herd that has not substantially increased in recent

years (Fig 4.2). We used the same matrix approach described above, but rather than

Page 155: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

141

decrement vital rates based on inbreeding depression, we increased them in accordance

with two levels of enhanced heterozygosity. We first simulated the growth rate of Mono

Basin if average heterozygosity (currently h = 0.43) was equal to that of the source herd,

Baxter (h = 0.48), our measure of conducting genetic rescue “within” the Sierra Nevada

range. Second, we modeled the growth rate of Mono Basin if average heterozygosity

could be boosted to 0.59, the mean heterozygosity of 8 other populations of Rocky

Mountain and desert bighorn sheep (O. c. nelsoni) reported in Forbes and Hogg (1999).

In practice, given inherently low genetic variation in Sierra Nevada bighorn sheep,

achieving this heterozygosity level would require translocations from other bighorn

subspecies and represented the effects of using “outside” genetic variation to conduct

genetic rescue.

RESULTS

Genetic Variation

Of the 29 loci genotyped, 4 were monomorphic (FCB128, FCB266, IFNG and

OINF), while the 25 polymorphic loci had 2 to 5 alleles per locus (Appendix E; mean =

2.84, SE = 0.17). Only 10 alleles had frequencies <5% in any population, and only 3

private alleles were detected, one in each of the Baxter, Wheeler, and Mono Basin

populations. Of the polymorphic loci, 17 met neutral expectations (12 neutral loci, 5

potentially adaptive; Appendix D) and were used to calculate metrics of genetic variation,

regardless of whether they were originally considered neutral or potentially adaptive.

Locus-specific expected heterozygosity ranged from 0.09 (KERA) to 0.65

(MAF36; Appendix E). Using all loci (monomorphic and polymorphic) to estimate

population-specific expected heterozygosity, values ranged from 0.33 to 0.39 (Table 4.1).

Page 156: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

142

Using only polymorphic loci they ranged between 0.40 and 0.48 (Table 4.1). Values were

lowest in the reintroduced population Langley, and highest in the source population

Baxter (Table 4.1). Only Langley had significantly reduced heterozygosity from the

source herd (Wilcoxon signed rank test p = 0.02). Allelic richness was similar among

populations, with the mean number of alleles per locus ranging from 2.35 to 2.59. After

accounting for sample size differences, only the reintroduced Wheeler population was

significantly different from the source population (Wilcoxon signed rank test p = 0.03).

All herds showed significant genetic differentiation (all Fisher’s exact tests χ2 > 117, df =

34, p < 0.01) with FST values ranging from 0.04 to 0.08 (Table 4.2; global FST = 0.06).

The least differentiation was between Baxter and the reintroduced herds, while there was

greater differentiation among the reintroduced herds.

Detecting Inbreeding Depression

Individual multilocus heterozygosity was between 0.24 and 0.76 with a mean at

Baxter of 0.47 (SE = 0.020), at Langley of 0.41 (SE = 0.022), at Mono Basin of 0.48 (SE

= 0.024), and at Wheeler of 0.44 (SE = 0.016). Multilocus mean d2 ranged between 1.65

and 52.0 with a mean at Baxter of 26.35 (SE = 2.16), at Langley of 24.81 (SE = 2.40), at

Mono Basin of 25.58 (SE = 2.05), and at Wheeler of 24.54 (SE = 1.53). Of the 127

bighorn sheep monitored for survival, there were 51 deaths. Cox proportional hazards

analyses identified the best non-genetic model of adult survival to include only age as a

covariate (Table 4.3), with mortality risk increasing in older animals (Table 4.4).

Multilocus measures of genetic variation did not improve the fit of the survival model,

and h of AE16 was the only significant locus-specific effect (Appendix F). We obtained

194 observations of annual lambing status from 75 radio-collared females. Our best non-

Page 157: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

143

genetic fecundity model included age and age2 as covariates (Tables 4.3 and 4.4). The

annual probability of having a lamb was lower during very young and old ages, peaking

at intermediate ages. The addition of multilocus h improved model fit, as individuals

across all populations with higher heterozygosity had higher probabilities of successfully

reproducing (Fig. 4.3; Tables 4.3 and 4.4). Multilocus d2 did not improve the non-genetic

model, nor did genetic variation measures of any individual loci (Appendix F).

Effect of Inbreeding Depression on Population Dynamics

Effective population size was estimated to be 10.7 for Mono Basin and 13.2 for

Langley, reasonable estimates given 2008 ground counts were 32 for Mono Basin and 78

for Langley. For these Ne values genetic drift would decrease heterozygosity by

0.020/generation in Mono Basin and by 0.015/generation in Langley. Coupling these

losses-of-heterozygosity to our field-based inbreeding estimates translated into a 1.2%

decrease in annual fecundity/generation for Mono Basin and a 1.4% decrease for

Langley.

When mean fecundity values were decremented in population projection models,

growth rates did not appreciably decline (Table 4.5). Even after 10 generations (60

years), costs of inbreeding were only estimated to decrease λs by 0.7% for Mono Basin

(from 1.019 to 1.012) and by 0.5% for Langley (from 1.180 to 1.174). Stochastic lambda

increased by ≤ 0.7% when we simulated “within Sierra Nevada” genetic management at

Mono Basin by increasing mean heterozygosity to 0.48 (Table 4.5), yielding only modest

increases in median predicted population sizes (Fig. 4.4). Lambda increased by

approximately 1% when we modeled an increase in mean heterozygosity of 0.59,

Page 158: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

144

simulating the addition of “outside” genetic variation from other subspecies (Table 4.5),

resulting in dramatic increases in long-term population sizes (Fig. 4.4).

DISCUSSION

Despite strong evidence of detectable inbreeding depression in bighorn sheep

fecundity rates (Fig. 4.3), inbreeding costs are not likely to inhibit population recovery in

the short term. Although the population ecology literature has accepted as mainstream the

idea that all vital rates are not created equal in their effects on population growth (Morris

and Doak 2002, Mills 2007), this principle has not been applied to most studies of

inbreeding depression. Rather, such studies often infer from statistically significant HFCs

that population growth or viability is being compromised. Fecundity, for example, is

often a target of studies on inbreeding depression (Ralls et al. 1988, Heath et al. 2002,

Ortego et al. 2007), and following convention, our observation of inbreeding in bighorn

sheep fecundity rates would be taken as a conservation alarm. By applying the observed

fecundity decrement to models of population growth, however, we found that variation in

other vital rates ameliorated the genetic effects on reproduction (Table 4.5). This is not to

say that genetic factors are not important to endangered populations, but that their

influence will depend on the values of key vital rates driving population performance.

This distinction between detecting inbreeding depression in a vital rate and inbreeding

depression in population growth underscores the importance of assessing field-based

inbreeding costs relative to population dynamics to most effectively manage small and

endangered populations (Oostermeijer et al. 2003, Beissinger et al. 2008).

Initiating genetic rescue activities within Sierra Nevada bighorn sheep

populations was also not expected to have a substantial benefit in spurring bighorn sheep

Page 159: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

145

population growth rates (Fig. 4.4). Significant increases in population size only occurred

if we simulated heterozygosity levels representative of Rocky Mountain or desert bighorn

sheep populations (Fig. 4.4), a management option that is not being considered at this

time. It is important to recognize that our simulations did not model heterosis or hybrid

vigor effects, where fitness traits may more dramatically improve with an influx of new

genetic material, particularly if populations have a fixed genetic load (Tallmon et al.

2004). Because Sierra Nevada bighorn sheep populations have inherently low genetic

variation and similar allele frequencies, such demographic effects are not expected unless

bighorn from other subspecies are translocated into the range (Hogg et al. 2006). As a

result, bringing in “outside” genetic variation could have a greater effect on populations

than we predicted, rendering our estimates conservative. Currently, populations of Sierra

Nevada bighorn sheep that are increasing have vital rates comparable to other bighorn

populations (Johnson et al. 2010), suggesting that a deleterious fixed load is not

hampering recovery at this time. Given, however, that this subspecies already has

reduced genetic variation, a novel disease event or change in environmental conditions

could more dramatically affect these populations than otherwise expected. To maximize

the adaptive potential of this subspecies, management strategies should focus on

maintaining genetic variation by restoring gene flow among existing herds.

While we predicted that inbreeding depression would have a negligible effect on

near-term recovery, patterns were consistent with genome-wide inbreeding expectations.

Fecundity was associated with multilocus heterozygosity even after accounting for locus-

specific effects. We had expected to find more locus-specific effects, given that the

potentially adaptive markers we tested have been associated with fitness traits in other

Page 160: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

146

ungulates. The only marker significantly associated with SNBS vital rates was h of

AE16, which improved the fit of the survival model. AE16 is supposedly a neutral

marker. This locus-specific effect may either reflect physical linkage to a gene under

selection, that this marker has been inappropriately assumed neutral, or a chance

outcome. Furthermore, while there was strong evidence for HFCs in fecundity we did not

detect HFCs in adult survival. This follows observations that inbreeding depression is

stronger in younger age classes than older ones (David and Jarne 1997, Cohas et al.

2009), as individuals with unfit genotypes are likely eliminated from the population early

in life.

In contrast to multilocus h, multilocus d2 was not associated with bighorn sheep

vital rates. Mean d2 is thought to measure the genetic distance between the gametes that

derived an individual, reflecting the extent of outbreeding (Coulson et al. 1998). While

the relevance of d2 in inbreeding studies has been questioned (Hedrick et al. 2001), the

metric has been associated with fitness traits in several species (Coulson et al. 1999,

Hoglund et al. 2002, Da Silva et al. 2009). Investigators have suggested that h tends to

outperform d2 in detecting inbreeding depression (Slate and Pemberton 2002, Coltman

and Slate 2003), and that d2

may be most informative in populations that have

experienced recent admixture or long-distance immigrants (Da Silva et al. 2009). Given

that our populations are assumed to have been demographically isolated for several

generations, our results follow expected patterns.

While inbreeding studies have questioned the use of molecular markers as a proxy

for the pedigree inbreeding coeffient (f; Balloux et al. 2004, Slate et al. 2004), the

demographic characteristics of small and endangered populations likely increase the

Page 161: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

147

reliability of HFCs in elucidating detrimental inbreeding effects (Grueber et al. 2008).

Endangered populations are often bottlenecked, with low genetic variation, and few

founders. They also have increased levels of inbreeding, variation in pedigree f, linkage

disequilibrium, and environmental stress; all characteristics found to increase the strength

of the relationship between pedigree f and multilocus h (Bierne et al. 2000, Balloux et al.

2004, Aparicio et al. 2007, Grueber et al. 2008, Hansson and Westerberg 2008, Ruiz-

Lopez et al. 2009). Additionally, bighorn sheep populations are small, highly structured,

and polygynous, features that further reinforce the relevance of HFCs in studies of

inbreeding depression (Balloux et al. 2004, Mainguy et al. 2009). While pedigree data are

clearly the gold standard for such studies, molecular data can still provide important

information about small and endangered populations for which little long-term pedigree

data may exist.

Measures of genetic variation and differentiation of bighorn sheep herds closely

matched known population histories. Sierra Nevada bighorn sheep have the smallest

population size and distribution of any bighorn subspecies and as expected, their genetic

variation was lower than that reported for Rocky Mountain or desert bighorn sheep

populations (Forbes et al. 1995, Boyce et al. 1997, Forbes and Hogg 1999, Gutiérrez-

Espeleta et al. 2000). Genetic variation in Sierra Nevada bighorn was just slightly higher

than values reported for Red Rock captive breeding facility in New Mexico (Gutiérrez-

Espeleta et al. 2000, U.S. Fish and Wildlife Service 2007) and similar to those at the

National Bison Range, where inbreeding depression was detected in multiple fitness traits

(Hogg et al. 2006). Congruent with the translocation record, Baxter, the source

population, had the highest genetic variation, with the reintroduced populations having

Page 162: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

148

lower variation. While FST values were significant among herds it is uncertain whether

population differentiation is an artifact of founder effects or genetic drift.

While we expected that Sierra Nevada bighorn sheep should be highly susceptible

to inbreeding depression, measured inbreeding costs are not expected to hinder

conservation objectives in the next few decades, given similar conditions. We highlight

the importance of not only detecting HFCs in endangered populations, but coupling field-

based estimates of inbreeding depression to vital rates driving population growth. Using

this approach, particularly while focusing on vital rates most influential to population

performance, will be highly beneficial for guiding conservation decisions about

threatened and endangered species. Given that genetic effects within these populations -

either negative (inbreeding depression) or positive (genetic rescue) - are not expected to

significantly influence Sierra Nevada bighorn sheep populations in the coming years, we

recommend the Recovery Program focus on 1) maintaining genetic variation by

enhancing gene flow among existing populations, and 2) on non-genetic management

activities predicted to yield greater population gains in the near-term (i.e. disease

prevention, predator removal, and habitat enhancement projects; Bouzat et al. 2009); both

strategies that will minimize future losses of genetic variation in this subspecies.

ACKNOWLEDGEMENTS

We thank all the people that collected vital rate data on Sierra Nevada bighorn

sheep including L. Bowermaster, M. Cahn, K. Chang, K. Ellis, J. Erlenbach, D. German,

B. Gonzalez, A. Feinberg, J. Fusaro, L. Greene, D. Jensen, M. Kiner, L. Konde, K. Knox,

B. Pierce, C. Schroeder and others. We thank Landells Aviation and Quicksilver Air for

safely conducting bighorn sheep helicopter captures. We are grateful to S. Amish, D.

Page 163: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

149

Pletscher, M. Hebblewhite, and H. Robinson for providing constructive discussion and

comments on this manuscript. We would also like to thank the California Dept. of Fish &

Game and the Canon Science Scholars Program for supporting this work.

LITERATURE CITED

Aparicio JM, Ortego J, Cordero PJ (2007) Can a simple algebraic analysis predict

markers-genome heterozygosity correlations? J Hered 98:93-96.

Acevedo-Whitehouse K, et al. (2006) Contrasting effects of heterozygosity on survival

and hookworm resistance in California sea lion pups. Mol Ecol 15:1973-1982.

Balloux F, Amos W, Coulson T (2004) Does heterozygosity estimate inbreeding in real

populations? Mol Ecol 13:3021-3031.

Beissinger SR, Wunderle JM, Meyers JM, Saether BE, Engen S (2008) Anatomy of a

bottleneck: diagnosing factors limiting population growth in the Puerto Rican

Parrot. Ecol Monogr 78:185-203.

Bierne N, Tsitrone A, David P (2000) An inbreeding model of associative overdominance

during a population bottleneck. Genetics 155:1981-1990.

Bijlsma R, Bundgaard J, Boerema AC (2000) Does inbreeding affect the extinction risk

of small populations?: predictions from Drosophila. J Evol Biol 13:502-514.

Bleich VC, Wehausen JD, Jones KR, Weaver RA (1990) Status of bighorn sheep in

California, 1989 and translocations from 1971 through 1989. Desert Bighorn

Council Trans 1990:24-26.

Bouzat JL, et al. (2009) Beyond the beneficial effects of translocations as an effective

tool for the genetic restoration of isolated populations. Conserv Genet 10:191-

201.

Page 164: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

150

Boyce WM, et al. (1997) Genetic variation of major histocompatibility complex and

microsatellite loci: a comparison in bighorn sheep. Genetics 145:421-433.

Burnham KP, Anderson DR (2002) Model selection and multimodel inference: a

practical information-theoretic approach (Springer, New York, USA), 4th Ed.

Charlesworth D, Charlesworth B (1987) Inbreeding depression and its evolutionary

consequences. Annu Rev Ecol Syst 18:237-268.

Chapman JR, Nakagawa S, Coltman DW, Slate J, Sheldon BC (2009) A quantitative

review of heterozygosity-fitness correlations in animal populations. Mol Ecol

18:2746-2765.

Cleves M, Gould W, Gutierrez RG, Marchenko YU (2008) An introduction to survival

analysis using Stata (Stata Press, College Station, USA), 2nd Ed.

Cohas A, Bonenfant C, Kempenaers B, Allaine D (2009) Age-specific effect of

heterozygosity on survival in alpine marmots, Marmota marmota. Mol Ecol

18:1491-1503.

Coltman DW, Slate J (2003) Microsatellite measures of inbreeding: a meta-analysis.

Evolution 57:971-983.

Coulson T, Albon S, Slate J, Pemberton J (1999) Microsatellite loci reveal sex-dependent

responses to inbreeding and outbreeding in red deer calves. Evolution 53:1951-

1960.

Coulson TN, et al. (1998) Microsatellites reveal heterosis in red deer. Proc R Soc London

Ser B 265:489-495.

Cox DR (1972) Regression models and life-tables (with discussion). J Roy Statist Soc Ser

B 34:187-220.

Page 165: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

151

Da Silva A, Luikart G, Yoccoz NG, Cohas A, Allaine D (2006) Genetic diversity-fitness

correlation revealed by microsatellite analyses in European alpine marmots

(Marmota marmota). Conserv Genet 7:371-382.

Da Silva A, et al. (2009) Heterozygosity-fitness correlations revealed by neutral and

candidate gene markers in roe deer from a long-term study. Evolution 63:403-417.

David P, Jarne P (1997) Context-dependent survival differences among electrophoretic

genotypes in natural populations of the marine bivalve Spisula ovalis. Genetics

146:335-344.

Fieberg J, DelGiudice GD (2009) What time is it? Choice of time origin and scale in

extended proportional hazards models. Ecology 90:1687-1697.

Forbes SH, Hogg JT, Buchanan FC, Crawford AM, Allendorf FW (1995) Microsatellite

evolution in congeneric mammals: domestic and bighorn sheep. Mol Biol Evol

12:1106-1113.

Forbes SH, Hogg JT (1999) Assessing population structure at high levels of

differentiation: microsatellite comparisons of bighorn sheep and large carnivores.

Anim Conserv 2:223-233.

Grueber CE, Wallis GP, Jamieson IG (2008) Heterozygosity-fitness correlations and their

relevance to studies on inbreeding depression in threatened species. Mol Ecol

17:3978-3984.

Gutiérrez-Espeleta GA, Kalinowski ST, Boyce WM, Hedrick PW (2000) Genetic

variation and population structure in desert bighorn sheep: implications for

conservation. Conser Genet 1:3-15.

Hansson B, Westerberg L (2008) Heterozygosity-fitness correlations within inbreeding

Page 166: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

152

classes: local or genome-wide effects? Conserv Genet 9:73-83.

Heath DD, et al. (2002) Relationships between heterozygosity, allelic distance (d2), and

reproductive traits in Chinook salmon, Oncorhynchus tshawytscha. Can J Fish

Aquat Sci 59:77-84.

Hedrick P, Fredrickson R, Ellegren H (2001) Evaluation of d2, a microsatellite measure

of inbreeding and outbreeding, in wolves with a known pedigree. Evolution

55:1256-1260.

Hedrick PW, Kalinowski ST (2000) Inbreeding depression in conservation biology. Annu

Rev Ecol Syst 31:139-162.

Hogg JT, Forbes SH, Steele BM, Luikart G (2006) Genetic rescue of an insular

population of large mammals. Proc Royal Soc Ser B 273: 1491-1499.

Hoglund J, et al. (2002) Inbreeding depression and male fitness in black grouse. Proc

Royal Soc Ser B 269:711-715.

Johnson HE, Mills LS, Wehausen J, Stephenson TR (2010) Population-specific vital rate

contributions influence management of an endangered ungulate. Ecol Appl.

20:1753-1765.

Johnson HE, Mills LS, Wehausen J, Stephenson TR (In Press) Combining ground

count, telemetry, and mark-resight data to infer dynamics in an endangered

species. J. Appl Ecol.

Keller LF, Waller DM (2002) Inbreeding effects in wild populations. Trends Ecol Evol

17:230-241.

Lacy RC, Alaks G, Walsh A (1996) Hierarchical analysis of inbreeding depression in

Peromyscus polionotus. Evolution 50:2187-2200.

Page 167: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

153

Lacy RC, Ballou JD (1998) Effectiveness of selection in reducing the genetic load in

populations of Peromyscus polionotus during generations of inbreeding.

Evolution 52:900-909.

Lesbarreres D, Primmer SR, Laurila A, Merila J (2005) Environmental and population

dependency of genetic variability-fitness correlations in Rana temporaria. Mol

Ecol 14:311-323.

Luikart G, England PR, Tallmon D, Jordan S, Taberlet P (2003) The power and promise

of population genomics: from genotyping to genome typing. Nature Rev Genet

4:981-994.

Mainguy J, Côté SD, Coltman DW (2009) Multilocus heterozygosity, parental

relatedness and individual fitness components in a wild mountain goat, Oreamnos

americanus population. Mol Ecol 18:2297-2306.

McCullagh P, Nelder JA (1989) Generalized linear models (Chapman & Hall, London,

UK), 2nd Ed.

Mills LS (2007) Conservation of wildlife populations: demography, genetics and

management (Blackwell Publishing, Malden, USA).

Mills LS, Smouse PE (1994) Demographic consequences of inbreeding in remnant

populations. Am Nat 144:412-431.

Mitton JB (1993) in Theory and data pertinent to the relationship between heterozygosity

and Fitness: The Natural History of Inbreeding and Outbreeding, ed Thornhill

NW (University of Chicago Press, Chicago, USA), pp 17-41.

Morris WF, Doak DF (2002) Quantitative conservation biology: theory and practice of

population viability (Sinauer Associates, Sunderland, USA).

Page 168: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

154

Newman D, Pilson D (1997) Increased probability of extinction due to decreased genetic

effective population size: experimental populations of Clarkia pulchella.

Evolution 51:354-362.

Oostermeijer JGB, Luijten SH, den Nijs JCM (2003) Integrating demographic and

genetic approaches in plant conservation. Biol Cons 112:389-398.

Ortego J, Calabuig G, Cordero PJ, Aparicio JM (2007) Egg production and individual

genetic diversity in lesser kestrels. Mol Ecol 16:2383-2392.

Rabe-Hesketh S, Skrondal A (2008) Multilevel and longitudinal modeling using Stata

(Stata Press, College Station, USA), 2nd Ed.

Ralls K, Ballou JD, Templeton A (1988) Estimates of lethal equivalents and the cost of

inbreeding in mammals. Conserv Biol 2:185-193.

Roff DA (2002) Inbreeding depression: tests of the overdominance and partial dominance

hypotheses. Evolution 56:768-775.

Ruiz-López M, Roldán ERS, Espeso G, Gomendio M (2009) Pedigrees and

microsatellites among endangered ungulates: what do they tell us? Mol Ecol

18:1352-1364.

Saccheri I, et al. (1998) Inbreeding and extinction in a butterfly metapopulation. Nature

392:491-494.

Slate J, et al. (2004) Understanding the relationship between the inbreeding coefficient

and multilocus heterozygosity: theoretical expectations and empirical data.

Heredity 93:255-265.

Slate J, Pemberton JM (2002) Comparing molecular measures for detecting inbreeding

depression. J Evol Biol 15:20-31.

Page 169: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

155

StataCorp (2007) Stata User’s Guide. Release 10.0 (Stata Press, College Station, USA).

Tallmon DA, Luikart G, Waples RS (2004) The alluring simplicity and complex reality

of genetic rescue. Trends Ecol Evol 19:489-496.

U.S. Fish and Wildlife Service (2007) Recovery Plan for the Sierra Nevada bighorn

sheep. Sacramento, California, USA.

Waples RS (2006) A bias correction for estimates of effective population size based on

linkage disequilibrium at unlinked gene loci. Conserv Genet 7:167-184.

Wehausen JD (1980) Sierra Nevada bighorn sheep: history and population ecology. Ph.D

Dissertation, University of Michigan, Ann Arbor, Michigan, USA.

Wehausen JD (1996) Effects of mountain lion predation on bighorn sheep in the Sierra

Nevada and Granite Mountains of California. Wildl Soc Bull 24:471-479.

Wehausen JD, Ramey RR, II, Epps CW (2004) Experiments in DNA extraction and PCR

amplification from bighorn sheep feces: the importance of DNA extraction

method. J. Hered 95:503-509.

Westemeier RL, et al. (1998) Tracking the long-term decline and recovery of an isolated

population. Science 282:1695-1698.

Wright S (1951) The genetical structure of populations. Ann Eug 15:323-354.

Page 170: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

156

Table 4.1. Estimates of genetic variation for Sierra Nevada bighorn sheep populations, CA. Individuals captured between 1999

and 2009 (n) were used to calculate observed (Ho), expected heterozygosity (He), and the mean number of alleles (A) for each

population. Heterozygosity estimates are provided from the 25 loci that were polymorphic (p) and from the polymorphic and 4

monomorphic loci (p+m).

Population History n Ho (p) He (p) Ho (p+m) He (p+m) A

Baxter Source 26 0.468 0.482 0.379 0.390 2.59

Langley Translocated 21 0.412 0.403 0.333 0.326 2.35

Mono Basin Translocated 29 0.477 0.433 0.386 0.350 2.53

Wheeler Translocated 52 0.442 0.429 0.358 0.347 2.59

MEAN 128 0.450 0.436 0.364 0.353 2.52

Page 171: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

157

Table 4.2. Pairwise FST values for Sierra Nevada bighorn sheep populations, CA.

Baxter Langley Mono Wheeler

Baxter 0 - - -

Langley 0.0517 0 - -

Mono 0.0589 0.0750 0 -

Wheeler 0.0430 0.0818 0.0667 0

Page 172: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

158

Table 4.3. Model selection metrics for adult survival and fecundity rates of Sierra Nevada

bighorn sheep, CA. Metrics include the number of model parameters (p), log likelihood

(LL), AICC, ΔAICC, and model weight values. The best non-genetic model (NGM) is

listed in bold. Multilocus heterozygosity (h) and mean d2 were added to the NGM.

Significant genetic models are also listed in bold.

Fitness Trait Model p LL AICC ΔAICC Weight

Survival (n = 549)

Non-Genetic ModelsA

Age 1 -205.52 413.04 0 0.32

Age+Age2 2 -205.34 414.70 1.66 0.14

Age+Sex 2 -205.51 415.05 2.01 0.12

Age+Pop 4 -303.72 415.51 2.47 0.09

Age+Sex+Pop 5 -203.63 417.37 4.33 0.04

Sex 1 -210.20 422.41 9.37 0.00

Pop 3 -208.67 423.39 10.35 0.00

Sex+Pop 4 -208.65 425.38 12.34 0.00

Best Non-Genetic Model + Genetic Factors

NGM + d2 + Pop + (d

2 x Pop) 8 -199.124 414.51 1.47 0.15

NGM + H 2 -205.495 415.01 1.97 0.12

NGM + d2 2 -204.52 421.68 8.64 0.00

NGM + H + Pop + (H x Pop) 8 -203.311 422.89 9.85 0.00

Page 173: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

159

Fecundity (n = 194)

Non-Genetic Models

Intercept+Age+Age2 4 -125.12 258.46 1.86 0.20

Intercept 2 -129.15 262.36 5.76 0.03

Intercept+Age 3 -128.43 262.98 6.38 0.02

Intercept+Age+Age2+Pop 7 -124.88 264.36 7.76 0.01

Intercept+Pop 5 -128.90 268.11 11.51 0.00

Best Non-Genetic Model + Genetic Factors

NGM + H 5 -123.14 256.60 0 0.52

NGM + d2 5 -124.05 258.42 1.82 0.21

NGM + H + Pop + (H x Pop) 11 -121.45 266.36 9.76 0.00

NGM + d2 + Pop + (d

2 x Pop) 11 -123.36 270.18 13.58 0.00

ACox survival models do not include an intercept term.

Page 174: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

160

Table 4.4. Parameter coefficients (± SE) of best non-genetic and genetic models for adult

survival and female fecundity.

Non-Genetic Model Genetic Model

Model Parameter β SE p-value β SE p-value

Adult SurvivalA

Age 0.14 0.04 0.002 NA NA NA

Female Fecundity

Constant -1.04 0.94 0.266 -2.68 1.31 0.041

Age 0.61 0.30 0.042 0.65 0.30 0.031

Age2 -0.05 0.02 0.023 -0.05 0.02 0.017

H NA NA NA 3.42 1.79 0.056

ACox survival models do not include and intercept term.

Page 175: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

161

Table 4.5. Means and variances of stochastic population growth rates for the Mono Basin

and Langley populations of Sierra Nevada bighorn sheep under current conditions

(Baseline) and given the expected effects of inbreeding depression (ID). We also provide

growth rates from simulated genetic management at Mono Basin where heterozygosity

(h) was increased from 0.43 to 0.48 and 0.59.

Population Baseline ID h = 0.48 h = 0.59

Mono Basin

1 generation 1.023 (3.7e-3) 1.020 (3.2e-3) 1.031 (3.1e-3) 1.033 (4.3e-3)

5 generations 1.026 (2.8e-3) 1.021 (5.8e-3) 1.027 (3.8e-3) 1.031 (5.0e-3)

10 generations 1.019 (1.2e-2) 1.012 (1.3e-2) 1.020 (1.2e-2) 1.030 (8.8e-3)

Langley

1 generation 1.189 (4.8e-4) 1.188 (4.6e-4)

5 generations 1.181 (7.7e-5) 1.178 (6.8e-5)

10 generations 1.180 (3.4e-5) 1.174 (3.3e-5)

Page 176: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

162

Figure 4.1. Location of Mono Basin, Wheeler, Baxter and Langley populations of Sierra

Nevada bighorn sheep, CA.

Page 177: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

163

Figure 4.2. Number of adult females in the Wheeler, Langley, Baxter and Mono Basin

populations of Sierra Nevada bighorn sheep, 1998-2008. Numbers are based on annual

minimum counts except for Wheeler and Langley from 2006 to 2008 which are based on

mark-resight estimates.

Year

1998 2000 2002 2004 2006 2008

Nu

mb

er o

f A

du

lt F

ema

les

0

10

20

30

40

50

Mono Basin

Wheeler

Baxter

Langley

Page 178: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

164

Figure 4.3. Annual fecundity (± SE) for adult female Sierra Nevada bighorn sheep as a

function of multilocus heterozygosity. Predictions are based on a mixed effects logistic

regression model, holding all other effects constant.

Heterozygosity

0.0 0.2 0.4 0.6 0.8 1.0

An

nu

al P

robab

ilit

y o

f H

avin

g a

Lam

b

0.0

0.2

0.4

0.6

0.8

1.0

Page 179: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

165

Figure 4.4. Predicted size of the Mono Basin bighorn sheep population over 10

generations (60 years) if heterozygosity (h) remains unchanged (h = 0.43; Baseline), if

inbreeding depression continues to reduce h and fecundity (Inbreeding Depression), if

average h was increased to 0.48 (the mean value of the source population Baxter) and if

average h was increased to 0.59 (the mean value of other Rocky Mountain and desert

bighorn sheep populations).

0

20

40

60

80

100

120

140

160

180

1 2 3 4 5 6 7 8 9 10

Me

dia

n P

red

icte

d P

op

ula

ito

n S

ize

Generation

Baseline

Inbreeding Depression

Heterozygosity = 0.48

Heterozygosity = 0.59

Page 180: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

166

APPENDIX H. Identifying loci meeting neutral expectations and characterizing genetic

variation of Sierra Nevada bighorn sheep.

Methods

Sierra Nevada bighorn sheep persist in small, bottlenecked, reintroduced

populations we so expected that some loci would not meet neutral expectations. Loci or

locus combinations that were repeatedly flagged as outliers in multiple populations were

conservatively removed from statistical analysis characterizing genetic variation. We first

tested for the presence of null alleles and genotyping error using the programs

Microchecker (Van Oosterhout et al. 2004) and Dropout (McKelvey and Schwartz 2005).

We then performed exact tests for Hardy-Weinberg equilibrium (HWE) on each locus

within each population using GENEPOP v. 4.0 (Rousset 2008). To identify gametic

disequilibrium we used GENEPOP to test all locus combinations within populations.

We also tested whether loci were under selection or out of mutation-drift

equilibrium. We used FST outlier tests in the program LOSITAN (Antao et al. 2008) to

identify loci that may be under positive or balancing selection, manifested through

excessively high or low FST values relative to mean FST (Beaumont and Balding 2004).

Loci identified as being under selection were removed from additional analyses. We also

used the program BOTTLENECK to identify loci within our populations severely out of

mutation-drift equilibrium (Cornuet and Luikart 1996). This program detects loci that

have an excess of evenness in allele frequencies or in rare alleles, while detecting

signatures of recent population bottlenecks. We used two mutation models, a stepwise

model (SMM) and a two-phase model (TPM) with 80% SSM and 20% multi-step

mutations (as recommended by Piry et al. 1999).

Page 181: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

167

Using loci that met assumptions of neutrality, whether they were initially

categorized as “neutral” or “adaptive,” we calculated statistics characterizing genetic

variation of bighorn sheep populations. We calculated observed and expected

heterozygosity using GenAlEx6.2 (Peakall and Smouse 2006). We also calculated allelic

richness (the average number of alleles per locus or A), accounting for differences in

population sample size using the program FSTAT (Goudet 1995). For both

heterozygosity and allelic richness we tested whether reintroduced bighorn sheep

populations had significantly reduced genetic variation than the source herd (Baxter). We

also used FSTAT to quantify the genetic divergence between pairs of populations using

FST values (Nei 1978). After calculating all these statistics with only those microsatellite

markers that behaved neutrally, we recalculated these same metrics using all polymorphic

loci, regardless of whether they met neutral assumptions, to assess the importance of

meeting such assumptions when characterizing population-level genetic variation and

divergence.

Results

All 128 individuals were successfully genotyped at all loci, except for one

individual not typed at the HH64 locus. Genotyping error was not identified as a problem

as the only locus flagged as having a null allele in multiple populations was HH64. After

accounting for multiple comparisons, HH64 was also the only locus that significantly

deviated from HWE (p < 0.001) and therefore was removed from further analyses.

The locus pairs MHC1/OLA and MHC1/TGLA387 showed significant gametic

disequilibrium in all 4 populations (p < 0.01), leading us to remove MHC1 from all

analyses relying on assumptions of neutrality. There was also evidence that

Page 182: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

168

disequilibrium occurred in locus combinations MAF36/JMP29 and HH62/SOMA in 2

populations so we conservatively removed JMP29 and SOMA from further analyses.

From the total 2,400 tests conducted, 39 locus pairs were also significant for gametic

disequilibrium in single populations. Because significant tests are expected by chance,

particularly with small, bottlenecked populations, we did not remove additional loci.

FST outlier and bottleneck tests identified additional loci that did not meet neutral

expectations. The marker CP20 was identified as being under positive selection (p =

0.969), and markers AE16 and MMP9 as being under balancing selection (p = 0.044 and

p = 0.006, respectively). Given both SMM and TPM models in BOTTLENECK, loci in

all of our populations were significantly out of mutation-drift equilibrium (Wilcoxon tests

p < 0.003) showing the heterozygosity excess expected with bottlenecked populations.

We removed OLA and AE16, as these loci significantly (p < 0.05) deviated from

mutation-drift equilibrium in 2 and 3 populations, respectively.

Measures of genetic variation are reported in the manuscript text. Calculating the

same population level statistics using all our polymorphic genetic markers, regardless of

whether they met assumptions of neutrality, there were no qualitative differences in

heterozygosity, allelic diversity, or FST patterns (global FST = 0.067; Tables 4.A.1-2).

Literature Cited

Antao, T., A. Lopes, R.J. Lopes, A. Beja-Pereira, and G. Luikart. 2008. LOSITAN: a

workbench to detect molecular adaptation based on a Fst-outlier method. BMC

Bioinformatics 9:323.

Beaumont, M.A., and D.J. Balding. 2004. Identifying adaptive genetic divergence among

populations from genome scans. Molecular Ecology 13:969-980.

Page 183: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

169

Cornuet J.M. and Luikart G. 1996. Description and power analysis of two tests for

detecting recent population bottlenecks from allele frequency data. Genetics

144:2001-2014.

Goudet J. 1995. FSTAT (vers. 1.2): a computer program to calculate F-statistics. Journal

of Heredity 86: 485-486.

McKelvey, K.S., and M.K. Schwartz. 2005. DROPOUT: a program to identify problem

loci and samples for noninvasive genetic samples in a capture-mark-recapture

framework. Molecular Ecology Notes 5:716-718.

Nei, M. 1978. Estimation of average heterozygosity and genetic distance from a small

number of individuals. Genetics 89:538-590.

Peakall, R. and P.E. Smouse. 2006. GENALEX 6: genetic analysis in Excel. Population

genetic software for teaching and research. Molecular Ecology Notes 6: 288-295.

Piry, S., G. Luikart, and J.-M. Cornuet. 1999. Bottleneck: a computer program for

detecting recent reductions in effective population size from allele frequency data.

Journal of Heredity 90:502-503.

Rousset, F. 2008. GENEPOP’007: a complete re-implementation of the GENEPOP

software for Windows and Linux. Molecular Ecology Resources 8:103-106.

Van Oosterhout, C., W.F. Hutchinson, and D.P.M. Willis, and P. Shipley. 2004. MICRO-

CHECKER: software for identifying and correcting genotyping errors in

microsatellite data. Molecular Ecology Notes 4: 535-538.

Page 184: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

170

Table 4.A.1. Comparison of population-level genetic variation and differentiation statistics for all polymorphic microsatellite markers

(neutral and potentially adaptive), regardless of meeting neutrality assumptions in Sierra Nevada bighorn sheep populations, CA.

Individuals captured between 1999 and 2009 (n) were used to calculate observed (Ho), expected heterozygosity (He), and the mean

number of alleles (A) for each population. Heterozygosity estimates are provided from only polymorphic loci (p) and from

polymorphic and monomorphic loci (p+m).

Population History n Ho (p) He (p) Ho (p+m) He (p+m) A

Baxter Source 26 0.466 0.477 0.402 0.411 2.63

Langley Translocated 21 0.450 0.427 0.388 0.368 2.40

Mono Basin Translocated 29 0.474 0.458 0.409 0.395 2.54

Wheeler Translocated 52 0.465 0.457 0.401 0.394 2.55

MEAN 128 0.464 0.455 0.400 0.392 2.53

Page 185: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

171

Table 4.A.2. Pairwise FST values of Sierra Nevada bighorn sheep populations, CA, using

all polymorphic markers, regardless of meeting the assumptions of neutrality.

Baxter Langley Mono Wheeler

Baxter 0 - - -

Langley 0.0667 0 - -

Mono 0.0681 0.0797 0 -

Wheeler 0.0379 0.0813 0.0695 0

Page 186: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

172

APPENDIX I. Locus-specific genetic variation of the 25 polymorphic microsatellite markers genotyped on Sierra Nevada bighorn

sheep including the total number of alleles observed (A), allele size ranges, observed (Ho) and expected heterozygosity (He), FIS and

FST.

Locus A Allele size range (bp) HO HE FIS FST Reference

MAF65 2 118 - 134 0.5708 0.4787 -0.192 0.014 Buchanan et al. 1992

MAF209 3 108 - 122 0.3919 0.3619 -0.083 0.087 Buchanan and Crawford 1992a

FCB304 2 143 - 147 0.4386 0.4578 0.042 0.069 Buchanan and Crawford 1993

FCB11 2 124 - 128 0.4086 0.3861 -0.058 0.054 Buchanan and Crawford 1993

MAF36 5 90 - 106 0.666 0.6473 -0.029 0.083 Swarbrick et al. 1991

MAF33 3 124 - 128 0.6761 0.6368 -0.062 0.04 Buchanan and Crawford 1992b

MAF48 3 123 - 127 0.4748 0.5162 0.074 0.120 Buchanan et al. 1991

AE16 3 85 - 95 0.6237 0.6248 0.002 0.023 Penty et al. 1993

HH47 2 135 - 137 0.4662 0.4585 -0.017 0.075 Henry et al. 1993

HH62 3 108 - 114 0.565 0.5769 0.021 0.073 Ede et al. 1994

CP20 3 80 - 96 0.5622 0.5091 -0.106 0.190 Ede et al. 1995

Page 187: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

173

HH64 5 119 - 123 0.2157 0.351 0.385 0.032 Henry et al. 1993

FCB193 4 103 - 115 0.5921 0.5412 -0.094 0.049 Buchanan and Crawford 1993

BM4513 3 135 - 153 0.5101 0.5347 0.046 0.087 Bishop et al. 1994

AE129 3 179 - 187 0.2636 0.2722 0.032 0.015 Penty et al. 1993

JMP29 3 134 - 138 0.431 0.3937 -0.095 0.134 Crawford et al. 1995

TGLA387 3 144 - 150 0.3705 0.3974 0.068 0.037 Georges and Massey 1992

TCRB 2 171 - 175 0.5515 0.4601 -0.199 0.017 Crawford et al. 1995

KERA 2 177 - 179 0.0914 0.0853 -0.072 0.022 J.F. Maddox, unpublished

ADC 2 91 - 95 0.2449 0.2274 -0.077 0.065 Wood and Phua 1994

SOMA 2 112 -116 0.5128 0.4535 -0.131 0.065 Lucy et al. 1998

MMP9 3 184 - 190 0.591 0.6068 0.026 0.012 Maddox 2001

BL4 3 156 - 162 0.3639 0.3854 0.056 0.038 Bishop et al. 1994

MCHI 3 192 - 196 0.5237 0.5442 0.038 0.057 Groth and Weatherall 1994

OLA 2 272 - 286 0.4877 0.4717 -0.034 0.054 Schwaiger et al. 1993

Page 188: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

174

Literature Cited

Bishop, M.D., S.M. Kappes, J.W. Keele, R.T. Stone, S.L.F. Sunden, G.A. Hawkins, S.S.

Toldo, R. Fries, M.D. Grosz, J. Yoo, and C.W. Beattie. 1994. A genetic linkage

map for cattle. Genetics 136:619–639.

Buchanan, F.C., and A.M. Crawford. 1992a. Ovine dinucleotide repeat polymorphism at

the MAF209 locus. Animal Genetics 23:183.

Buchanan, F.C., and A.M. Crawford. 1992b. Ovine dinucleotide repeat polymorphism at

the MAF33 locus. Animal Genetics 23:186.

Buchanan, F.C., P.A. Swarbrick, and A.M. Crawford. 1991. Ovine dinucleotide repeat

polymorphism at the MAF48 locus. Animal Genetics 22:379-380.

Buchanan, F.C., P.A. Swarbrick, and A.M. Crawford. 1992. Ovine dinucleotide repeat

polymorphism at the MAF65 locus. Animal Genetics 23:85.

Buchanan, F.C., and A.M. Crawford. 1993. Ovine microsatellites at the OarFCB11,

OarFCB128, OarFCB193, OarFCB266, and OarFCB304 loci. Animal Genetics

24:145.

Crawford, A.M., K.G. Dodds, A.J. Ede, C.A. Pierson, G.W. Montgomery, H.G.

Garmonsway, A.E. Beattie, K. Davies, J.F. Maddox, S.W. Kappes, R.T. Stone,

T.C. Nguyen, J.M. Penty, E.A. Lord, J.E. Broom, J. Buitkamp, W. Schwaiger,

J.T. Epplen, P. Matthew, M.E. Matthews, D.J. Hulme, K.J. Beh, R.A. McGraw,

and C.W. Beattie. 1995. An autosomal genetic linkage map of the sheep genome.

Genetics 140:703-724.

Ede, A.J., C.A. Pierson, and A.M. Crawford. 1995. Ovine microsatellites at the OarCP9,

Page 189: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

175

OarCP16, OarCP20, OarCP21, OarCP23 and OarCP26 loci. Animal Genetics

26:129-130.

Ede, A.J., C.A. Peirson, H. Henry, and A.M. Crawford. 1994. Ovine microsatellites at the

OarAE64, OarHH22, OarHH56, OarHH62, and OarVH4 loci. Animal Genetics

25:51–52.

Georges, M., and J. Massey. 1992. Polymorphic DNA markers in Bovidae. In WO Publ.

No. 92/13120. Geneva: World Intellectual Property Organization.

Groth, D.M., and J.D. Weatherall. 1994. Dinucleotide repeat polymorphism within the

ovine major histocompatibility complex class I region. Animal Genetics 25:61.

Henry, H.M., J.M. Pentry, C.A. Pierson, and A.M. Crawford. 1993. Ovine microsatellites

at the OarHH35, OarHH41, OarHH44, OarHH47, and OarHH64 loci. Animal

Genetics 24:222.

Lucy, M.C., G.S. Johnson, H. Shibuya, C.K. Boyd, and W.O. Herring. 1998. Rapid

communication: polymorphic (GT)n microsatellite in the bovine somatotropin

receptor gene promoter. Journal of Animal Science 76:2209-2210.

Maddox, J.F. 2001. Mapping the matrix metalloproteinase 9 (MMP9) gene and the

BL1071 microsatellite to ovine chromosome 13 (OAR13). Animal Genetics

32:329-331.

Penty, J.M., H.M. Henry, A.J. Ede, and A.M. Crawford. 1993. Ovine microsatellites at

the OarAE16, OarAE54, OarAE119 and OarAE129 loci. Animal Genetics

24:219.

Schwaiger, F.-W., J. Buitkamp, E. Weyers, and J.T. Epplen. 1993. Typing of artiodactyl

Page 190: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

176

MHC-DRB genes with the help of intronic simple repeated DNA sequences.

Molecular Ecology 2:55-59.

Swarbrick P.A., F.C. Buchanan, and A. Crawford. 1991. Ovine dinucleotide repeat

polymorphism at the MAF36 locus. Animal Genetics 22:377–378.

Wood, N.J., and S.H. Phua. 1994. A dinucleotide repeat polymorphism at the adenylate

cyclase activating polypeptide locus in sheep. Animal Genetics 24:329.

Page 191: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

177

APPENDIX J. AICc values from survival and fecundity models that include locus-

specific heterozygosity and d2, in addition to age and age

2, the covariates identified in the

best non-genetic model. No locus-specific model improves the fit of the non-genetic

model (NGM). ΔAICc reports the difference between the AICc of the locus-specific

model and the NGM.

Survival Reproduction

AICc ΔAICc AICc ΔAICc

NGM 413.04 0.00 258.46 0.00

Heterozygosity

MAF65 414.73 1.69 259.26 0.80

MAF209 414.94 1.90 259.29 0.83

FCB304 413.66 0.62 259.05 0.59

FCB11 415.01 1.97 260.26 1.80

MAF36 414.55 1.51 260.27 1.81

MAF33 413.21 0.17 260.16 1.70

MAF48 414.05 1.01 258.30 -0.16

HH47 414.80 1.76 258.34 -0.12

HH62 413.08 0.04 260.34 1.88

FCB193 414.34 1.30 260.45 1.99

BM4513 413.15 0.11 260.57 2.11

AE129 415.00 1.96 260.56 2.10

AE16 409.90 -3.17 259.78 1.33

CP20 414.67 1.60 260.57 2.11

Page 192: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

178

JMP29 413.80 0.73 259.94 1.48

TGLA387 414.40 1.36 260.46 2.00

TCRB 413.66 0.62 260.56 2.10

KERA 415.05 2.01 260.54 2.08

ADC 414.02 0.98 260.03 1.57

SOMA 412.07 -0.97 258.44 -0.02

MMP9 414.99 1.95 259.68 1.22

BL4 414.91 1.87 260.22 1.76

MHCI 415.01 1.97 260.51 2.05

OLA 413.27 0.23 259.95 1.49

Mean d2

MAF65 414.73 1.69 259.26 0.80

MAF209 414.93 1.89 259.12 0.66

FCB304 413.66 0.62 259.05 0.59

FCB11 415.01 1.97 260.26 1.80

MAF36 412.06 -0.98 260.50 2.04

MAF33 411.84 -1.20 260.25 1.79

MAF48 414.93 1.89 260.55 2.09

HH47 414.80 1.76 258.34 -0.12

HH62 414.15 1.11 260.55 2.09

FCB193 413.35 0.31 259.70 1.24

BM4513 412.02 -1.02 260.47 2.01

AE129 415.04 2.00 260.54 2.08

Page 193: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

179

AE16 413.22 0.14 258.20 -0.26

CP20 412.55 -0.53 260.04 1.58

JMP29 413.77 0.69 259.94 1.48

TGLA387 414.74 1.70 260.46 2.00

TCRB 413.66 0.62 260.56 2.10

KERA 415.05 2.01 260.54 2.08

ADC 414.02 0.98 260.03 1.57

SOMA 412.07 -0.97 260.57 2.11

MMP9 415.00 1.96 260.25 1.79

BL4 414.88 1.84 260.52 2.06

MHCI 415.04 2.00 260.51 2.05

OLA 413.27 0.23 259.95 1.49

Page 194: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

180

CHAPTER 5

APPARENT COMPETITION LIMITS THE RECOVERY OF AN

ENDANGERED UNGULATE: QUANTIFYING THE DIRECT AND

INDRECT EFFECTS OF PREDATION

HEATHER E. JOHNSON, University of Montana, Wildlife Biology Program, College

of Forestry and Conservation, Missoula, MT 59812, USA

MARK HEBBLEWHITE, University of Montana, Wildlife Biology Program, College

of Forestry and Conservation, Missoula, MT 59812, USA

THOMAS R. STEPHENSON, Sierra Nevada Bighorn Sheep Recovery Program,

California Department of Fish and Game, 407 West Line Street, Bishop, CA

93514, USA

DAVE GERMAN, Sierra Nevada Bighorn Sheep Recovery Program, California

Department of Fish and Game, 407 West Line Street, Bishop, CA 93514, USA

BECKY M. PIERCE, Sierra Nevada Bighorn Sheep Recovery Program, California

Department of Fish and Game, 407 West Line Street, Bishop, CA 93514, USA

ABSTRACT

Predation is a dominant force that shapes the demography, behavior, and

distribution of prey populations, through both direct and indirect effects. These effects

can disproportionately impact small and endangered prey populations when generalist

predators are numerically linked to more common primary prey. Apparent competition,

the term for this phenomenon, has been increasingly implicated in the declines of small

Page 195: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

181

prey populations and in inhibiting conservation efforts of endangered prey species. We

examined the potential for apparent competition to limit the recovery of federally

endangered Sierra Nevada bighorn sheep, the rarest subspecies of mountain sheep in

North America. Using a combination of demographic, spatial, habitat, and observation

data we assessed whether mountain lion predation was having direct and indirect effects

on bighorn sheep as a consequence of their spatial overlap with mule deer. In accordance

with the apparent competition hypothesis we predicted that bighorn sheep populations

with greater spatial overlap with deer would exhibit 1) higher rates of lion predation, 2)

lower rates of annual survival, 3) lion-kills in close proximity to deer winter ranges, and

4) stronger antipredator behaviors. We found evidence that lion predation directly

affected bighorn sheep demographic rates and elicited indirect antipredator behaviors, as

populations with higher predation rates selected safer habitat and utilized larger group

sizes. Results supported the predictions of the apparent competition hypothesis with

predation having a larger effect on bighorn sheep populations with greater spatial overlap

with deer. We also found that the influence of asymmetric predation was highly spatially

and temporally variable, with spatial variation driven by landscape-scale differences in

habitat availability and temporal variation driven by finer-scale forage availability. While

evidence suggests that elevated predation is limiting recovery success in some bighorn

sheep herds, it is not limiting in all herds, providing key implications for the conservation

of existing populations and the reintroduction of new populations. Management strategies

for endangered species should incorporate the spatial distributions of competitors and

predators to reduce the potential for apparent competition to hijack conservation success.

Page 196: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

182

INTRODUCTION

Predation is a dominant selective force in shaping the demography, behavior, and

distribution of prey populations, impacting prey species at multiple scales and through

various mechanisms (Lima and Dill 1990). In their most widely recognized role predators

affect prey populations through direct consumption, with the degree of their impact

depending on the rate of predation, whether predation is compensatory, and the sex/stage

classes that are killed (Mills 2007). Predators also exert indirect effects on prey

populations, as animals exhibit a variety of antipredator behavioral strategies which

confer fitness costs (Creel and Christianson 2008). For example, prey may avoid

preferred habitats that have a high risk of predation (Mao et al. 2005, Fortin et al. 2005),

thereby reducing resource acquisition, and thus, decreasing survival and reproductive

success (Wehausen 1996, Schmitz et al. 1997, Nelson et al. 2004, Cresswell 2008).

Recent studies have found that the demographic consequences of these indirect effects

may be as great or even greater than direct predation itself (Preisser et al. 2005, Creel and

Christianson 2008).

These effects of predators can disproportionately impact the dynamics of small

and endangered prey populations (Sinclair et al. 1998, McLellan et al. 2010, DeCesare et

al. In Press). This commonly occurs when ≥ 1 primary prey species supports the

numerical response of a predator that is shared with an endangered, secondary prey

species. Under these conditions, the opportunistic take of the secondary prey can yield

dramatic population declines (Roemer et al. 2002, Bryant and Page 2005, Wittmer et al.

2005a). Holt (1977) termed this phenomenon “apparent competition,” as the

asymmetrical influence of predation on primary and secondary prey can appear as if the

Page 197: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

183

species were in direct competition with one another. While prey population declines may

initially occur due to habitat loss, exotic species introductions, or disease, once

populations become small they are highly vulnerable to predators subsidized by other

prey. Depending on the functional response of predators to prey (Holling 1959), apparent

competition can lead to either the extirpation of secondary prey populations (Type II

response) or trap them in a “predator pit,” where low numbers of the prey remain but the

population is inhibited from recovering (Type III response; Sinclair et al. 1998, Messier

1994). As a result, the direct effects of asymmetric predation via apparent competition is

increasingly implicated in hindering conservation and recovery efforts of endangered

prey species (Sinclair et al. 1998, Roemer et al. 2002, Wittmer et al. 2005a, Angulo et al.

2007, DeCesare et al. In Press).

Studies of apparent competition in endangered species have traditionally focused

on only the direct effects of predation, even though the ecological literature has

demonstrated that indirect predation effects may be just as important. This disconnect

likely stems from the difficulty in obtaining demographic costs of indirect effects of

predation in natural field settings and the perception that direct effects have more

immediate influence on population performance (Creel and Christianson 2008). If,

however, indirect effects are just as significant as direct effects (Preisser et al. 2005) their

detection could be critical for developing effective management strategies for endangered

prey. For example, species that strongly avoid areas of high predator density may have

low direct mortality but still significantly benefit from a predator removal program.

We examined the direct and indirect effects of predation on federally endangered

Sierra Nevada bighorn sheep (Ovis canadensis sierra; hereafter bighorn sheep), the rarest

Page 198: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

184

subspecies of mountain sheep in North America (U.S. Fish and Wildlife Service 2007).

Populations were initially reduced in the 1800’s due to disease, market hunting, and

competition with domestic livestock. More recently, however, population declines have

been attributed to predation by mountain lions (Felis concolor; hereafter lions; Wehausen

1996, U.S. Fish and Wildlife Service 2007). While there are estimated to be <400 bighorn

sheep in the Sierra Nevada (CDFG unpublished data), thousands of mule deer

(Odocoileus hemionus) winter in close proximity to endangered herds serving as the

primary prey source for lions (Pierce et al. 1999, Pierce et al. 2000). The spatial

proximity between bighorn sheep and deer creates the potential for apparent competition

to reduce bighorn sheep population growth rates and limit recovery success.

The bighorn-deer-lion ecosystem in the eastern Sierra exhibits several

characteristics classically associated with apparent competition (Holt 1977, Holt and

Lawton 1994, Chaneton and Bonsall 2000, Chase et al. 2002, Harmon and Andow 2004,

Chesson and Kuang 2008, DeCesare et al. In Press). Mountain lions are generalist

predators with high mobility between the ranges of each prey species, there appears to be

overlap in winter ranges of deer and bighorn sheep, and deer have a much larger

population size than bighorn sheep. Additionally, roughly 75% of mortalities of collared

bighorn sheep occur during winter months when there is high spatial proximity to deer,

and thus, mountain lions (Fig. 5.1). While some bighorn herds consistently winter in

close proximity to deer and lions, other herds exist in areas of low deer and lion density.

This spatial variation has created uncertainty about the demographic effect of lion

predation on bighorn sheep, how it may differentially influence isolated herds, and the

utility of lion removal as a recovery strategy. Given that lions are a protected species in

Page 199: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

185

California (Torres et al. 1996), clarifying the role of predation on bighorn sheep is critical

because lion removal is a highly controversial management tool.

Our objective was to examine whether asymmetric predation was limiting bighorn

sheep recovery by testing 4 predictions of the apparent competition hypothesis related to

the direct and indirect effects of lion predation. We evaluated these predictions in 4

isolated bighorn sheep populations which span the geographic range of the subspecies

(Fig. 5.2), have exhibited widely varying population dynamics (Johnson et al. 2010), and

encompass a range of deer and lion densities. If apparent competition was responsible for

direct effects of lion predation on bighorn sheep we predicted that herds with greater

spatial overlap with deer would have 1) higher rates of lion predation, and 2) lower rates

of annual survival (Holt 1977, Holt and Lawton 1994). We also predicted that 3) lion-

killed bighorn sheep would occur in close proximity to deer winter ranges (James et al.

2004, McLoughlin et al. 2005). If apparent competition has generated indirect predation

effects we predicted that populations with greater overlap with deer would 4) exhibit

stronger antipredator behavior (Lima and Dill 1990). Because bighorn sheep are

dependent upon vigilance and their use of rugged, rocky terrain to escape and evade

predators (Geist 1970), we expected bighorn herds with higher overlap with deer to either

avoid areas of high lion use (Wehausen 1996), select “safer” terrain (more steep and

rugged; Hamel and Côté 2007; Schroeder et al. 2010), or employ larger group sizes in

areas of increased predation risk (Fortin and Fortin 2009).

Page 200: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

186

METHODS

Study Area and Populations

The Sierra Nevada mountain range forms the eastern backbone of California and

is approximately 650 km long and ranges from 75 to 125 km wide (Hill 1975). Historical

and current distributions of bighorn sheep include only the southern half of the Sierra

Nevada, where geologic processes have created the highest mountains and the most

alpine habitat. Bighorn sheep spend summers in the alpine along the crest of the Sierra

Nevada and winter either in the alpine or at lower elevations typically east of the crest,

inhabiting elevations ranging from 1,525 to >4,000 m (U.S. Fish & Wildlife Service

2007). Climate in the Sierra Nevada is characterized by relatively dry conditions in

summer (May-Sept), with most of the annual precipitation received as snow in winter

(Nov-Apr), varying considerably by year. There is a strong rain shadow effect in

precipitation east of the Sierra crest resulting in open, xeric vegetation communities. Low

elevations (1,500-2,500 m) are characterized by Great Basin desert sagebrush-bitterbrush

scrub; mid-elevations (2,500-3,300 m) by pinyon-juniper woodland, sub-alpine meadows,

and forests; and high elevations (>3,300 m) by sparse alpine vegetation including

occasional meadows. Virtually all bighorn sheep habitat is public land, managed

primarily by Yosemite and Sequoia-Kings Canyon National Parks, and Inyo and Sierra

National Forests.

We evaluated the 4 bighorn sheep populations for which extensive demographic

and habitat data have been collected: Mt. Warren and Mt. Gibbs (Mono Basin), Wheeler

Ridge (Wheeler), Mt. Baxter and Sawmill Canyon (Baxter), and Mount Langley

(Langley); situated north to south along the Sierra Nevada crest (Fig. 5.2). These herds

Page 201: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

187

represent approximately 85% of all bighorn sheep in the subspecies, exhibit high spatial

and temporal variation in population trends (Johnson et al. 2010), and are geographically

isolated. Deer herds that winter in close proximity to bighorn sheep are located in hunt

zones X12, X9A, and X9B. The local mountain lion population is closely tied to these

deer herds, migrating seasonally with them as the deer comprise their primary food

source (Pierce et al. 1999, Pierce et al. 2000).

Quantifying Spatial Overlap in Bighorn and Deer Winter Ranges

We used bighorn sheep and deer locations collected between Dec 1st and Apr 30

th

to quantify the degree of spatial overlap between the winter ranges of these prey species.

In October 2007 and 2008 we captured bighorn sheep using a net-gun fired from a

helicopter (Krausman et al. 1985; University of Montana Institutional Animal Care and

Use Protocol 024-07MHWB-071807). We deployed 32 global-positioning-system (GPS)

collars on adult females. Five females were collared in Mono Basin, 6 in Wheeler, 8 in

Langley, and 13 in Baxter representing approximately 45%, 17%, 24%, and 34% of the

total adult females in each herd, respectively. Collars were programmed to collect either

3 locations/day (00:00, 08:00, 16:00) or 6 locations/day (00:00, 04:00, 08:00, 12:00,

16:00, 20:00) and were manufactured by the companies Advanced Telemetry Systems,

North Star, Lotek Wireless Inc., and Televilt International. Eleven females were collared

during both the winters of 2008 (Dec 2007 – Apr 2008) and 2009 (Dec 2008 – Apr 2009)

and we treated data from each animal/year separately, as there were substantial

differences in habitat use patterns among years. Using data from 2008 and 2009 we

obtained a total of 7 animal- and year-specific data sets from Mono Basin, 10 from

Wheeler, 10 from Langley, and 17 from Baxter to characterize bighorn sheep winter

Page 202: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

188

range use. Collars collected a total of 21,350 locations with a GPS fix rate of 87%.

Missing GPS fixes of <10% are not expected to dramatically alter inferences about

habitat selection (D’Eon 2003, Frair et al. 2004). Because we used locations to delineate

general winter range use patterns and coarse differences in habitat selection among

populations (described below) we assumed that our fix rate was adequate and did not

correct for missing locations.

To delineate deer winter ranges we compiled locations from annual helicopter

surveys that occurred in January and March from 2001 through 2009 (CDFG unpublished

data). Helicopter survey locations were not recorded for deer in Round Valley (a subset

of deer in the X9 hunt zone), so we obtained spatial data from GPS collars deployed on

80 deer in that area between 2002 and 2009. Deer were captured and collared biannually

with net-gun helicopter operations (Idaho State University Institutional Animal Care and

Use Protocol 650-0410). Deer GPS collars were manufactured by Advanced Telemetry

Solutions and Televilt International and were programmed to collect locations every 2

hours or every 7 hours on revolving schedules. Because the number of GPS locations

from Round Valley was an order of magnitude larger than helicopter survey locations

from the other deer herds, we randomly selected 5% of the GPS locations (similar to

sample sizes of helicopter surveys) to estimate the winter range of this herd. In total we

used 5,278 locations to estimate deer winter ranges along the eastern Sierra.

We determined the degree of spatial overlap in bighorn sheep and mule deer

winter ranges by creating contour polygons around the 95% probability density

distribution of locations for each species. We first estimated an appropriate smoothing

factor (h) for each data set using likelihood cross-validation (CVh; Horne and Garton

Page 203: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

189

2006) in the program Animal Space Use (Horne and Garton 2007). We then used a kernel

density estimator (KDE; Worton 1989) in Hawthtools 3.27 (Beyer 2004) with the

designated h value to calculate a spatial probability density function for each prey

species. From those functions, we generated 95% volume contours for bighorn sheep and

deer (containing approximately 95% of the locations used to create the kernel density

estimate), and considered this to be the delineated winter range for each species. For each

bighorn sheep population we calculated the area (km2) of winter range overlap with deer,

assuming that risk of lion predation was a function of the absolute area of overlap rather

than the relative area. We made this assumption because the absolute area of overlap

relates directly to differences in the number of deer in close proximity to bighorn sheep,

and thus, to the expected numerical response of mountain lions.

Measuring Direct Effects of Predation

To assess the influence of direct predation on bighorn sheep we used data

collected from radio-collared individuals. We began deploying very-high-frequency

(VHF) collars in Wheeler in 1999, in Mono Basin in 2002, in Baxter in 2003, and in

Langley in 2004. Since then, bighorn sheep have been captured 1-2 times/year, collaring

a total of 53 sheep in Wheeler (26 females, 27 males), 39 in Mono Basin (17 females, 22

males), 44 in Baxter (36 females and 8 males), and 26 in Langley (24 females, 2 males).

After collars were deployed, individuals were monitored at least twice/month by ground

and aerial telemetry to determine survival and cause-specific mortality rates. After a

collar was heard emitting a mortality signal, ground crews investigated the location to

determine the cause of death based on evidence of predation, accidents, and nutritional

condition.

Page 204: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

190

To evaluate our first prediction of the apparent competition hypothesis, that

greater spatial overlap with deer would result in higher lion predation, we calculated

cause-specific mortality rates for each population using nonparametric cumulative

incidence functions (Heisey and Patterson 2006). Animals entered the study following a

staggered entry design (based on initial capture date) and exited when they died or were

censored. Six animals were censored due to collar failure, and 94 animals were censored

for being alive at end of the study (1 March 2010). Given the small number of collared

bighorn sheep in each population/year, we only calculated mean annual mortality rates

based on a biological year from April 15th

to the following April 14th

(the start of the

lambing season). Competing risks, or causes of mortality, were classified as lion

predation, physical injury (namely from falls and rock-slides), other, and unknown. The

“other” category included mortality agents that were rarely identified in the dataset such

as old age, road kill, and malnutrition.

We tested the second prediction, that bighorn sheep herds with higher deer

overlap would have lower survival rates, using known-fate telemetry data. We estimated

mean annual survival for each population using nonparametric Kaplan Meier models

(Pollock et al. 1989). We followed the same guidelines as in the cause-specific mortality

analysis relative to staggered entry, censoring, and calculating annual rates based on a

biological bighorn sheep year. To assess the predictions of the apparent competition

hypothesis we then correlated population-specific survival rates to areas of deer overlap.

To evaluate the third prediction, that lion predation on bighorn sheep would occur

near deer winter ranges, we compared the distance-to-deer-winter-range of 52 confirmed

lion-killed sheep to the proportion of bighorn sheep winter range in proximity to deer.

Page 205: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

191

Twenty-one lion-kills were collared bighorn sheep and 31 were un-collared. Un-collared

kills were found opportunistically in the field while conducting population surveys,

habitat sampling, investigating collared bighorn sheep mortalities, and from kill sites

from collared lions. We conducted a Spearman’s rank correlation test comparing 5 area-

adjusted frequency bins of distance-to-deer values in bighorn sheep winter range to the

number of lion-kills within the same frequency bins (Boyce et al. 2002, Hebblewhite and

Merrill 2007). If lion-kills occurred in areas close to deer we expected a negative

correlation between the frequency of kills and the distance-to-deer-range. We also

conducted the same correlation test using only the 21 collared bighorn sheep lion-kills.

All analyses were conducted in StataMP 11.0 (StataCorp 2009).

Measuring Indirect Effects of Predation

We examined the indirect effects of apparent competition by testing our fourth

prediction that bighorn sheep populations with greater overlap with deer would exhibit

stronger antipredator behavior. To determine whether risk of predation induced bighorn

sheep to avoid areas of high lion density or select safer terrain we estimated a mixed-

effects winter resource selection model (RSF) for each herd based on a use-availability

design (Manly et al. 2002). We focused on winter because that is when most lion

mortalities occur (Fig. 5.1). RSF models were generated from GPS collar data collected

during the winters of 2008 and 2009 (described in Quantifying Spatial Overlap). Habitat

attributes of each GPS location were compared to 3 randomly selected locations within

available habitat. For each population, available habitat was delineated from combining

the 100% winter minimum convex polygons (MCPs) of each collared female/year in the

herd (McLoughlin et al. 2008). This represented 3rd

order selection (Johnson 1980),

Page 206: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

192

assessing the selection of individual bighorn sheep based on population-level availability

(Erickson et al. 2001).

Because the ratio of used/unused locations is unknown in a use-available design,

we employed the exponential approximation to the logistic model (Johnson et al. 2006) to

compare used and available locations to estimate a relative probability of use (w(x)):

w(x) = exp(β1x1 + β2x2 . . . + βpxp + γ0j)

as a function of habitat covariates (xi), their respective selection coefficients (βi), and a

random intercept for each animal- and year-specific dataset (γ0j). We included the random

intercept to account for autocorrelation within individuals and differences in sample size

among animals (Gillies et al. 2006). Note that the fixed intercept estimated from logistic

regression is dropped by convention (because the ratio of used/unused is unknown) but

the calculation of a mixed-effects logistic regression model with random intercepts will

often change the fixed-effect coefficients (Gillies et al. 2006). A coefficient > 0 indicated

selection for a habitat covariate, whereas a coefficient of < 0 indicated avoidance, with

values estimated from covariate availability.

Due to small numbers of GPS-collared bighorn sheep in each population, we only

included covariates found most important in bighorn sheep habitat studies (Smith et al.

1991; Bleich et al. 1997; McKinney et al. 2003; DeCesare and Pletscher 2006, Bleich et

al. 2008), focusing on factors related to topography, vegetation, and risk of predation.

Topographic variables included elevation, slope, aspect, and terrain ruggedness, derived

from 30 m USGS Digital Elevation Models (DEMs). Elevation and slope values were

generated directly from DEMs. We coded aspect as a continuous variable from -1 to 1

following Cushman and Wallin (2002). We estimated terrain ruggedness using an index

Page 207: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

193

developed by Sappington et al. (2007) which incorporates heterogeneity in slope and

aspect and has been found important in bighorn sheep habitat selection.

To account for vegetation we included a categorical variable for forested land

cover types, as they are strongly avoided by bighorn sheep (Risenhoover and Bailey

1985, Smith et al. 1991, DeCesare and Pletscher 2006). We used the dominant vegetation

class in U.S. Forest Service Calveg maps (www.fs.fed.us/r5/rsl/projects/mapping) to

categorize pixels as either forested or non-forested (the reference class). We also included

the Normalized Difference Vegetation Index (NDVI) in habitat models, a satellite-driven

measure of primary productivity (Huete et al. 2002). NDVI values were obtained every

16 days from 250 m2 MODIS images. NDVI values should serve as a proxy for forage

quality, and have been correlated to demographic and habitat use patterns of several other

ungulate species (Pettorelli et al. 2005, Pettorelli et al. 2007, Hebblewhite et al. 2008,

Hamel et al. 2009). Because NDVI values varied through the winter in conjunction spring

green-up (which begins at low elevations in Feb; Greene et al. In Preparation), each GPS

location was attributed with the NDVI value from the satellite image closest to the date

the location was recorded. Available locations were randomly assigned a date and

attributed with the corresponding NDVI value, allowing selection for forage quality to

vary dynamically throughout the winter.

We defined risk of mountain lion predation for bighorn sheep as the relative

probability of encountering a lion during prime hunting hours (Lima and Dill 1990).

Investigators have found that encounter rates and kill rates can be quite different for

coursing predators (i.e. wolves), with kill success mediated by habitat conditions and

prey vulnerability (Husseman et al. 2003, Hebblewhite et al. 2005, Kauffman et al. 2007).

Page 208: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

194

There is no evidence to suggest this is the case for more effective ambush predators, like

mountain lions (Husseman et al. 2003), so we assumed that encountering a lion while it is

hunting equated to risk of lion predation. To estimate predation risk we used GPS

locations from 21 collared lions. Lions were captured in and around bighorn sheep habitat

to monitor their impact on endangered bighorn sheep populations. Mountain lions were

captured with hounds between 2002 and 2009 using methods described in Davis et al.

(1996). Each lion was fitted with a GPS collar manufactured by North Star, Lotek

Wireless Inc., or Televilt International and programmed to collect locations every 4, 6, or

8 hours on a revolving schedule. We only used winter locations (Dec-Apr) collected from

1 hour pre-sunset to 1 hour post-sunrise (Pierce et al. 1998). Within these specifications

we excluded “clusters” of nighttime locations indicative of kill/feeding sites, keeping

only the first location to represent lion hunting occurrence. From the remaining 5,673

available locations we estimated a KDE (as estimated for bighorn sheep and deer

locations) to calculate the spatial probability density function of lions adjacent to bighorn

sheep winter ranges. We validated our lion predation risk layer with the 52 out-of-sample

lion-killed bighorn sheep (Hebblewhite and Merril 2007). We used a Spearman’s rank

correlation test to compare the area-adjusted frequency of predation risk values in

bighorn sheep winter range (5 bins) to the number of lion-killed sheep within the same

frequency bins (Boyce et al. 2002). Our index of risk and kill sites had a correlation

coefficient of 0.872 (p = 0.054), indicating that our layer strongly reflected risk of

predation.

We examined habitat covariates for collinearity to determine that no two variables

were highly related using correlation coefficients (r > |0.6|) and variance inflation factors

Page 209: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

195

(VIF > 5; Menard 1995). Elevation and predation risk were confounded and negatively

correlated with one another (r2 for different populations ranged from 0.52 to 0.65), as

predation risk increased in low elevation areas and decreased in high elevation areas. We

removed elevation from multivariate analyses that included predation risk, and evaluated

selection for elevation in a separate univariate analysis for each population. We

conducted univariate tests of all habitat variables (Hosmer and Lemeshow 2000), using a

cut-off value of p = 0.1 (based on Wald z statistics) for inclusion into habitat models. We

also modeled all effects as linear as univariate tests revealed no non-linear functions.

To test whether lion predation influenced bighorn sheep habitat selection we first

fit a baseline model that included only topographic (except elevation) and vegetation

covariates. We then added predation risk to baseline models to determine whether the

inclusion of risk improved model fit, and the direction and magnitude of its effect.

Finally, if risk did improve model fit, we tested whether selection for risk varied over the

course of the winter by including a risk by date interaction. Greene et al. (In Preparation)

found that forage quality was high on low elevation bighorn sheep winter ranges by the

end of the winter. Within the onset of green-up, we expected high forage benefits in risky

areas (being at low elevation) may create temporal variation in the relationship between

selection and risk.

Model selection was conducted using Akaike’s information criterion (Burnham

and Andersen 2002) with the correction for small sample sizes (AICC) based on the

number of GPS collared females/year/population. We validated the predicative power of

the best model for each herd with cross-validation (Boyce et al. 2002) using out-of-

sample GPS locations collected during the winters of 2002 through 2006. We randomly

Page 210: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

196

selected 1,000 GPS locations collected at Wheeler, Baxter, and Langley, and 700

locations from Mono Basin (as there were fewer available locations), which were

obtained from 13, 4, 6, and 3 adult females, respectively.

As a second measure of antipredator responses to predation risk, we tested

whether bighorn sheep employed larger group sizes as an antipredator strategy. We used

linear regression to determine whether group size was a function of predation risk. Group

size information was obtained from ground observations of bighorn sheep where the

composition and location of each group was recorded. To conduct this analysis we only

had adequate winter observation data for the Wheeler and Baxter populations. Winter

observations were recorded at Wheeler between 2001 and 2009 and at Baxter between

2002 and 2009. All analyses were conducted in StataMP 11.0 (StataCorp 2009).

RESULTS

Spatial Overlap of Bighorn and Deer Winter Ranges

The CVh values used to generate kernel density models from bighorn sheep and

deer locations were 64.83 and 597.53, respectively. Given the 95% kernel volume

contours, the amount of spatial overlap between bighorn sheep and deer winter ranges

varied considerably by population. There was no winter range overlap with deer in Mono

Basin, 4.39 km2 overlap with deer at Wheeler, 6.24 km

2 overlap with deer at Baxter, and

1.39 km2 overlap with deer at Langley (Fig. 5.2).

Direct Effects of Predation

Of the 172 bighorn sheep collared, 62 died during the course of the study (37

females and 25 males); 17 in Mono Basin, 19 in Wheeler, 19 in Baxter, and 7 in Langley.

Thirty-nine (63%) were attributed to a known cause while 23 (37%) were classified as

Page 211: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

197

unknown. Across all populations 22 deaths were assigned to lion predation, 8 to physical

injury, and 9 to other factors. Cumulative annual hazard rates for lion predation were 0 in

Mono Basin (no lion predation detected), 0.03 in Langley, 0.05 in Wheeler, and 0.12 in

Baxter (Table 5.1), conservative estimates given that many unknown deaths were also

likely due to lion predation. As predicted from the apparent competition hypothesis, rates

of lion predation were positively correlated with the amount of spatial overlap between

bighorn sheep and deer winter ranges (r2 = 0.94, p = 0.06, n = 4; Table 5.1). In Baxter

and Wheeler, lion predation was the dominant mortality factor, while unknown factors

were responsible for most mortality in Mono Basin and Langley. Mono Basin had the

highest rate of mortality by “other” factors, which included 3 deaths attributed to

malnutrition, 1 death to coyote predation, and 1 death to road kill; the only instances

those factors were recorded. Across all populations, mean annual bighorn sheep survival

was 0.87 (SE = 0.02). Annual survival was 0.80 in Baxter, 0.82 in Mono Basin, 0.89 in

Langley, and 0.90 in Wheeler (Table 5.1). Although Baxter, the population with the

greatest overlap with deer, also had the lowest annual survival, generally survival rates

were not correlated to the area of deer spatial overlap (r2 = -0.16, p = 0.84, n = 4).

As predicted by the apparent competition hypothesis, we found that the number of

lion-killed bighorn sheep was inversely related to distance-to-deer winter range (52

collared and un-collared lion-kills Spearman’s rank correlation = -0.90, p = 0.037, n = 5;

21 collared bighorn sheep lion kills correlation = -0.87, p = 0.054, n = 5). Of all known

lion mortalities, 58% occurred within delineated deer winter range and 79% occurred

within 500 m of deer winter range (see example of locations of lion-kills at Baxter in Fig.

5.3).

Page 212: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

198

Indirect Effects of Predation

Independent of predation risk, bighorn sheep during winter in all populations

avoided forested areas, selected for SSW aspects, and avoided areas of higher NDVI

(Table 5.2). NDVI did not appear to adequately track forage quality for bighorn sheep

during winter as higher values were often associated with forest cover, areas avoided by

bighorn sheep. From the univariate analysis on selection for elevation, we found that

populations overlapping deer range selected for lower elevations, while Mono Basin,

with no overlap with deer, selected for high elevations (Fig. 5.4; Table 5.2). Bighorn

sheep populations selected for elevation in accordance with the amount they overlapped

with deer range, such that Baxter selected the lowest elevations and had the greatest

overlap with deer, followed by Wheeler, Langley and Mono Basin, respectively.

For Wheeler, Baxter, and Langley, the populations that experienced quantifiable

risk of lion predation, models that included risk fit significantly better than baseline

models including only topographic and vegetation characteristics (Table 5.3).

Paradoxically, however, for all 3 populations there was no evidence that bighorn sheep

avoided areas of high lion density, but that they selected positively for areas used by lions

(Table 5.2; Fig. 5). Selection for risky habitat was greatest in Langley, the population

with the least overlap with deer, and lower in Wheeler and Baxter, populations with

higher overlap with deer (Fig. 5). When we included an interaction term to test whether

selection for risk varied over the course of the winter, we found that the interaction

significantly improved the fit of the Baxter model (Table 5.3). Bighorn sheep at Baxter

increased their selection for areas of high predation risk as the winter progressed and

avoided areas of low risk (Fig.5.6).

Page 213: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

199

While bighorn sheep did not avoid areas of high predation risk, populations with

higher overlap with deer did select safer terrain, as predicted from the apparent

competition hypothesis. Of the 3 populations with lion predation, bighorn sheep selected

for slope and terrain ruggedness proportional to the amount of their spatial overlap with

deer (Fig. 5). Bighorn sheep in Mono Basin, which had no overlap with deer, had

intermediate selection for slope, but showed the strongest selection for terrain ruggedness

(Table 5.2).

All habitat selection models had high predictive power when tested against out-

of-sample GPS locations. Within 10 area-adjusted frequency bins of predicted habitat

quality, Spearman rank correlations between expected and observed probabilities of

selection were 0.94 for Mono Basin (p < 0.001), 0.91 for Wheeler (p < 0.001), 0.98 for

Baxter (p < 0.001), and 0.82 for Langley (p = 0.004).

To test whether bighorn sheep exhibited increasing group sizes as an antipredator

strategy we used 631 winter observations of bighorn sheep from Wheeler and 159

observations from Baxter. At Wheeler there was a positive relationship between group

size and predation risk (F = 60.33, p < 0.001, r2 = 0.09), while at Baxter the relationship

was not significant (F = 1.57, p = 0.21, r2 = 0.01).

DISCUSSION

As predicted by the apparent competition hypothesis, we found that endangered

bighorn sheep populations with high spatial overlap with deer exhibited higher rates of

lion predation, lion-kills within and adjacent to deer ranges, and stronger antipredator

behavior. Mountain lions had both direct and indirect effects on bighorn sheep

populations, influencing demographic rates, habitat use patterns, and grouping behaviors.

Page 214: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

200

We also found that the impacts of asymmetric predation were highly spatially and

temporally variable. Spatial variation appeared to be generated by landscape-scale

differences in selection for elevation which dictated the area of overlap with deer.

Temporal variation occurred in response to changes in finer-scale forage availability with

the onset of spring green-up. The repercussions of asymmetric predation limits recovery

in some herds, but not in others, providing key implications for conservation strategies of

this subspecies. Although this study only evaluated 4 bighorn sheep populations (out of a

total of 6 populations in the subspecies), the suite of detailed demographic, habitat, and

observation data on an endangered species coupled with spatial data on the primary prey

and shared predator provides a unique opportunity to elucidate the causes and

consequences of apparent competition in a natural system.

Several lines of evidence support our predictions that the direct effects of lion

predation on bighorn sheep were triggered by spatial overlap with deer. Rates of

mountain lion predation occurred in direct accordance with the amount of overlap

between bighorn sheep and deer winter ranges (Table 5.1). Furthermore, approximately

80% of lion killed bighorn sheep were located within 500 m of deer winter range (Fig.

5.3), even though only about 20% of the available winter bighorn sheep habitat was

within that distance. Of the populations with lion predation, the herd (Baxter) with the

greatest overlap with deer also had the lowest survival rates. This evidence of direct

predation, particularly on adults, is of concern as Johnson et al. (2010) found that adult

survival explained most of the variation in growth rates of bighorn sheep in the Sierra

Nevada, with low and highly variable rates responsible for population declines. We

estimated population-specific adult survival rates as low as 80% with 12% of annual

Page 215: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

201

mortality conservatively attributed to lion predation; rates that are unsustainable for

persistence (Gaillard et al. 2000, Johnson et al. 2010). For long-lived species, direct

predation on adults can inflict a dramatic toll, by decreasing mean adult survival rates

with high proportional effects on population growth (Bryant and Page 2005, Owen-Smith

and Mason 2005, Wittmer et al. 2005b, Angulo et al. 2007), increasing the variation in

those rates (Gasaway et al. 1992), or both.

Results from analyses on the direct effects of predation revealed some patterns

that were unexpected by our apparent competition predictions. Mono Basin had no

overlap with deer or any confirmed lion predation during this study, yet it had the second

lowest rate of annual survival. This is the smallest herd, with only 35 individuals

accounted for in the winter of 2009. It appears that stochastic factors may be largely

responsible for low demographic rates as Johnson et al. (In Press; In Preparation) found

that Allee effects, environmental stochasticity, and inbreeding depression were driving

variation in survival and reproductive rates in this herd. Another unexpected result was

that Langley, with low rates of predation and minimal overlap with deer, had an annual

survival rate comparable to Wheeler, a herd with higher rates of predation and deer

overlap. We expect that the low survival rate at Langley is largely a function of sample

size as this herd had the fewest collared individuals (and only 7 deaths) and was

monitored for the shortest number of years. Based on count-data, Johnson et al. (2010)

reported that the mean adult female survival rate between 1997 and 2007 was 0.98 for

this herd, closer to expectations based on minimal overlap with deer and lion populations.

Recent studies have illustrated that the indirect effects of predation can have just

as great of an influence on prey populations as direct predation itself (Schmitz et al. 2007,

Page 216: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

202

Preisser et al. 2005, Creel et al. 2007, Cresswell 2008). Indeed, Wehausen (1996)

suggested that Sierra Nevada bighorn sheep abandoned low elevation winter ranges in the

1980’s and early 1990’s due to high rates of lion predation. By remaining at high

elevations, he concluded that bighorn sheep had reduced forage quality, which in turn,

depressed recruitment rates and contributed to population declines. We found, however,

that bighorn sheep did not avoid areas of high predation risk, but actually selected for

areas of risk (Fig. 5). Our results corroborate several other recent studies showing that

ungulates select habitat primarily based on topography and vegetation, not by avoiding

predation risk (Walker et al. 2007, Kittle et al. 2008, Valeix et al. 2009). In a study of

several African ungulates Valeix et al. (2009) found that browsers avoided predation risk

while grazers did not. They speculated that diet constraints on grazers may inhibit them

from spatially avoiding areas of high risk if those same areas also have high forage

quality. We expect that this is also the case for bighorn sheep in the Sierra Nevada as

areas of high risk are also low in elevation, have higher forage quality (Greene et al. In

Preparation), and often overlap with deer. Positive coefficients for predation risk

probably reflect selection for desired vegetation, not risk itself, a relationship that was not

adequately captured by NDVI with 250m2 pixels. It appears that in our arid study system,

during winter months, NDVI was more strongly correlated with forested vegetation

types, not grasses, forbs and shrubs consumed by bighorn sheep.

Although bighorn sheep selected for areas of higher risk of lion predation, this

selection was temporally variable over the course of the winter. A key result from our

habitat analysis was that bighorn sheep in Baxter, the herd with the highest predation rate

and overlap with deer, dramatically increased their use of risky areas as the winter

Page 217: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

203

progressed (Fig. 5.6). In early winter, bighorn sheep showed no preference for risky

habitat as the energetic requirements for pregnant ungulates are minimal, body condition

is adequate, and forage quality is low at all elevations (Perkins et al. 1998, Parker et al.

2009, CDFG unpublished data). By early spring (Feb-Apr in our study area), however,

ungulates are at their poorest body condition and have high energetic costs associated

with the last trimester of pregnancy. Simultaneously green-up is commencing on low

elevation ranges (Greene et al. In Preparation), drawing bighorn sheep down into areas of

overlap with deer and lions. This situation exacerbates the impact of apparent

competition as the spatial overlap between bighorn sheep and deer increases during

critical late winter months. It also demonstrates that asymmetric predation can be

temporally variable, only having demographic repercussions during a specific time period

(Fig. 5.1).

While bighorn sheep do not avoid predation risk, they do appear to perceive and

attempt to mediate their risk through various antipredator behaviors (Lima and Dill 1990,

Gude et al. 2006). Populations with predation selected “safer” terrain in relation to their

overlap with deer (Fig. 5). Such selection patterns may allow bighorn sheep to forage in

areas inhabited by lions while mitigating risk at finer spatial scales than we measured

(Poole et al. 2007, Halofsky and Ripple 2008, Hebblewhite and Merrill 2009). Ungulates

also have been observed to forage in larger groups when risk of predation is high (Isvaran

2007, Fortin and Fortin 2009, Schroeder et al. 2010). We found that group sizes at

Wheeler did indeed increase in areas of higher risk, while results were inconclusive at

Baxter. This difference may be a function of the available data as observations at

Wheeler spanned a wide range of risk values (mean = 7.91 ± 0.45) while limitations in

Page 218: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

204

sightability resulted in observations at Baxter that were all primarily collected in high risk

zones (mean = 15.65 ± 0.74). While the potential fitness costs of these antipredator

behaviors are unknown in our system, as separating the indirect and direct effects of

predation is not possible, other studies suggest that selection for steeper, rugged terrain

and increasing group size could significantly reduce forage intake for ungulates (Fortin et

al. 2004, Hamel and Côté 2007, Isvaran 2007).

Contrary to the other herds, bighorn sheep in Mono Basin did not exhibit expected

patterns of habitat selection. Given that bighorn sheep in Mono Basin had no overlap

with deer and no measurable lion predation we expected them to select benign terrain

relative to other herds. Instead, they showed the strongest selection for terrain ruggedness

(Table 5.2). This behavior is likely a function of their use of high elevation habitat, while

all other populations use lower elevations (Fig. 5.4). In a post-hoc analysis we calculated

the amount of low elevation (< 2745 m), non-forested habitat available within population

MCPs from the GPS collar data. Mono Basin had 0.1 km2

of low elevation habitat

available, Wheeler had 26.0 km2, Baxter had 32.8 km

2 and Langley had 5.4 km

2. Thus, it

appears that, within their home ranges, populations selected for low elevation habitat in

accordance to its relative availability, showing a positive functional response for low

elevation winter range (Mysterud and Ims 1998, Osko et al. 2004). Bighorn sheep in

Mono Basin had virtually no low elevation winter habitat within their MCP area, and as a

result, they persisted through the winter at high elevations. High elevation habitat consists

of wind-blown slopes free from snow, which are also steep and rugged (Walker et al.

2007).

Page 219: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

205

The availability of low elevation habitat appears to drive the direct and indirect

effects of apparent competition. Regressing the availability of low elevation habitat for

each population against its respective spatial overlap with deer we found almost a perfect

1:1 relationship (r2 = 0.98, F = 103.1, p = 0.01). Thus, with greater availability of low

elevation habitat, bighorn sheep increase their selection for those areas, enlarge their

spatial overlap with deer, amplify their risk of predation, and are ultimately taxed by

direct and indirect fitness consequences. The powerful role of apparent competition on

bighorn sheep populations stresses the need to incorporate biotic interactions into habitat

evaluations (Araujo and Luoto 2007, Ritchie et al. 2009). Topographic and vegetation

characteristics alone are not adequate to identify high quality habitat for bighorn sheep in

the Sierra Nevada, given the influence of deer and lions in shaping their demography.

Instead, habitat must be assessed from the perspective of the complete ecological niche,

encompassing both abiotic and biotic habitat factors to account for interactions in habitat

selection and asymmetric predation (Hirzel and Le Lay 2008).

Effects of apparent competition have strong implications for conservation and

management. For endangered populations hindered by asymmetric predation (i.e.

Baxter), managers can eliminate the predators and potentially the primary prey (Lessard

et al. 2005, DeCesare et al. In Press), recognizing that predators will likely need to be

removed first to avoid enhanced predation on endangered species (Courchamp et al.

2003, Collins et al. 2009). Given the high societal value of deer for human harvest in the

eastern Sierra, lion removal is the best short-term management option for increasing

bighorn sheep survival and boosting population growth rates (Johnson et al. 2010),

acknowledging that this could also contribute to higher deer densities. Because lion

Page 220: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

206

predation only influences bighorn sheep during winter months in specific areas of deer

overlap, lion control can be targeted temporally or at the individual lion level (Sanz-

Aguilar et al. 2009), minimizing the controversy regarding control actions among the

broader California public. Such lion control should be viewed as a short-term strategy for

maximizing growth rates of bighorn sheep populations to 1) restore large, healthy herds,

and 2) supply source stock for translocations and reintroductions. As mangers implement

recovery goals and reintroduce new populations of bighorn sheep they must also account

for the spatial distribution of deer and lions and temporal shifts in bighorn sheep habitat

selection. Reintroduction sites should either be in areas where low elevation winter range

is de-coupled from deer and lion populations or areas where high elevation habitat can

successfully sustain bighorn sheep throughout the winter. Such places can serve as

refugia for bighorn sheep recovery, minimizing the need for predator control and

continual management (Sinclair et al. 1998, Scott et al. 2005). Finally, adult survival

should be closely monitored in all bighorn sheep populations, as it is the most critical

vital rate for detecting unsustainable rates of predation and directing management

activities.

The combination of having abundant deer, generalist lion predators, and overlap

in winter ranges of bighorn sheep and deer has created the perfect storm for apparent

competition to limit population performance of endangered bighorn sheep. While initially

bighorn sheep were reduced by disease and overexploitation, several lines of evidence

now suggest that recovery of some populations is hampered by lions. As both native and

non-native prey increase and expand their distributions (Côté et al. 2004, Spear and

Chown 2009) the effects of apparent competition will continue to threaten small and

Page 221: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

207

endangered prey populations (Seip et al. 1992, Robinson et al. 2002, Wittmer et al. 2005).

Studies of endangered species typically focus on only the direct effects of apparent

competition, but we observed indirect antipredator behavior that may also inflict

significant fitness costs. Additionally, we found that the effects of asymmetric predation

were both spatially and temporally variable, with spatial variation driven by landscape-

scale habitat availability and temporal variation driven by finer-scale forage availability.

Management strategies for small populations should consider both the habitat selection

patterns of prey and the spatial dynamics of competitors and predators to reduce the

potential for apparent competition to hijack conservation success.

ACKNOWLEDGEMENTS

We are grateful to all the people who have collected data on bighorn sheep, mule

deer, and mountain lions including L. Bowermaster, J. Davis, V. Davis, K. Ellis, J.

Erlenbach, A. Feinberg, J. Fusaro, D. Jensen, M. Kiner, K. Knox, K. Monteith, J.

Ostergard, P. Partridge, B. Pierce, C. Schroeder, D. Spitz, A. Stephenson, J. Wehausen,

and others. We thank Landells Aviation and Quicksilver Air for safely conducting

numerous bighorn sheep and deer helicopter captures. We are grateful to L.S. Mills, M.

Mitchell, D. Pletscher, G. Luikart, and H. Robinson for providing constructive discussion

and comments on this manuscript. CDFG and the Canon National Parks Scholars

Program provided funding for this work.

LITERATURE CITED

Angulo E, Roemer G.W., Berec L., Gascoigne J., and Courcamp F. 2007. Double Allee

effects and extinction in the Island Fox. Conservation Biology 21:1082-1091.

Araujo M.B., and Luoto M. 2007. The importance of biotic interactions for modelling

Page 222: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

208

species distributions under climate change. Global Ecology and Biogeography

16:743-753.

Bleich, V.C., Bowyer T.R., and Wehausen J.D. 1997. Sexual segregation in mountain

sheep: resources or predation? Wildlife Monograph 134:1-50.

Bleich V.C., Johnson H.E., Holl S.A., Konde L., Torres S.G., and Krausman P.R. 2008.

Fire history in a chaparral ecosystem: implications for conservation of a native

ungulate. Rangeland Ecology and Management 61:571-579.

Boyce M.S., Vernier P.R., Nielsen S.E., and Schmiegelow F.K.A. 2002. Evaluating

resource selection functions. Ecological Modelling 157:281-300.

Beyer, H.L. 2004. Hawth's Analysis Tools for ArcGIS. Available at

http://www.spatialecology.com/htools.

Bryant A.A., and Page R.E. 2005. Timing and causes of mortality in the endangered

Vancouver Island marmot (Marmota vancouverensis). Canadian Journal of

Zoology 83:674-682.

Burnham K.P., and Anderson D.R. 2002. Model selection and multimodel inference: a

practical information-theoretic approach. Fourth Edition. Springer, New York,

USA.

Chaneton E.J., and Bonsall M.B. 2000. Enemy-mediated apparent competition: empirical

patterns and the evidence. Oikos 88:380-394.

Chase J.M., Abrams P.A., Grover S., Diehl J.P., Chesson P., Holt R.D., Richards S.A.,

Nisbet R.M., and Case T.J. 2002. The interaction between predation and

competition: a review and synthesis. Ecology Letters 5:302-315.

Chesson P., and Kuang J.J. 2008. The interaction between predation and competition.

Page 223: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

209

Nature 456:235–238.

Courchamp F., Woodroffe R., and Roemer G. 2003. Removing protected populations to

save endangered species. Science 302:1532.

Collins P.W., Latta B.C., and Roemer G.W. 2009. Does the order of invasive species

removal matter? The case of the eagle and the pig. Plos One 4:e7005.

Côté S.D., Rooney T.P., Tremblay J.P., Dussault C., and Waller D.M. 2004. Ecological

impacts of deer overabundance. Annual Review of Ecology Evolution and

Systematics 35:113-147.

Creel S., Christianson D., Liley S., and Winnie J.A. 2007. Predation risk affects

reproductive physiology and demography of elk. Science 315:960.

Creel S., and Christianson D. 2008. Relationships between direct predation and risk

effects. Trends in Ecology and Evolution 23:194-201.

Cresswell W. 2008. Non-lethal effects of predation in birds. Ibis 150:3-17.

Cushman S.A., and Wallin D.O. 2002. Separating the effects of environmental, spatial

and disturbance factors on forest community structure in the Russian Far East.

Forest Ecology and Management 168:201-215.

Davis J.L., Chetkiewicz C.-L.B., Bleich V.C., Raygorodetsky G., Pierce B.M., Ostergard

J.W., and Wehausen J.D. 1996. A device to safely remove immobilized mountain

lions from trees and cliffs. Wildlife Society Bulletin 24:537-539.

D’Eon R.G. 2003. Effects of a stationary GPS fix-rate bias on habitat selection analyses.

Journal of Wildlife Management 67:858-863.

DeCesare N., Hebblewhite M., Robinson H., and Musiani M. In Press. Endangered,

Page 224: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

210

apparently: the role of apparent competition in endangered species conservation.

Animal Conservation. Doi:10.1111/j.1469-1795.2009.00328.x

DeCesare N., and Pletscher D. 2006. Movements, connectivity, and resource selection of

Rocky Mountain bighorn sheep. Journal of Mammology 87:531-538.

Erickson W.P., McDonald T.L., Gerow K.G., Howlin S., and Kern J.W. 2001. Statistical

issues in resource selection studies with radio-marked animals. Pages 209-242 in

J.J. Millspaugh and J.M. Marzluff, editors. Radio-tracking and animal

populations. Academic Press, New York, USA.

Fortin D., Beyer H.L., Boyce M.S., Smith D.W., Duchesne T., and Mao J.S. 2005.

Wolves influence elk movements: behavior shapes a trophic cascade in

Yellowstone National Park. Ecology 86:1320-1330.

Fortin D., Boyce M.S., Merrill E.H., and Fryxell J.M. 2004. Foraging costs of vigilance

in large mammalian herbivores. Oikos 107:172-180.

Fortin D., and Fortin M.E. 2009. Group-size dependent association between food

profitability, predation risk and distribution of free-ranging bison. Animal

Behaviour 78:887-892.

Frair J.L., S.E. Nielsen, Merrill E.H., Lele S.R., Boyce M.S., Munro R.H.M., Stenhouse

G.B., and Beyer H.L. 2004. Removing GPS collar bias in habitat selection

studies. Journal of Applied Ecology 41:201-212.

Gaillard, J.-M., Festa-Bianchet M., Yoccoz N.G., Loison A., and Toïgo C. 2000.

Temporal variation in fitness components and population dynamics of large

herbivores. Annual Review of Ecology and Systematics 31:367-393.

Gasaway W.C., Boertje R.D., Grangaard D.V., Kelleyhouse D.G., Stephenson R.O., and

Page 225: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

211

Larsen D.G. 1992. The role of predation in limiting moose at low densities in

Alaska and Yukon and implications for conservation. Wildlife Monographs

120:1-59.

Geist V. 1970. Mountain Sheep. University of Chicago Press, Illinois, USA.

Gillies C.S., Hebblewhite M., Nielsen S.E., Krawchuk M.A., Aldridge C.L., Frair J.L.,

Saher D.J., Stevens C.E., and Jerde C.L. 2006. Application of random effects to

the study of resource selection by animals. Journal of Animal Ecology 75:887-

898.

Greene L., Stephenson T., and Hebblewhite M. Short-term effects of wildfire on the

winter range of Sierra Nevada bighorn sheep. In Preparation.

Gude J.A., Barrott R.A., Borkowski J.J., and King F. 2006. Prey risk allocation in a

grazing ecosystem. Ecological Applications 16:285-298.

Halofsky J.S., and Ripple W.J. 2008. Fine-scale predation risk on elk after wolf

reintroduction in Yellowstone National Park, USA. Oecologia 155:869-877.

Hamel S., and Côté S.D. 2007. Habitat use patterns in relation to escape terrain: are

alpine ungulate females trading off better foraging sites for safety? Canadian

Journal of Zoology 85:933-943.

Hamel S., Garel M., Festa-Bianchet M., Gaillard J.M., and Côté S.D. 2009. Spring

normalized difference vegetation index (NDVI) predicts annual variation in

timing of peak faecal crude protein in mountain ungulates. Journal of Applied

Ecology 46:582-589.

Harmon J.P., and Andow D.A. 2004. Indirect effects between shared prey: predictions for

biological control. BioControl 49:605-626.

Page 226: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

212

Hebblewhite M., and Merrill E.H. 2007. Multiscale wolf predation risk for elk: does

migration reduce risk? Oecologia 152:377-387.

Hebblewhite M., and Merrill E.H. 2009. Trade-offs between risk and forage differ

between migrant strategies in a migratory ungulate. Ecology 90:3445-3454.

Hebblewhite M., Merrill E., McDermid G. 2008. A multi-scale test of the forage

maturation hypothesis in a partially migratory ungulate population. Ecological

Monographs 78:141-166.

Hebblewhite M., Merrill E.H., and McDonald T.L. 2005. Spatial decomposition of

predation risk using resource selection functions: an example in a wolf-elk

predator-prey system. Oikos 111:101-111.

Heisey D.M., and Patterson B.R. 2006. A review of methods to estimate cause-specific

mortality in presence of competing risks. Journal of Wildlife Management

70:1544-1555.

Hill M. 1975. Geology of the Sierra Nevada. The University of California Press,

California, USA.

Hirzel A.H., and Le Lay G. 2008. Habitat suitability modeling and niche theory. Journal

of Applied Ecology 45:1372-1381.

Holling, C.S. 1959. The components of predation as revealed by a study of small-

mammal predation of the European pine sawfly. The Canadian Entomologist

91:293–320.

Holt R.D. 1977. Predation, apparent competition and the structure of prey communities.

Theoretical Population Biology 12:197-229.

Holt R.D., and Lawton J.H. 1994. The ecological consequences of shared natural

Page 227: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

213

enemies. Annual Review of Ecology and Systematics 25:495-520.

Horne J.S., and Garton E.O. 2006. Likelihood cross-validation versus least squares cross-

validation for choosing the smoothing parameter in kernel home-range analysis.

Journal of Wildlife Management 70:641-648.

Horne J.S., and Garton E.O. 2007. Animal Space Use 1.3

http://www.cnr.uidaho.edu/population_ecology/animal_space_use.htm.

Hosmer D.W., and Lemeshow S. 2000. Applied Logistic Regression. Second Edition.

John Wiley and Sons, New York, USA.

Huete A., Didan K., Miura T., Rodriguez E.P., Gao X., and Ferreira L.G. 2002. Overview

of the radiometric and biophysical performance of the MODIS vegetation indices.

Remote sensing of environment 83:195-213.

Husseman J.S., Murray D.L., Power G., Mack C., Wenger C.R., and Quigley H. 2003.

Assessing differential prey selection patterns between two sympatric large

carnivores. Oikos 101:591-601.

Isvaran K. 2007. Intraspecific variation in group size in the blackbuck antelope: the roles

of habitat structure and forage at different spatial scales. Oecologia 154:435-444.

James A.R.C., Boutin S., Hebert D.M., and Rippin A.B. 2004. Spatial separation of

caribou from moose and its relation to predation by wolves. Journal of Wildlife

Management 68:799-809.

Johnson D.H. 1980. The comparison of usage and availability measurements for

evaluating resource preference. Ecology 61:65-71.

Johnson C.J., Nielsen S.E., Merrill E.H., McDonald T.L., and Boyce M.S. 2006.

Page 228: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

214

Resource selection functions based on use-availability data: theoretical

motivations and evaluation methods. Journal of Wildlife Management 70:347-

357.

Johnson H.E., Mills L.S., Stephenson T.S., and Wehausen J.D. 2010. Population-specific

vital rate contributions influence management of an endangered ungulate.

Ecological Applications 20:1753-1765.

Johnson H.E., Mills L.S., Wehausen J.D., and Stephenson T.S. Combining minimum

count, telemetry, and mark-resight data to infer population dynamics in an

endangered species. In Press: Journal of Applied Ecology.

Johnson H.E., Mills L.S., Wehausen J.D., and Stephenson T.S. From lambs to lambda:

effects of inbreeding depression on recovery of endangered bighorn sheep. In

Preparation.

Kauffman M.J., Varley N., Smith D.W., Stahler D.R., MacNulty D.R., and Boyce M.S.

2007. Landscape heterogeneity shapes predation in a newly restored predator-prey

system. Ecology Letters 10:690-700.

Kittle A.M., Fryxell J.M., Desy G.E., and Hamr J. 2008. The scale-dependent impact of

wolf predation risk on resource selection by three sympatric ungulates. Oecologia

157:163-175.

Krausman P.R., Hervert J.J., and Ordway L.L. 1985. Capturing deer and mountain

sheep with a net-gun. Wildlife Society Bulletin 13:71-73.

Lessard R.B., Martell S.J.D., Walters C.J., Essington T.E., and Kitchell J.F. 2005. Should

ecosystem management involve active control of species abundances? Ecology

and Society 10:1-23.

Page 229: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

215

Lima S.L., and Dill L.M. 1990. Behavioral decisions made under the risk of predation: a

review and prospectus. Canadian Journal of Zoology 68:619-640.

Manly B.F.J., McDonald L.L., Thomas D.L., McDonald T.L., and Erickson W.P. 2002.

Resource selection by animals; statistical design and analysis for field studies.

Second Edition. Kluwer Academic Publishers, Massachussetts, USA.

Mao J.S., Boyce M.S., Smith D.W., Singer F.J., Vales D.J., Vore J.M., and Merrill E.H.

2005. Habitat selection by elk before and after wolf reintroduction in Yellowstone

National Park. Journal of Wildlife Management 69:1691-1707.

McKinney T., Boe S.R., and deVos J.C. Jr. 2003. GIS-based evaluation of escape terrain

and desert bighorn sheep populations in Arizona. Wildlife Society Bulletin

31:1129-1236.

McLellan B.N., Serrouya R., Wittmer H.U., and Boutin S. 2010. Predator-mediated Allee

effects in multi-prey systems. Ecology 91:286-292.

McLoughlin P.D., Dunford J.S., and Boutin S. 2005. Relating predation mortality to

broad-scale habitat selection. Journal of Animal Ecology 74:701-707.

McLoughlin P.D., Coulson T., and Clutton-Brock T. 2008. Cross-generational effects of

habitat and density on life history in red deer. Ecology 89:3317-3326.

Menard S. 1995. Applied logistic regression analysis. Sage Publications, California,

USA.

Messier, F. 1994. Ungulate population models with predation: a case study with the

North American moose. Ecology 75:478-488.

Mills L.S. 2007. Conservation of wildlife populations: demography, genetics and

management. Blackwell Publishing, Massachusetts, USA.

Page 230: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

216

Mysterud A., and Ims R.A. 1998. Functional responses in habitat use: availability

influences relative use in trade-off situations. Ecology 79:1435-1441.

Nelson E.H., Matthews C.E., and Rosenheim J.A. 2004. Predators reduce prey population

growth by inducing changes in prey behavior. Ecology 85:1853-1858.

Osko T.J., Hiltz M.N., Hudson R.J., and Wasel S.M. 2004. Moose habitat preferences in

response to changing availability. Journal of Wildlife Management 68:576-584.

Owen-Smith N., and D.R. Mason. 2005. Comparative changes in adult vs. juvenile

survival affecting population trends of African ungulates: Journal of Animal

Ecology 74:762-773.

Parker K.L., Barboza P.S., and Gillingham M.P. 2009. Nutrition integrates environmental

responses of ungulates. Functional Ecology 23:57-69.

Perkins P.J., Smith K.S., and Mautz W.W. 1998. The energy costs of gestation in white-

tailed deer. Canadian Journal of Zoology 76:1091-1097.

Pettorelli N., Vik J.O., Mysterud A., Gaillard J.-M., Tucker C.J., and Stenseth N.C. 2005.

Using the satellite-derived NDVI to assess ecological responses to environmental

change. Trends in Ecology and Evolution 20:503-510.

Pettorelli N., Pelletier F., Von Hardenberg A., Festa-Bianchet M., and Côté S.D. 2007.

Early onset of vegetation growth vs. rapid green-up: impacts on juvenile mountain

ungulates. Ecology 88:381-390.

Pierce B.M., Bleich V.C., Chetkiewicz C.L.B., and Wehausen J.D. 1998. Timing of

feeding bouts of mountain lions. Journal of Mammalogy 79:222-226.

Pierce B.M., Bleich V.C., and Bowyer R.T. 2000. Social organization of mountain lions:

does a land-tenure system regulate population size? Ecology 81:1533-1543.

Page 231: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

217

Pierce B.M., Bleich V.C., Wehausen J.D., and Bowyer R.T. 1999. Migratory patterns of

mountain lions: implications for social regulation and conservation. Journal of

Mammalogy 80:986-992.

Pollock K.H., Winterstein S.R., and Bunck C.M. 1989. Survival analysis in telemetry

studies: the staggered entry design. Journal of Wildlife Management 53:7-15.

Poole K.G., Serrouya R., and Stuart-Smith K. 2007. Moose calving strategies in interior

ecosystems. Journal of Mammalogy 88:139-150.

Preisser E.L., Bolnick D.I., and Benard M.F. 2005. Scared to death? The effects of

intimidation and consumption in predator-prey interactions. Ecology 86:501-509.

Risenhoover K.L., and Bailey J.A. 1985. Foraging ecology of mountain sheep:

implications for habitat management. Journal of Wildlife Management 49:797-

804.

Ritchie E.G., Martin J.K., Johnson C.N., and Fox B.J. 2009. Separating the influences of

environment and species interactions on patterns of distribution and abundance:

competition between large herbivores. Journal of Animal Ecology 78:724-731.

Robinson H.S., Wielgus R.B., Gwilliam J.C. 2002. Cougar predation and population

growth of sympatric mule deer and white-tailed deer. Canadian Journal of

Zoology 80:556-568.

Roemer G.W., Donlan C.J., and Courchamp F. 2002. Golden eagles, feral pigs, and

insular carnivores: how exotic species turn native predators into prey. Proceedings

of the National Academy of Sciences of the United States of America 99:791-796.

Sanz-Aguilar A., Martínez-Abraín A., Tavecchia G., Mínguez E., and Oro D. 2009.

Page 232: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

218

Evidence-based culling of a facultative predator: efficacy and efficiency

components. Biological Conservation 142:424-431.

Sappington J.M., Longshore K.M., and Thompson D.B. 2007. Quantifying landscape

ruggedness for animal habitat analysis: a case study using bighorn sheep in the

Mojave desert. Journal of Wildlife Management 71:1419-1426.

Schmitz O.J., Beckerman A.P., and Obrien K.M. 1997. Behaviorally mediated trophic

cascades: effects of predation risk on food web interactions. Ecology 78:1388-

1399.

Schroeder C.A., R.T. Bowyer, V.C. Bleich, and T.R. Stephenson. 2010. Sexual

segregation in Sierra Nevada bighorn sheep, Ovis canadensis sierrae:

ramifications for conservation. Arctic, Antarctic and Alpine Research. In Press.

Scott J.M., Goble D.D., Wiens J.A., Wilcove D.S., Bean M., and Male T. 2005. Recovery

of imperiled species under the Endangered Species Act: the need for a new

approach. Frontiers in Ecology and the Environment 3:383-389.

Seip D.R. 1992. Factors limiting woodland caribou populations and their

interrelationships with wolves and moose in southeastern British Columbia.

Canadian Journal of Zoology 70:1494-1503.

Sinclair A.R.E., Pech R.P., Dickman C.R., Hik D., Mahon R., and Newsome A.E. 1998.

Predicting effects of predation on conservation of endangered prey. Conservation

Biology 12:564-575.

Smith T.S., Flinders J.T., and Winn D.S. 1991. A habitat evaluation procedure for Rocky

Mountain bighorn sheep in the intermountain west. Great Basin Naturalist

51:205-225.

Page 233: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

219

Spear D., and Chown S.L. 2009. Non-indigenous ungulates as a threat to biodiversity.

Journal of Zoology 279:1-17.

StataCorp. 2009. Stata Statistical Software: Release 11.0. Stata Corporation, Texas, USA.

Torres S., Mansfield T.M., Foley J.E., Lupo T., and Brinkhaus A. 1996. Mountain lion

and human activity in California: testing speculations. Wildlife Society Bulletin

24:451-460.

U.S. Fish and Wildlife Service. 2007. Recovery Plan for the Sierra Nevada bighorn

sheep. Sacramento, California, USA.

Valeix M., Loveridge A.J., Chamaillé-Jammes S., Davidson Z., Murindagomo F., Fritz

H., and Macdonald D.W. 2009. Behavioral adjustments of African herbivores to

predation risk by lions: spatiotemporal variations influence habitat use. Ecology

90:23-30.

Walker A.B.D., Parker K.L., Gillingham M.P., Gustine D.D., and Lay R.J. 2007. Habitat

selection by female Stone’s sheep in relation to vegetation, topography, and risk

of predation. Ecoscience 14:55-70.

Wehausen J.D. 1996. Effects of mountain lion predation on bighorn sheep in the Sierra

Nevada and Granite Mountains of California. Wildlife Society Bulletin 24:471-

479.

Wittmer H.U., Sinclair A.R.E., and McLellan B.N. 2005a. The role of predation in the

decline and extirpation of woodland caribou. Oecologia 144:257-267.

Wittmer H.U., McLellan B.N., Seip D.R., Young J.A., Kinley T.A., Watts G.S., and

Hamilton D. 2005b. Population dynamics of the endangered mountain ecotype of

Page 234: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

220

woodland caribou (Rangifer tarandus caribou) in British Columbia, Canada.

Canadian Journal of Zoology 83:407-418.

Worton B.J. 1989. Kernel methods for estimating the utilization distribution in home-

range studies. Ecology 70:164-168.

Page 235: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

221

Table 5.1. Mean annual probabilities of cause-specific mortality, known-fate survival, and overlap with deer winter range for Sierra

Nevada bighorn sheep populations, CA.

Cause-Specific Mortality

Population Lion Physical Injury Other Unknown

Survival Deer Overlap (km2)

Mono Basin None 0.01 (0.01) 0.06 (0.02) 0.11 (0.03)

0.82 (0.04)

0

Wheeler 0.05 (0.02) 0.01 (0.01) 0.01 (0.01) 0.04 (0.01)

0.90 (0.02)

4.39

Baxter 0.12 (0.04) 0.04 (0.02) 0.02 (0.01) 0.03 (0.02)

0.80 (0.04)

6.24

Langley 0.03 (0.02) 0.02 (0.02) 0 0.06 (0.03)

0.89 (0.04)

1.39

Page 236: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

222

Table 5.2. Habitat model coefficient estimates (and SE) from the top model of each population of Sierra Nevada bighorn sheep, CA.

Coefficient Mono Basin Wheeler Baxter Langley

Multivariate Model

Slope 0.0498 (0.0029) 0.0569(0.0025) 0.0627 (0.0014) 0.0397 (0.0016)

Aspect 0.83 (0.05) 0.69 (0.03) 0.82 (0.02) 0.94 (0.03)

Ruggedness 39.50 (2.32) 8.31 (1.06) 14.34 (0.69) 6.20 (1.14)

Forest -1.41 (0.12) -0.67 (0.09) -0.87 (0.05) -1.33 (0.08)

NDVI -0.0002 (<0.0001) -4.0e-5 (<0.0001) -0.0002 (<0.0001) -0.0002 (<0.0001)

Risk NA 0.0663 (0.0023) -0.0125 (0.0044) 1.2202 (0.0514)

Time NA NA -0.1601 (0.0054) NA

Risk*Time NA NA 0.0238 (0.0007) NA

Univariate Model

Elevation 0.0031 (0.0001) -0.0008 (<0.0001) -0.0009 (<0.0001) -0.0002 (<0.0001)

Page 237: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

223

Table 5.3. Comparison of models for winter habitat selection of Sierra Nevada bighorn

sheep populations with baseline parameters alone (slope, aspect, terrain ruggedness,

forest, and NDVI), with the addition of lion predation risk (R), and with a risk by date

interaction (R*D).

Population & Model Pseudo r2 k LL AICC ΔAICC

Mono Basin

BaselineA 0.120 5 -4063 NA

A NA

Wheeler

Baseline 0.061 5 -6852 13733 928

Baseline + R 0.126 6 -6381 12805 0

Baseline + R + R*D 0.131 8 -6344 12853 48

Baxter

Baseline 0.123 5 -20627 44731 5792

Baseline + R 0.210 6 -20130 40284 1345

Baseline + R + R*D 0.237 8 -19451 38939 0

Langley

Baseline 0.108 5 -9484 18997 679

Baseline + R 0.139 6 -9151 18318 0

Baseline + R + R*D 0.140 8 -9145 18454 136

AModel selection was not conducted for Mono Basin as there was no detectable risk of

lion predation.

Page 238: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

224

Figure 5.1. Mortalities of radio-collared Sierra Nevada bighorn sheep that have occurred

each month of the year from 2002 to 2010.

0

2

4

6

8

10

12

14

16

18

20

1 2 3 4 5 6 7 8 9 10 11 12

Nu

mb

er o

f M

ort

alit

ies

Month

Page 239: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

225

Figure 5.2. Location of Mono Basin, Wheeler, Baxter, and Langley populations of Sierra

Nevada bighorn sheep, CA. Green areas represent the winter minimum convex polygons

for bighorn sheep in each population and yellow areas delineate the winter ranges of

adjacent mule deer herds.

Page 240: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

226

Figure 5.3. Locations of bighorn sheep mortalities in the Baxter population by mountain

lion predation. Green areas delineate the winter range of bighorn sheep and yellow areas

delineate the winter range of mule deer.

Page 241: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

227

Figure 5.4. Probabilities of selection for elevation by Sierra Nevada bighorn sheep

populations, CA. The lowest elevation available to bighorn sheep in Mono Basin is

approximately 2,500 m.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1500 2000 2500 3000 3500 4000

Pro

bab

ility

of

Sele

ctio

n

Elevation

Mono Basin

Baxter

Wheeler

Langley

Page 242: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

228

Figure 5.5. Probability of selection for A) predation risk, B) slope, and C) terrain ruggedness for populations of Sierra Nevada bighorn

sheep wintering in close proximity to deer herds. The Baxter population has the greatest amount of spatial overlap with deer, followed

by Wheeler, and Langley, respectively. At Langley, risk of predation is only modeled for values <5 as this encompassed the range of

possible values.

A B C

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 10 20 30 40

Pro

bab

ility

of

Sele

ctio

n

Risk of Predation

BaxterWheelerLangley

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

10 30 50 70

Slope

Baxter

Wheeler

Langley0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 0.05 0.1 0.15 0.2 0.25

Terrain Ruggedness

BaxterWheelerLangley

Page 243: Escaping the extinction vortex: identifying factors ... · Johnson, Heather E., Ph.D., Spring 2010 Wildlife and Fisheries Biology Escaping the extinction vortex: identifying factors

229

Figure 5.6. Probability of selection for areas of low, medium, and high risk of lion

predation by Sierra Nevada bighorn sheep in the Baxter population over the course of the

winter (Dec-Apr).

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1-Dec 31-Dec 30-Jan 1-Mar 31-Mar 30-Apr

Pro

bab

ility

of

Sele

ctio

n

Time

High Risk

Medium Risk

Low Risk