<<

Clemson University TigerPrints

All Dissertations Dissertations

May 2020

Individual, Population, and Community-Level Drivers of ( jubatus) Population Dynamics

Laura Gigliotti Clemson University, [email protected]

Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations

Recommended Citation Gigliotti, Laura, "Individual, Population, and Community-Level Drivers of Cheetah (Acinonyx jubatus) Population Dynamics" (2020). All Dissertations. 2597. https://tigerprints.clemson.edu/all_dissertations/2597

This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by an authorized administrator of TigerPrints. For more information, please contact [email protected].

INDIVIDUAL, POPULATION, AND COMMUNITY-LEVEL DRIVERS OF CHEETAH (ACINONYX JUBATUS) POPULATION DYNAMICS

A Dissertation Presented to the Graduate School of Clemson University

In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Wildlife and Fisheries Biology

by Laura Christine Gigliotti May 2020

Accepted by: Dr. David Jachowski, Committee Chair Dr. Robert Baldwin Dr. Michael Childress Dr. Beth Ross

ABSTRACT

Ecological processes can operate at different scales; individual characteristics can scale-up and affect individual performance, which in turn can influence population and community-level processes. Similarly, processes at the community or population level can affect individual characteristics. For mesopredators, the majority of prior research has focused on how top-down regulation by apex predators affects population dynamics. In contrast, less is known about how mesopredators can be affected by processes happening at the individual, population, and community-level simultaneously. I studied a population of (Acinonyx jubatus) in the Mun-Ya-Wana Conservancy, South Africa to better understand how ecological processes across multiple scales can affect mesopredators. In

Chapter 1, I investigated population-level drivers of survival, reproduction, and recruitment of cheetahs using a 25-year dataset. I found that demographic drivers were complex and context dependent. Specifically, cheetah monthly survival was best described by density and prey density, but opposite of predicted relationships; both adults and cubs had the highest survival when lion densities were highest and prey densities were lowest. I found that there were no strong drivers of litter size, but that cheetahs had the highest recruitment during times of low cheetah density and low prey density. Next, in Chapter 2 I considered how individual habitat use of cheetahs can scale- up and influence population survival. I assessed habitat use at short-term and long-term scales in relation to lion density, prey density, and habitat complexity and used these spatial covariates to predict survival. I found that over both the short-term and the long- term, cheetah survival was highest in areas with open vegetation, and that over the long-

ii term cheetah survival was lowest in areas of high lion density. In Chapter 3, I examined how spatial and temporal variation in predation risk, as well as habitat complexity, can influence cheetah anti-predator behaviors. Using a playback experiment, I manipulated short-term risk in areas of varying long-term risk and assessed cheetah behavioral responses. I found that cheetah vigilance was not associated with long-term predation risk, but that cheetahs responded to short-term risk by being vigilant or fleeing.

Additionally, habitat complexity affected cheetah anti-predator behaviors, with cheetah more vigilant in open areas and more likely to flee from lion sounds in closed vegetation and from sounds in open vegetation. Finally, in Chapter 4 I investigated how habitat disturbance can affect carnivore coexistence and suppression. I used prescribed burning to experimentally increase prey densities and monitored how individual , as well as large carnivores and small carnivores as a whole, respond to burning. I found that some large and small carnivores increased use of burned areas post-fire, but that most carnivores were unaffected by burning. Small carnivores may have experienced a suppression of opportunity, where they were not able to benefit from increased prey in burned areas because of high lion use in these areas. Collectively, my research highlights the need to consider multiple scales of ecological processes to understand mesopredator population dynamics. Specifically, I show that top-down effects on mesopredators are context-dependent and depend on the scale of investigation, so understanding how multiple factors simultaneously affect mesopredator populations is critical.

iii

ACKNOWLEDGMENTS

As with any research project, this dissertation was made possible by the support of a large number of people. First, I’d like to thank my PhD advisor, David Jachowski, who gave me the opportunity to work on this project. He provided me with the independence and opportunities that I needed to grow as a scientist, while always being available to offer advice and support. Most importantly, he believed in my project even during the times when I didn’t think that things would work out. I am honored to be the first PhD student to graduate from the Jachowski Lab.

Next, I’d like to thank my committee members who helped me develop and refine my research: Rob Baldwin, Michael Childress, and Beth Ross. I appreciate the time that they spent discussing my project with me, as well as helping me figure out various methods and analyses. Thanks especially to Beth for her help with the analysis for my fourth chapter.

I have several collaborators who helped me at different points in this research.

Rob Slotow was instrumental in initiating this project and gave me helpful feedback on my ideas and results throughout my entire PhD. Luke Hunter generously allowed me to use some of the data the he collected for his own PhD research and provided useful comments on my first two chapters. I am privileged to have been able to continue the work that he started on Phinda’s cheetah population. Julien Fattebert provided me with some of the prey data that I used for my first and second chapters and was able to give insight into my study system before I was able to visit there myself. I am also appreciative of his support throughout the entire analysis and writing process. Finally, my

iv fourth chapter would not have been possible without the help and support of Gonçalo

Curveira-Santos. From teaching me everything about camera trapping, to being willing to share cameras and discuss ideas, to being a constant source of encouragement during the many times that something went wrong with this chapter, I am grateful that we had the opportunity to work together and I look forward to many more collaborations in the future.

This project would not have been successful without the support of many individuals in South Africa. Collectively, I’d like to thank the Mun-Ya-Wana

Conservancy and Phinda Private Game Reserve for allowing me to conduct this research, and the numerous employees and volunteers who were involved with collecting data over the years, which comprised the long-term datasets that I used for my research. Simon

Naylor entrusted me with these incredible datasets and I also appreciate his enthusiasm for incorporating research into the day-to-day management of the reserve. I’d especially like to thank Craig Sholto-Douglas, without whom I probably wouldn’t have a dissertation. Craig made sure that all the logistical components of my fieldwork ran smoothly, was always available to answer my frequent questions about data, and was always willing to “make a plan” for any situation. Additionally, thanks to Charli De Vos,

Richard Steyn, Winston Pretorius, and Rigardt Hoffman who did the bulk of data collection for my third chapter and kept my camera traps running when I wasn’t in the field. I appreciate their willingness to teach me about South Africa and Phinda, and for their friendship during my times in the field. J.R. Janse van Rensburg collected some of the prey data that I used in my analyses and made sure that several of my camera traps

v didn’t burn during my second year of fieldwork. Thanks also to Adam Bowes who helped me set up all my camera traps during my second field season and the many volunteers who tagged along for camera set-up or checks and cheetah playbacks. Overall, I’m grateful that I got the chance to work in such an incredible place, especially one that places a high value on research and long-term data collection.

Thanks to Zoey Chapman, Kayla Goodman, Juliana Humphreys, Abby Johanson,

Logan Miller, Kara Rhodes, Sarah Stewart, Martha Stowasser, and Sara Westwood for their help in processing my camera trap photos. I also thank the Creative Inquiry program at Clemson University for the opportunity to mentor this group of dedicated students.

My time at Clemson was enhanced by getting to know the graduate students and faculty within my department. I’d especially like to thank the past and present members of the Jachowski Lab. I loved being a part of such a supportive lab and I appreciate all the feedback and friendship that they provided me.

My second and fourth chapters were supported by a Kaplan Graduate Award from

Panthera. Additionally, I received funding in the form of an R.C. Edwards Graduate

Fellowship from Clemson University, a Columbus Townsend Graduate Fellowship from the College of Agriculture, Forestry, and Life Sciences at Clemson University, and an

Alexander Family Fellowship from the Department of Forestry and Environmental

Conservation at Clemson University. I am grateful for the support, which allowed me to pursue this degree.

Finally, I’d like to thank my friends and family. In particular, my parents instilled in me a love of learning and have always been my biggest supporters.

vi

TABLE OF CONTENTS

Page

TITLE PAGE ...... i

ABSTRACT ...... ii

ACKNOWLEDGMENTS ...... iv

LIST OF TABLES ...... ix

LIST OF FIGURES ...... x

CHAPTER

1. CONTEXT-DEPENDENCY OF TOP-DOWN, BOTTOM-UP, AND DENSITY-DEPENDENT INFLUENCES ON CHEETAH DEMOGRAPHY ...... 1

Abstract ...... 1 Introduction ...... 2 Methods...... 5 Results ...... 13 Discussion ...... 15 Conclusions ...... 21 References ...... 22

2. HABITAT COMPLEXITY AND LIFETIME PREDATION RISK INFLUENCE MESOPREDATOR SURVIVAL IN A MULTI- PREDATOR SYSTEM ...... 40

Abstract ...... 40 Introduction ...... 41 Methods...... 44 Results ...... 49 Discussion ...... 51 References ...... 55

vii

Table of Contents (Continued)

Page

3. SHORT-TERM PREDATION RISK AND HABITAT COMPLEXITY INFLUENCE CHEETAH ANTI-PREDATOR BEHAVIORS ...... 71

Abstract ...... 71 Introduction ...... 72 Methods...... 75 Results ...... 81 Discussion ...... 83 References ...... 89

4. EFFECTS OF PRESCRIBED BURNING ON WILDLIFE COMMUNITY DYNAMICS ...... 102

Abstract ...... 102 Introduction ...... 103 Methods...... 106 Results ...... 111 Discussion ...... 113 References ...... 119

APPENDICES ...... 132

A1: Data collection and analysis timelines ...... 133 A2: Estimating prey density in the study area ...... 134 A3: Effects of time and sex on cheetah resighting and survival rates ...... 144 A4: Supplementary results of cheetah survival analysis...... 148 B1: Stability of cheetah home ranges ...... 149 B2: Effects of spatial covariates at multiple scales on cheetah survival ...... 152 B3: Effects of EVI on cheetah kill sites and locations where cheetahs were killed ...... 155 C1: Determining the best model structure to analyze multi-species responses to prescribed burning ...... 157 C2: Camera trap data summary...... 162

viii

LIST OF TABLES

Table Page

1.1 Model selection results for cheetah litter size regression models, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018...... 34

1.2 Model selection results for cheetah recruitment regression models, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018...... 35

1.3 Model selection results for structural multi-state joint live-encounter dead-recovery survival models for cheetahs to incorporate variation by time and sex for resighting (p) and survival (S) rates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018...... 36

1.4 Model selection results for multi-state joint live-encounter dead-recovery survival models for cheetahs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018 ...... 37

2.1 Model selection results for multi-state joint live-encounter dead-recovery spatially-explicit survival models for cheetahs with seasonal spatial covariates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018...... 66

2.2 Model selection results for multi-state joint live-encounter dead-recovery spatial-explicit survival models for cheetahs with spatial covariates averaged across individual cheetahs’ lifetimes, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018...... 67

4.1 Group-level covariate estimates describing changes in intensity of use related to prescribed burning, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019...... 128

ix

LIST OF FIGURES

Figure Page

1.1 Probability of cub recruitment in relation to cheetah density and prey density, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018 ...... 38

1.2 Monthly survival for a) adult and b) cub cheetahs in relation to prey density and lion density during dry seasons of years with average resighting rates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018 ...... 39

2.1 Monthly survival of adult and cheetah cubs in relation to short-term (seasonal) average Enhanced Vegetation Index (EVI) within an individual’s home range, Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI...... 68

2.2 Monthly survival of adult and cheetah cubs in relation to a) long- term (lifetime) average Enhanced Vegetation Index (EVI) within an individual’s home range while holding lion density constant at an average value, and b) long-term (lifetime) average probability of encountering a lion within an individual’s home range while holding EVI constant at an average value, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI ...... 69

2.3 Relative probability of cheetah kill site occurrence in relation to Enhanced Vegetation Index (EVI), Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI...... 70

3.1 Relationship between average Enhanced Vegetation Index (EVI) within 50 meters of playback location and baseline vigilance for female cheetahs with cubs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa. Shaded regions represent 85% CI ...... 98

x

List of figures (continued)

Figure Page

3.2 Relationship between playback type and cheetah post-playback a) vigilance or b) probability of fleeing, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018-2019. Error bars represent SE ...... 99

3.3 Differences in post-predator (lion and leopard) playback vigilance behavior between female cheetahs with cubs and males, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa. Error bars represent SE ...... 100

3.4 Probability of cheetahs fleeing following lion and leopard playbacks based on average Enhanced Vegetation Index (EVI) within 50 meters of playback location, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa. Shaded regions represent 85% CI ...... 101

4.1 Locations of prescribed burned and camera trap sites, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019...... 129

4.2 Species-specific covariates estimates (mean ± 95% Credible Interval) for a) time period, b) treatment, and c) the interaction between time period and treatment on intensity of use, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019 ...... 130

4.3 Estimated intensity of use (mean ± 95% CI) for a) buffalo, b) bushpig, c) giraffe, d) impala, e) nyala, f) warthog, g) wildebeest, h) zebra, i) lion, j) honey , and k) side-striped jackal in relation to prescribed burning, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018-2019...... 131

xi

CHAPTER 1

CONTEXT-DEPENDENCY OF TOP-DOWN, BOTTOM-UP, AND DENSITY- DEPENDENT INFLUENCES ON CHEETAH DEMOGRAPHY

ABSTRACT

1. Research on drivers of demographic rates has mostly focused on top predators and

their prey, and comparatively less research has considered the drivers of

mesopredator demography. Of those limited studies, most focused on top-down

effects of apex predators on mesopredator population dynamics, whereas studies

investigating alternative mechanisms are less common.

2. In this study, we tested hypotheses related to top-down, bottom-up, and density-

dependent regulation of demographic rates in an imperiled mesopredator, the

cheetah (Acinonyx jubatus).

3. We used a 25-year dataset of lion density, cheetah density, and prey density from

the Mun-Ya-Wana Conservancy in South Africa and assessed the effects of top-

down, bottom-up, and density-dependent drivers on cheetah survival and

reproduction.

4. In contrast to the top-down and bottom-up predications, both adult and juvenile

cheetahs experienced the lowest survival during months with high prey densities

and low lion densities. We only observed support for a density-dependent

response in juvenile cheetahs, where they had a higher probability of reaching

independence during times with low cheetah density, and low prey density. We

did not identify any strong drivers of litter size.

1

5. Collectively, our results indicate that high apex predator abundance might not

always have negative effects on mesopredator populations, and suggest that

context dependency in top-down, bottom-up, and density-dependent factors may

regulate demographic rates of cheetahs and other mesopredators. Our results

highlight the complexities of population-level drivers of cheetah demographic

rates and the importance of considering multiple hypotheses of mesopredator

population regulation.

INTRODUCTION

Understanding drivers of population dynamics is a central theme of ecology, and can help inform the conservation and management of imperiled species. Historically, research has attempted to identify top-down, bottom-up, and density-dependent influences on survival, reproduction, and overall population growth in a variety of systems (Hairston, Smith, & Slobodkin, 1960; Hanski, 1990; McNaughton, Oesterheld,

Frank, & Williams, 1989). However, most previous research on population dynamics has focused on classic predator-prey systems, in which linkages exist between predators and their prey, and between prey and primary producers that they consume (Owen-Smith,

Mason, & Ogutu, 2005; Sinclair & Krebs, 2002). In systems with higher-ranking apex predators and lower-ranking, subordinate predators (hereafter, ‘mesopredators’), the population dynamics of all trophic levels might differ from those in which there is only a single predator. Mesopredators often compete with apex predators for resources, and apex predators and mesopredators sometimes exhibit intraguild predation, in which the

2 apex predator kills the mid-ranking predator (Palomares & Caro, 1999). As a result, the suite of factors influencing mesopredator population dynamics might be more complex than those in simple predator-prey systems.

Mesopredator population dynamics are frequently attributed to the processes of mesopredator suppression or mesopredator release, in which the density or abundance of an apex predator affects the population size, distribution, or behavior of the mesopredator

(Prugh et al., 2009; Ritchie & Johnson, 2009). However, the effects of apex predators on mesopredator populations remain equivocal in the literature (Crimmins et al., 2016; Gehrt

& Prange, 2007). In addition, studies on mesopredator suppression or release often do not consider other mechanisms of population regulation for mesopredators, although research suggests that factors such as environmental productivity or prey availability can modulate mesopredator suppression (Elmhagen & Rushton, 2007; Greenville, Wardle, Tamayo, &

Dickman, 2014; Pasanen-Mortensen et al., 2017). As a result, investigations into mesopredator demography should, ideally, compare the strength and interaction of top- down, bottom-up, and density-dependent drivers simultaneously.

The cheetah (Acinonyx jubatus) is an ideal species with which to investigate ecological drivers of mesopredator population regulation. Although cheetahs are large in body size, they are mid-ranking predators in the African carnivore community (Swanson et al., 2014; Vanak et al., 2013), and thus might be influenced by both top-down and bottom-up factors. Cheetahs are subordinate to ( leo) that predate on them

(Laurenson, 1994; Mills & Mills, 2014). However, while predation by lions has been found to be the main cause of natural mortality in cheetahs in the Serengeti (Laurenson,

3

1994), a recent analysis of cheetah population sizes in relation to lion population sizes suggests that predation by, and competition with, lions might not scale-up to population- level effects on cheetahs (Swanson et al., 2014). Therefore, other factors such as prey or conspecific densities could be at play in affecting specific demographic processes in cheetahs (Durant, Kelly, & Caro, 2004).

Additionally, cheetahs are a species of conservation concern (Durant et al., 2017;

Weise et al., 2017) and future management activities could influence the sustained persistence of this species. Therefore, understanding the top-down, bottom-up, and density-dependent factors associated with cheetah population regulation is critical.

Previous work has examined demographic trends in cheetah populations in large unfenced protected areas (Durant et al., 2004; Kelly et al., 1998), and in agricultural- dominated areas (Marker, Dickman, Jeo, Mills, & Macdonald, 2003), but little research exists on the drivers of cheetah demography in fenced reserves, beyond purely descriptive statistics (Bissett & Bernard, 2011). Understanding drivers of cheetah demography in fenced reserves is particularly important because they represent an important component for the persistence of cheetahs (Buk, van der Merwe, Marnewick,

& Funston, 2018; Durant et al., 2017). These fenced populations are also subjected to management-induced changes in apex predator, prey, and conspecific densities (Balme,

Slotow, & Hunter, 2009; Kettles & Slotow, 2009), which provides an ideal situation to test multiple hypotheses related to drivers of mesopredator demography.

We used a 25-year dataset of cheetah reproduction and survival to investigate support for three competing hypotheses of mesopredator population regulation: 1)

4 cheetah reproduction and survival would be driven by lion densities (top-down regulation; Hairston et al., 1960), 2) cheetah reproduction and survival would be driven by prey densities (bottom-up regulation; McNaughton et al., 1989), and 3) cheetah reproduction and survival would be driven by cheetah densities (density-dependent regulation; Hanski, 1990). Under the top-down hypothesis we predicted that cheetah litter sizes, survival, and recruitment would be negatively related to lion density. Under the bottom-up hypothesis we predicted that cheetah litter sizes, survival, and recruitment would be positively related to prey density. Under the density-dependent hypothesis we predicted that cheetah litter sizes, survival, and recruitment would be negatively related to cheetah density. In addition, we predicted that drivers of demographic rates might be context-dependent, in that top-down, bottom-up, and density-dependent factors would interact. By understanding drivers of cheetah demography, we can better identify factors that might promote high rates of cheetah population growth, and develop a greater understanding of the complex factors that might regulate mesopredator populations.

METHODS

Study Area

We studied cheetah demographics in Mun-ya-wana Conservancy (Phinda Private

Game Reserve), in northern KwaZulu-Natal, South Africa from 1992 – 2018. The elevation of Mun-ya-wana Conservancy ranges from 4 to 201 meters above sea level, and the dominant vegetation type is broad-leaf woodland, with open and semi- open wooded-grasslands interspersed throughout the reserve. The climate is subtropical

5 with warm, dry winters (April – September) and hot, humid summers (October – March).

The average annual rainfall is 550 mm, with the majority of rain falling in the summer

(Janse van Rensburg, McMillan, Giżejewska, & Fattebert, 2018). The Mun-ya-wana

Conservancy is surrounded by electrified game fencing, and has grown in size as adjacent reserves have joined the Conservancy. From 1990 – 2004 the study area was 170 km2 in area, after which time a fence was removed and the Conservancy expanded to 235 km2

(Druce, Pretorius, & Slotow, 2008). Cheetahs and lions were reintroduced into the reserve in 1992 and have been monitored since (Hunter, 1998; Hunter et al., 2007; this study). The average cheetah density (0.10 cheetahs/km2) and lion density (0.11 lions/km2) in our study area was similar to that of other fenced reserves (0.001 - 0.29 cheetahs/km2,

Buk et al., 2018; 0.02 - 0.17 lions/km2, Miller & Funston, 2014) , but higher than the densities of cheetahs and lions in unfenced protected areas such as Serengeti National

Park (0.06 lions/km2, Bauer & Van Der Merwe, 2004; 0.005 cheetah/km2, Durant et al.,

2017).

Carnivore monitoring

To ensure even monitoring, we subdivided the reserve into seven sections, and trained monitors typically drove the roads in each section at least once a week. In addition, monitors frequently followed-up on sightings reported by game rangers conducting game drives within the reserve. Cheetahs and lions can be individually recognized using their spot patterns, whisker spots, and scars, which allowed us to monitor the populations based on sightings alone (Caro, 1994). We only included data

6 from sightings where the identity of the was known with complete certainty. We obtained an average of 32 ± 2 cheetah sightings per month and an average of 18 ± 1 lion pride sightings per month. We used data from 1992 – 2018 for general descriptive statistics (litter sizes, causes of death), but restricted our analyses of drivers of litter size, recruitment, and survival to 2008 – 2018 (Figure A1.1).

When cheetahs or lions were observed, we recorded the location, behavior, and number of individuals present. We divided the total monthly cheetah or lion population size, including cubs, by the total area of the reserve to obtain monthly densities.

However, we occasionally were not sure of the status of a cheetah or lion if we were unable to sight it, or determine if it was dead. Therefore, we removed individuals from the monthly population count if they had not been seen in 6 months, given that the probability of survival is <0.005 if an animal is not seen for that long (this study). Cubs are rarely included in density estimates for large ; however, we included them in our analyses because they comprised a large portion of the total felid biomass present and because older cubs are functional similar to adults in their food requirements. Our study area, along with most small fenced reserves, actively manages their lion populations through removals, introductions, and female contraception (Ferreira & Hofmeyr, 2014; S.

M. Miller et al., 2013). Therefore, fluctuations in the lion density within our study area were primarily the result of management actions, which allowed to us focus specifically on drivers of cheetah demographics, without simultaneously assessing drivers of changes in lion density.

7

Prey density

We estimated prey density in the reserve by collecting distance sampling data on impala (Aepyceros melampus) and nyala (Tragelaphus angasii) during the dry season

(April – September), and the wet season (October – March) from 2010 – 2015 (Appendix

A2). We limited our prey analyses to these species because they comprised 82% of cheetah kills in the study area (Hunter, 1998). We estimated prey abundance using hierarchical distance sampling models with spatial covariates on both the abundance and detection processes (Royle, Dawson, & Bates, 2004; Sillett et al., 2012), and used our top model to extrapolate prey abundance over our entire study period (Appendix A2). We divided the total seasonal prey abundances by the total area of the reserve to obtain seasonal densities.

Cheetah reproductive output

We studied top-down, bottom-up, and density-dependent effects on cheetah litter sizes. Reproduction in carnivores can be influenced by top-down, bottom-up, and density-dependent drivers by the mechanisms of food limitation, female body condition, or stress (Creel, Creel, Mills, & Monfort, 1997; Fuller & Sievert, 2001; Tannerfeldt &

Angerbjorn, 1998). When a female cheetah was first seen with a new litter, we estimated the age of the cubs based on their size, following Caro (1994). Because cheetahs are hard to locate when they are denning, our litter sizes were typically counts of cubs after they had emerged from dens. Previous research suggests that some cub mortality occurs while still in the den (Laurenson, 1994; Mills & Mills, 2014) so we recognize that our counts

8 might be biased low. However, post-emergence counts of litters have been used in similar studies of cheetah reproductive output (Bissett & Bernard, 2011; Kelly et al., 1998).

We analyzed drivers of cheetah litter size using generalized linear models with a generalized Poisson error distribution (Kendall & Wittmann, 2010) in Program R

(Version 3.5.3; R Core Team, 2019). Litter sizes can be influenced by environmental conditions pre-conception and during gestation (Lack, 1948). Cheetahs can give birth at any time during the year and cheetah gestation lasts approximately 3 months (Kelly et al.,

1998); therefore, we calculated the average of our covariates of interest in the 6 months prior to a litter being born to incorporate effects during the pre-gestation and gestation time periods. We specified nine a priori models based on our hypotheses of interest related to litter size and considered covariates of monthly lion density, monthly prey density, monthly cheetah density, as well as additive models, and models with interactions between pairs of covariates. Prior to this analysis, and all subsequent analyses, we assessed collinearity between continuous covariates using a Pearson correlation to determine if any should be excluded from analysis (|r|> 0.75). For all analyses, we compared models using Akaike’s Information Criterion corrected for sample size (AICc; Burnham and Anderson, 2002), considered models within 2 ΔAICc of the top model to be competitive, and evaluated if covariates were informative by calculating 85% confidence intervals (Arnold, 2010).

9

Cheetah Recruitment

We analyzed the probability of cubs reaching adulthood. Independence in cheetahs is typically marked by an abrupt and clear separation from the mother (Hunter,

1998), so we calculated age at independence to the nearest month after it was unequivocal that separation from the mother had occurred. Based on this criterion, we also calculated the percentage of litters with at least 1 cub reaching independence, and the percentage of total cubs reaching independence. For each cub, we treated recruitment as a binary variable and used logistic regression to evaluate the relationship between recruitment and top-down, bottom-up, and density-dependent drivers. We averaged lion density, prey density, and cheetah density across the entire time-period when a cheetah was a cub. Because cubs from the same litter might represent non-independent samples

(Pettorelli & Durant, 2007), we included litter as a random effect. We specified nine a priori models and considered covariates of lion density, prey density, cheetah density, as well as additive models, and models with interactions between pairs of covariates.

Cheetah Survival

We analyzed drivers of cheetah monthly survival from February 2009 to March

2018. Because the cheetah population was intensively monitored and the reserve was surrounded by electric game fencing, we were able to determine the cause and time of death for most cheetahs. When a dead cheetah was recovered, we attempted to determine the cause of death by examining the carcass and the surrounding area for tracks and scat.

For each month, we recorded if individual cheetahs were sighted or recovered dead as

10 adults or cubs. If a cheetah was removed from the reserve for management purposes, we censored that individual from analyses. Because lion and cheetah density can vary greatly within a season, and because cubs can be born and become independent at any time during the year, we conducted our analysis on a monthly timescale to best reflect the conditions that might be driving survival. We calculated the average prey density for each season, and the average lion density and cheetah density for every month, and used these values as covariates. In addition, we calculated the average prey density in the 6 months prior to every month to investigate a potential resource time-lag effect.

Model Structure

We analyzed cheetah survival using multi-state joint live-encounter dead-recovery models (Barker, White, & McDougall, 2005) using the rmark R package (Laake, 2013).

This model made use of our frequent resightings and mortality data, and also allowed for survival estimation based on individuals that were never recorded as dead. In addition, because juvenile cheetahs stay with their mothers for variable amounts of time (Kelly et al., 1998), we could not incorporate a standard age structure into our models. Thus, we used a multi-state approach to estimate survival for both cubs and adults simultaneously.

We specified the two model states as cub (juvenile cheetahs dependent on their mother) and adult (cheetahs that were independent from their mother). The study area is surrounded by electric game fencing that cheetahs very rarely penetrated. Accordingly, we did not incorporate immigration or emigration into our models.

11

Hypothesis Testing

We performed model selection in a multi-step approach to determine the appropriate model structure, before testing for covariate relationships (Cubaynes et al.,

2014; Doherty, White, & Burnham, 2012). We first tested for seasonal or yearly effects on resighting rates (p) while holding survival rates (S) and reporting rates I constant, and ranked models using AICc. Using our top resighting model, we next tested for effects of season and sex on survival rates, while holding reporting rate constant. It is hard to determine the sex of young cheetah cubs, so we only considered the effects of sex for adult cheetahs. Because of convergence issues, we were unable to test for yearly effects on survival. However, given pronounced differences in climatic conditions between wet and dry seasons at our study area, we felt that seasonal variation would be the more important temporal driver of survival. To aid in model convergence for our covariate models, we simplified our best structural model by grouping time periods or sexes that did not differ.

Finally, using our top simplified structural model, we tested for the effects of environmental covariates on cheetah survival. We developed 12 a priori models based on our hypotheses of interest that included covariates of monthly lion density, monthly prey density, monthly cheetah density, average prey density in the six months prior, as well as additive and multiplicative models with the same covariates. Because adults and cubs are known to have different survival rates (Kelly et al., 1998), we did not consider any models in which state was not included.

12

RESULTS

Reproductive parameters

We collected data on 61 cheetah litters from 1992 – 2018. The average litter size was 3.26 ± 0.17 and the average age at first reproduction for females was 28.1 ± 1.8 months. Litters became independent at an average of 16.7 ± 1.0 months, and mothers had an inter-birth interval of 19.4 ± 2.4 months. Based on litters from 2008 – 2018, litter size variation was best described by the null model, but also by models including lion density and prey density (Table 1.1). However, the confidence intervals of the parameter estimates for lion density (β = 3.71; 85% CI = -1.75 to 7.71) and prey density (β = -0.01;

85% CI = -0.03 to 0.01) overlapped 0, indicating that they were uninformative.

Cheetah Recruitment

We included 119 cubs from 40 litters from 2008 – 2018 in our analysis of recruitment. At least one cub reached independence in 56.7% of litters, and 41.5% of all cubs reached independence. Probability of recruitment was best described by a model with an interaction between cheetah density and prey density (Table 1.2). Based on this model, cheetahs had the highest probability of reaching independence if they were cubs during times of low prey density and low cheetah density (Figure 1.1).

Survival

Of the 239 cheetahs monitored over the course of the study (1992 – 2018), 43.9%

(n = 105) died from unknown causes, 21.8% (n = 52) had a known cause of death, and

13

9.2% (n = 22) were still alive at the completion of this study. In addition, 52 cheetahs were translocated to other reserves and 8 cheetahs were known to have escaped the reserve in the months immediately following reintroduction.

Predation accounted for 84.6% (n = 44) of known deaths, accounting for 93.8%

(n=30) of cub mortality and 70% (n = 14) of adult mortality. Of the predation deaths, lions accounted for 63.3% (n = 19) of cub predation and 35.7% (n = 5) of adult predation, accounted for 3.3% (n = 1) of cub predation and 21.4% (n = 3) of adult predation, accounted for 7.1% (n = 1) of adult predation, and unknown predators accounted for 20.0% (n = 6) of cub predation and 7.1% (n = 1) of adult predation. Adult cheetahs killed other cheetahs in 28.6% (n = 4) of adult predation deaths and 13.3% (n =

4) of cub predation deaths. Other sources of mortality included vehicle collisions (n = 2), cub abandonment (n = 1), injury (n = 2), and snaring (n = 3).

We included 138 cheetahs in our survival analysis from for a total of 110 months.

Our best structural model included effects of year and season on resighting rates and effects of season and sex on survival (Table 1.3). Resighting rates were similar in most years (Figure A3.3) with the exception of a lower average resighting rate in 2014 (0.48;

85% CI = 0.42 – 0.53), and a higher average resighting rate in 2018 (0.95; 85% CI = 0.87

– 1.00). However, resighting rates were similar between wet seasons (0.60; 85% CI =

0.58 – 0.62) and dry seasons (0.63; 85% CI = 0.61 – 0.66; Figure A3.4). Based on the top structural model, survival did not differ between males (0.98; 85% CI = 0.969 – 0.987) and females (0.96; 85% CI = 0.953 – 0.974; Figure A3.5). Across all cheetahs, survival was higher in wet seasons (0.96; 85% CI = 0.954 – 0.972) compared to dry seasons (0.94;

14

95% CI = 0.926 – 0.948; Figure A1.6). Based on these results, our final structural model for evaluating environmental covariates included effects of year on resighting rates, and season on survival.

Survival was best described by a model with an interaction between monthly lion density and monthly prey density (Table 1.4). In contrast to predictions of top-down and bottom-up regulation, adults and cubs both experienced the lowest survival rates in months with high prey density and low lion density (Figure 1.2). At the lowest lion density (0.06 lions/km2), the odds of cub survival decreased by 27.6% for every 1-unit increase in prey density. At the highest prey density (46 prey/km2), the odds of cub survival increased 35.4% for every 1-unit increase in lion density (Figure A4.1).

Similarly, at the lowest lion density, the odds of adult survival decreased by 21.4% for every 1-unit increase in prey density and at the highest prey density, the odds of adult survival increased by 44.1% for every 1-unit increase in lion density (Figure A4.1).

Based on the top model, adult cheetahs had an average monthly survival rate of 0.97

(85% CI = 0.95 – 0.98), whereas cheetah cubs had an average monthly survival rate of

0.91 (85% CI = 0.88 – 0.94).

DISCUSSION

We found evidence suggesting that cheetah demographic rates, particularly survival and recruitment, varied in their sensitivity to top-down, bottom-up, and density- dependent factors, although not always in the manner classically predicted in predator- prey systems. In particular, the recruitment, reproduction, and survival rates of cheetahs

15 did not appear to be negatively affected by high lion densities, which does not follow the predictions of mesopredator suppression (Prugh et al., 2009; Ritchie & Johnson, 2009). In contrast, survival was typically higher for both cubs and adults during periods of high lion density, but the strength of that effect depended on the density of prey. Further, a density-dependent effect on recruitment was observed, but only during periods of low prey availability. Collectively, our results highlight the context-dependency in population-level top-down, bottom-up and density-dependent drivers of cheetah demographic rates, and the importance of simultaneously considering multiple mechanistic hypotheses of mesopredator regulation (Elmhagen & Rushton, 2007;

Pasanen-Mortensen et al., 2017).

Reproduction

Our findings suggest that cheetah litter size was not as sensitive to top-down, bottom-up or density-dependent factors as other demographic parameters. There are likely several reasons for why we did not identify any strong drivers of cheetah litter size.

First, we might not have seen density-dependent changes in litter size because density- dependent reproduction often occurs as a result of resource depletion or poor body condition of females (Fuller & Sievert, 2001). Although prey densities in our study area fluctuated, due to intensive management, there were no prolonged periods of low prey densities. As a result, even at times of high cheetah density there likely was not high competition for food resources. Second, similar to other carnivores, cheetahs might share a common optimal litter size which maximizes fitness, but does not vary according to

16 environmental conditions (Gaillard, Nilsen, Odden, Andrén, & Linnell, 2014). Thus, the lack of long-term low prey densities might explain why we also did not see bottom-up influences on cheetah litter sizes. Finally, the management of lions in our study area could explain why we did not observe top-down drivers of cheetah litter sizes. Predators and predation risk have been found to reduce reproductive rates and litter sizes in some species (Karels, Byrom, Boonstra, & Krebs, 2000; Korpimaki, Norrdahl, & Valkama,

1994). However, in our study area the lion population was intensively managed within the reserve, and it might be managed at densities that are too low to affect cheetah reproductive rates.

Recruitment

In contrast to litter size, we did see factors influencing the probability of cubs becoming independent and recruiting into the population. During times with the highest cheetah densities, cubs had a very low probability of reaching independence, regardless of the prey density (Figure 1.1). This relationship suggests that resource limitations likely were not driving the observed density-dependent recruitment at the highest cheetah densities. Instead, space limitation might be affecting patterns that we observed. Density- dependence resulting from space limitations rather than prey availability has been observed in other carnivore species such as (Cubaynes et al., 2014), lions (Kissui

& Packer, 2004), and leopards (Balme et al., 2013). Reduced recruitment because of density-dependent space use might be more pronounced when prey resources are limited because there is more competition for high quality habitats. Indeed, we found that

17 cheetah recruitment was most sensitive to density-dependence when prey densities were low. By contrast, cheetahs experienced high recruitment during times of low prey density and low cheetah density. Not only is density-dependence heightened when prey is low, but at low densities prey were likely not distributed in dense aggregations. The lack of large prey aggregations could have decreased the probability of cheetahs encountering lions (Mark Hebblewhite & Pletscher, 2002), thus increasing the probability of cheetahs surviving long enough to become independent.

Survival

Although resources and apex predators affected cheetah survival under certain conditions, the overall direction of response was typically opposite of what is predicted for classic predator-prey systems. In particular, we did not observe negative effects of apex predators on mesopredators, as predicted by the mesopredator suppression hypothesis (Prugh et al., 2009; Ritchie & Johnson, 2009). Our findings contrast previous work suggesting that lions suppress cheetah populations (Chauvenet, Durant, Hilborn, &

Pettorelli, 2011; Laurenson, 1994), and rather provides support to recent work which indicates that lions might not have substantial effects on the persistence of cheetahs

(Swanson et al., 2014). Much of the research that found significant top-down effects on cheetah dynamics was conducted in large, unfenced protected areas with open habitats

(Durant et al., 2004; Kelly et al., 1998; Laurenson, 1994), which highlights the need to understand variation in drivers of cheetah demography, especially in fenced populations.

Whereas cheetahs in the Serengeti sometimes move great distances to follow migratory

18 prey (Durant, Caro, Collins, Alawi, & Fitzgibbon, 1988), and are able to immigrate and emigrate, the fenced boundaries of our study area resulted in little seasonal home range shifts, and a lack of long-distance dispersal. Contrary to leopards (Fattebert, Balme,

Dickerson, Slotow, & Hunter, 2015), the inability of cheetah to disperse out of the fenced reserve, along with high local concentrations of predators and prey, may result in the changing community-level spatial relationships and behavior that we hypothesize were the mechanisms behind our observed survival trends. Additionally, our study area contains areas of dense vegetation, which could reduce cheetah mortality from lions by acting as a predation refuge (Mills & Mills, 2014), whereas the open plains of the

Serengeti do not offer substantial cover for hiding cheetahs. Research in other systems has also found a lack of strong evidence for top-down regulation of mesopredator populations. For example, recovering ( lupus) populations in Wisconsin did not limit the abundance of (Canis latrans), potentially because of prey availability or habitat arrangements that benefited coyotes (Crimmins & Van Deelen,

2019). Similarly, research in Australia suggests that dingos (Canis lupus dingo) do not exclude feral cats ( catus) from areas, although top-down effects might be context- dependent (Allen, Allen, & Leung, 2015).

Our results add to the growing literature suggesting that top-down regulation of mesopredators might not be ubiquitous, and that bottom-up or density-dependent factors can modulate the strength of top-down effects (Elmhagen & Rushton, 2007; Pasanen-

Mortensen et al., 2017). Under classical predator-prey theory, top-down influences should be greatest when predators are abundant and bottom-up conditions are limited

19

(Leibold, 1989). By contrast, our results suggest that when lions were at their lowest densities and prey was most readily available, cheetahs exhibited their lowest survival.

While seemingly counterintuitive, we believe there are several behavioral trade-offs that potentially explain this pattern. First, cheetahs and lions have been found to use the same general areas on a landscape scale, particularly during times of high prey densities when prey form into large aggregations (Durant, 1998; Vanak et al., 2013). This shared space- use could have increased intra-guild predation rates. However, cheetahs use fine-scale temporal partitioning to avoid interactions with lions (Broekhuis, Cozzi, Valeix, Mcnutt,

& Macdonald, 2013; Rostro-García, Kamler, & Hunter, 2015; Swanson, Arnold,

Kosmala, Forester, & Packer, 2016; Vanak et al., 2013). During periods of low lion density, cheetahs might reduce their fine-scale partitioning with lions, and thus have higher encounter rates. Similarly, our observed pattern of cheetah survival could be related to changes in cheetah vigilance behavior. Mesopredators can adjust their anti- predator behaviors under varying levels of predation risk from apex predators. For example, coyotes spend more time vigilant while feeding in areas of high wolf activity

(Switalski, 2003). Thus, lower cheetah survival during periods of low lion density might be related to decreased temporal avoidance and vigilance behavior, which could have increased their risk of attack by lions. A second explanation might be that the behavior of lions influenced cheetah survival patterns. During times of high prey densities, lions might not have needed to invest as much time and energy on finding and killing prey.

Therefore, they could have afforded to spend more time on territorial behaviors and actively pursuing cheetahs, which in turn could reduce cheetah survival rates.

20

Finally, our observed patterns of survival might also reflect changes in the overall predator community. Although we were only able to include lion density in our models, leopard abundance increased and spotted hyenas (Crocuta crocuta) persisted at low densities within the study area during our study period (Balme et al., 2009). During times of high prey density, multiple top-predator species might have used areas of highly aggregated prey, and thus increased encounter rates with cheetahs. In addition, lions occasionally kill and compete with leopards and hyenas (Balme, Miller, Pitman, &

Hunter, 2017; Trinkel & Kastberger, 2005), and leopards, similar to cheetahs, have been found to use fine-scale spatial partitioning to avoid interactions with lions (du Preez,

Hart, Loveridge, & Macdonald, 2015; Vanak et al., 2013). Therefore, when lion densities were low and prey densities were high, leopard and hyenas might have used areas typically used by lions and imposed top-down effects on cheetahs, either through direct interactions or exploitative competition. Similarly, in many systems worldwide subordinate large carnivores have been found to fill the functional role of extirpated or declining apex predator populations and induced top-down effects on mesopredator populations (Gompper, 2002; Letnic, Ritchie, & Dickman, 2012; Oakwood, 2000; Ralls

& White, 1995).

CONCLUSIONS

We highlight the complexities in understanding the ecological drivers of mesopredator population dynamics and the importance of investigating support for multiple, often interacting hypotheses on multiple demographic rates. We found support

21 for hypotheses related to top-down, bottom-up, and density-dependent drivers of cheetah survival and recruitment. However, density-dependent recruitment was the only finding that corresponded with classical predictions. Conversely, we found that apex predators and prey affected cheetah survival in the opposite direction to classical linear interpretations based on top-down and bottom-up hypotheses. In addition, we found that different demographic rates differed in their sensitivity to top-down, bottom-up, and density-dependent drivers. We show that high apex predator abundance does not always have negative effects on mesopredator populations, and demonstrate that the processes of mesopredator suppression and release is not be universal across all species and systems.

REFERENCES

Allen, B. L., Allen, L. R., & Leung, L. K. P. (2015). Interactions between two naturalised

invasive predators in Australia: Are feral cats suppressed by dingoes? Biological

Invasions, 17(2), 761–776. doi:10.1007/s10530-014-0767-1

Arnold, T. W. (2010). Uninformative parameters and model selection using Akaike’s

Information Criterion. Journal of Wildlife Management, 74(6), 1175–1178.

doi:10.2193/2009-367

Balme, G. A., Batchelor, A., De Woronin Britz, N., Seymour, G., Grover, M., Hes, L., …

Hunter, L. T. B. (2013). Reproductive success of female leopards Panthera

pardus: The importance of top-down processes. Review, 43(3), 221–

237. doi:10.1111/j.1365-2907.2012.00219.x

22

Balme, G. A., Miller, J. R. B., Pitman, R. T., & Hunter, L. T. B. (2017). Caching reduces

kleptoparasitism in a solitary, large felid. Journal of Animal Ecology, 86(3), 634–

644. doi:10.1111/1365-2656.12654

Balme, G. A., Slotow, R., & Hunter, L. T. B. (2009). Impact of conservation

interventions on the dynamics and persistence of a persecuted leopard (Panthera

pardus) population. Biological Conservation, 142(11), 2681–2690.

doi:10.1016/j.biocon.2009.06.020

Barker, R. J., White, G. C., & McDougall, M. (2005). Movement of Paradise Shelduck

between molt sites: a joint multistate-dead recovery mark-recapture model.

Journal of Wildlife Management, 69(3), 1194–1201. doi:10.2193/0022-

541X(2005)069[1194:MOPSBM]2.0.CO;2

Bauer, H., & Van Der Merwe, S. (2004). Inventory of free-ranging lions Panthera leo in

Africa. Oryx, 38(1), 26–31. doi:10.1017/S0030605304000055

Bissett, C., & Bernard, R. T. F. (2011). Demography of cheetahs in fenced reserves in

South Africa: Implications for conservation. South African Journal of Wildlife

Research, 41, 181–191. doi:10.3957/056.041.0205

Broekhuis, F., Cozzi, G., Valeix, M., Mcnutt, J. W., & Macdonald, D. W. (2013). Risk

avoidance in sympatric large carnivores: Reactive or predictive? Journal of

Animal Ecology, 82(5), 1098–1105. doi:10.1111/1365-2656.12077

23

Buk, K. G., van der Merwe, V. C., Marnewick, K., & Funston, P. J. (2018). Conservation

of severely fragmented populations: lessons from the transformation of

uncoordinated reintroductions of cheetahs (Acinonyx jubatus) into a managed

metapopulation with self ‑ sustained growth. Biodiversity and Conservation, 1–

31. doi:10.1007/s10531-018-1606-y

Burnham, K. P., & Anderson, D. R. (2002). Model selection and multimodel inference: A

practical information-theoretic approach. Springer Science & Business Media.

Caro, T. M. (1994). Cheetahs of the Serengeti Plains. Chicago: The University of

Chicago Press.

Chauvenet, A. L. M., Durant, S. M., Hilborn, R., & Pettorelli, N. (2011). Unintended

consequences of conservation actions: Managing disease in complex ecosystems.

PLoS ONE, 6(12). doi:10.1371/journal.pone.0028671

Creel, S., Creel, N. M., Mills, M. G. L., & Monfort, S. L. (1997). Rank and reproduction

in cooperatively breeding African wild : Behavioral and endocrine correlates.

Behavioral Ecology, 8(3), 298–306. doi:10.1093/beheco/8.3.298

Crimmins, S. M., & Van Deelen, T. R. (2019). Limited evidence for mesocarnivore

release following wolf recovery in Wisconsin, USA. Wildlife Biology, 1, 1–7.

doi:10.2981/wlb.00511

Crimmins, S. M., Walleser, L. R., Hertel, D. R., Mckann, P. C., Rohweder, J. J., &

Thogmartin, W. E. (2016). Relating mesocarnivore relative abundance to

anthropogenic land-use with a hierarchical spatial count model. Ecography, 39(6),

524–532. doi:10.1111/ecog.01179

24

Cubaynes, S., Macnulty, D. R., Stahler, D. R., Quimby, K. A., Smith, D. W., & Coulson,

T. (2014). Density-dependent intraspecific aggression regulates survival in

northern Yellowstone wolves (Canis lupus). Journal of Animal Ecology, 83(6),

1344–1356. doi:10.1111/1365-2656.12238

Doherty, P. F., White, G. C., & Burnham, K. P. (2012). Comparison of model building

and selection strategies. Journal of Ornithology, 152(SUPPL. 2), 317–323.

doi:10.1007/s10336-010-0598-5

Druce, H. C., Pretorius, K., & Slotow, R. (2008). The response of an elephant population

to conservation area expansion: Phinda Private Game Reserve, South Africa.

Biological Conservation, 141(12), 3127–3138. doi:10.1016/j.biocon.2008.09.024 du Preez, B., Hart, T., Loveridge, A. J., & Macdonald, D. W. (2015). Impact of risk on

animal behaviour and habitat transition probabilities. Animal Behaviour, 100, 22–

37. doi:10.1016/j.anbehav.2014.10.025

Durant, S. M. (1998). Competition refuges and coexsistance: An example from Serengeti

carnivores. Journal of Animal Ecology, 67, 370–386. doi:10.1046/j.1365-

2656.1998.00202.x

Durant, S. M., Caro, T. M., Collins, D. A., Alawi, R. M., & Fitzgibbon, C. D. (1988).

Migration patterns of Thomson’s gazelles and cheetahs on the Serengeti Plains.

African Journal of Ecology, 26, 257–268.

Durant, S. M., Kelly, M., & Caro, T. M. (2004). Factors affecting life and death in

Serengeti cheetahs: Environment, age, and sociality. Behavioral Ecology, 15(1),

11–22. doi:10.1093/beheco/arg098

25

Durant, S. M., Mitchell, N., Groom, R., Pettorelli, N., Ipavec, A., Jacobson, A. P., …

Young-Overton, K. (2017). The global decline of cheetah Acinonyx jubatus and

what it means for conservation. Proceedings of the National Academy of

Sciences, 114(3), 528–533. doi:10.1073/pnas.1611122114

Elmhagen, B., & Rushton, S. P. (2007). Trophic control of mesopredators in terrestrial

ecosystems: Top-down or bottom-up? Ecology Letters, 10(3), 197–206.

doi:10.1111/j.1461-0248.2006.01010.x

Fattebert, J., Balme, G., Dickerson, T., Slotow, R., & Hunter, L. (2015). Density-

dependent natal dispersal patterns in a leopard population recovering from over-

harvest. PLoS ONE, 10(4). doi:10.1371/journal.pone.0122355

Ferreira, S. M., & Hofmeyr, M. (2014). Managing charismatic carnivores in small areas:

large felids in South Africa. South African Journal of Wildlife Research,

44(April), 32–42. doi:10.3957/056.044.0102

Fuller, T., & Sievert, P. (2001). Carnivore demography and the consequences of changes

in prey availability. In J. L. Gittleman, S. M. Funk, D. Macdonald, & R. K.

Wayne (Eds.), Carnivore conservation (pp. 163–179). Cambridge, UK:

Cambridge University Press.

Gaillard, J. M., Nilsen, E. B., Odden, J., Andrén, H., & Linnell, J. D. C. (2014). One size

fits all: Eurasian females share a common optimal litter size. Journal of

Animal Ecology, 83(1), 107–115. doi:10.1111/1365-2656.12110

26

Gehrt, S. D., & Prange, S. (2007). Interference competition between coyotes and

: A test of the mesopredator release hypothesis. Behavioral Ecology,

18(1), 204–214. doi:10.1093/beheco/arl075

Gompper, M. (2002). Top carnivores in the suburbs? Ecological and conservation issues

raised by colonization of northeastern North America by coyotes. BioScience,

52(2), 185–190.

Greenville, A. C., Wardle, G. M., Tamayo, B., & Dickman, C. R. (2014). Bottom-up and

top-down processes interact to modify intraguild interactions in resource-pulse

environments. Oecologia, 175(4), 1349–1358. doi:10.1007/s00442-014-2977-8

Hairston, N. G., Smith, F. E., & Slobodkin, L. B. (1960). Community structure,

population control, and competition. American Naturalist, 94(879), 421–425.

doi:10.1086/282146

Hanski, I. (1990). Density dependence, regulation and variability in animal populations.

Philosophical Transactions of the Royal Society B, 330(1257), 141–150.

doi:10.1098/rstb.2010.0098

Hebblewhite, M., & Pletscher, D. H. (2002). Effects of elk group size on predation by

wolves. Canadian Journal of Zoology, 80(5), 800–809. doi:10.1139/z02-059

Hunter, L. T. B. (1998). The behavioural ecology of reintroduced lions and cheetahs in

the Phinda Resource Reserve, KwaZulu-Natal, South Africa. PhD Thesis, 1–206.

27

Hunter, L. T. B., Pretorius, K., Carlisle, L. C., Rickelton, M., Walker, C., Slotow, R., &

Skinner, J. D. (2007). Restoring lions Panthera leo to northern KwaZulu-Natal,

South Africa: Short-term biological and technical success but equivocal long-term

conservation. Oryx, 41(2), 196–204. doi:10.1017/S003060530700172X

Janse van Rensburg, J., McMillan, M., Giżejewska, A., & Fattebert, J. (2018). Rainfall

predicts seasonal home range size variation in nyala. African Journal of Ecology,

56(2), 418–423. doi:10.1111/aje.12455

Karels, T. J., Byrom, A. E., Boonstra, R., & Krebs, C. J. (2000). The interactive effects of

food and predators on reproduction and overwinter survival of arctic ground

squirrels. Journal of Animal Ecology, 69(2), 235–247. doi:10.1046/j.1365-

2656.2000.00387.x

Kelly, M. J., Laurenson, M. K., Fitzgibbon, C. D., Collins, D. A., Durant, S. M., Frame,

G. W., … Caro, T. M. (1998). Demography of the Serengeti cheetah (Acinonyx

jubatus) population: the first 25 years. Journal of Zoology, 224(4), 473–488.

Kendall, B. E., & Wittmann, M. E. (2010). A stochastic model for annual reproductive

success. The American Naturalist, 175(4), 461–468. doi:10.1086/650724

Kettles, R., & Slotow, R. (2009). Management of free-ranging lions on an enclosed game

reserve. South African Journal of Wildlife Research, 39(1), 23–33.

Kissui, B. M., & Packer, C. (2004). Top-down population regulation of a top predator:

lions in the Ngorongoro Crater. Proceedings of the Royal Society B: Biological

Sciences, 271(1550), 1867–1874. doi:10.1098/rspb.2004.2797

28

Korpimaki, E., Norrdahl, K., & Valkama, J. (1994). Reproductive investment under

fluctuating predation risk : Microtine rodents and small mustelids. Evolutionary

Ecology, 8, 357–368. doi:10.1007/BF01238188

Laake, J. L. (2013). RMark: An R Interface for Analysis of Capture-Recapture Data with

MARK. Version 2.2.4.

Lack, D. (1948). The significance of litter-size. doi:10.1111/j.1474-919X.1947.tb04155.x

Laurenson, M. K. (1994). High juvenile mortality in cheetahs (Acinonyx jubatus) and its

consequences for maternal care. Journal of the Zoological Society of London,

234(1), 387–408. doi:10.1111/j.1469-7998.1994.tb04855.x

Leibold, M. A. (1989). Resource edibility and the effects of predators and productivity on

the outcome of trophic interactions. The American Naturalist, 134(6), 922–949.

doi:10.1086/285022

Letnic, M., Ritchie, E. G., & Dickman, C. R. (2012). Top predators as biodiversity

regulators: The dingo Canis lupus dingo as a case study. Biological Reviews,

87(2), 390–413. doi:10.1111/j.1469-185X.2011.00203.x

Marker, L. L., Dickman, A. J., Jeo, R. M., Mills, M. G. L., & Macdonald, D. W. (2003).

Demography of the Namibian cheetah, Acinonyx jubatus jubatus. Biological

Conservation, 114(3), 413–425. doi:10.1016/S0006-3207(03)00069-7

McNaughton, S. J., Oesterheld, M., Frank, D. A., & Williams, K. J. (1989). Ecosystem-

level patterns of primary productivity and herbivory in terrestrial habitats. Nature,

341, 189–92. doi:10.1038/340301a0

29

Miller, S. M., Bissett, C., Burger, A., Courtenay, B., Dickerson, T., Druce, D. J., …

Zimmermann, D. (2013). Management of reintroduced lions in small, fenced

reserves in South Africa: An assessment and guidelines. South African Journal of

Wildlife Research, 43(2), 138–154. doi:10.3957/056.043.0202

Miller, S. M., & Funston, P. J. (2014). Rapid growth rates of lion (Panthera leo)

populations in small, fenced reserves in South Africa: A management dilemma.

South African Journal of Wildlife Research, 44(1), 43–55.

doi:10.3957/056.044.0107

Mills, M. G. L., & Mills, M. E. J. (2014). Cheetah cub survival revisited: a re-evaluation

of the role of predation, especially by lions, and implications for conservation.

Journal of Zoology, 292(2), 136–141. doi:10.1111/jzo.12087

Oakwood, M. (2000). Reproduction and demography of the northern quoll, Dasyurus

hallucatus, in the lowland savanna of northern Australia. Australian Journal of

Zoology, 48(5), 519–539. doi:10.1071/ZO00028

Owen-Smith, N., Mason, D. R., & Ogutu, J. O. (2005). Correlates of survival rates for 10

African ungulate populations: Density, rainfall and predation. Journal of Animal

Ecology, 74(4), 774–788. doi:10.1111/j.1365-2656.2005.00974.x

Palomares, F., & Caro, T. M. (1999). Interspecific killing among mammalian carnivores.

The American Naturalist, 153(5), 492–508. doi:10.1086/303189

30

Pasanen-Mortensen, M., Elmhagen, B., Lindén, H., Bergström, R., Wallgren, M., van der

Velde, Y., & Cousins, S. A. O. (2017). The changing contribution of top-down

and bottom-up limitation of mesopredators during 220 years of land use and

climate change. Journal of Animal Ecology, 86(3), 566–576. doi:10.1111/1365-

2656.12633

Pettorelli, N., & Durant, S. M. (2007). Family effects on early survival and variance in

long-term reproductive success of female cheetahs. Journal of Animal Ecology,

76(5), 908–914. doi:10.1111/j.1365-2656.2007.01266.x

Prugh, L. R., Stoner, C. J., Epps, C. W., Bean, W. T., Ripple, W. J., Laliberte, A. S., &

Brashares, J. S. (2009). The rise of the mesopredator. BioScience, 59(9), 779–

791. doi:10.1525/bio.2009.59.9.9

R Core Team. (2019). R: A language and environment for statistical computing. Vienna,

Austria: R Foundation for Statistical Computing.

Ralls, K., & White, P. J. (1995). Predation on San Joaquin kit by larger canids.

Journal of Mammalogy, 76(3), 723–729.

Ritchie, E. G., & Johnson, C. N. (2009). Predator interactions, mesopredator release and

biodiversity conservation. Ecology Letters, 12(9), 982–998. doi:10.1111/j.1461-

0248.2009.01347.x

Rostro-García, S., Kamler, J. F., & Hunter, L. T. B. (2015). To kill, stay or flee: the

effects of lions and landscape factors on habitat and kill site selection of cheetahs

in South Africa. Plos One, 10(2), e0117743. doi:10.1371/journal.pone.0117743

31

Royle, J. A., Dawson, D. K., & Bates, S. (2004). Modelling abundance effects in distance

sampling. Ecology, 85(6), 1591–1597. doi:10.1890/03-3127

Sillett, S., Chandler, R. B., Royle, J. A., Kéry, M., Morrison, S. a, Sillett, T. S., & Kery,

M. (2012). Hierarchical distance sampling models to estimate population size and

habitat-specific abundance of an island endemic. Ecological Applications, 22(7),

1997–2006. doi:10.1890/11-1400.1

Sinclair, A. R. E., & Krebs, C. J. (2002). Complex numerical responses to top–down and

bottom–up processes in vertebrate populations. Philosophical Transactions of the

Royal Society of London. Series B: Biological Sciences, 357(1425), 1221–1231.

doi:10.1098/rstb.2002.1123

Swanson, A., Arnold, T., Kosmala, M., Forester, J., & Packer, C. (2016). In the absence

of a “landscape of fear”: How lions, hyenas, and cheetahs coexist. Ecology and

Evolution, 1–12. doi:10.1002/ece3.2569

Swanson, A., Caro, T., Davies Mostert, H., Mills, M. G. L., Macdonald, D. W., Borner,

M., … Packer, C. (2014). Cheetahs and wild dogs show contrasting patterns of

suppression by lions. Journal of Animal Ecology, 1418–1427. doi:10.1111/1365-

2656.12231

Switalski, T. A. (2003). foraging ecology and vigilance in response to gray wolf

reintroduction in Yellowstone National Park. Canadian Journal of Zoology, 81(6),

985–993. doi:10.1139/z03-080

Tannerfeldt, M., & Angerbjorn, A. (1998). Fluctuating resources and the evolution of

litter size in the arctic . Oikos, 83(3), 545–559.

32

Trinkel, M., & Kastberger, G. (2005). Competitive interactions between spotted hyenas

and lions in the , Namibia. African Journal of Ecology, 43,

220–224.

Vanak, A. T., Fortin, D., Thaker, M., Ogden, M., Owen, C., Greatwood, S., & Slotow, R.

(2013). Moving to stay in place: Behavioral mechanisms for coexistence of

African large carnivores. Ecology, 94(11), 2619–2631. doi:10.1890/13-0217.1

Weise, F. J., Vijay, V., Jacobson, A. P., Schoonover, R. F., Groom, R. J., Horgan, J., …

Pimm, S. L. (2017). The distribution and numbers of cheetah (Acinonyx jubatus)

in southern Africa. PeerJ, 5, e4096. doi:10.7717/peerj.4096

33

TABLES

Table 1.1. Model selection results for cheetah litter size regression models, Mun-Ya- Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. a b c Model AICc ΔAICC -2×ln(L) w k

Null 155.2 0 -76.55 0.33 1

Lion density 156.2 1.04 -75.97 0.19 2

Prey density 156.4 1.25 -76.07 0.18 2

Cheetah density 157.3 2.15 -76.52 0.11 2

Prey density + lion density 158.4 3.19 -75.89 0.07 3

Prey density + cheetah density 158.4 3.25 -75.92 0.06 3

Prey density + lion density + cheetah density 160.5 5.33 -75.74 0.02 4

Prey density * cheetah density 160.5 5.35 -75.75 0.02 4

Prey density * lion density 160.8 5.61 -75.89 0.02 4

a Log-likelihood b Akaike model weight c Number of model parameters

34

Table 1.2. Model selection results for cheetah recruitment regression models, Mun-Ya- Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. a b c AICc ΔAICC -2×ln(L) w k Cheetah density* prey density 115.96 0 -52.72 0.34 5

Cheetah density + prey density 116.00 0.04 -53.83 0.33 4

Cheetah density + lion density + prey density 117.93 1.96 -53.70 0.13 5

Cheetah density + lion density 119.04 3.08 -55.35 0.07 4

Cheetah density * lion density 119.78 3.81 -54.62 0.05 5

Cheetah density 120.92 4.95 -57.35 0.03 3

Lion density * prey density 121.15 5.19 -55.31 0.03 5

Prey density 121.96 5.99 -57.87 0.02 3

Lion density + prey density 122.40 6.44 -57.03 0.01 4

Lion density 124.09 8.12 -58.94 0.01 3

Null 131.37 15.41 -63.63 0.00 2

a Log-likelihood b Akaike model weight c Number of model parameters

35

Table 1.3. Model selection results for structural multi-state joint live-encounter dead- recovery survival models for cheetahs to incorporate variation by time and sex for resighting (p) and survival (S) rates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018.

a b c Parameter Model AICc ΔAICC -2×ln(L) w k p p(year+season) 3570.39 0 3540.02 0.66 15

p(year) 3571.70 1.31 3543.38 0.34 14

p(season) 3584.35 13.96 3572.29 0.00 6

p(null) 3584.37 13.98 3574.33 0.00 5

S S(sex+season) 3551.37 0 3533.23 0.80 9

S(sex) 3554.09 2.72 3537.98 0.20 8

S(season) 3578.76 27.39 3564.67 0.00 7

S(null) 3584.37 33.01 3574.33 0.00 5

a Log-likelihood b Akaike model weight c Number of model parameters

36

Table 1.4. Model selection results for multi-state joint live-encounter dead-recovery survival models for cheetahs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. States in the model include cubs (juveniles dependent on their mothers) and adults (non-juveniles). All models include effects of year on recovery rates and season on survival rates. a b c Model AICc ΔAICC -2×ln(L) w k

S(state:lion density * state:prey density) 3538.66 0.00 3518.49 0.94 10

S(state:cheetah density + state:lion density + 3545.57 10.23 3525.41 0.03 10 state:prey density)

S(state:prey density + state:lion density) 3546.69 10.54 3530.58 0.02 8

S(state:prey density) 3549.04 13.76 3536.98 0.01 6

S(state:prey density + state:cheetah density) 3549.68 16.06 3533.57 0.00 8

S(state:prey density * state:cheetah density) 3550.12 16.35 3529.95 0.00 10

S(state:average prey density 6mos prior + 3567.74 29.52 3551.63 0.00 8 state:cheetah density)

S(state:average prey density 6mos prior + 3569.32 33.48 3553.21 0.00 8 state:lion density)

S(state:average prey density 6mos prior) 3573.93 35.45 3561.86 0.00 6

S(state:lion density) 3577.17 36.19 3565.11 0.00 6

S(state:cheetah density) 3580.14 43.16 3568.08 0.00 6

S(state) 3584.37 45.42 3574.33 0.00 5

a Log-likelihood b Akaike model weight c Number of model parameters

37

FIGURES

Figure 1.1. Probability of cub recruitment in relation to cheetah density and prey density, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018.

38

Figure 1.2. Monthly survival for a) adult and b) cub cheetahs in relation to prey density and lion density during dry seasons of years with average resighting rates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018.

39

CHAPTER 2

HABITAT COMPLEXITY AND LIFETIME PREDATION RISK INFLUENCE MESOPREDATOR SURVIVAL IN A MULTI-PREDATOR SYSTEM

ABSTRACT

Variability in habitat selection can lead to differences in individual or population- level fitness; however limited research exists on how the habitat selection of mid-ranking predators can influence population-level processes in multi-predator systems. For mid- ranking, or mesopredators, differences in habitat use might have strong demographic effects because mesopredators need to avoid apex predators, while still finding and killing prey. Additionally, many mesopredators use fine-scale partitioning to coexist with apex predators, but little is known about how these avoidance behaviors might lead to long-term demographic consequences. We studied spatially-explicit survival of cheetahs

(Acinonyx jubatus) in the Mun-Ya-Wana Conservancy, South Africa, to test hypotheses related to spatial influences of predation risk, prey availability, and vegetation complexity, on mesopredator survival. For each monitored cheetah, we estimated lion encounter risk, prey density, and vegetation complexity within their home range, on short-term (seasonal) and long-term (lifetime) scales and used multi-state live-encounter dead-recovery models to estimate survival based on these covariates. Survival was lowest for adult cheetahs and cubs in areas with high vegetation complexity on both seasonal and lifetime scales. Additionally, cub survival was negatively related to the long-term risk of encountering a lion. We suggest that complex habitats are only beneficial to mesopredators when they are able to effectively find and hunt prey, and show that spatial

40 drivers of survival for mesopredators can vary temporally, with short-term survival driven by prey acquisition and long-term survival influenced by both top-down and bottom-up factors. We show for the first time that long-term coexistence with apex predators can reduce the survival of mesopredators in a natural system. Collectively, our research illustrates that individual variation in mesopredator habitat use can scale-up and have population-level effects.

INTRODUCTION

Understanding how individual habitat use can influence fitness can offer insight into the structure of food webs and predator-prey interactions (Schmitz, Miller, Trainor,

& Abrahms, 2017). Spatial and temporal variation in resources, as well as the ability of individuals to find and use these resources, can lead to differences in individual survival and reproduction, which in turn can scale up to population-level effects (Gaillard et al.,

2010; van Noordwijk & de Jong, 1986). In predator-prey systems, environmental features that influence the predator’s ability to find or kill prey can affect the survival or reproduction of the predator (Kosterman, Squires, Holbrook, Pletscher, & Hebblewhite,

2018; Mosser, Fryxell, Eberly, & Packer, 2009). On the other hand, environmental features that influence predator avoidance, predator detection, the prey’s ability to escape predators, or the prey’s ability to find food, can affect the survival or reproduction of the prey (DeCesare et al., 2014; M. Hebblewhite, Merrill, & McDonald, 2005; McLoughlin,

Dunford, & Boutin, 2005). Identifying the connections between habitat use and

41 demography is important for understanding the fitness costs and benefits of habitats, although this understanding is lacking for systems with multiple predators.

In systems with multiple predators, the interaction between habitat use and demography could be particularly complex. Subordinate predators, or mesopredators, need to select habitats that will allow them to obtain prey, while still avoiding predation by top predators (Ritchie & Johnson, 2009). In some systems, apex predators and mesopredators occur in the same general habitats, but mesopredators use fine-scale spatial or temporal partitioning to reduce encounter rates (Torretta, Serafini, Puopolo, &

Schenone, 2016; Vanak et al., 2013). Although this partitioning might reduce short-term mortality risk for mesopredators, the longer-term risk of co-occurring with apex predators could reduce the fitness of the mesopredator through non-consumptive effects, such as reduced foraging opportunities or shifts into non-optimal habitats (Preisser, Bolnick, &

Benard, 2005). However, limited research exists on the demographic consequences of mesopredator-apex predator cooccurrence.

Additionally, habitat characteristics can modulate the habitat-survival relationship in systems with multiple predators. Most of what we know about the effects of habitat use on demography in natural systems comes from top predators (e.g., lions, Panthera leo;

Mosser et al. 2009, , Lynx lynx; Basille et al. 2013) or prey (e.g., woodland caribou, Rangifer tarandus caribou; DeCesare et al. 2014, roe deer, Capreolus capreolus;

McLoughlin et al. 2007), whereas effects of habitat use on subordinate predator demography have typically only been studied in experimental systems with aquatic or insect species (Finke & Denno, 2002; Janssen, Sabelis, Magalhães, & Van, 2007). Theory

42 predicts that mesopredators will experience reduced mortality from apex predators (i.e., intraguild predation or intraspecific killing) in areas with high habitat complexity, because of lower encounter rates (Janssen et al., 2007). A greater understanding of how habitat characteristics can influence mesopredator survival in natural systems is critical.

We studied the interaction between habitat use and mesopredator demography in a natural system, using the cheetah (Acinonyx jubatus) as our focal species. Cheetahs are subordinate to lions, and the majority of cheetah mortality is from lion predation

(Laurenson, 1994). In addition to direct predation, lions can steal prey from cheetahs

(Hunter, Durant, & Caro, 2007), and can affect the habitat use and behavior of cheetahs

(Hilborn et al., 2018; Swanson et al., 2016). In turn, these non-consumptive effects related to predation risk might affect the long-term survival and fitness of cheetahs.

Previous research suggests that cheetahs and lions exhibit high levels of home range overlap, but that cheetah use fine-scale spatial partitioning to avoid the short-term risk of lion predation (Broekhuis et al., 2013; Dröge, Creel, Becker, & M’soka, 2016; Vanak et al., 2013). However, how this partitioning might indirectly affect the survival of cheetahs in the long-term is unknown. In addition, dense vegetation is hypothesized as another mechanism of lion-cheetah coexistence by acting as a predation refuge (Mills & Mills,

2014), but research has not been conducted to test the linkage between vegetation and cheetah survival.

We investigated support for three competing hypotheses of spatial drivers of mesopredator survival: 1) mesopredator survival would be driven by the risk of encountering top predators (top-down spatial regulation), 2) mesopredator survival would

43 be driven by prey densities (bottom-up spatial regulation), and 3) mesopredator survival would be driven by vegetation complexity (habitat complexity risk mediation hypothesis). Under the spatial top-down hypothesis we predicted that cheetah survival would be negatively related to the probability of encountering lions. Under the spatial bottom-up hypothesis we predicted that cheetah survival would be positively related to spatial prey density. Under the habitat complexity risk mediation hypothesis, we predicted that cheetah survival would be positively related to vegetation complexity. By studying spatial influences of mesopredator demography, we can better understand factors that structure food webs with multiple predators, and in turn prioritize habitat features that promote the coexistence of multiple predator species.

METHODS

Study Area

We studied cheetah demographics in Mun-Ya-Wana Conservancy (Phinda Private

Game Reserve), in northern KwaZulu-Natal, South Africa, from 2008 – 2018. The dominant vegetation type is broad-leaf woodland, with open grasslands and semi-open wooded-grasslands interspersed throughout the reserve. The elevation of the Conservancy ranges from 4 to 201 meters above sea level. The climate is subtropical with warm, dry winters (April – September) and hot, humid summers (October – March), with the majority of rain falling in the summer (Janse van Rensburg et al., 2018). The Mun-Ya-

Wana Conservancy is surrounded by electrified game fencing, and has grown in size as adjacent reserves have joined the Conservancy. From 2008 – 2017 the study area was 235

44 km2 in area, after which internal fences were removed and the Conservancy expanded to

285 km2. Cheetahs and lions were reintroduced into the reserve in 1992 and have been monitored since (Hunter, 1998; Hunter et al., 2007; this study).

Carnivore Monitoring

We monitored the cheetah and lion populations by subdividing the reserve into seven sections. Trained monitors usually drove the roads in each section at least once a week. In addition, monitors frequently followed-up on sightings reported by game rangers conducting game drives within the reserve. Cheetahs and lions can be individually recognized using their spot patterns, whisker spots, and scars, which allowed us to monitor the populations based on sightings alone (Caro, 1994). We obtained an average of 32 ± 2 cheetah sightings per month and an average of 18 ± 1 lion pride sightings per month. When cheetahs or lions were observed, we recorded the location, behavior, and number of individuals present.

Cheetah Habitat Use

We quantified coarse-scale cheetah habitat use by estimating lifetime home ranges for individual cheetahs. Although small-scale differences in habitat use might occur seasonally, cheetahs in our study area had stable home ranges across their lifetimes

(Appendix B1). However, cheetahs will often shift home ranges when they become independent from their mothers, so for individuals that were included in the study as both cubs and adults, we estimated cub and adult home ranges separately. We estimated home

45 ranges by calculating a utilization distribution (UD) using a fixed-kernel estimator and the plug-in method of bandwidth selection (Gitzen, Millspaugh, & Kernohan, 2006). For each cheetah’s home range, we extracted time-varying covariates of lion encounter risk, prey spatial density, and vegetation complexity (see below). To account for temporal differences in spatial drivers of survival, we extracted covariates within home ranges corresponding to each season, and also averaged covariates within home ranges across the lifetime of individual cheetahs. To identify the spatial scale most influential to survival, we extracted these covariates within the 50%, 75%, and 95% UD isopleths and compared the resulting survival patterns at each scale (Appendix B2).

Lion Encounter Risk

We estimated lion encounter risk by analyzing the spatial distribution of lions in each season (Moll, Killion, Montgomery, Tambling, & Hayward, 2016; Thaker et al.,

2011). We collected sightings data on the location of lion prides, rather than individual lions, from 2000-2019. For each pride of lions in a given season, we calculated a utilization distribution (UD) using a fixed-kernel estimator using the plug-in method of bandwidth selection (Gitzen et al., 2006). To account for differences in pride size, we multiplied the UD for each pride by the average number of lions in that pride within a specific season (Kauffman et al., 2007). To obtain a reserve-level measure of lion encounter risk, we added the individual pride UDs and rescaled the resulting values such that a value of 0 indicated no risk of encounter, and 1 indicated the highest risk of encounter.

46

Prey Spatial Density

We estimated spatial variation in prey density in the reserve by collecting distance sampling data on impala (Aepyceros melampus) and nyala (Tragelaphus angasii) during the dry season (April – September), and the wet season (October – March) from 2010 –

2015 (Chapter 1). We limited our prey analyses to these species because they comprised

82% of cheetah kills in the study area (Hunter, 1998). We estimated prey abundance using hierarchical distance sampling models with spatial covariates on both the abundance and detection processes, and used our top model to extrapolate prey abundance over our entire study period (Chapter 1). Our resulting prey density rasters depicted the average number of prey within 400 m2 cells for each season of each year.

Vegetation Complexity

We incorporated spatial variation in vegetation complexity into our cheetah survival models. Enhanced Vegetation Index (EVI) has been found to be correlated with vegetation structure in Africa, with open areas having low EVI values, and areas with dense vegetation having high EVI values (Tsalyuk, Kelly, & Getz, 2017). In addition,

EVI is sensitive to changes in rainfall. Thus, using EVI as an index for vegetation complexity also allowed us to incorporate changes in greenness, which could affect visibility. We obtained (EVI) data at a 250 m resolution

(https://lpdaac.usgs.gov/data_access/data_pool) and calculated seasonal EVI values on a yearly basis by averaging EVI values across the entirety of a season.

47

Cheetah Spatially-Explicit Survival

We analyzed spatial drivers of cheetah monthly survival from February 2009 to

March 2018. When a dead cheetah was recovered, we attempted to determine the cause of death by examining the carcass and the surrounding area for tracks and scats. For each month, we recorded if individual cheetahs were sighted or recovered dead as adults or cubs. If a cheetah was removed from the reserve for management purposes, we censored that individual animal from analyses. Because lion and cheetah density can vary greatly within a season, and because cubs can be born and become independent at any time during the year, we conducted our analysis on a monthly timescale to best reflect the conditions that might be driving survival.

Model Structure

We analyzed cheetah survival using multi-state joint live-encounter dead-recovery models (Barker et al., 2005) using the rmark R package (Laake, 2013). This model made use of our frequent re-sightings and mortality data, and allowed for survival estimation based on individuals with unknown fates. Additionally, because juvenile cheetahs stay with their mothers for variable amounts of time (Kelly et al., 1998), we could not incorporate a standard age structure into our models. Thus, we used a multi-state approach to estimate survival for both cubs and adults simultaneously. We specified the two model states as cub (juvenile cheetahs dependent on their mother) and adult

(cheetahs that were independent from their mother) and did not incorporate immigration or emigration because our population was a closed population.

48

Hypothesis Testing

We previously determined that cheetah survival was best described using a structural model with resighting rate varying by year, and survival varying by season

(Chapter 1). Therefore, we used the same structural model for these analyses to test for the effects of spatial covariates on cheetah survival. We considered spatial covariates on two temporal scales: short-term (seasonal), and long-term (spatial covariates within a home range averaged over an individual’s lifetime). For each temporal scale, we developed 11 a priori models based on our hypotheses of interest that included covariates of average lion encounter risk, average prey density, and average EVI, as well as additive and multiplicative models with the same covariates. Because adults and cubs are known to have different survival rates (Kelly et al., 1998), we did not consider any models in which state was not included. We compared models using Akaike’s Information Criterion corrected for sample size (AICc; Burnham and Anderson, 2002), considered models within 2 ΔAICc of the top model to be competitive, and evaluated if covariates were informative by calculating 85% confidence intervals (Arnold, 2010).

RESULTS

Short-term spatial drivers of survival

We included 133 cheetahs in our survival analyses for a total of 110 months.

Cheetah survival was most sensitive to short-term environmental conditions within the

50% home range contour and the same top model was supported regardless of home range contour (Appendix B2). Thus, we only present the results from the 50% HR

49 models. At the short-term time scale, cheetahs exhibited variation in environmental conditions within their home range with regards to EVI (mean = 0.29; range = 0.14 –

0.44), lion encounter risk (mean = 0.36; range = 0.16 – 0.90), and prey density (mean =

37.8 prey/km2, range = 24.0 – 80.3 prey/km2).

For our short-term survival models, survival was best described by the average

EVI within the core of a cheetah’s home range (Table 2.1). For both adults and cubs, EVI had a negative influence on survival, with higher survival occurring at the lowest EVI values (Figure 2.1). Cub survival was most sensitive to changes in EVI, with the probability of surviving decreasing 2.8% for every 1-unit increase in EVI.

Long-term spatial drivers of survival

Similar to short-term survival, we assessed lifetime environmental covariates using the 50% HR contour. At the lifetime scale, cheetahs exhibited variation in environmental conditions within their home range with regards to EVI (mean = 0.3; range

= 0.19 – 0.42), lion encounter risk (mean = 0.37; range = 0.13 – 0.84), and prey density

(mean = 40.4 prey/km2, range = 25.8 – 75.7 prey/km2).

When considering environmental covariates across individuals’ lifetimes, survival was best described by a model including the average EVI within a cheetah’s home range during their lifetime, and the average risk of encountering a lion within a cheetah’s home range during their lifetime (Table 2.2). Models including EVI and lion encounter risk separately were also competitive (Table 2.2). Based on the top model, lifetime EVI had a

50 negative influence on adult and cub survival, and lifetime lion encounter risk had a negative influence on cub survival, but no influence on adult survival (Figure 2.2).

DISCUSSION

We evaluated support for effects of spatial influences of top-down predation risk, bottom-up prey availability, and habitat complexity, and found the most consistent support for survival being influenced by habitat complexity across multiple temporal scales. However, our results contradict the habitat complexity risk mediation hypothesis, which predicts that mesopredators should experience increased survival in areas of high habitat complexity (Janssen et al., 2007). Instead, our results show that subordinate predators do not always benefit from structurally complex habitats, potentially because subordinate predators might only benefit from habitat complexity if they are able to avoid predation, and effectively obtain prey, in complex habitats. In predator-prey systems, using specific areas as refuges from predation can come at a cost to the prey, because although they might reduce predation risk, the resource availability of the refuge might be lower than non-refuge areas because of increased competition or sub-optimal conditions

(Donelan, Grabowski, & Trussell, 2016; Orrock, Preisser, Grabowski, & Trussell, 2013).

In systems of multiple predators, the quality of a predation refuge habitat might be related to the subordinate predator’s ability to find and hunt prey, which is a function of the subordinate predator’s hunting mode (Miller, Ament, & Schmitz, 2014).

There are two main explanations as to why we did not find support for the habitat complexity risk mediation hypothesis in our study system. First, vegetation complexity

51 might increase or reduce the probability of cheetahs being predated upon. Ambush predators in a variety of systems have been found to have enhanced hunting abilities in structurally-complex areas because of decreased sight lines for prey (Blake & Gese,

2016; Michel & Adams, 2009). Although lions use a variety of habitat types, they kill prey more frequently in areas of dense vegetation (Davies, Tambling, Kerley, & Asner,

2016; Hopcraft, Sinclair, & Packer, 2005). Closed habitat types could act as a predation refuge for cheetahs by providing cover to enhance concealment from lions, but these habitats could also increase predation risk because of the hunting preferences of lions.

Conversely, open habitat types might reduce the probability of predation by improving cheetahs’ ability to detect nearby lions, compared to closed habitats (Camp, Rachlow,

Woods, Johnson, & Shipley, 2012). However, when we analyzed locations where cheetahs were killed by other predators (Appendix B3), we did not find evidence to suggest that vegetation complexity increased the risk of cheetahs dying, but we cannot entirely exclude this mechanism as driving our observed patterns of spatial survival.

Therefore, it seems that cheetahs experience predation independent of vegetation complexity and higher predation in areas of higher EVI might not be the mechanism driving our observed patterns of spatial survival.

Second, vegetation complexity might influence cheetahs’ hunting ability, which in turn could affect survival. Cheetahs are coursing predators, as opposed to ambush predators, and can reach high speeds when chasing prey (Wilson et al., 2013). Therefore, open areas could improve cheetah hunting success by allowing cheetahs to see prey easier and facilitating high-speed chases. Although cheetahs are able to hunt in areas of

52 dense vegetation (Rostro-García et al., 2015), they are more likely to initiate hunts, and have higher hunting success, in open habitats (Mills, Broomhall, du Toit, & Toit, 2004).

Prey availability can be an important driver of carnivore demography (Fuller & Sievert,

2001), so the use of areas to facilitate hunting, rather than the density of prey themselves

(Balme, Hunter, & Slotow, 2007), could influence mesopredator survival. Indeed, when we analyzed cheetah kill site locations (Appendix B3) we found that cheetah kill sites were more likely to be located in areas with low EVI (Figure 2.3).

In addition to habitat complexity affecting cheetah survival, we found that cheetah survival was also influenced by duration of exposure to top-down predation risk. Our results indicate that long-term risk of encountering a lion, rather than short-term risk of encountering a lion, influenced cheetah survival, with cubs having a lower probability of survival if there was a higher probability of encountering lions in their home range during their lifetime. We likely did not observe effects of lion encounter risk on short-term cheetah survival because cheetahs have adapted behaviors to minimize short-term predation risk (Swanson et al., 2016). For example, cheetahs use the same general areas as apex predators such as lions and use fine-scale spatial partitioning to reduce the probability of lion encounters (Broekhuis et al., 2013; Dröge et al., 2016; Vanak et al.,

2013). The use of spatial or temporal partitioning by mesopredators affected by apex predators has been found in a variety of other systems, such as red foxes ( Vulpes) avoiding coyotes (Canis latrans) in North America (Gosselink, Van Deelen, Warner, &

Joselyn, 2003), and European ( meles) avoiding wolves (Canis lupus) in

Italy (Torretta et al., 2016).

53

Although fine-scale partitioning or other predator avoidance behaviors might be beneficial to reduce short-term risk for mesopredators, our results show that the long- term risk of co-occurring with an apex predator can negatively influence mesopredator survival. In the long-term, the risk of encountering lions could be associated with the direct effects of predation, with an increased probability of antagonistic encounters

(Palomares & Caro, 1999). Long-term risk can also be associated with non-consumptive effects of predation related to reduced foraging (Brown, 1999), or reduced parental care

(Dudeck, Clinchy, Allen, & Zanette, 2018). In the absence of direct predation, the long- term risk of predation has been found to cause changes in the morphology (Relyea,

2001), behavior (Suraci, Clinchy, Dill, Roberts, & Zanette, 2016; Valeix et al., 2009), physiology (Clinchy et al., 2011; Sheriff, Krebs, & Boonstra, 2009), and demography

(LaManna & Martin, 2016; Travers, Clinchy, Zanette, Boonstra, & Williams, 2010) of a variety of prey species. Our results build on the growing literature on long-term risk of predation in natural systems to demonstrate how long-term predation risk can affect the demography of mesopredators, rather than just of prey.

Understanding spatial variation in survival can help inform wildlife conservation actions by focusing efforts on environmental factors that improve the survival of imperiled species. Specific for cheetahs in southern Africa, bush encroachment has caused the transition from open grasslands to closed habitats dominated by woody plants

(Roques, O’Connor, & Watkinson, 2001). Bush encroachment can be caused by a number of factors including climate, fire, and herbivore distributions, but is predicted to increase based on future climate change models (Tews & Jeltsch, 2004). Based on our

54 results, increased bush encroachment could be detrimental to cheetah populations that do not have adequate open areas for hunting. Thus, the persistence of this species in the southern portion of their range could be improved by prescribed burning or mechanical vegetation removal in order to maintain open habitats (Joubert, Smit, & Hoffman, 2012;

Lohmann, Tietjen, Blaum, Joubert, & Jeltsch, 2014).

Our research shows how individual space use of mesopredators can scale-up and influence population-level processes, and illustrates the importance of understanding spatial drivers of survival on different temporal scales (Moll et al., 2017; Prugh et al.,

2019). We propose that, in systems with apex predators and mesopredators, the survival of mesopredators in the short-term is driven by vegetative complexity likely associated with prey acquisition, whereas long-term survival depends on both top-down and bottom- up influences. Additionally, our results show that complex habitats might only be beneficial for mesopredators when they allow mesopredators to avoid apex predators, and effectively find and hunt prey, at the same time. Understanding how individual space use can influence population-level processes of mesopredators can offer insight into how communities with multiple predators are structured and can provide recommended conservation actions to ensure the future persistence of mesopredator species.

REFERENCES

Arnold, T. W. (2010). Uninformative parameters and model selection using Akaike’s

Information Criterion. Journal of Wildlife Management, 74(6), 1175–1178.

doi:10.2193/2009-367

55

Balme, G., Hunter, L. T. B., & Slotow, R. (2007). Feeding habitat selection by hunting

leopards Panthera pardus in a woodland savanna: prey catchability versus

abundance. Animal Behaviour, 74(3), 589–598. doi:10.1016/j.anbehav.2006.12.014

Barker, R. J., White, G. C., & McDougall, M. (2005). Movement of Paradise Shelduck

between molt sites: a joint multistate-dead recovery mark-recapture model. Journal

of Wildlife Management, 69(3), 1194–1201. doi:10.2193/0022-

541X(2005)069[1194:MOPSBM]2.0.CO;2

Basille, M., Moorter, B. Van, Herfindal, I., Martin, J., Linnell, J. D. C., Odden, J., …

Gaillard, J. (2013). Selecting habitat to survive: The impact of road density on

survival in a large carnivore. PloS One, 8(7). doi:10.1371/journal.pone.0065493

Blake, L. W., & Gese, E. M. (2016). Resource selection by : Influence of

behavioral state and season. Journal of Wildlife Management, 80(7), 1205–1217.

doi:10.1002/jwmg.21123

Broekhuis, F., Cozzi, G., Valeix, M., Mcnutt, J. W., & Macdonald, D. W. (2013). Risk

avoidance in sympatric large carnivores: Reactive or predictive? Journal of Animal

Ecology, 82(5), 1098–1105. doi:10.1111/1365-2656.12077

Brown, J. S. (1999). Vigilance, patch use and habitat selection: Foraging under predation

risk. Evolutionary Ecology Research, 1(1), 49–71.

Burnham, K. P., & Anderson, D. R. (2002). Model selection and multimodel inference: A

practical information-theoretic approach. Springer Science & Business Media.

56

Camp, M. J., Rachlow, J. L., Woods, B. A., Johnson, T. R., & Shipley, L. A. (2012).

When to run and when to hide: The influence of concealment, visibility, and

proximity to refugia on perceptions of risk. Ethology, 118(10), 1010–1017.

doi:10.1111/eth.12000

Caro, T. M. (1994). Cheetahs of the Serengeti Plains. Chicago: The University of

Chicago Press.

Clinchy, M., Zanette, L., Charlier, T. D., Newman, A. E. M., Schmidt, K. L., Boonstra,

R., & Soma, K. K. (2011). Multiple measures elucidate glucocorticoid responses to

environmental variation in predation threat. Oecologia, 166(3), 607–614.

doi:10.1007/s00442-011-1915-2

Davies, A. B., Tambling, C. J., Kerley, G. I. H., & Asner, G. P. (2016). Effects of

vegetation structure on the location of lion kill sites in African thicket. PLoS ONE,

11(2), 1–20. doi:10.1371/journal.pone.0149098

DeCesare, N. J., Hebblewhite, M., Bradley, M., Hervieux, D., Neufeld, L., & Musiani,

M. (2014). Linking habitat selection and predation risk to spatial variation in

survival. Journal of Animal Ecology, 83(2), 343–352. doi:10.1111/1365-2656.12144

Donelan, S. C., Grabowski, J. H., & Trussell, G. C. (2016). Refuge quality impacts the

strength of nonconsumptive effects on prey. Ecology, 98(2), 403–411.

doi:10.1002/ecy.1647

Dröge, E., Creel, S., Becker, M. S., & M’soka, J. (2016). Spatial and temporal avoidance

of risk within a large carnivore guild. Ecology and Evolution, 7, 189–199.

doi:10.1002/ece3.2616

57

Dudeck, B. P., Clinchy, M., Allen, M. C., & Zanette, L. Y. (2018). Fear affects parental

care, which predicts juvenile survival and exacerbates the total cost of fear on

demography. Ecology, 99(1), 127–135. doi:10.1002/ecy.2050

Finke, D. L., & Denno, R. F. (2002). Intraguild predation diminished in complex-

structured vegetation: Implication for prey suppression. Ecology, 83(3), 643–652.

doi:10.1890/0012-9658(2002)083[0643:IPDICS]2.0.CO;2

Fuller, T., & Sievert, P. (2001). Carnivore demography and the consequences of changes

in prey availability. In J. L. Gittleman, S. M. Funk, D. Macdonald, & R. K. Wayne

(Eds.), Carnivore conservation (pp. 163–179). Cambridge, UK: Cambridge

University Press.

Gaillard, J.-M., Hebblewhite, M., Loison, A., Fuller, M., Powell, R., Basille, M., & Van

Moorter, B. (2010). Habitat-performance relationships: finding the right metric at a

given spatial scale. Philosophical Transactions of the Royal Society of London.

Series B, Biological Sciences, 365(1550), 2255–2265. doi:10.1098/rstb.2010.0085

Gitzen, R. A., Millspaugh, J. J., & Kernohan, B. J. (2006). Bandwidth selection for fixed-

kernel analysis of animal utilization distributions. Journal of Wildlife Management,

70(5), 1334–1344. doi:10.2193/0022-541x(2006)70[1334:bsffao]2.0.co;2

Gosselink, T. E., Van Deelen, T. R., Warner, R. E., & Joselyn, M. G. (2003). Temporal

habitat partitioning and spatial use of coyotes and red foxes in East-Central Illinois.

The Journal of Wildlife Management, 67(1), 90–103.

58

Hebblewhite, M., Merrill, E. H., & McDonald, T. L. (2005). Spatial decomposition of

predation risk using resource selection functions: An example in a wolf-elk

predator-prey system. Oikos, 111(1), 101–111. doi:10.1111/j.0030-

1299.2005.13858.x

Hilborn, A., Pettorelli, N., Caro, T., Kelly, M. J., Laurenson, M. K., & Durant, S. M.

(2018). Cheetahs modify their prey handling behavior depending on risks from top

predators. Behavioral Ecology and Sociobiology.

Hopcraft, J. G. C., Sinclair, A. R. E., & Packer, C. (2005). Planning for success:

Serengeti lions seek prey accessibility rather than abundance. Journal of Animal

Ecology, 74(3), 559–566. doi:10.1111/j.1365-2656.2005.00955.x

Hunter, J. S., Durant, S. M., & Caro, T. M. (2007). To flee or not to flee: Predator

avoidance by cheetahs at kills. Behavioral Ecology and Sociobiology, 61, 1033–

1042. doi:10.1007/s00265-006-0336-4

Hunter, L. T. B. (1998). The behavioural ecology of reintroduced lions and cheetahs in

the Phinda Resource Reserve, KwaZulu-Natal, South Africa. PhD Thesis, 1–206.

Hunter, L. T. B., Pretorius, K., Carlisle, L. C., Rickelton, M., Walker, C., Slotow, R., &

Skinner, J. D. (2007). Restoring lions Panthera leo to northern KwaZulu-Natal,

South Africa: Short-term biological and technical success but equivocal long-term

conservation. Oryx, 41(2), 196–204. doi:10.1017/S003060530700172X

Janse van Rensburg, J., McMillan, M., Giżejewska, A., & Fattebert, J. (2018). Rainfall

predicts seasonal home range size variation in nyala. African Journal of Ecology,

56(2), 418–423. doi:10.1111/aje.12455

59

Janssen, A., Sabelis, M. W., Magalhães, S., & Van, T. (2007). Habitat structure affects

intraguild predation. Ecology, 88(11), 2713–2719.

Joubert, D. F., Smit, G. N., & Hoffman, M. T. (2012). The role of fire in preventing

transitions from a grass dominated state to a bush thickened state in arid savannas.

Journal of Arid Environments, 87, 1–7. doi:10.1016/j.jaridenv.2012.06.012

Kauffman, M. J., Varley, N., Smith, D. W., Stahler, D. R., MacNulty, D. R., & Boyce, M.

S. (2007). Landscape heterogeneity shapes predation in a newly restored predator-

prey system. Ecology Letters, 10(8), 690–700. doi:10.1111/j.1461-

0248.2007.01059.x

Kelly, M. J., Laurenson, M. K., Fitzgibbon, C. D., Collins, D. A., Durant, S. M., Frame,

G. W., … Caro, T. M. (1998). Demography of the Serengeti cheetah (Acinonyx

jubatus) population: the first 25 years. Journal of Zoology, 224(4), 473–488.

Kosterman, M. K., Squires, J. R., Holbrook, J. D., Pletscher, D. H., & Hebblewhite, M.

(2018). Forest structure provides the income for reproductive success in a southern

population of . Ecological Applications, 28(4), 1032–1043.

doi:10.1002/eap.1707

Laake, J. L. (2013). RMark: An R Interface for Analysis of Capture-Recapture Data with

MARK. Version 2.2.4.

LaManna, J. A., & Martin, T. E. (2016). Costs of fear: Behavioural and life-history

responses to risk and their demographic consequences vary across species. Ecology

Letters, 19(4), 403–413. doi:10.1111/ele.12573

60

Laurenson, M. K. (1994). High juvenile mortality in cheetahs (Acinonyx jubatus) and its

consequences for maternal care. Journal of the Zoological Society of London,

234(1), 387–408. doi:10.1111/j.1469-7998.1994.tb04855.x

Lohmann, D., Tietjen, B., Blaum, N., Joubert, D. F., & Jeltsch, F. (2014). Prescribed fire

as a tool for managing shrub encroachment in semi-arid savanna rangelands. Journal

of Arid Environments, 107, 49–56. doi:10.1016/j.jaridenv.2014.04.003

McLoughlin, P. D., Dunford, J. S., & Boutin, S. (2005). Relating predation mortality to

broad-scale habitat selection. Journal of Animal Ecology, 74(4), 701–707.

doi:10.1111/j.1365-2656.2005.00967.x

McLoughlin, P. D., Gaillard, J., Boyce, M. S., Bonenfant, C., Messier, F., Duncan, P., …

Klein, F. (2007). Lifetime reproductive success and composition of the home range

in a large herbivore. Ecology, 88(12), 3192–3201. doi:10.1890/06-1974.1

Michel, M. J., & Adams, M. M. (2009). Differential effects of structural complexity on

predator foraging behavior. Behavioral Ecology, 20(2), 313–317.

doi:10.1093/beheco/arp005

Miller, J. R. B., Ament, J. M., & Schmitz, O. J. (2014). Fear on the move: Predator

hunting mode predicts variation in prey mortality and plasticity in prey spatial

response. Journal of Animal Ecology, 83(1), 214–222. doi:10.1111/1365-

2656.12111

61

Mills, M. G. L., Broomhall, L. S., du Toit, J. T., & Toit, J. T. (2004). Cheetah Acinonyx

jubatus feeding ecology in the Kruger National Park and a comparison across

African savanna habitats: is the cheetah only a successful hunter on open

plains? Wildlife Biology, 10(3), 177–186. doi:10.1016/j.tibtech.2011.06.009

Mills, M. G. L., & Mills, M. E. J. (2014). Cheetah cub survival revisited: a re-evaluation

of the role of predation, especially by lions, and implications for conservation.

Journal of Zoology, 292(2), 136–141. doi:10.1111/jzo.12087

Moll, R. J., Killion, A. K., Montgomery, R. A., Tambling, C. J., & Hayward, M. W.

(2016). Spatial patterns of African ungulate aggregation reveal complex but limited

risk effects from reintroduced carnivores. Ecology, 97(5), 1123–1134.

doi:10.1890/15-0707.1/suppinfo

Moll, R. J., Redilla, K. M., Mudumba, T., Muneza, A. B., Gray, S. M., Abade, L., …

Montgomery, R. A. (2017). The many faces of fear: a synthesis of the

methodological variation in characterizing predation risk. Journal of Animal

Ecology, 86(4), 749–765. doi:10.1111/1365-2656.12680

Mosser, A., Fryxell, J. M., Eberly, L., & Packer, C. (2009). Serengeti real estate: Density

vs. fitness-based indicators of lion habitat quality. Ecology Letters, 12(10), 1050–

1060. doi:10.1111/j.1461-0248.2009.01359.x

Orrock, J. L., Preisser, E. L., Grabowski, J. H., & Trussell, G. C. (2013). The cost of

safety: refuges increase the impact of predation risk in aquatic systems. Ecology,

94(3), 573–579. doi:10.1890/12-0502.1

62

Palomares, F., & Caro, T. M. (1999). Interspecific killing among mammalian carnivores.

The American Naturalist, 153(5), 492–508. doi:10.1086/303189

Preisser, E. L., Bolnick, D. I., & Benard, M. F. (2005). Scared to death? The effects of

intimidation and consumption in predator–prey interactions. Ecology, 86(2), 501–

509.

Prugh, L. R., Sivy, K. J., Mahoney, P. J., Ganz, T. R., Ditmer, M. A., van de Kerk, M., …

Montgomery, R. A. (2019). Designing studies of predation risk for improved

inference in carnivore-ungulate systems. Biological Conservation, 232(October

2018), 194–207. doi:10.1016/j.biocon.2019.02.011

Relyea, R. A. (2001). Morphological and behavioral plasticity of larval anurans in

response to different predators. Ecology, 82(2), 523–540.

Ritchie, E. G., & Johnson, C. N. (2009). Predator interactions, mesopredator release and

biodiversity conservation. Ecology Letters, 12(9), 982–998. doi:10.1111/j.1461-

0248.2009.01347.x

Roques, K. G., O’Connor, T. G., & Watkinson, A. R. (2001). Dynamics of shrub

encroachment in an African savanna: Relative influences of fire, herbivory, rainfall

and density dependence. Journal of Applied Ecology, 38(2), 268–280.

doi:10.1046/j.1365-2664.2001.00567.x

Rostro-García, S., Kamler, J. F., & Hunter, L. T. B. (2015). To kill, stay or flee: the

effects of lions and landscape factors on habitat and kill site selection of cheetahs in

South Africa. Plos One, 10(2), e0117743. doi:10.1371/journal.pone.0117743

63

Schmitz, O. J., Miller, J. R. B., Trainor, A. M., & Abrahms, B. (2017). Toward a

community ecology of landscapes: predicting multiple predator–prey interactions

across geographic space. Ecology, 98(9), 2281–2292. doi:10.1002/ecy.1916

Sheriff, M. J., Krebs, C. J., & Boonstra, R. (2009). The sensitive hare: Sublethal effects

of predator stress on reproduction in snowshoe hares. Journal of Animal Ecology,

78(6), 1249–1258. doi:10.1111/j.1365-2656.2009.01552.x

Suraci, J. P., Clinchy, M., Dill, L. M., Roberts, D., & Zanette, L. Y. (2016). Fear of large

carnivores causes a trophic cascade. Nature Communications, 7, 1–7.

doi:10.1038/ncomms10698

Swanson, A., Arnold, T., Kosmala, M., Forester, J., & Packer, C. (2016). In the absence

of a “landscape of fear”: How lions, hyenas, and cheetahs coexist. Ecology and

Evolution, 6(23), 8534–8545. doi:10.1002/ece3.2569

Tews, J., & Jeltsch, F. (2004). Modelling the impact of climate change on woody plant

population dynamics in South African savanna. BMC Ecology, 4(17), 1–12.

doi:10.1186/1472-6785-4-17

Thaker, M., Vanak, A. T., Owen, C. R., Ogden, M. B., Niemann, S. M., & Slotow, R.

(2011). Minimizing predation risk in a landscape of multiple predators: effects on

the spatial distribution of African ungulates. Ecology, 92(2), 398–407.

doi:10.1890/10-0126.1

Torretta, E., Serafini, M., Puopolo, F., & Schenone, L. (2016). Spatial and temporal

adjustments allowing the coexistence among carnivores in Liguria (N-W Italy). Acta

Ethologica, 19(2), 123–132. doi:10.1007/s10211-015-0231-y

64

Travers, M., Clinchy, M., Zanette, L., Boonstra, R., & Williams, T. D. (2010). Indirect

predator effects on clutch size and the cost of egg production. Ecology Letters,

13(8), 980–988. doi:10.1111/j.1461-0248.2010.01488.x

Tsalyuk, M., Kelly, M., & Getz, W. M. (2017). Improving the prediction of African

savanna vegetation variables using time series of MODIS products. ISPRS Journal

of Photogrammetry and Remote Sensing, 131, 77–91. doi:10.1007/s11065-015-

9294-9.Functional

Valeix, M., Loveridge, A. J., Chamaille-Jammes, S., Davidson, Z., Murindagomo, F.,

Fritz, H., & MacDonald, D. W. (2009). Behavioral adjustments of African

herbivores to predation risk by lions: Spatiotemporal variations influence habitat

use. Ecology, 90(1), 23–30. doi:10.1890/08-0606.1 van Noordwijk, A. J., & de Jong, G. (1986). Acquisition and allocation of resources:

Their influence on variation in life history. The American Naturalist, 128(1), 137–

142.

Vanak, A. T., Fortin, D., Thaker, M., Ogden, M., Owen, C., Greatwood, S., & Slotow, R.

(2013). Moving to stay in place: Behavioral mechanisms for coexistence of African

large carnivores. Ecology, 94(11), 2619–2631. doi:10.1890/13-0217.1

Wilson, A. M., Lowe, J. C., Roskilly, K., Hudson, P. E., Golabek, K. A., & McNutt, J.

W. (2013). Locomotion dynamics of hunting in wild cheetahs. Nature, 498(7453),

185–189. doi:10.1038/nature12295

65

TABLES

Table 2.1. Model selection results for multi-state joint live-encounter dead-recovery spatial-explicit survival models for cheetahs with seasonal spatial covariates, Mun-Ya- Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. States in the model include cubs (juveniles dependent on their mothers) and adults (non-juveniles). All models include effects of year on recovery rates and season on survival rates. a b c Model AICc ΔAICC -2×ln(L) w k

S(state:EVI) 3601.15 0 3583.02 0.55 9

S(state:lion + state:EVI) 3603.51 2.36 3581.32 0.17 11

S(state:lion * state:EVI) 3604.06 2.91 3577.80 0.13 13

S(state:prey + state:EVI) 3604.60 3.45 3582.41 0.10 11

S(state:EVI + state:lion + state:prey) 3607.08 5.93 3580.82 0.03 13

S(state:prey * state:EVI) 3607.31 6.16 3581.05 0.03 13

S(state:lion * state:prey) 3613.35 12.20 3587.09 0.00 13

S(state:prey) 3617.96 16.81 3599.84 0.00 9

S(state) 3619.33 18.19 3603.23 0.00 8

S(state:lion) 3620.62 19.47 3602.49 0.00 9

S(state:lion+state:prey) 3621.51 20.36 3599.32 0.00 11

a Log-likelihood b Akaike model weight c Number of model parameters

66

Table 2.2. Model selection results for multi-state joint live-encounter dead-recovery spatial-explicit survival models for cheetahs with spatial covariates averaged across individual cheetahs’ lifetimes, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. States in the model include cubs (juveniles dependent on their mothers) and adults (non-juveniles). All models include effects of year on recovery rates and season on survival rates. a b c Model AICc ΔAICC -2×ln(L) w k

S(state:lion + state:EVI) 3608.85 0 3586.66 0.30 11

S(state:EVI) 3609.06 0.21 3590.93 0.27 9

S(state:lion) 3610.45 1.60 3592.32 0.13 9

S(state:prey * state:EVI) 3611.52 2.67 3585.26 0.08 13

S(state:lion * state:EVI) 3611.82 2.97 3585.56 0.07 13

S(state:EVI + state:lion + state:prey) 3612.36 3.51 3586.10 0.05 13

S(state:prey + state:EVI) 3612.44 3.59 3590.25 0.05 11

S(state:lion * state:prey) 3613.40 4.55 3587.14 0.03 13

S(state:lion + state:prey) 3614.41 5.56 3592.22 0.02 11

S(state:prey) 3618.74 9.89 3600.61 0.00 9

S(state) 3619.33 10.49 3603.23 0.00 8

a Log-likelihood b Akaike model weight c Number of model parameters

67

FIGURES

Figure 2.1. Monthly survival of adult and cheetah cubs in relation to short term (seasonal) average Enhanced Vegetation Index (EVI) within an individual’s home range, Mun-ya- wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI.

68

Figure 2.2. Monthly survival of adult and cheetah cubs in relation to a) long-term (lifetime) average Enhanced Vegetation Index (EVI) within an individual’s home range while holding lion density constant at an average value, and b) long-term (lifetime) average probability of encountering a lion within an individual’s home range while holding EVI constant at an average value, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI.

69

Figure 2.3. Relative probability of cheetah kill site occurrence in relation to Enhanced Vegetation Index (EVI), Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. Shaded regions represent 85% CI.

70

CHAPTER 3

SHORT-TERM PREDATION RISK AND HABITAT COMPLEXITY INFLUENCE CHEETAH ANTI-PREDATOR BEHAVIORS

ABSTRACT

The risk of predation by top predators can influence the behavior of species in lower trophic levels. Animals can reduce predation risk by increasing anti-predator behaviors in areas of long-term risk (the risky places hypothesis), exhibit more anti- predator behaviors when exposed to an immediate risk (the risky times hypothesis), or vary responses to short-term risk based on long-term risk (the risky times and risky places hypothesis or the predation risk allocation hypothesis). Most research on responses to spatial and temporal variation in predation risk has come from systems with herbivore prey and single predators, whereas very little research has been conducted on mesopredator behavioral responses to predation risk. We studied anti-predator behaviors of cheetahs (Acinonxy jubatus) exposed to long-term and short-term predation risk of lions (Panthera leo) and leopards (Panthera pardus) to better understand risk perception of mesopredators by using a playback experiment to manipulate short-term predation risk in areas of differing long-term predation risk. We found that cheetah anti-predator behavior varied based on spatial and temporal factors. Specifically, we did not find support for the risky places hypothesis; cheetah vigilance was not influenced by long- term risk. On the contrary, we found support for the risky times hypothesis; cheetah were more vigilant and more likely to fleeing following lion and leopard playbacks.

Additionally, we did not find support for the risky times and risky places hypothesis or

71 the predation risk allocation hypothesis; cheetah vigilance or probability of fleeing following a predator playback were not associated with long-term predation risk. Finally, we found that habitat complexity interacts with both long-term and short-term risk to affect cheetah anti-predator behaviors. Cheetahs had higher baseline vigilance in areas of open vegetation, whereas cheetahs were more likely to flee from lion sounds in areas of dense vegetation and from leopards in areas of open vegetation. Collectively, we found that cheetah vary behaviors based on short-term risk and habitat characteristics, but that long-term risk does not influence cheetah anti-predator behaviors. Our results offer insight into risk perception by mesopredators and highlight that mesopredator responses to predation risk can differ from prey responses to risk.

INTRODUCTION

Predators can influence lower trophic levels through direct predation as well as indirectly through changes in behavior (Creel and Christianson 2008). These predator- induced changes in behavior can scale-up and affect population demography (LaManna and Martin 2016), and structure entire ecosystems (Ford et al. 2014, Suraci et al. 2016).

As predator species worldwide are undergoing population declines and range contractions

(Ripple et al. 2014, Wolf and Ripple 2017), it is important to understand how predators influence the demography and behavior of lower trophic levels, through either direct predation or non-consumptive effects.

Animals under the risk of predation can reduce predation risk through behaviors such as minimizing the amount of time in risky areas (Thaker et al. 2011), increasing

72 vigilance or alarm behavior (Lima 1987), or fleeing from risky situations (Cooper 2003,

Stankowich and Coss 2007). These anti-predator behaviors can be beneficial for avoiding predation, but can also come at energetic or fitness costs because of trade-offs with foraging or parental care (Lima 1987). Therefore, animals need to accurately assess predation risk in order to maximize the benefits of anti-predator behaviors.

Perception of predation risk and associated anti-predator behaviors can depend on a variety of factors. Animals might exhibit more anti-predator behaviors in areas of high long-term predation risk (i.e. the “risky places” hypothesis; Creel et al., 2008).

Alternatively, animals might exhibit more anti-predator behaviors when exposed to an immediate risk (i.e. the “risky times” hypothesis; Creel et al., 2008). Additionally, the long-term risk of predation can affect how animals respond to short-term predation risk, either by exhibiting higher responses to short-term risk in areas of high-long term risk

(i.e. the "risky times and risky places" hypothesis; Dröge et al. 2017), or by exhibiting the highest response to predation risk when predation risk is variable, compared to when predation risk is either uniformly high or low (predation risk allocation hypothesis; Lima

& Bednekoff, 1999). In addition, environmental characteristics such as habitat complexity can influence the responses of prey to predation risk. Although dense vegetation might serve as a refuge for prey species, it might also reduce visibility, which in turn could increase predation risk (Camp et al. 2012, McCormick and Löonnstedt

2013).

Most of our understanding of anti-predator behaviors and risk perception comes from studies of herbivore prey species (Dröge et al. 2017). Considerably less research has

73 been conducted on anti-predator behaviors and risk perception of mesopredators.

Mesopredators need to avoid the risk of predation by top predators while still finding and catching prey. However, dissimilar to prey species, carnivorous mesopredators might not often experience trade-offs between anti-predator behavior and foraging because they hunt relatively infrequently compared to foraging herbivores (Jeschke 2007).

Additionally, because apex predators typically prey on herbivores rather than mesopredators (Palomares and Caro 1999), the behavioral responses of mesopredators to long-term or short-term predation risk might differ from responses of herbivore prey.

Therefore, a greater understanding on how mesopredators perceive and respond to variation in predation risk is needed.

The large carnivore guild of South Africa offers an ideal system to investigate risk perception in mesopredators. Cheetahs (Acinonyx jubatus) are subordinate to lions

(Panthera leo) and leopard (Panthera pardus), and the majority of cheetah mortality is from lions and leopards (Laurenson 1994). Previous research has shown that cheetahs respond to short-term risk from apex predators, through behavioral adjustments (Durant

2000), and respond to long-term risk by altering fine-scale habitat use patterns (Swanson et al. 2014). However, little is known about how cheetahs respond behaviorally to both long-term and short-term risk simultaneously. Additionally, recent research has shown that cheetah survival is driven by both long-term and short-term habitat use in relation to vegetation complexity and predation risk (Chapter 2). Therefore, spatial variation in anti- predator behaviors might influence spatial patterns of demography.

74

Here we use long-term data from an intensely studied cheetah population in a study area with known spatial variation in long-term predation risk, and manipulated short-term predation risk to investigate how mesopredators such as cheetahs perceive predation risk. Specifically, we tested three main hypotheses related to risk perception:

1) Risky places hypothesis: Cheetahs perceive predation risk of apex predators

using long-term predator cues

2) Risky times hypothesis: Cheetahs perceive predation risk of apex predators

using short-term predator cues

3) Risky times and risky places hypothesis or predation risk allocation

hypothesis: Cheetahs perceive risk of predation by apex predators using a

combination of long-term and short-term predator cues

Additionally, we investigated how the effects of sex and habitat complexity interact with long-term and short-term predation risk to influence cheetah anti-predator behaviors. By identifying factors that influence anti-predator behaviors in cheetahs, we can obtain a greater understanding of how mesopredators perceive predation risk, which in turn can be used to help understand the role of apex carnivores in ecological systems.

METHODS

Study Area

We studied cheetah perception of predation risk in Mun-Ya-Wana Conservancy

(Phinda Private Game Reserve), in KwaZulu-Natal, South Africa. The Conservancy is

285 km2 and is surrounded by electrified game fencing. The dominant vegetation type is

75 broad-leaf woodland, with open grasslands and semi-open wooded-grasslands interspersed throughout the reserve. The elevation of the Conservancy ranges from 4 to

201 meters above sea level. The climate is subtropical with warm, dry winters (April –

September) and hot, humid summers (October – March), with the majority of rain falling in the summer. Cheetahs and lions were reintroduced into the reserve in 1992 and have been monitored since (Hunter, 1998; Hunter et al., 2007). Leopards have historically occurred in the study area and currently exist at a stable population density (Balme et al.

2010). Although spotted hyenas (Crocuta crocuta) also occur at low densities in the study area and compete with cheetahs for resources, they rarely kill cheetahs in our study population (Chapter 1), so we did not consider them in our study.

Playback Sounds

We tested hypotheses related to mesopredator perception of predation risk by conducting a playback experiment. We used two predator sounds (lion and leopard) and a control sound (African hoopoe, Upupa africana). We selected the hoopoe as our control because they are active throughout the day, are found across the entirety of our study area, are neither potential predators nor prey for cheetahs, and have similar sound qualities to lions and leopards. We acquired sound clips from online audio databases and edited them to 10 second exemplars. For lion and leopard exemplars, we used contact call vocalizations of single animals. We ensured that all exemplars across all treatments and controls were similar in regards to sound frequency and amplitude (Blumstein et al.

2008). All playbacks were conducted using a portable speaker (Klipsch Groove,

76

Indianapolis, IN, USA) at a volume of approximately 100 dB, which is within the range of naturally occurring lion and leopard sounds (Grinnell and McComb 2001, Schel and

Zuberbühler 2009).

Data Collection

We collected data from February 2018 to September 2019. Data collection was approved by Clemson University IACUC (no. AUP2017-077). When we encountered a non-sleeping adult cheetah, we recorded the GPS location of the animal, the number of individuals (including cubs) present, the distance of the vehicle to the animal, the weather, the wind speed (categorized into high, medium, and low categories), and identified any other animals or vehicles within the area.

If multiple adult cheetahs were at the same sighting, we randomly selected one to be the focal animal. In addition, we randomly selected a treatment and exemplar for each playback trial. Using a handheld video camera (Sony HDRCX405 Handycam, Tokyo,

Japan) we recorded the baseline behavior of the focal animal for five minutes. We then played a 10 second exemplar and recorded an additional five minutes of video post- playback. To ensure that cheetahs were not habituated or negatively affected by the playbacks, we only conducted trials on individual cheetahs a maximum of two times per week and we did not conduct playbacks on successive days. We also did not collect data on cheetah cubs, or cheetahs that were at kill sites.

77

Environmental Covariates

For each playback location we created a 50-meter buffer and extracted environmental covariates. We selected 50 m as our buffer distance because cheetahs are likely able to perceive conditions within this distance, as evidenced by their ability to locate prey 50 m away (Eaton 1970). We estimated lion predation risk by creating seasonal utilization distributions of lion sightings data (see Chapter 2 for full methods).

We estimated leopard predation risk by collecting sightings data from trained guides from

2018-2019 and creating a utilization distribution following the same methods as we used to quantify lion risk. We considered both lion and leopard risk separately, and also added utilization distributions of lions and leopards to create a composite of total predation risk.

Finally, we used Enhanced Vegetation Index (EVI) as an index for vegetation complexity because it is correlated with vegetation structure in Africa, with open areas having low

EVI values, and areas with dense vegetation having high EVI values (Tsalyuk et al.

2017).

Video Analysis

We analyzed videos using the behavioral analysis program Jwatcher (Blumstein and Daniel 2007). Following Caro (1994), we defined cheetah vigilance behavior as time spent visually scanning the surrounding area. To this end, we considered lying alert, sitting, and standing body positions as vigilance (Caro 1994). For each trial, we calculated the proportion of time vigilant pre- and post-playback. We also recorded if the

78 focal cheetah fled the playback location, which we defined as moving out of sight, or greater than 100 m.

Data Analysis

To ensure that additional factors were not influencing responses to playback sounds, we first ran analyses to assess the effects of wind, distance to focal animal, and time of day on vigilance, both pre- and post-playback. Because our response variables were bounded between zero and one, we ran beta regression models with each potential confounding covariate in Program R (R Core Team 2019) using the betareg package

(Cribari-Neto and Zeileis 2010) with pre-playback vigilance or post-playback vigilance as the response variable. For these analyses, and all subsequent analyses, we assessed significance using an α value of 0.05.

Risky Places Hypothesis

To assess how long-term predation risk influences anti-predator behavior, we analyzed baseline vigilance behavior before any playback sounds were played. Therefore, any behavioral responses would have been a result of background or long-term spatial covariates. For these analyses, we included data from all trials, regardless of the sound that was eventually played. We used beta regression models to compare the proportion of time vigilant with the long-term risk of predation. We ran models with lion risk and leopard risk separately, as well as a model with the cumulative risk of lions and leopards combined. We also used beta regression models to determine if the proportion of baseline

79 time vigilant was associated with EVI. Because males and females might have different responses to long-term risk, we included sex as a factor in all models.

Risky Times Hypothesis

To assess how short-term predation risk influences anti-predator behavior, we compared the proportion of time vigilant pre-playback to the proportion of time vigilant post-playback using beta regression. We also used logistic regression to determine the influence of playbacks on the probability of fleeing. We tested for differences in short term perception of predation risk by sex, using only data from lion and leopard trials.

Because of a low sample size of female cheetahs without cubs, we only included males and females with cubs in our analysis including sex.

Risky Times and Risky Places Hypothesis or Predation Risk Allocation Hypothesis

We assessed how long-term predation right might influence perception of short- term predation risk. Using only data from lion and leopard trials, we used beta regression models to compare the proportion of time vigilant post-playback and long-term predation risk. In addition to testing for a linear relationship between long-term predation risk and proportion of time vigilant post-playback, we assessed a quadratic model, which would be consistent with the predictions of the predation risk allocation hypothesis. Finally, we tested the hypothesis that vegetation complexity would influence short-term responses to risky situations by analyzing the influence on EVI on post-playback vigilance using beta regression and the probability of fleeing using logistic regression.

80

RESULTS

We collected data on 16 individual cheetahs for a total of 78 trials. The trials consisted of 19 hoopoe playbacks, 21 leopard playbacks, 23 lion playbacks, and 15 trials where we only collected pre-playback data. There was no relationship between pre- playback vigilance and wind speed (plow wind-high wind = 0.82; pmedium wind-high wind = 0.26; plow wind-medium wind = 0.35), distance from observer to cheetah (p = 0.98), or time of day (p =

0.64), and post-playback vigilance and wind speed (plow wind-high wind = 0.82; pmedium wind-high wind = 0.28; plow wind-medium wind = 0.18), distance from observer to cheetah (p = 0.76), or time of day (p = 0.77); therefore we were able to include data from all trials in our subsequent analyses.

Risky Places Hypothesis

The long-term predation risk of lions and leopards did not influence baseline levels of cheetah vigilance regardless of whether we considered only the risk of lions (p =

0.95), only the risk of leopards (p = 0.93), or the cumulative risk of lions and leopards (p

= 0.88). Habitat complexity and predation risk were not strongly correlated (|r| = 0.41).

Habitat complexity influenced baseline levels cheetah vigilance for female cheetahs with cubs (p = 0.04), with female cheetahs exhibiting higher proportions of time vigilant in open areas (Figure 2.1), although this same relationship was not significant for males

(p=0.08). There was not a significant interaction between cumulative risk of lion and leopards and habitat complexity on vigilance for all cheetah combined (p = 0.80), or when cheetahs were differentiated by sex (p = 0.77).

81

Risky Times Hypothesis

Across all individuals, cheetahs increased vigilance following lion (p=0.006) and leopard (p<0.001) playbacks, compared to hoopoe playbacks (Figure 2.2), but post- playback vigilance behavior did not differ among lion and leopard trials (p = 0.26).

Similarly, cheetahs were more likely to flee following lion (p = 0.04) and leopard (p =

0.006) playbacks, compared to hoopoe playbacks (Figure 2.3), but probability of fleeing did not differ among lion and leopard trials (p = 0.23).

When comparing post-playback vigilance behavior by sex, female cheetahs with cubs were more vigilant following lion and leopard playbacks than male cheetahs (p =

0.03), with females with cubs spending 99.9% (SE = 0.07) of time vigilant following a predator playback and males spending 73.8% (SE = 7.5) of time vigilant following a predator playback (Figure 2.4). There was no difference in the probability of fleeing following lion and leopard playbacks between males and females (p = 0.26).

Risky Times and Places Hypothesis or Predation Risk Allocation Hypothesis

Long-term predation risk did not influence post-playback vigilance in a linear fashion as predicted by the risky times and places hypothesis, regardless of if we considered lion and leopard risk combined (p = 0.45), lion risk (p = 0.98) or leopard risk

(p = 0.35). Similarly, long-term predation risk did not influence post-playback vigilance in a quadratic fashion as predicted by the predation risk allocation hypothesis, regardless of if we considered lion and leopard risk combined (p = 0.60), lion risk (p = 0.82), or leopard risk (p = 0.86).

82

Long-term predation risk also did not influence the probability of fleeing post- playback, in a linear fashion as predicted by the risky times and places hypothesis, regardless of if we considered lion and leopard risk combined (p = 0.27), lion risk (p =

0.19), or leopard risk (p = 0.70). Similarly, long-term predation risk did not influence the probability of fleeing post-playback in a quadratic fashion as predicted by the predation risk allocation hypothesis, regardless of if we considered lion and leopard risk combined

(p = 0.97), lion risk (p = 0.70), or leopard risk (p = 0.51).

There was no relationship between habitat complexity and the proportion of time that cheetahs were vigilant following lion and leopard playbacks (p = 0.66). However, cheetahs (both male and female) were more likely to flee following a lion sound in areas of higher EVI (p = 0.03), and following a leopard sound in areas of lower EVI (p = 0.03;

Figure 2.4).

DISCUSSION

We found that cheetah anti-predator behavior varied based on spatial and temporal factors. Our results build on previous research on cheetah anti-predator behaviors which found that cheetahs respond to short-term risk of lions and hyenas

(Durant 2000), to reveal that long-term predation risk of lions and leopards does not influence cheetah anti-predator behaviors, but that habitat complexity does play a role in cheetah risk perception. Not only do our results offer insight into drivers of cheetah anti- predator behaviors, but they also indicate that anti-predator behaviors may differ across trophic levels. In particular, differences between our results and studies of risk perception

83 in herbivore prey highlights the need to consider how top-down effects on behavior might vary among multiple trophic levels within a system.

We found that long-term predation risk did not influence cheetah vigilance, indicating that the risky places hypothesis does not always adequately describe anti- predator behaviors. There are three possible explanations as to why we did not find support for the risky places hypothesis. First, differences in the ecology of herbivore prey and mesopredators could explain our observed lack of support for this hypothesis.

Several studies of prey vigilance in relation to long-term predator risk in a variety of systems have found higher anti-predator behaviors in areas of high long-term predation risk (Lung and Childress 2007, Valeix et al. 2009, Donadio and Buskirk 2016). Unlike prey (Lima 1987), there likely is not a strong vigilance-foraging trade-off for mesopredators that do not forage continuously, so they can exhibit some level of anti- predator behavior regardless of the background level of risk without incurring negative energetic consequences. Indeed, we found that most cheetahs exhibited some baseline vigilance behavior regardless of background levels of risk.

Second, differences between prey and mesopredators ecology in relation to lethal encounters with apex predators might lead to our results showing that long-term predation risk does not influence cheetah anti-predator behaviors. Even in areas of high long-term predation risk, encounters between apex predators and mesopredators do not always result in mortality (Linnell and Strand 2000). For example, wolves (Canis lupus) often encounter coyotes (Canis latrans) at kill sites, but the majority of these encounter do not end in coyote mortality (Merkle et al. 2009). In our system, there is typically a

84 large amount of spatial overlap between cheetahs, lions, and leopards (Swanson et al.

2014) and cheetahs are adept at fleeing from apex predators (Hunter et al. 2007).

Additionally, cheetahs use fine-scale spatial adjustments to avoid encounters with lions and leopards (Broekhuis et al. 2013, Vanak et al. 2013). Therefore, cheetah responses to predation risk are likely more reactive than predictive and thus increased anti-predator behavior in areas of high risk does not provide a large benefit.

Third, the perception of long-term risk requires the ability to assess predation risk

(Weissburg et al. 2014), which necessitates reliable cues of risk as well as the cognitive ability of the prey or mesopredator to associate these cues with risk (Gaynor et al. 2019).

We used robust datasets of lion and leopard locations, but other proxies for risk such as predator kills sites or predator scat and markings might have been stronger cues for cheetahs to assess long-term risk (Cornhill and Kerley 2020). Similarly, lions usually vocalize daily as a means of long-distance communication (Grinnell and McComb 2001) and therefore cheetahs may have been using actual audio cues to assess where lions actually were on any given day.

Similar to previous research in the Serengeti (Durant 2000), we found that cheetahs exhibited strong anti-predator behaviors to short-term risk. When faced with an immediate threat, the benefits of anti-predator behavior likely outweigh any potential costs (Lima and Dill 1990). The majority of cheetah mortality, especially for cubs, is a result of predation by lions and leopards (Laurenson 1994); thus it make sense that cheetahs would respond strongly to lion and leopard sounds. Because cubs are particularly vulnerable to predation, our finding that female cheetahs with cubs exhibited

85 higher vigilance than males in response to short-term risk make sense. Similar to our results, previous research on cheetah behavior at kill sites in the Serengeti found that female cheetahs with cubs spent more time vigilant than feeding (Hilborn et al. 2018).

Increased anti-predator behaviors by females with offspring has been observed in a variety of herbivore prey (Burger and Gochfeld 1994, Toïgo 1999), likely as a means of females prioritizing offspring safety to maximize lifetime reproductive success. Thus, the effects of parental status on anti-predator behaviors seems to hold for both herbivore prey and mesopredators.

As opposed to research on herbivore prey, we did not find support to suggest that long-term predation risk affects cheetah responses to short-term risk. Herbivores such as elk have been found to alter vigilance based on exposure to predators consistent with the predation risk allocation hypothesis; elk were more vigilant in areas with intermittent exposure to wolves, compared to areas with high wolf exposure (Creel et al. 2008).

Recent work in Namibia has also shown that ungulates exhibit the highest level of anti- predator behaviors when they are exposed to short-term risk in areas of high long-term risk (Dröge et al. 2017). Regardless of the relationship between long-term and short-term risk, herbivore prey seem to modify behaviors based on a combination of long-term and short-term risk, whereas we did not find this relationship for cheetahs. Similar to our results related to the risky places hypothesis, our failure to find support for either the risky times and places hypothesis or the predation risk allocation hypothesis could be related to the inability of cheetahs to assess background risk, or could be related to ecological differences between herbivore prey and mesopredators, particularly the lack of

86 vigilance-foraging tradeoff for cheetahs. Overestimation of risk is more beneficial than underestimating risk (Bouskila and Blumstein 1992), therefore responding to playbacks regardless of background risk level would have likely been the most profitable for cheetahs, given that there likely wasn’t a large energy tradeoff for these anti-predator behaviors.

Habitat complexity has been found to influence habitat use and demography of mesopredators (Janssen et al. 2007, Chapter 2) and our results indicate that habitat complexity can also affect mesopredator behavior. Instead of responding to long-term risk, we found that cheetahs modified baseline vigilance behavior based on habitat characteristics; females spent the highest proportion of time vigilant in open areas. Other species such as warthogs (Phacochoerus africanus) and wildebeest (Connochaetes taurinus) exhibit higher vigilance in open areas, likely because of the risk of being spotted by nearby predators and the higher probability of detecting nearby predators

(Scheel 1993). Female cheetahs also likely follow this pattern because they are potentially more vulnerable in open habitats. Because lions and leopards do not typically target cheetahs as a source of food (Laurenson 1994), the probability of an attack on a cheetah likely only depends on a lion or leopard spotting a cheetah and having the time or energy to attack them. Therefore, in areas where apex predators can see cheetahs from a further distance, cheetahs benefit from increased vigilance because it allows them a greater change of detecting a predator before they are detected. The higher amounts of vigilance in open areas might scale-up and have population-level consequences. Both adult cheetahs and cubs experience higher survival in open areas (Chapter 2), which

87 might be caused in part by higher anti-predator behavior such as vigilance to offset the risk of predation.

Additionally, we found that habitat complexity affected how cheetahs differentially expressed flee responses to leopards compared to lions. Compared to vigilance, fleeing represents a greater cost because of higher energetic demands and leaving a potentially profitable area (Ydenberd and Dill 1986). Thus, animals should only flee when they experience the highest cost to their fitness. Lions are ambush predators that benefit from concealment in dense vegetation (Davies et al. 2016). Therefore, our results suggest it is likely that when cheetahs hear lion sounds, but cannot see them, the risk of staying is greater than the costs of leaving. Although leopards are also ambush predators (Balme et al. 2007), we found the opposite relationship between vegetation complexity and fleeing. However, as opposed to lions, leopards typically avoid open habitats (Balme et al. 2007). Therefore, cheetahs’ higher probability of fleeing from leopard sounds in open habitats could be a response to a novel predatory threat, given that they associate leopard sounds with predation risk, but they are hearing the sound in an area that they don’t associated with leopards. Several recent studies have demonstrated the need to consider the effects of multiple predators on prey behavior because of variation in responses to different predators, which do not always follow predicted relationships (Thaker et al. 2011, Creel et al. 2017). Because we found variation in mesopredator responses to different predators, our results suggest the important of also considering multi-predator effects on mesopredator species as well, rather than

88 attempting to simplify the complexity of multiple predator effects (Montgomery et al.

2019).

Collectively, our results offer insight into how mesopredators might perceive predation risk, and how risk and anti-predator behaviors are related to spatial and temporal factors. Many studies use vigilance or other anti-predator behaviors as a proxy for predation risk (Moll et al. 2017), but our research adds to the growing body of literature to suggest that risk and risk effects need to be considered across multiple spatial and temporal scales because behavioral responses do not always match the underlying risk of predation and habitat complexity can influence these responses (Moll et al. 2017,

Gaynor et al. 2019). Additionally, our study is one of the most comprehensive studies of risk perception in mesopredators. We show that the behavioral responses of mesopredators to short-term and long-term predation risk sometimes differ than what is observed in prey species, potentially because of differences in tradeoffs between energy and anti-predator behaviors. Therefore, understanding differences in non-consumptive effects among trophic levels, rather than assuming that top predators affect all trophic levels similarly, is critical.

REFERENCES

Balme, G. A., Slotow, R., & Hunter, L. T. B. (2010). Edge effects and the impact of non-

protected areas in carnivore conservation: Leopards in the Phinda-Mkhuze

Complex, South Africa. Animal Conservation, 13(3), 315–323.

doi:10.1111/j.1469-1795.2009.00342.x

89

Balme, G., Hunter, L. T. B., & Slotow, R. (2007). Feeding habitat selection by hunting

leopards Panthera pardus in a woodland savanna: prey catchability versus

abundance. Animal Behaviour, 74(3), 589–598.

doi:10.1016/j.anbehav.2006.12.014

Blumstein, D. T., Cooley, L., Winternitz, J., & Daniel, J. C. (2008). Do yellow-bellied

marmots respond to predator vocalizations? Behavioral Ecology and

Sociobiology, 62(3), 457–468. doi:10.1007/s00265-007-0473-4

Blumstein, D. T., & Daniel, J. C. (2007). Quantifying behavior the JWatcher way.

Sunderland, MA: Sinaeur Associates Inc.

Bouskila, A., & Blumstein, D. T. (1992). Rules of thumb for predation hazard

assessment: Predictions from a dynamic model. American Naturalist, 139(1),

161–176.

Broekhuis, F., Cozzi, G., Valeix, M., Mcnutt, J. W., & Macdonald, D. W. (2013). Risk

avoidance in sympatric large carnivores: Reactive or predictive? Journal of

Animal Ecology, 82(5), 1098–1105. doi:10.1111/1365-2656.12077

Burger, J., & Gochfeld, M. (1994). Vigilance in African : Differences among

mothers, other females, and males. Behaviour, 131(3), 153–169.

Camp, M. J., Rachlow, J. L., Woods, B. A., Johnson, T. R., & Shipley, L. A. (2012).

When to run and when to hide: The influence of concealment, visibility, and

proximity to refugia on perceptions of risk. Ethology, 118(10), 1010–1017.

doi:10.1111/eth.12000

90

Caro, T. M. (1994). Cheetahs of the Serengeti Plains. Chicago: The University of

Chicago Press.

Cooper, W. E. (2003). Effect of risk on aspects of escape behavior by a lizard,

Holbrookia propinqua, in relation to optimal escape theory. Ethology, 109(8),

617–626. doi:10.1046/j.1439-0310.2003.00912.x

Cornhill, K. L., & Kerley, G. I. H. (2020). Cheetah communication at scent-marking sites

can be inhibited or delayed by predators. Behavioral Ecology and Sociobiology,

74(2). doi:10.1007/s00265-020-2802-9

Creel, S., & Christianson, D. (2008). Relationships between direct predation and risk

effects. Trends in Ecology and Evolution, 23(4), 194–201.

doi:10.1016/j.tree.2007.12.004

Creel, S., Dröge, E., M’soka, J., Smit, D., Becker, M., Christianson, D., & Schuette, P.

(2017). The relationship between direct predation and antipredator responses: A

test with multiple predators and multiple prey. Ecology, 98(8), 2081–2092.

doi:10.1002/ecy.1885

Creel, S., Winnie, J. A., Christianson, D., & Liley, S. (2008). Time and space in general

models of antipredator response: tests with wolves and elk. Animal Behaviour,

76(4), 1139–1146. doi:10.1016/j.anbehav.2008.07.006

Cribari-Neto, F., & Zeileis, A. (2010). Beta Regression in R. Journal of Statistical

Software, 34(2), 1–24. doi:10.18637/jss.v069.i12

91

Davies, A. B., Tambling, C. J., Kerley, G. I. H., & Asner, G. P. (2016). Effects of

vegetation structure on the location of lion kill sites in African thicket. PLoS

ONE, 11(2), 1–20. doi:10.1371/journal.pone.0149098

Donadio, E., & Buskirk, S. W. (2016). Linking predation risk, ungulate antipredator

responses, and patterns of vegetation in the high Andes. Journal of Mammalogy,

97(3), 966–977. doi:10.1093/jmammal/gyw020

Dröge, E., Creel, S., Becker, M. S., & M’soka, J. (2017). Risky times and risky places

interact to affect prey behaviour. Nature Ecology & Evolution, 1(8), 1123–1128.

doi:10.1038/s41559-017-0220-9

Durant, S. M. (2000). Living with the enemy: avoidance of hyenas and lions by cheetahs

in the Serengeti. Behavioral Ecology, 11(6), 624–32.

doi:10.1093/beheco/11.6.624

Eaton, R. L. (1970). Hunting behavior of the cheetah. Journal of Wildlife Management,

34(1), 56–67.

Ford, A. T., Goheen, J. R., Otieno, T. O., Bidner, L., Isbell, L. A., Palmer, T. M., …

Pringle, R. M. (2014). Large carnivores make savanna tree communities less

thorny. Science, 346(6207), 346–349. doi:10.1126/science.1252753

Gaynor, K. M., Brown, J. S., Middleton, A. D., Power, M. E., & Brashares, J. S. (2019).

Landscapes of fear: Spatial patterns of risk perception and response. Trends in

Ecology and Evolution, 34(4), 355–368. doi:10.1016/j.tree.2019.01.004

92

Grinnell, J., & McComb, K. (2001). Roaring and social communication in african lions:

The limitations imposed by listeners. Animal Behaviour, 62(1), 93–98.

doi:10.1006/anbe.2001.1735

Hilborn, A., Pettorelli, N., Caro, T., Kelly, M. J., Laurenson, M. K., & Durant, S. M.

(2018). Cheetahs modify their prey handling behavior depending on risks from

top predators. Behavioral Ecology and Sociobiology.

Hunter, J. S., Durant, S. M., & Caro, T. M. (2007). To flee or not to flee: Predator

avoidance by cheetahs at kills. Behavioral Ecology and Sociobiology, 61, 1033–

1042. doi:10.1007/s00265-006-0336-4

Janssen, A., Sabelis, M. W., Magalhães, S., & Van, T. (2007). Habitat structure affects

intraguild predation. Ecology, 88(11), 2713–2719.

Jeschke, J. M. (2007). When carnivores are “full and lazy.” Oecologia, 152(2), 357–364.

doi:10.1007/s00442-006-0654-2

LaManna, J. A., & Martin, T. E. (2016). Costs of fear: Behavioural and life-history

responses to risk and their demographic consequences vary across species.

Ecology Letters, 19(4), 403–413. doi:10.1111/ele.12573

Laurenson, M. K. (1994). High juvenile mortality in cheetahs (Acinonyx jubatus) and its

consequences for maternal care. Journal of the Zoological Society of London,

234(1), 387–408. doi:10.1111/j.1469-7998.1994.tb04855.x

Lima, S. L. (1987). Vigilance while feeding and its relation to the risk of predation.

Journal of Theoretical Biology, 124(3), 303–316. doi:10.1016/S0022-

5193(87)80118-2

93

Lima, S. L., & Bednekoff, P. A. (1999). Temporal variation in danger drives antipredator

behavior: The predation risk allocation hypothesis. The American Naturalist,

153(6), 649–659. doi:10.1086/303202

Lima, S. L., & Dill, L. M. (1990). Behavioral decisions made under the risk of predation:

a review and prospectus. Canadian Journal of Zoology, 68(4), 619–640.

doi:10.1139/z90-092

Linnell, J. D. C., & Strand, O. (2000). Interference interactions,co-existence and

conservation of mammalian carnivores. Biodiversity Research, 6, 169–176.

doi:10.1046/j.1472-4642.2000.00069.x

Lung, M. A., & Childress, M. J. (2007). The influence of conspecifics and predation risk

on the vigilance of elk (Cervus elaphus) in Yellowstone National Park.

Behavioral Ecology, 18(1), 12–20. doi:10.1093/beheco/arl066

McCormick, M. I., & Löonnstedt, O. M. (2013). Degrading habitats and the effect of

topographic complexity on risk assessment. Ecology and Evolution, 3(12), 4221–

4229. doi:10.1002/ece3.793

Merkle, J. A., Stahler, D. R., & Smith, D. W. (2009). Interference competition between

gray wolves and coyotes in Yellowstone National Park. Canadian Journal of

Zoology, 87(1), 56–63. doi:10.1139/Z08-136

Moll, R. J., Redilla, K. M., Mudumba, T., Muneza, A. B., Gray, S. M., Abade, L., …

Montgomery, R. A. (2017). The many faces of fear: a synthesis of the

methodological variation in characterizing predation risk. Journal of Animal

Ecology, 86(4), 749–765. doi:10.1111/1365-2656.12680

94

Montgomery, R. A., Moll, R. J., Say-Sallaz, E., Valeix, M., & Prugh, L. R. (2019). A

tendency to simplify complex systems. Biological Conservation, 233, 1–11.

doi:10.1016/j.biocon.2019.02.001

Palomares, F., & Caro, T. M. (1999). Interspecific killing among mammalian carnivores.

The American Naturalist, 153(5), 492–508. doi:10.1086/303189

R Core Team. (2019). R: A language and environment for statistical computing. Vienna,

Austria: R Foundation for Statistical Computing.

Ripple, W. J., Estes, J. A., Beschta, R. L., Wilmers, C. C., Ritchie, E. G., Hebblewhite,

M., … Wirsing, A. J. (2014). Status and ecological effects of the world’s largest

carnivores. Science, 343(6167), 1241484. doi:10.1126/science.1241484

Scheel, D. (1993). Watching for lions in the grass: the usefulness of scanning and its

effects during hunts. Animal Behavior, 46, 695–704.

Schel, A. M., & Zuberbühler, K. (2009). Responses to leopards are independent of

experience in Guereza colobus monkeys. Behaviour, 146(12), 1709–1737.

Stankowich, T., & Coss, R. G. (2007). Effects of risk assessment, predator behavior, and

habitat on escape behavior in Columbian black-tailed deer. Behavioral Ecology,

18(2), 358–367. doi:10.1093/beheco/arl086

Suraci, J. P., Clinchy, M., Dill, L. M., Roberts, D., & Zanette, L. Y. (2016). Fear of large

carnivores causes a trophic cascade. Nature Communications, 7, 1–7.

doi:10.1038/ncomms10698

95

Swanson, A., Caro, T., Davies Mostert, H., Mills, M. G. L., Macdonald, D. W., Borner,

M., … Packer, C. (2014). Cheetahs and wild dogs show contrasting patterns of

suppression by lions. Journal of Animal Ecology, 83, 1418–1427.

doi:10.1111/1365-2656.12231

Thaker, M., Vanak, A. T., Owen, C. R., Ogden, M. B., Niemann, S. M., & Slotow, R.

(2011). Minimizing predation risk in a landscape of multiple predators: effects on

the spatial distribution of African ungulates. Ecology, 92(2), 398–407.

doi:10.1890/10-0126.1

Toïgo, C. (1999). Vigilance behavior in lactating female Alpine ibex. Canadian Journal

of Zoology, 77(7), 1060–1063. doi:10.1139/z99-081

Tsalyuk, M., Kelly, M., & Getz, W. M. (2017). Improving the prediction of African

savanna vegetation variables using time series of MODIS products. ISPRS

Journal of Photogrammetry and Remote Sensing, 131, 77–91.

doi:10.1007/s11065-015-9294-9.Functional

Valeix, M., Fritz, H., Loveridge, A. J., Davidson, Z., Hunt, J. E., Murindagomo, F., &

Macdonald, D. W. (2009). Does the risk of encountering lions influence African

herbivore behaviour at waterholes? Behavioral Ecology and Sociobiology, 63(10),

1483–1494. doi:10.1007/s00265-009-0760-3

Vanak, A. T., Fortin, D., Thaker, M., Ogden, M., Owen, C., Greatwood, S., & Slotow, R.

(2013). Moving to stay in place: Behavioral mechanisms for coexistence of

African large carnivores. Ecology, 94(11), 2619–2631. doi:10.1890/13-0217.1

96

Weissburg, M., Smee, D. L., & Ferner, M. C. (2014). The sensory ecology of

nonconsumptive predator effects. American Naturalist, 184(2), 141–157.

doi:10.1086/676644

Wolf, C., & Ripple, W. J. (2017). Range contractions of the world’s large carnivores.

Royal Society Open Science, 4(7), 1–11. doi:10.1098/rsos.170052

Ydenberd, R. C., & Dill, L. M. (1986). The economics of fleeing from predators.

Advances in the Study of Behavior, 16(C), 229–249.

97

FIGURES

Figure 3.1. Relationship between average Enhanced Vegetation Index (EVI) within 50 meters of playback location and baseline vigilance for female cheetahs with cubs, Mun- Ya-Wana Conservancy, KwaZulu-Natal, South Africa. Shaded regions indicate 85% confidence intervals.

98

Figure 3.2. Relationship between playback type and cheetah post-playback a) vigilance or b) probability of fleeing, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018-2019. Error bars represent SE.

99

Figure 3.3. Differences in post-predator (lion and leopard) playback vigilance behavior between female cheetahs with cubs and males, Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa. Error bars represent SE.

100

Figure 3.4. Probability of cheetahs fleeing following lion and leopard playbacks based on average Enhanced Vegetation Index (EVI) within 50 meters of playback location, Mun- Ya-Wana Conservancy, KwaZulu-Natal, South Africa. Shaded regions represent 85% CI.

101

CHAPTER 4

EFFECTS OF PRESCRIBED BURNING ON WILDLIFE COMMUNITY DYNAMICS

ABSTRACT

Fires are common in many ecosystems worldwide and are also a common management tool. Although the responses of herbivores to fire have been well-studied, less research has addressed how carnivores respond to fire. In particular, little is known about how fires affect multiple carnivore species simultaneously, and how changes in prey abundance as a result of fire has the potential to influence carnivore coexistence or suppression. We used prescribed burning to experimentally increase prey densities and monitored carnivore intensity of use using camera traps with a Before-After-Control-

Impact study design. We analyzed the camera trap data using community N-mixture models to understand how individual species, as well as large carnivores and small carnivores as a whole respond to burning. We found that although the intensity of use of several important prey species increased following burns, the responses of carnivores to prescribed burning were variable. Lions, honey badgers, and side-striped jackals increased intensity of use in burned area following the fires, whereas all other large and small carnivore intensity of use was not affected by burning. Our results indicate that fire does not promote carnivore coexistence by creating conditions for all carnivores to increase use of burned areas, but that it also likely does not result in numerical suppression of subordinate predators. Instead, fires might cause a suppression of opportunities for subordinate predators because they need to avoid lions rather than take

102 advantage of increased hunting opportunities in recently burned areas. In systems with a high diversity of carnivore species, complexity in species-specific responses to fire is to be expected.

INTRODUCTION

Fires are a common occurrence throughout most of the world, either as natural wildfires, or management-focused prescribed burning (Bond & Keeley 2005). Not only can fires causes changes in the physical landscape of an area, but they can also affect wildlife distributions, habitat use, and abundance (He et al. 2019, Nimmo et al. 2019).

Herbivore responses to fire have been well-studied, and species in a variety of systems have been found to increase their use of burned areas because of increases in nutrients and forage quality (Archibald et al. 2005). In contrast, our understanding of carnivore responses of fire is limited and mainly based on studies from Australia and North

America, and there is considerable variability in how different carnivore species, or different populations of the same species respond to fire (Geary et al. 2019).

Some carnivores have been found to increase use of burned areas following fires in response to higher prey densities (Dees et al. 2001, Robinson et al. 2012), or higher hunting success because of prey vulnerability (Leahy et al. 2015, McGregor et al. 2016).

For example, feral cats (Felis catus) have been found to predate upon rodents in Australia following fires because of increased vulnerability of prey in burned areas (Leahy et al.

2015), and have been found to make long-distance movements to select for burned areas

(McGregor et al. 2016). Carnivores can also have neutral responses to burning because of

103 their adaptability to use a wide variety of habitats, and lack of preference for specific habitat characteristics (Nimmo et al. 2019). Finally, some carnivores might have a negative response to fire because of reduced concealment for ambush predators (Eby et al. 2013), or lack of suitable vegetation for den sites (Sweitzer et al. 2016). For example, fishers (Pekania pennanti) in the northwestern United States are negatively associated with fires because of their requirement for unburned mature forests for denning (Sweitzer et al. 2016). Most research on carnivore responses to fire has focused on single species’ responses to changes in prey or habitat characteristics as a result of fire, whereas less research has been conducted on responses to fire by multiple carnivore species simultaneously.

Because there are no clear patterns on how carnivores as a whole respond to fire and because most research is limited to a single species, there is also limited information on how fire might affect interspecific interactions within the carnivore guild. One possible community-level response to fires is promoting the coexistence of multiple carnivore species. With an abundance of prey following fires, competition for resources among predators should decrease (Périquet et al. 2015). Evidence for fires promoting carnivore coexistence includes research in the Serengeti, which found that both large and small carnivores were more abundant in areas following fires compared to pre-burn conditions (Green et al. 2015). Similarly, a high diversity of both large and small predators was recorded in recently-burned areas in Turkey (Soyumert et al. 2019).

Alternatively, fires might create conditions that lead to mesopredator suppression, in which subordinate predators are negatively affected by the presence, abundance, or

104 behavior of apex predators. If apex predators and mesopredators are attracted to localized areas of high prey density as a result of burning, there could be a higher probability of antagonistic encounters, thereby numerically reducing mesopredator populations.

Additionally, fires might cause a spatial suppression of mesopredators whereby subordinate predators avoid recently burned areas as a means of avoiding apex predators using those areas (Ritchie & Johnson 2009). This has been observed in Australia, where dingos (Canis dingo) are attracted to burned areas, which in turn leads to avoidance of burned areas by red foxes (Vulpes vulpes), and in North America where coyotes (Canis latrans) have higher abundances near recently burned stands, but some smaller carnivores such as racoons ( lotor) prefer areas far from recent burns potentially because of predator avoidance (Jorge et al. 2020). More research is needed on how fires can influence carnivore community dynamics, especially in complex systems in which multiple carnivores exists in different trophic levels, which could drive both competition and suppression.

The southern African system is ideal to investigate effects of fire on carnivore community dynamics. Fire has historically been present in the system and is frequently used as a management tool (Brocket et al. 2001), and herbivore responses to burning are well-described. In particular, many large-bodied grazer species such as zebra (Equus quagga), impala (Aepyceros melampus) and wildebeest (Connochaetes taurinus), as well as several rodent species increase use of burned areas (Tomor & Owen-Smith 2002,

Monadjem & Perrin 2003, Archibald et al. 2005), whereas browsers such as black rhinos

(Diceros bicornis) often do not show preference for recently burned areas (Anderson et

105 al. 2020). In this system, there is a diversity of carnivore species, including large carnivores that prey upon larger-bodied prey species and impose top-down effects on smaller carnivores, and small carnivores that prey upon smaller-bodied prey species, or scavenge from the kills of larger carnivores.

We experimentally manipulated prey densities using prescribed burning to better understand how carnivore communities are influenced by fire. We tested two competing hypotheses related to the effect of fire on carnivore community dynamics:

1. Fire promotes carnivore coexistence: We predicted that apex predators and

mesopredators would have higher intensity of use following prescribed burns

2. Fire promotes mesopredator suppression: We predicted that apex predators

would have higher intensity of use following prescribed burns and that

mesopredators would have lower intensity of use following prescribed burns.

By studying the responses to fire of multiple carnivore species simultaneously, we can gain a better understanding of how fires might affect carnivore coexistence and suppression.

METHODS

Study Area

We studied wildlife community responses to prescribed burning in the Mun-Ya-

Wana Conservancy (Phinda Private Game Reserve), in KwaZulu-Natal, South Africa.

The Conservancy is 285 km2 and is surrounded by electrified game fencing. The dominant vegetation type is broad-leaf woodland, with open grasslands and semi-open

106 wooded-grasslands interspersed throughout the reserve. The elevation of the Conservancy ranges from 4 to 201 meters above sea level. The climate is subtropical with warm, dry winters (April – September) and hot, humid summers (October – March), with most rain falling in the summer.

Fire was historically present in this system and since the early 1990’s prescribed burns have been conducted as a form of management. Prescribed burns occur towards the end of the dry season (April – September) with the goals of decreasing woody underbrush encroachment, reducing the fuel load and promoting vegetation re-growth (C. Sholto-

Douglas, pers. comm.). For this study, prescribed burns were conducted on 4-5

September 2018 and 16 August 2019 or 2-3 September 2019. Burns were an average of

5.4 km2 (range: 0.1 – 38.6 km2) and distributed throughout the study area to create a mosaic of burned and unburned areas (Figure 4.1).

Data Collection

We conducted a camera trap survey from 7 August 2018 – 26 October 2018 and 8

August 2019 – 4 November 2019. During both years of the survey, we used a Before-

After-Control-Impact design, whereby cameras were placed at treatment sites and control sites and monitored both before and after the burns occurred. We collected data for approximately 1 month prior to burns, and 2 months after the burns occurred. We placed a 1.25 km2 grid across the entire study area and tried to optimize camera placement within the grid to achieve a balance of treatment and control sites, such that all cameras were an average of 1.1 km (SE = 0.15 km) from the nearest camera. We attempted to

107 stratify sampling sites, such that treatment and control points were similar in terms of habitat characteristics and other factors that might influence wildlife space use such as elevation and distance to water. We had burn histories for 85% of all sites going back to

1994. For sites with known burn histories, our control points were burned an average of

10.1 years prior to sampling (range: 4 – 25 years), and our treatment sites were burned an average of 7.1 years prior to being burned for this research (range: 2 – 21 years). In 2018 we had 73 camera sites active and in 2019 we had 74 camera sites active. We used

Cuddeback white flash camera traps (Professional Color model 1347, Cuddeback, Green

Bay, WI) set at 30 cm in height on trees or stakes, and approximately 2-m from roads or heavily used game trails. We did not use any type of bait or attractant at the camera trap sites.

Because we predicted that burning might change vegetation characteristics, which in turn could influence detection rates, we calculated a visibility metric for each camera.

We measured from the camera to the closest source of visual obstruction up to 15m at 10º increments from 70º to 110º, with 90º corresponding to measurements perpendicular to the camera. We took these measurements before and after the burns and averaged the site-specific measurements for each treatment period (Appendix C1).

Data Processing

We identified all identifiable species present in each photo. We excluded any blank photos, unidentifiable photos, or photos of vehicles and humans from analyses. We manually counted the daily number of animals of each species at each camera site.

108

Because the majority of species did not exhibit individual markings, we attempted to account for individual animals that were included in multiple successive photos by using a 30-minute window to demarcate independent events. Within each event, we used information such as sex, animal size, and direction of travel to count the number of individual animals. Although we recognize that this method of counting could induce bias in traditional abundance-estimation models because a single animal could have been counted multiple times a day, we were interested in estimating visitation rates, rather than true abundance.

Data Analysis

To assess changes in carnivore community dynamics in relation to burning, we used community N-mixture models (Yamaura et al. 2012) to estimate intensity of use.

Because many of the species that we were interested in monitoring were widespread in our study area, an occupancy analysis would not have allowed us to detect changes in response to burning. However, we wanted to incorporate detection probabilities into our analysis, rather than using a simple count metric. Our modelling framework makes use of repeated counts to estimate detection probabilities. However, instead of using this model to estimate abundance, we chose to model daily counts as a measure of relative abundance or daily visitation rates similar to Keim et al. (2019), but with the addition of incorporating detection.

An advantage of using community N-mixture models is that covariates can be estimated for species with low sample sizes by borrowing information from species with

109 larger sample sizes, while investigating group-level effects (Zipkin et al. 2009, Pacifici et al. 2014). We modeled the effects of three separate groups, which we categorized based on our predicted responses to burning: large carnivores, small carnivores, and prey. We predicted that prey species would be affected by changes in vegetation, large carnivores would be affected by changes in the prey species that we were monitoring, and small carnivores would be affected by changes in smaller prey items that we didn’t monitor such as rodents and amphibians. Within each group we also modeled individual species effects. As is typical for BACI experiments, our model for assessing changes to prescribed burning included an effect of treatment (burn vs. control), time period (pre- burn vs. post-burn), and an interaction between treatment and time period. Although the individual treatment and time period covariates could offer insight into the ecology of the system, we were mainly interested in the interaction covariate, which indicates changes in intensity of use as a result of treatments while accounting for potential time-period differences. We initially modeled detection as a function of visibility index. However, this model did not perform as well as a model without detection covariates (Appendix

C1), so our final model formulation only included covariates on the abundance portion of the model.

We formulated the model such that the abundance of species i at site j is a Poisson random variable based on expected abundance λi,j:

푁푖,푗~푃표푖푠푠표푛(휆푖,푗)

110

The expected abundance λi,j is based on our site-level covariates of interest:

log(휆푖,푗) = 훽0푖 + 훽1푖 ∗ 푇푟푒푎푡푚푒푛푡푗 + 훽2푖 ∗ 푃푒푟푖표푑푗 + 훽3푖 ∗ 푇푟푒푎푡푚푒푛푡푗 ∗ 푃푒푟푖표푑푗

Under this formulation, the β covariates in the model are species specific, but are drawn from group-level normal distributions with mean μ and variance σ2 for group g:

2 훽0푖,푔 ~ 푁표푟푚푎푙(휇훽0 푔, 휎 훽0푔 )

We ran our models using Markov Chain Monte Carlo (MCMC) simulation in

JAGS (Plummer 2003) using the jagsUI package (Kellner 2015) for Program R (Version

3.5.3; R Core Team 2019). As priors, we used N(0, 0.1) for all mean parameters and unif(0, 3) for all standard deviation parameters. We ran the models using three parallel chains with a length of 50,000 after discarding 100,000 iterations for adaptation and burn-in, and a thinning rate of 5. We assessed model convergence using the Gelman-

Rubin statistic (Gelman et al. 2013) and visual inspection of traceplots. We considered covariates to be significant if the 95% credible interval did not overlap zero, and marginally significant if the 90% credible interval did not overlap zero.

RESULTS

We collected data during 4,014 camera trap nights in 2018 and 5,165 camera trap nights in 2019 for a total of 9,179 camera trap nights. Across the entire survey, we

111 recorded 48 species totaling 76,035 detections (Appendix C2). We excluded ten species from our analyses because we obtained less than 20 captures each over the entire sampling period (, Caracal caracal; greater cane rat, Thryonomys swinderianus; greater galago, Otolemur crassicaudatus; leopard tortoise, Stigmochelys pardalis; mountain reedbuck, Redunca fulvorufula; rock monitor, Varanus albigularis; steenbok,

Raphicerus campestris; striped , striatus; suni, Neotragus moschatus,

African , Felis lybica). We excluded domestic dogs (Canis lupus familiaris), domestic cats (Felis catus), and goats (Capra aegagrus hircus) because they were rarely detected and do not occur on the landscape naturally. We also excluded any bird species other than guineafowl because they are not reliable detected using camera traps. Finally, we excluded African wild dogs (Lycaon pictus) which are not found in our study area but occasionally enter for brief periods from neighboring reserves.

At the group level, the intensity of use of carnivores and mesocarnivores was not affected by time period, treatment, or the interaction between time period and treatment

(Table 4.1). At the individual species level, lions showed a significant effect of the interaction between time period and treatment (Figure 4.2), with higher intensity of use in burned areas following prescribed burns (Figure 4.3). Honey badgers (Mellivora capensis) and side-striped jackals (Canis adustus) showed a marginally significant effect of the interaction between time period and treatment (Figure 4.2), with higher intensity of use in burned areas following prescribed burns (Figure 4.3). All other large and small carnivore species did not exhibit any significant interaction between time period and treatment (Figure 4.2).

112

The intensity of use of all herbivore species combined was not affected by time period or treatment, but did show a significant interaction between time period and treatment (Table 4.1). Several individual herbivore species including buffalo (Syncerus caffer), giraffe (Giraffa Camelopardalis), impala (Aepyceros melampus), warthog

(Phacochoerus africanus), wildebeest (Connochaetes taurinus), and zebra (Equus quagga) showed a significant effect of the interaction between time period and treatment

(Figure 4.2), with higher intensity of use in burned areas following prescribed burns

(Figure 4.3). Nyala (Tragelaphus angasii) was the only species to respond negatively to prescribed burning, with marginally lower intensity of use in burned areas following the prescribed burns (Figure 4.3).

DISCUSSION

Similar to previous research in southern Africa, several species of herbivores increased use of burned areas post-fire, which offered the opportunity to test hypotheses related to the effects of fire on carnivore communities. Specifically, we found that several large-bodied grazers such as impala and zebra had higher intensity of use in recently burned areas, likely because of increased foraging opportunities, whereas grazers such as elephants (Loxodonta africana) and black rhinos did not adjust their intensity of use in relation to burning. Because many of the grazers that we found to significantly respond to burning comprise a large portion of carnivore diets in this system (Hayward & Kerley

2008), we were able to assess how fire can influence carnivore coexistence and suppression. We found variability in responses of carnivores in a multi-carnivore system

113 to fire. Specifically, we did not find support for the hypothesis that fire can promote coexistence or numerical suppression in carnivore communities, but did find support to suggest that burning can lead to suppression of opportunities for mesopredators. Our results highlight the need to better understand how fires can differentially influence carnivores in multi-carnivore systems.

Although burning did not affect the intensity of use of all large carnivores, we found that several carnivore species at both the apex (e.g., lions) and mesocarnivore (i.e., side-striped jackals and honey badgers) trophic levels had higher intensity of use in areas following burning. In this system, lions’ preferred prey are wildebeest, warthog, nyala, and zebra (Hunter 1998), and all of these species, with the exception of nyala, increased their use of areas that were recently burned. Thus, lions likely positively responded to increased prey availability post-burn, similar to research in Australia that found that feral cats used recently burned areas to benefit from increased hunting opportunities

(McGregor et al. 2016). This is in contrast to previous research in the Serengeti which found that lions avoided burned areas during the day because reductions in vegetation cover might have reduced hunting success (Eby et al. 2013). In our system, there are more woodland habitats with understory shrubs that do not burn, compared to more open vegetation types found in the Serengeti; thus some vegetative cover remained in most areas that were burned in our study area, which could have been used for stalking by lions. Additionally, by using camera traps we were able to collect data at all times of the day. Even if vegetative cover was reduced from the fires, lions are effective hunters at night (Van Orsdol 1984), which might explain why our results deviate from Eby et al.

114

(2013) who only considered daytime locations of lions. We also found that side-striped jackals and honey badgers had marginally positive responses to burning, suggesting that small carnivores can also increase use of areas post-burn. Both side-striped jackals and honey badgers are generalists and can feed opportunistically on a variety of prey

(Atkinson et al. 2002, Begg et al. 2003). These two species might have been able to benefit from small mammals, amphibians, and insects that were killed during the burns or scavenging from the remains of lion kills in these areas.

Although lions and two species of small carnivores did respond positively to burning, we did not find strong support for the hypothesis that fires promote carnivore coexistence. Contrary to our predictions for this hypothesis, the majority of large and small carnivores did not exhibit any significant response to prescribed burning and associated increase in prey. In other systems worldwide, some carnivore species have been found to exhibit similar neutral responses to fire (Geary et al. 2019). For example, (Lynx rufus) used burned areas as frequently as unburned areas following a fire in

California (Borchert 2012), and red foxes (Vulpes vulpes) did not change in abundance in response to fire in Australia (Green & Sanecki 2006). One potential explanation for neutral responses to burning is that many carnivore species have large home ranges, and can make use of a variety of prey species (Santos et al. 2014, Nimmo et al. 2019).

Because they aren’t entirely dependent on the areas that were burned or one specific prey item, the benefits of using recently burned areas might be negligible. In addition, the post-burn responses of smaller species such as rodents and amphibians, which can comprise a large portion of smaller carnivores’ diets (Caro & Stoner 2003), could explain

115 our results. Previous research in Africa has found that several rodent species increase in abundance following burns (Cheeseman & Delany 1979, Monadjem & Perrin 2003) or maintain stable population sizes (MacFadyen et al. 2012), whereas other research suggests that some rodent species are negatively affected by fire (Monadjem & Perrin

2003). Therefore, if small mammals did not respond positively to the burns, we might not have seen increased use by small carnivores. However, research also suggests that burning creates conditions that are conducive to hunting small mammals such as rodents

(Conner et al. 2011, Leahy et al. 2015); therefore even if rodent populations were reduced from the fires we might still expect to see increased used of burned areas by small carnivores because of greater hunting success.

We also did not find support for the hypothesis that fires promote mesopredator suppression. For the majority of mesopredator species we did not observe increased use of burned areas, but we also did not observe a decrease in use of these areas. In this system several species of mesopredator use fine-scale spatial or temporal partitioning to avoid apex predators (Ramesh et al. 2017); therefore mesopredators might have spatially or temporally adjusted activity patterns to account for higher lion use in burned areas without needing to completely avoid the burned areas. Additionally, we only monitored the community for approximately 2 months following the burns. Although increased use of burned areas by predators and prey is typically most intense for the first few months post-burn (Green et al. 2015), it could take longer to detect numerical decreases in mesopredator populations. Lastly, the diversity of carnivores in our systems could have mitigated strong evidence of inter-specific competition in response to increased prey

116 availability at burn sites. Previous research in Australia found that dingos increased use of burned areas, which in turn indirectly affected red foxes through avoidance behaviors

(Geary et al. 2018). By contrast, in our system there were several large carnivores that hunt larger bodied prey species, as well as several small carnivores that feed on smaller prey species. With such a diversity of predator and prey species, the top-down effects of increased lion use in burned areas might be diluted (Finke & Denno 2004, Haswell et al.

2017).

Although our results do not indicate numerical suppression of mesopredators, our results suggest that there was a suppression of opportunity for subordinate predators to take advantage of an increase in prey density following fire. For other large carnivores such as leopards, cheetahs, and spotted hyenas, the increase in use of burned areas by several prey species would have created more hunting opportunities. Even though these carnivore species are large in body size, lions can impose top-down effects on each of them through direct and indirect interactions (Laurenson 1994, Trinkel & Kastberger

2005, du Preez et al. 2015). Therefore, because lions increased use of burned areas following fires, the other large carnivores might have not increased as a means of avoiding interactions with lions and therefore fail to benefit from higher prey densities in the burned areas. Similar to large carnivores, we found support that fire might result in suppression of opportunity for small carnivores (with the exception of honey badgers and side-stiped jackals). Even though lions do not directly compete with small carnivore for prey, lions have been found to kill most of our studied mesopredator species (Caro &

Stoner 2003, Donadio & Buskirk 2006), so increased used of burned areas following fires

117 could be detrimental to small carnivores also using those areas. In the Serengeti, honey badgers do not avoid lions in space or time, potentially because of their aggressive nature

(Allen et al. 2018), which could also be an explanation as to how honey badgers were able to increase use of burned areas, even with higher lion intensity of use. Side-stiped jackals exhibit variability in activity patterns in different parts of their range (Fuller et al.

1989, Loveridge & Macdonald 2003); thus this temporal flexibility could be a means for them to use burned areas while avoiding lions. Indirect effects of top predators on lower trophic levels have been found to be more important than direct predation in structuring some systems (Preisser et al. 2005) so future research is needed on understanding if this suppression of opportunity for non-dominate predators following fire can scale-up and have demographic or population-level consequences.

Collectively, our results highlight the complexities of understanding community- wide responses to disturbances such as fire. Whereas coexistence or suppression post- burn might follow expected predictions in simple predator-prey systems (e.g., Geary et al.

2018), our results suggest that in systems with multiple carnivore and prey species intraguild interactions can create complex interactions in response to burning. In addition to fire, there are other conditions where localized increases in resources could affect carnivore community dynamics. For example, mass mortality events of prey species could create localized resource subsidies, which in turn could cause changes in carnivore coexistence or suppression (Polis et al. 1997). Similarly, increases in human food subsidies have been found to affect predator abundances and space use, which in turn can affect coexistence with and suppression of mesopredators (Newsome et al. 2015). As

118 human impacts on environments continue to increase, it is important to understand how environmental changes can affect the underlying ecology of a system. Specifically for fire, when managers use prescribed burning as a means of habitat management or providing foraging opportunities for herbivores, they need to be aware of potential community-level consequences and variation in carnivore responses to fire.

REFERENCES

Allen, M. L., Peterson, B., & Krofel, M. (2018). No respect for apex carnivores:

Distribution and activity patterns of honey badgers in the Serengeti. Mammalian

Biology, 89, 90–94. doi:10.1016/j.mambio.2018.01.001

Anderson, T. M., Ngoti, P. M., Nzunda, M. L., Griffith, D. M., Speed, J. D. M., Fossøy,

F., … Graae, B. J. (2020). The burning question: Does fire affect habitat selection

and forage preference of the black rhinoceros Diceros bicornis in East African

savannahs? Oryx, 54(2), 234–243. doi:10.1017/S0030605318000388

Archibald, S., Bond, W. J., Stock, W. D., & Fairbanks, D. H. K. (2005). Shaping the

landscape: Fire-grazer interactions in an African savanna. Ecological

Applications, 15(1), 96–109. doi:10.1890/03-5210

Atkinson, R. P. D., Macdonald, D. W., & Kamizola, R. (2002). Dietary opportunism in

side-striped jackals Canis adustus Sundevall. Journal of Zoology, 257(1), 129–

139. doi:10.1017/S0952836902000729

119

Begg, C. M., Begg, K. S., Du Toit, J. T., & Mills, M. G. L. (2003). Sexual and seasonal

variation in the diet and foraging behaviour of a sexually dimorphic carnivore, the

(Mellivora capensis). Journal of Zoology, 260(3), 301–316.

doi:10.1017/S0952836903003789

Bond, W. J., & Keeley, J. E. (2005). Fire as a global “herbivore”: The ecology and

evolution of flammable ecosystems. Trends in Ecology and Evolution, 20(7),

387–394. doi:10.1016/j.tree.2005.04.025

Borchert, M. I. (2012). Mammalian carnivore use of a high-severity burn in conifer

forests in the San Bernardino Mountains of southern California, USA. Hystrix,

23(2), 50–56. doi:10.4404/hystrix-23.2-5610

Brocket, B. H., Biggs, H. C., & Van Wilgen, B. W. (2001). A patch mosaic burning

system for conservation areas in southern African savannas. International Journal

of Wildland Fire, 10(2), 169–183. doi:10.1071/WF01024

Caro, T. M., & Stoner, C. J. (2003). The potential for interspecific competition among

African carnivores. Biological Conservation, 110(1), 67–75. doi:10.1016/S0006-

3207(02)00177-5

Cheeseman, C. L., & Delany, M. J. (1979). The population dynamics of small rodents in

a tropical African grassland. Journal of Zoology, 188(4), 451–475.

doi:10.1111/j.1469-7998.1979.tb03429.x

Conner, L. M., Castleberry, S. B., & Derrick, A. M. (2011). Effects of mesopredators and

prescribed fire on hispid cotton rat survival and cause-specific mortality. Journal

of Wildlife Management, 75(4), 938–944. doi:10.1002/jwmg.110

120

Dees, C. S., Clark, J. D., & Van Manen, F. T. (2001). Florida panther habitat use in

response to prescribed fire. The Journal of Wildlife Management, 65(1), 141–147.

Donadio, E., & Buskirk, S. W. (2006). Diet, morphology, and interspecific killing in

. The American Naturalist, 167(4), 524–536. doi:10.2307/3844707 du Preez, B., Hart, T., Loveridge, A. J., & Macdonald, D. W. (2015). Impact of risk on

animal behaviour and habitat transition probabilities. Animal Behaviour, 100, 22–

37. doi:10.1016/j.anbehav.2014.10.025

Eby, S., Mosser, A., Swanson, A., Packer, C., & Ritchie, M. E. (2013). The impact of

burning on lion Panthera leo habitat choice in an African savanna. Current

Zoology, 59(3), 335–339.

Finke, D. L., & Denno, R. F. (2004). Predator diversity dampens trophic cascades.

Nature, 429(6990), 407–410. doi:10.1038/nature02554

Fuller, T. K., Biknevicius, A. R., Kat, P. W., Van Valkenburgh, B., & Wayne, R. K.

(1989). The ecology of three sympatric jackal species in the Rift Valley of Kenya.

African Journal of Ecology, 27(4), 313–323. doi:10.1111/j.1365-

2028.1989.tb01025.x

Geary, W. L., Doherty, T. S., Nimmo, D. G., Tulloch, A. I. T., & Ritchie, E. G. (2019).

Predator responses to fire: A global systematic review and meta-analysis. Journal

of Animal Ecology, (October 2019), 1–17. doi:10.1111/1365-2656.13153

121

Geary, W. L., Ritchie, E. G., Lawton, J. A., Healey, T. R., & Nimmo, D. G. (2018).

Incorporating disturbance into trophic ecology: Fire history shapes mesopredator

suppression by an apex predator. Journal of Applied Ecology, 55(4), 1594–1603.

doi:10.1111/1365-2664.13125

Gelman, A., Carlin, J. B., Stern, H. S., & Rubin, D. (2013). Bayesian Data Analysis.

Boca Raton, FL: Chapman and Hall/CRC Press.

Green, D. S., Roloff, G. J., Heath, B. R., & Holekamp, K. E. (2015). Temporal dynamics

of the reponses by African mammals to prescribed fire. The Journal of Wildlife

Management, 79(2), 235–242. doi:10.1002/jwmg.827

Green, K., & Sanecki, G. (2006). Immediate and short-term responses of bird and

mammal assemblages to a subalpine wildfire in the Snowy Mountains, Australia.

Austral Ecology, 31(6), 673–681. doi:10.1111/j.1442-9993.2006.01629.x

Haswell, P. M., Kusak, J., & Hayward, M. W. (2017). Large carnivore impacts are

context-dependent. Food Webs, 12, 3–13. doi:10.1016/j.fooweb.2016.02.005

Hayward, M. W., & Kerley, G. I. H. (2008). Prey preferences and dietary overlap

amongst Africa’s large predators. African Journal of Wildlife Research, 38(2),

93–108. doi:10.3957/0379-4369-38.2.93

He, T., Lamont, B. B., & Pausas, J. G. (2019). Fire as a key driver of Earth’s biodiversity.

Biological Reviews, 94(6), 1983–2010. doi:10.1111/brv.12544

Hunter, L. T. B. (1998). The behavioural ecology of reintroduced lions and cheetahs in

the Phinda Resource Reserve, KwaZulu-Natal, South Africa. PhD Thesis, 1–206.

122

Jorge, M. H., Garrison, E. P., Conner, L. M., & Cherry, M. J. (2020). Fire and land cover

drive predator abundances in a pyric landscape. Forest Ecology and Management,

461, 117939. doi:10.1016/j.foreco.2020.117939

Keim, J. L., Lele, S. R., DeWitt, P. D., Fitzpatrick, J. J., & Jenni, N. S. (2019). Estimating

the intensity of use by interacting predators and prey using camera traps. Journal

of Animal Ecology, 88(5), 690–701. doi:10.1111/1365-2656.12960

Kellner, K. F. (2015). jagsUI: A wrapper around rjags to streamline JAGS analyses.

Laurenson, M. K. (1994). High juvenile mortality in cheetahs (Acinonyx jubatus) and its

consequences for maternal care. Journal of the Zoological Society of London,

234(1), 387–408. doi:10.1111/j.1469-7998.1994.tb04855.x

Leahy, L., Legge, S. M., Tuft, K., McGregor, H. W., Barmuta, L. A., Jones, M. E., &

Johnson, C. N. (2015). Amplified predation after fire suppresses rodent

populations in Australia’s tropical savannas. Wildlife Research, 42(8), 705–716.

doi:10.1071/WR15011

Loveridge, A. J., & Macdonald, D. W. (2003). Niche separation in sympatric jackals

(Canis mesomelas and Canis adustus). Journal of Zoology, 259(2), 143–153.

doi:10.1017/S0952836902003114

MacFadyen, D. N., Avenant, N. L., van der Merwe, M., & Bredenkamp, G. J. (2012).

The influence of fire on rodent abundance at the N’washitshumbe enclosure site,.

African Zoology, 47, 138–146.

123

McGregor, H. W., Legge, S., Jones, M. E., & Johnson, C. N. (2016). Extraterritorial

hunting expeditions to intense fire scars by feral cats. Scientific Reports, 6, 1–7.

doi:10.1038/srep22559

Monadjem, A., & Perrin, M. (2003). Population fluctuations and community structure of

small mammals in a Swaziland grassland over a three-year period. African

Zoology, 38(1), 127–137. doi:10.1080/15627020.2003.11657200

Newsome, T. M., Dellinger, J. A., Pavey, C. R., Ripple, W. J., Shores, C. R., Wirsing, A.

J., & Dickman, C. R. (2015). The ecological effects of providing resource

subsidies to predators. Global Ecology and Biogeography, 24(1), 1–11.

doi:10.1111/geb.12236

Nimmo, D. G., Avitabile, S., Banks, S. C., Bliege Bird, R., Callister, K., Clarke, M. F.,

… Bennett, A. F. (2019). Animal movements in fire-prone landscapes. Biological

Reviews, 94(3), 981–998. doi:10.1111/brv.12486

Pacifici, K., Zipkin, E. F., Collazo, J. A., Irizarry, J. I., & Dewan, A. (2014). Guidelines

for a priori grouping of species in hierarchical community models. Ecology and

Evolution, 4(7), 877–888. doi:10.1002/ece3.976

Périquet, S., Fritz, H., & Revilla, E. (2015). The Lion King and the Queen: Large

carnivore interactions and coexistence. Biological Reviews, 90(4), 1197–1214.

doi:10.1111/brv.12152

Plummer, M. (2003). JAGS: A program for analysis of Bayesian graphical models using

Gibbs sampling. Proceedings of the 3rd International Workshop on Distributed

Statistical Computing (DSC 2003), 20–22. doi:10.1.1.13.3406

124

Polis, G. A., Anderson, W. B., & Holt, R. D. (1997). Toward an integration of landscape

and food web ecology: The dynamics of spatially subsidized food webs. Annual

Review of Ecology and Systematics, 28, 289–316.

doi:10.1146/annurev.ecolsys.28.1.289

Preisser, E. L., Bolnick, D. I., & Benard, M. F. (2005). Scared to death? The effects of

intimidation and consumption in predator–prey interactions. Ecology, 86(2), 501–

509.

R Core Team. (2019). R: A language and environment for statistical computing. Vienna,

Austria: R Foundation for Statistical Computing.

Ramesh, T., Kalle, R., & Downs, C. T. (2017). Staying safe from top predators: patterns

of co-occurrence and inter-predator interactions. Behavioral Ecology and

Sociobiology, 71(2). doi:10.1007/s00265-017-2271-y

Ritchie, E. G., & Johnson, C. N. (2009). Predator interactions, mesopredator release and

biodiversity conservation. Ecology Letters, 12(9), 982–998. doi:10.1111/j.1461-

0248.2009.01347.x

Robinson, H. S., Hebblewhite, M., DeCesare, N. J., Whittington, J., Neufeld, L., Bradley,

M., & Musiani, M. (2012). The effect of fire on spatial separation between wolves

and caribou. Rangifer, (20), 277–294. doi:10.7557/2.32.2.2276

Santos, X., Mateos, E., Bros, V., Brotons, L., De Mas, E., Herraiz, J. A., … Viñolas, A.

(2014). Is response to fire influenced by dietary specialization and mobility? A

comparative study with multiple animal assemblages. PLoS ONE, 9(2), 1–10.

doi:10.1371/journal.pone.0088224

125

Soyumert, A., Ertürk, A., & Tavşanoğlu, Ç. (2019). Fire-created habitats support large

mammal community in a Mediterranean landscape. Mammal Research.

doi:10.1007/s13364-019-00473-y

Sweitzer, R. A., Furnas, B. J., Barrett, R. H., Purcell, K. L., & Thompson, C. M. (2016).

Landscape fuel reduction, forest fire, and biophysical linkages to local habitat use

and local persistence of fishers (Pekania pennanti) in Sierra Nevada mixed-

conifer forests. Forest Ecology and Management, 361, 208–225.

doi:10.1016/j.foreco.2015.11.026

Tomor, B. M., & Owen-Smith, N. (2002). Comparative use of grass regrowth following

burns by four ungulate species in the Nylsvley Nature Reserve, South Africa.

African Journal of Ecology, 40(2), 201–204. doi:10.1046/j.1365-

2028.2002.00350.x

Trinkel, M., & Kastberger, G. (2005). Competitive interactions between spotted hyenas

and lions in the Etosha National Park, Namibia. African Journal of Ecology, 43,

220–224.

Van Orsdol, K. G. (1984). Foraging behaviour and hunting success of lions in Queen

Elizabeth National Park, Uganda. African Journal of Ecology, 22(2), 79–99.

doi:10.1111/j.1365-2028.1984.tb00682.x

Yamaura, Y., Royle, J. A., Shimada, N., Asanuma, S., Sato, T., Taki, H., & Makino, S.

(2012). Biodiversity of man-made open habitats in an underused country: A class

of multispecies abundance models for count data. Biodiversity and Conservation,

21(6), 1365–1380. doi:10.1007/s10531-012-0244-z

126

Zipkin, E. F., Dewan, A., & Andrew Royle, J. (2009). Impacts of forest fragmentation on

species richness: A hierarchical approach to community modelling. Journal of

Applied Ecology, 46(4), 815–822. doi:10.1111/j.1365-2664.2009.01664.x

127

TABLES

Table 4.1. Group-level covariate estimates describing changes in intensity of use related to prescribed burning, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019. Group Covariate β SD 95% CI

Large carnivore Time period -0.001 3.184 -6.30 – 6.17

Treatment -0.007 3.148 -6.23 – 6.13

Time period*treatment -0.002 3.138 -6.11 – 6.15

Small carnivore Time period 0.011 3.135 -6.18 – 6.13

Treatment 0.002 3.172 -6.22 – 6.17

Time period*treatment -0.024 3.153 -6.20 – 6.17

Prey Time period -0.082 0.054 -0.19 – 0.02

Treatment -0.039 0.082 -0.20 – 0.12

Time period*treatment 0.183 0.069 0.05 – 0.32

128

FIGURES

Figure 4.1. Locations of prescribed burned and camera trap sites, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019.

129

Figure 4.2. Species-specific covariates estimates (mean ± 95% Credible Interval) for a) time period, b) treatment, and c) the interaction between time period and treatment on intensity of use, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019.

130

Figure 4.3. Estimated intensity of use (mean ± 95% CI) for a) buffalo, b) bushpig, c) giraffe, d) impala, e) nyala, f) warthog, g) wildebeest, h) zebra, i) lion, j) honey badger, and k) side-striped jackal in relation to prescribed burning, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018-2019.

131

APPENDICES

132

APPENDIX A1

Data collection and analysis timelines

Figure A1.1 Time periods corresponding with data collection of lion, cheetah, and prey covariates, as well as time periods used in analyses of cheetah reproduction and survival, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 1992 – 2018.

133

APPENDIX A2

Estimating prey density in the study area

Prey density estimation methods

We set 3 – 6 transects per season ranging from 2.75 – 14 km that encompassed a range of habitat types within the reserve (Figure A2.1). We drove these transects 4 – 6 times per season, and restricted our sampling to three hours after sunrise or three hours before sunset. When an individual or group of a prey species were sighted, we stopped the vehicle and measured the distance and compass bearing to the individual animal or center of the animal group. We also recorded the number of individuals present, and sex and age composition of the group (defined as a cluster of conspecifics within a 30 m radius; Jathanna et al. 2003). Using distance, bearing and vehicle coordinates we then calculated the actual relocation coordinates of each observed individual or group.

We calculated the distance to water using an existing shapefile of water bodies in the study area. For the dry season, we only included dams and perennial water bodies, whereas we included all water bodies in the wet season. We obtained elevation data based on a 30 m digital elevation model. For our vegetation covariate we used Enhanced

Vegetation Index (EVI) data at a 500 m resolution

(https://lpdaac.usgs.gov/data_access/data_pool). We calculated seasonal EVI values on a yearly basis by averaging EVI values across the entirety of a season.

We estimated spatially-explicit abundance of impala and nyala using hierarchical distance sampling models, with spatial covariates on both the abundance and detection

134 processes (Royle, Dawson, & Bates, 2004; Sillett et al., 2012) using the R package unmarked (Fiske & Chandler, 2011). We divided each transect into 400-meter segments and created a 200-meter buffer on each side of each segment to create 400 x 400 m sampling units. Within each sampling unit, we calculated the average seasonal EVI, distance to water, and elevation. For each sampling unit, we summed the number of animals of each species falling within 20-meter distance classes, up to 200 meters.

Because sampling units were surveyed varying numbers of times, we pooled sightings data within a season and included an offset term of the log of the number of occasions that a unit was sampled. We trained our model using data collected over 6 years (2010 –

2015).

For both the wet and dry seasons, we first fit the detection component of the model while holding abundance constant and considered effect of EVI on detection using half-normal, or hazard detection function. We compared detection models using Akaike’s

Information Criterion corrected for sample size (AICc; Burnham and Anderson 2002) and we considered models within 2 AICc of the top model to be competitive. Once we found our top detection model for each species in each season, we evaluated models with EVI, distance to water, and elevation as covariates on the abundance component of the model, using a negative binomial formulation and evaluated models using AICc. We model averaged any models within 2 AICc. To assess the fit of our top seasonal models for each species, we used parametric bootstrapping with 1000 iterations and calculated a Freeman-

Tukey fit statistic, in which a P value >0.05 indicated a suitable fit (Sillett et al., 2012).

135

For each species, we used the top seasonal model to predict seasonal year-specific abundance and distribution estimates. If more than one model was supported, we predicted abundances using all competitive models and weighted our estimates based on the AIC weights. To predict prey abundance across the extent of the reserve, we created a regular grid of 400 x 400 meter squares and calculated the average covariates of interest for each grid cell for each year and season of interest. We used our top seasonal model for impala and nyala to estimate the abundance of these species across the extent of our grid in the years of interest (1992 – 2018). We summed the prey abundance in all grid cells to estimate the overall abundance within the reserve.

Prey density estimation results

Drivers of impala and nyala abundance varied seasonally. For both the wet and dry seasons, the detection of nyala was best described by EVI using a hazard detection function (Table A2.1), whereas the detection of impala was best described by EVI using a half-normal detection function (Table A2.2). The abundance of nyala during dry seasons was best described by EVI and elevation (Table A2.3), with more nyala found in areas with high EVI and low elevation (Table A2.5). The abundance of impala during dry seasons was best described by elevation, EVI, and distance to water (Table A2.4), with more impala found in areas with low elevation, low EVI, and located closer to water

(Table A2.5). The abundance of nyala during wet seasons was best described by EVI, and elevation (Table A2.3), with more nyala found in areas with high EVI, and low elevation

(Table A2.5). Finally, the abundance of impala during wet seasons was best described by

136

EVI and distance to water, with more impala found in areas with low EVI and located closer to water (Table A2.5). The bootstrap P values based on the Freeman-Tukey test statistics were P=0.06 for nyala in the dry season, P=0.23 for impala in the dry season,

P=0.09 for nyala in the wet season, and P=0.08 for impala in the wet season, indicating that all top models provided adequate fits.

References Burnham, K. P., & Anderson, D. R. (2002). Model selection and multimodel inference: A

practical information-theoretic approach. Springer Science & Business Media.

Fiske, I., & Chandler, R. B. (2011). Unmarked: An R Package for Fitting Hierarchical

Models of Wildlife Occurrence and Abundance. Journal of Statistical Software,

43(10), 1–23.

Jathanna, D., Karanth, K. U., & Johnsingh, A. J. T. (2003). Estimation of large herbivore

densities in the tropical forests of southern India using distance sampling. Journal of

Zoology, 261, 285–290. Doi:10.1017/S0952836903004278

Royle, J. A., Dawson, D. K., & Bates, S. (2004). Modelling abundance effects in distance

sampling. Ecology, 85(6), 1591–1597. Doi:10.1890/03-3127

Sillett, S., Chandler, R. B., Royle, J. A., Kéry, M., Morrison, S. a, Sillett, T. S., & Kery,

M. (2012). Hierarchical distance sampling models to estimate population size and

habitat-specific abundance of an island endemic. Ecological Applications, 22(7),

1997–2006. Doi:10.1890/11-1400.1

137

Figure A2.1. Driving transect routes during a) wet seasons, and b) dry seasons in relation to Enhanced Vegetation Index (EVI) as a seasonal composite from 2010-2014, Mun-ya- wana Conservancy, KwaZulu-Natal, South Africa.

138

Table A2.1. Model selection results for the seasonal detection of nyala based on spatially- explicit distance sampling models, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2010 – 2014.

Season Model AIC ΔAIC w K Dry EVI hazard -4222.72 0 1.00 5 Null hazard -4182.64 40.08 0.00 4 EVI halfnormal -3768.97 453.75 0.00 4 Null halfnormal -3715.79 506.93 0.00 3

Wet EVI hazard -1003.44 0 1.00 5 Null hazard -961.69 41.75 0.00 4 EVI halfnormal -560.65 442.79 0.00 4 Null halfnormal -503.51 499.93 0.00 3

139

Table A2.2 Model selection results for the seasonal detection of impala based on spatially-explicit distance sampling models, Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa, 2010 – 2014.

Season Model AIC ΔAIC w K Dry EVI halfnormal -2196.64 0 1.00 4 Null halfnormal -2054.56 142.08 0.00 3 EVI hazard -1.99 2194.65 0.00 5 Null hazard 3472.64 5669.28 0.00 3

Wet EVI halfnormal -3203.52 0 1.00 4 Null halfnormal -3169.93 33.59 0.00 3 Null hazard -1618.74 1584.78 0.00 4 EVI hazard -1616.74 1586.78 0.00 5

140

Table A2.3. Model selection results for the seasonal abundance of nyala based on spatially-explicit distance sampling models, Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa, 2010 – 2014.

Season Model AIC ΔAIC w K Dry EVI+elevation -4241.01 0 0.81 7 EVI -4238.15 2.85 0.19 6 Elevation -4224.43 16.57 0.00 6 Null -4222.72 18.28 0.00 5 EVI + distance to water -4189.83 51.18 0.00 7 EVI + elevation + distance to water -4188.93 52.08 0.00 8 Distance to water -4179.43 61.58 0.00 6

Wet EVI -1012.88 0 0.61 6 EVI + elevation -1011.13 1.75 0.25 7 EVI + distance to water -1009.6 3.28 0.12 7 Null -1003.44 9.44 0.01 5 EVI + elevation + distance to water -1003.3 9.58 0.01 8 Distance to water -1001.92 10.96 0.00 6 Elevation -1001.78 11.1 0.00 6

141

Table A2.4. Model selection results for the seasonal abundance of impala based on spatially-explicit distance sampling models, Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa, 2010 – 2014.

Season Model AIC ΔAIC w K Dry Null -2196.64 0 0.34 4 Elevation -2195.92 0.72 0.24 5 EVI + elevation + distance to water -2195.38 1.25 0.18 7 EVI -2194.67 1.97 0.13 5 EVI + elevation -2193.95 2.69 0.09 6 Distance to water -2190.97 5.67 0.02 5 EVI + distance to water -2188.74 7.9 0.01 6

Wet Distance to water -3839.43 0 0.54 5 EVI + distance to water -3837.88 1.56 0.25 6 EVI + elevation + distance to water -3836.31 3.12 0.11 7 Null -3834.42 5.01 0.04 4 EVI -3832.99 6.45 0.02 5 Elevation -3832.68 6.75 0.02 5 EVI + elevation -3831.61 7.83 0.01 6

142

Table A2.5. Parameter estimates on the abundance components of top seasonal spatially- explicit distance sampling models for nyala and impala, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2010 – 2014.

Species and Season Covariate Estimate SE Nyala (Dry Season) EVI 5.51 1.31 Elevation -0.006 0.003

Nyala (Wet Season) EVI 3.83 1.13 Elevation -0.002 0.003

Impala (Dry Season) EVI -4.21 2.7 Elevation -0.007 0.006

Distance to water -0.0003 0.0002

Impala (Wet Season) Distance to water -0.001 0.0004 EVI -0.76 0.23

143

APPENDIX A3

Effects of time and sex on cheetah resighting and survival rates

Figure A3.1. Annual probability of resighting cheetahs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2009 – 2018.

144

Figure A3.2. Seasonal probability of resighting cheetahs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2009 – 2018.

145

Figure A3.3. Average monthly survival of female vs. male cheetahs, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2009 – 2018.

146

Figure A3. 4. Average monthly survival of cheetahs during dry seasons (April – September) vs. wet seasons (October – March), Mun-Ya-Wana Conservancy, KwaZulu- Natal, South Africa, 2009 – 2018.

147

APPENDIX A4

Supplementary results of cheetah survival analysis

Figure A4.1. Monthly cheetah survival for a) adults in relation to prey density, b) adults in relation to lion density, c) cubs in relation to prey density, and d) cubs in relation to lion density, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018.

148

APPENDIX B1

Stability of cheetah home ranges

We analyzed lifetime variation in cheetah home ranges to determine if we could pool locations to estimate home ranges. For adult cheetahs with large sample sizes of locations (more than 30 locations in more than 2 years), we calculated lifetime home ranges, and yearly home ranges using a utilization distribution (UD) using a fixed-kernel estimator and the plug-in method of bandwidth selection (Gitzen et al. 2006). We quantified overlap between lifetime home ranges and yearly home ranges of individual cheetahs using Bhattacharyya’s affinity (BA; Bhattacharyya 1943) because this metric has been found to be appropriate for quantifying similarity between utilization distributions (Fieberg and Kochanny 2005). We calculated BA metrics for 50%, 75%, and 95% home range contours and considered a value of >0.60 to be indicative of high degrees of overlap. For all home range contours, cheetahs exhibited high overlap between yearly home ranges and lifetime home ranges (B1.1).

References

Bhattacharyya, A. 1943. On a measure of divergence between two statistical populations

defined by their probability distributions. Bulletin of the Calcutta Mathe- matical

Society 35:99–109.

Fieberg, J., and C. O. Kochanny. 2005. Quantifying home-range overlap: The importance

of the utilization distribution. Journal of Wildlife Management 69:1346–1359.

149

Gitzen, R. A., J. J. Millspaugh, and B. J. Kernohan. 2006. Bandwidth selection for fixed-

kernel analysis of animal utilization distributions. Journal of Wildlife Management

70:1334–1344.

150

Table B1.1. Lifetime and annual cheetah home range overlap based on Bhattacharyya’s affinity, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018.

Home range contour Mean (± SE) Bhattacharyya’s affinity statistic

50% 0.76 ± 0.02

75% 0.83 ± 0.01

95% 0.89 ± 0.01

151

APPENDIX B2

Effects of spatial covariates at multiple scales on cheetah survival

To determine what spatial scale was most related to cheetah survival, we ran survival models using spatial covariates extracted from 50%, 75%, and 95% home range contours. We calculated home ranges using a utilization distribution (UD) using a fixed- kernel estimator and the plug-in method of bandwidth selection (Gitzen et al. 2006) and extracted lion density, prey density, and vegetation density covariates using the 50%,

75% and 95% home range contours. We analyzed cheetah survival using multi-state joint live-encounter dead-recovery models (see Methods for full details) using the spatial covariates at the three scales. We compared models using Akaike’s Information Criterion corrected for sample size (AICc; Burnham and Anderson, 2002), considered models within 2 ΔAICc of the top model to be competitive

For all home range contours, short-term cheetah survival was most influenced by

EVI (Table B2.1). When comparing all models simultaneously, the 50% home range contour models provided the best fit for the data, with the 75% home range contour models being competitive (Table B2.1).

References

Burnham, K. P., and D. R. Anderson. 2002. Model selection and multimodel inference: A

practical information-theoretic approach. Springer Science & Business Media.

Gitzen, R. A., J. J. Millspaugh, and B. J. Kernohan. 2006. Bandwidth selection for fixed-

152 kernel analysis of animal utilization distributions. Journal of Wildlife Management

70:1334–1344.

153

Table B2.1. Model selection results to compare effects of covariates at multiple spatial scales using multi-state joint live-encounter dead-recovery spatial-explicit survival models for cheetahs with seasonal spatial covariates, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2008 – 2018. States in the model include cubs (juveniles dependent on their mothers) and adults (non-juveniles). All models include effects of year on recovery rates and season on survival rates.

Model (HR contour) AICc ΔAIC Weight k

EVI (50%) 3601.15 0 0.29 9 EVI (75%) 3602.17 1.04 0.17 9 EVI (95%) 3603.51 2.36 0.09 9

Lion+EVI (50%) 3603.51 2.36 0.09 11

Lion*EVI (50%) 3604.06 2.91 0.07 13 Prey+EVI (50%) 3604.60 3.45 0.05 11 Prey+EVI (75%) 3604.64 3.50 0.05 11 Prey+EVI (95%) 3605.36 4.21 0.04 11 Lion+EVI (75%) 3605.90 4.75 0.03 11 Lion*EVI (95%) 3606.28 5.13 0.02 13 Prey*EVI (75%) 3606.51 5.36 0.02 13 Prey*EVI (95%) 3606.86 5.69 0.02 13 EVI+Lion+Prey (50%) 3607.08 5.93 0.01 13 Lion+EVI (95%) 3607.18 6.03 0.01 11

Lion*EVI (75%) 3607.31 6.16 0.01 13 Prey*EVI (50%) 3607.31 6.16 0.01 13 EVI+Lion+Prey (75%) 3608.43 7.28 0.00 13 EVI+Lion+Prey (95%) 3608.96 7.81 0.00 13 Lion*Prey (50%) 3613.35 12.20 0.00 13 Prey (50%) 3617.96 16.81 0.00 9 Prey (75%) 3618.46 17.31 0.00 9 Prey (95%) 3618.49 17.34 0.00 9 Lion (75%) 3619.21 18.06 0.00 9 Null (75%) 3619.33 18.19 0.00 8

Null (50%) 3619.33 18.19 0.00 8

Null (95%) 3619.33 18.19 0.00 8 Lion+Prey (95%) 3619.97 18.82 0.00 11 Lion (95%) 3620.39 19.24 0.00 9 Lion (50%) 3620.62 19.47 0.00 9 Lion+Prey (75%) 3621.42 20.27 0.00 11 Lion+Prey (50%) 3621.51 20.36 0.00 11 Lion*Prey (75%) 3622.30 21.15 0.00 13 Lion*Prey (95%) 3622.84 21.69 0.00 11

154

APPENDIX B3

Effects of EVI on cheetah kill sites and locations where cheetahs were killed

We recorded the locations of cheetah kill sites (n = 1974), and sites that cheetahs were killed by other predators (n = 32), from 1996 – 2019. We located kill sites by following hunting animals, or through opportunistic sightings of carcasses or feeding animals. When we could determine the species of predator and prey, we recorded the location using a handheld GPS.

For each season, we generated random points equal to the number of kill site locations or the number of locations where cheetahs were killed. We extracted EVI values from the corresponding season, at each used site and random site. We used logistic regression models to assess the effect of EVI on the probability of use using the lm function in Program R (Version 3.5.3; R Core Team, 2019). We ran models separately for the cheetah kill site locations and the locations where cheetahs were killed. We evaluated if the EVI covariate was informative by calculating 85% confidence intervals, and considering it be informative if the confidence interval did not overlap zero (Arnold

2010). Locations of cheetah kill sites were negatively associated with EVI, with the highest probability of kills occurring in areas of low EVI. In contrast, there was no relationship between the locations where cheetahs were killed and EVI (βEVI = 3.78; 85%

C I= -1.47 to 9.27).

155

References

Arnold, T. W. 2010. Uninformative parameters and model selection using Akaike’s

Information Criterion. Journal of Wildlife Management 74:1175–1178.

R Core Team. 2019. R: A language and environment for statistical computing. R

Foundation for Statistical Computing, Vienna, Austria.

156

APPENDIX C1

Determining the best model structure to analyze multi-species responses to prescribed burning

We performed a pilot analysis to determine the best model structure to analyze multi-species responses to prescribed burning. Specifically, we tested the performance of two community N-mixture models (Yamaura et al., 2012) to determine if the addition of a detection covariate improved model performance. For both models, we modeled the abundance of species i at site j using a Poisson random variable based on expected abundance λi,j:

푁푖,푗~푃표푖푠푠표푛(휆푖,푗)

The expected abundance λi,j is based on our site-level covariates of interest:

log(휆푖,푗) = 훽0푖 + 훽1푖 ∗ 푇푟푒푎푡푚푒푛푡푗 + 훽2푖 ∗ 푃푒푟푖표푑푗 + 훽3푖 ∗ 푇푟푒푎푡푚푒푛푡푗 ∗ 푃푒푟푖표푑푗

Under this formulation, the β covariates in the model are species specific, but are drawn from group-level normal distributions with mean μ and variance σ2 for group g:

2 훽0푖,푔 ~ 푁표푟푚푎푙(휇훽0 푔, 휎 훽0푔 )

For the model including a detection covariate, we used the visibility index at each site during each time period (see Methods for full description). We modeled detection (p) such that the detection of species i, at site j, during replication k is modeled as:

157

푙표푔푖푡(푝푖,푗,푘) = 훼0푖 + 훼1푗푘 ∗ 푉푖푠푖푏푖푙푖푡푦

Similar to the abundance component of the model, the detection covariates were species- specific, but were drawn from a group-level normal distributions with mean μ and variance σ2 for group g:

2 훼0푖,푔 ~ 푁표푟푚푎푙(휇훼0 푔 , 휎 훼0푔 )

We ran the two models using Markov Chain Monte Carlo (MCMC) simulation in

JAGS (Plummer, 2003) using the jagsUI package (Kellner, 2015) for Program R (Version

3.5.3; R Core Team 2019). As priors, we used N(0, 0.1) for all mean parameters and unif(0, 3) for all standard deviation parameters. We ran the models using three parallel chains with a length of 50,000 after discarding 100,000 iterations for adaptation and burn-in, and a thinning rate of 5. We assessed model convergence using the Gelman-

Rubin statistic (Gelman, Carlin, Stern, & Rubin, 2013) and visual inspection of traceplots. We compared the two models using the deviance information criterion (DIC;

Spiegelhalter, Best, Carlin, & Van Der Linde, 2002).

Burning did affect visibility (Figure C1.1), with significantly higher site-level visibility at treatment sites post-burn compared to treatment sites pre-burn (p = 0.004), and control sites pre- (p < 0.001) and post-burn (p = 0.003). When comparing the model with the addition of visibility as a detection covariate, compared to the model without any

158 covariates on detection, the model without the inclusion of a detection covariate performed better (Table C1.1).

References

Gelman, A., Carlin, J. B., Stern, H. S., & Rubin, D. . (2013). Bayesian Data Analysis.

Boca Raton, FL: Chapman and Hall/CRC Press.

Kellner, K. F. (2015). jagsUI: A wrapper around rjags to streamline JAGS analyses.

Plummer, M. (2003). JAGS: A program for analysis of Bayesian graphical models using

Gibbs sampling. Proceedings of the 3rd International Workshop on Distributed

Statistical Computing (DSC 2003), 20–22. doi:10.1.1.13.3406

R Core Team. (2019). R: A language and environment for statistical computing. Vienna,

Austria: R Foundation for Statistical Computing.

Spiegelhalter, D. J., Best, N. G., Carlin, B. P., & Van Der Linde, A. (2002). Bayesian

measures of model complexity and fit. Journal of the Royal Statistical Society.

Series B: Statistical Methodology, 64(4), 583–616. doi:10.1111/1467-9868.00353

Yamaura, Y., Royle, J. A., Shimada, N., Asanuma, S., Sato, T., Taki, H., & Makino, S.

(2012). Biodiversity of man-made open habitats in an underused country: A class of

multispecies abundance models for count data. Biodiversity and Conservation,

21(6), 1365–1380. doi:10.1007/s10531-012-0244-z

159

Table C1.1. Model selection results for the best structural model to analyze multi-species responses to prescribed burning, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019.

Model DIC ΔDIC p(.) 186142 0

p(visibility) 186462 320

160

Figure C1.1. Average site visibility (mean ± SE) in relation to prescribed burning, Mun- Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019.

161

APPENDIX C2

Camera trap data summary

Table C2.1. List of species recorded during the camera trap survey, and number of photos per species, Mun-Ya-Wana Conservancy, KwaZulu-Natal, South Africa, 2018 – 2019.

Number of photos Common Name Scientific Name 2018 2019 Total Aardvark Orycteropus afer 63 70 133 Banded mungo 14 13 27 Bird spp Various 227 138 365 Black Rhino Diceros bicornis 30 36 66 African Buffalo Syncerus caffer 778 676 1454 Bushpig Potamochoerus larvatus 19 18 37 Caracal Caracal caracal 0 3 3 Chacma Baboon Papio ursinus 858 1755 2613 Cheetah Acinonyx jubatus 14 27 41 Domestic Canis lupus familiaris 3 0 3 Domestic Felis catus 12 0 12 Elephant Loxodonta africana 608 799 1407 Giraffe Giraffa camelopardalis 2947 1860 4807 Goat Capra aegagrus hircus 0 4 4 Greater Cane Rat Thryonomys swinderianus 0 1 1 Greater Galago Otolemur crassicaudatus 0 2 2 Greater Kudu Tragelaphus strepsiceros 1099 918 2017 Grey Duiker Sylvicapra grimmia 720 784 1504 Guineafowl Numida meleagris/Guttera pucherani 501 451 952 Hippopotamus Hippopotamus amphibius 62 36 98 Honey Badger Mellivora capensis 38 17 55 Impala Aepyceros melampus 7899 9067 16966 Large-spotted Genetta tigrina 557 349 906 Leopard Panthera pardus 223 204 427 Leopard tortoise Stigmochelys pardalis 7 4 11 Lion Panthera leo 355 149 504 Mountain Reedbuck Redunca fulvorufula 7 9 16 Nyala Tragelaphus angasii 5019 9994 15013 Porcupine Hystrix africaeaustralis 150 183 333 Red Duiker Cephalophus natalensis 139 815 954 Rock monitor Varanus albigularis 3 14 17

162

Scrub Hare Lepus saxatillis 228 74 302 Leptailurus serval 8 18 26 Side-striped Jackal Canis adustus 243 108 351 sanguinea 10 31 41 Spotted Hyaena Crocuta crocuta 196 272 468 Steenbok Raphicerus campestris 6 12 18 Ictonyx striatus 1 3 4 Suni Neotragus moschatus 2 17 19 Vervet Monkey Chlorocebus pygerythrus 322 344 666 Warthog Phacochoerus africanus 3310 1743 5053 Waterbuck Kobus ellipsiprymnus 83 111 194 White-tailed Mongoose Ichneumia albicauda 219 268 487 White Rhino Ceratotherium simum 740 529 1269 Lycaon pictus 1 4 5 Felis lybica 1 0 1 Wildebeest Connochaetes taurinus 2405 2289 4694 Zebra Equus quagga 7076 4613 11689

163