<<

University of Southampton

16th August 2016

FACULTY OF NATURAL AND ENVIRONMENTAL SCIENCES CENTRE FOR BIOLOGICAL SCIENCES

RESOURCE PARTITIONING BETWEEN SPOTTED HYENAS (CROCUTA CROCUTA) AND (PANTHERA LEO)

Arjun Dheer

A dissertation submitted in partial fulfilment of the requirements for the degree of M.Res. Wildlife Conservation.

1

As the nominated University supervisor of this M.Res. project by Arjun Dheer, I confirm that I have had the opportunity to comment on earlier drafts of the report prior to submission of the dissertation for consideration of the award of M.Res. Wildlife Conservation.

Signed………………………………….. UoS Supervisor’s name: Prof. C. Patrick Doncaster

As the nominated Marwell Wildlife supervisor of this M.Res. project by Arjun Dheer, I confirm that I have had the opportunity to comment on earlier drafts of the report prior to submission of the dissertation for consideration of the award of M.Res. Wildlife Conservation.

Signed…………………………………… MW Supervisor’s name: Dr. Zeke Davidson

2

Abstract The negative impact of anthropogenic activities on wildlife has led to protected areas being set aside to prevent human-wildlife conflict. These protected areas are often small and fenced in order to meet the needs of expanding human communities and to conserve wildlife. This creates challenges for the management of wide-ranging such as large carnivores, especially those that compete with one another for limited resources. This study focused on resource partitioning between GPS-GSM collared spotted hyenas (hereafter referred to as hyenas) and lions in Lewa Wildlife Conservancy (LWC) and Borana Conservancy (BC), . Scat analysis revealed that hyenas and lions show a high degree of dietary overlap, though hyenas have broader diets and feed on species, which lions completely avoid. Spatially, hyenas show stronger intraspecific avoidance and more exclusive territorial behaviour than do lions. Hyenas and lions have a high degree of spatial overlap, though lions may influence den site selection in hyenas. Both species are heavily nocturnal and crepuscular, though hyenas tend to travel significantly further at night than do lions. Hyenas and lions both display mixed results in dynamic spatiotemporal interactions, with 40% of hyena- pairs showing attraction and 70% showing simultaneous use of overlapping areas. All but one inter-clan hyena pairs show strong avoidance, though lion pairs were not as mutually repulsive. This shows that hyenas and lions may use one another as sources of food and that scavenging and kleptoparasitism likely play a role in their dynamic. The hyena population is suggested to be growing and healthy, though the lion population is of concern due to lower density, isolation, and low recruitment. Further investigation into human-carnivore conflict within surrounding communities, long-term demographic and behavioural trends of all members of the large carnivore guild, and the potential development of a dispersal-based metapopulation management scheme will allow for the continued persistence of large carnivores in the Lewa-Borana Landscape (LBL) and their coexistence with human communities.

Target journal: African Journal of Ecology

3

Table of Contents

1. Introduction ...... 9 1.1 Background ...... 9 1.2 The value of large carnivores ...... 9 1.3 Resource partitioning between carnivores ...... 10 1.4 Carnivores in small reserves ...... 11 1.5 LWC and BC ...... 11 2. Aims and objectives ...... 13 3. Materials and methods ...... 14 3.1 Study area ...... 14 3.1.1 Location ...... 14 3.1.2 Climate and biodiversity ...... 14 3.1.3 Communities and human influence ...... 15 3.2 Data collection ...... 18 3.3 Capture and collaring...... 18 3.3.1 Hyenas ...... 18 3.3.2 Lions ...... 21 3.4 Dietary partitioning ...... 22 3.4.1 Prey selection ...... 23 3.4.2 Dietary overlap ...... 23 3.4.3 Niche breadth ...... 24 3.5 Spatial partitioning ...... 24 3.5.1 Overlap ...... 25 3.5.2 Den activity ...... 25 3.5.3 Den buffers ...... 26 3.5.4 Community land use ...... 26 3.6 Temporal partitioning ...... 26 3.6.1 Activity budgets ...... 26 3.6.2 Dynamic analysis ...... 27 3.7 Demographics ...... 28 3.7.1 Hyenas ...... 28 3.7.2 Lions ...... 29 4. Results ...... 30 4.1 Dietary partitioning ...... 30 4.1.1 Raw proportions ...... 30 4.1.2 Relative proportional contribution...... 30 4.1.3 Jacobs’ index ...... 32 4.1.4 Pianka’s index ...... 33

4

4.1.5 Levins’ index ...... 33 4.2 Spatial partitioning ...... 34 4.2.1 Home range and core sizes ...... 34 4.2.2 Overlap ...... 37 4.2.3 Den activity ...... 41 4.2.4 Den buffers ...... 42 4.2.5 Community land use ...... 45 4.3 Temporal partitioning ...... 46 4.3.1 Activity budgets ...... 46 4.3.2 Dynamic analysis ...... 48 4.4 Demographics ...... 52 4.4.1 Identified individuals ...... 52 5. Discussion ...... 54 5.1 Collaring ...... 54 5.2 Dietary partitioning ...... 55 5.3 Spatial partitioning ...... 56 5.4 Temporal partitioning ...... 57 5.5 Demographics ...... 58 5.6 Conclusions and management recommendations ...... 59 Acknowledgments ...... 62 References ...... 64 Appendix A ...... 74 Appendix B ...... 75 Appendix C ...... 76

5

List of figures

Figure 3.1 Large scale map of LBL within Kenya………………………………………………..16

Figure 3.2 Small scale map of LBL and points of interest………………………………………...17

Figure 3.3 Leg-hold snare trap set for hyenas…………………………..…………...... 19

Figure 3.4 CH01 recovering post-immobilisation………………………………………………..20

Figure 3.5 DL01, the male lion in this study……………………………………………………...21

Figure 3.6 Examples of hyena and lion scats in the field………………………………………….23

Figure 3.7 Example of unique left and right spot patterns on BH01……………………………....29

Figure 4.1 Relative proportional contribution of prey items to hyena and lion diets……………...31

Figure 4.2 Jacobs’ index values………………………………………………………….……….33

Figure 4.3 Mean home range and core sizes……………………………………………………...34

Figure 4.4 Map of hyena and lion cores…………………………………………………..…...... 35

Figure 4.5 Map of hyena and lion home ranges………………………………………..………....36

Figure 4.6 Map of hyena home ranges and cores based on clan membership………………...…...39

Figure 4.7 Map of lion home ranges and cores based on pride membership……………………...40

Figure 4.8 Den distances to intraspecific and interspecific home ranges and cores…………...... 41

Figure 4.9 Den buffers in relation to hyena home ranges and cores…………………...……...... 43

Figure 4.10 Den buffers in relation to lion home ranges and cores………………………...……...44

Figure 4.11 Community use by hyenas and lions……………………………………………..…..45

Figure 4.12 Proportional hyena and lion activity budgets by time windows…………………...…46

Figure 4.13 24-hour chart of hyena and lion mean distance travelled…………………..……...... 47

Appendix A Trapping sites for hyenas and lions…………………………………………….…...74

Appendix C Collar deployment progression………………………………………….……...... 76

6

List of tables

Table 3.1 Collared hyenas by sex and clan membership………………………………………...22

Table 3.2 Collared lions by sex and pride membership…………...…………………………...... 22

Table 4.1 Raw frequencies of prey hairs in hyena and lion scats……………………………...... 30

Table 4.2 Relative proportional contribution………………………………………………...... 31

Table 4.3 Jacobs’ selection index and standard error values………………………………...... 32

Table 4.4 Levins’ index of niche breadth……………………………...………………………...33

Table 4.5 Average percent overlap based on intraspecific and interspecific relationships……...37

Table 4.6 Mann-Whitney U test results for overlap……………………………………………..38

Table 4.7 Average minimum distances based on den activity…………………………………...41

Table 4.8 Frequency of occurrence of hyena den buffers to cores and home ranges…………....42

Table 4.9 Average distances travelled at four separate time windows……………...…………...46

Table 4.10 Doncaster’s test results for hyena-lion pairs…………………………………………49

Table 4.11 Doncaster’s test results for hyena pairs……………………………………………...50

Table 4.12 Doncaster’s test result for a lion pair…………………………………………...... 50

Table 4.13 Minta’s test result for interspecific and intraspecific pairs…………………………..51

Table 4.14 Demographics of identified individuals by sex and age class……………………….53

Table B.1 Sample collaring data sheet………………………………………...………………....75

7

List of abbreviations

BC Borana Conservancy

CSV Comma-Separated Values

GIS Geographic Information System

GPS Global Positioning System

GR

IUCN International Union for Conservation of Nature

KWS Kenya Wildlife Service

LWC Lewa Wildlife Conservancy

NRT Northern Rangelands Trust

8

1. Introduction

1.1 Background

Global expansion of human populations has led to wildlife populations becoming increasingly confined to protected areas due to loss. Loss of habitat in turn has resulted in significant biodiversity loss and increased rates across taxa (Brooks et al. 2002). Over the past 500 years, it is estimated that human-driven extinction rates exceed those in the fossil record several hundred times (Dirzo and Raven 2003). Biodiversity is a driver of ecosystem change, resulting in reduction of natural resources available for human use. Loss of biodiversity in turn limits ecosystem services and function and results in less productive ecological communities (Cardinale et al. 2012).

In response, protected areas are set aside in the face of burgeoning human populations for conservation purposes and to meet the needs of expanding human communities. There is an increasing trend worldwide for these protected areas to be fenced to limit human-wildlife conflict (Packer et al. 2013). However, meta-analyses of protected areas reveal that relying only on large protected areas can have shortcomings in their effort to conserve biodiversity due to lack of funding, research on their wildlife populations, and administrative disorganisation (Mora and Sale 2011). A more holistic approach that includes connected smaller reserves with human cohabitation and direct oversight and management is a more plausible scenario for the future (Wackernagel et al. 2002).

This presents challenges for the management of wide-ranging animals that inherently exist at low densities, such as large carnivores, especially in these small reserves close to human habitation (Treves and Karanth 2003). Increased close management of such protected areas that can support viable populations of these predators appears to be the most viable conservation strategy for the future (Davies-Mostert et al. 2015) and can allow for the coexistence of human and carnivore populations.

1.2 The value of large carnivores

Apex predators exert top-down effects on lower trophic levels and allow for the proper functioning of natural systems. They regulate prey populations and in turn promote healthy

9 biomass of producers, which form the basis of a food web (Miller et al. 2001). The presence of large carnivores is an indicator of the ecological integrity of an ecosystem (Schmitz et al. 2000). They serve as conservation umbrellas for species occupying lower trophic levels and provide economic benefits due to their appeal to tourists (Linnell et al. 2000). While controversial, regulated trophy hunting of large carnivores in southern Africa has also provided economic benefits for local economies, including conservation charities (Lindsey et al. 2006).

However, large carnivore populations worldwide have declined as human populations have grown (Weber and Rabinowitz 1996). As they occupy the top trophic levels of their ecosystems, they are particularly sensitive to anthropogenic influences (Woodroffe and Ginsberg 1998). Understanding the ecological needs of large carnivores is therefore necessary to apply appropriate management policies for the benefit of protected areas. Behavioural studies, including analyses of diet and spatiotemporal activity, as well as demographic studies for population management, allow for the development of such policies. East Africa remains one of the last strongholds for large carnivores left on Earth, though there is a general population decline in all species of large carnivore in the region (Packer et al. 2013, Schuette et al. 2013).

1.3 Resource partitioning between carnivores

Research on large carnivore ecology in East Africa has been ongoing since the middle 20th century (Dalerum et al. 2008). Understanding competition and niche separation between carnivore species is needed to determine the roles that different members of the guild have. This in turn will allow for management decisions to be made at the species level, especially when study species are of conservation concern (Kamler et al. 2015). Carnivores that occupy similar niches coexist through resource partitioning (Kamler et al. 2012, Vieira and Port 2006). For example, in , bat-eared foxes, cape foxes, and black-backed jackals display dietary and spatiotemporal resource partitioning (Kamler et al. 2012). In , , hyenas altered their foraging behaviour and prey selection in response to an increase in lion population (Periquet et al. 2015). In northern , and African wild dogs display temporal and spatial avoidance of hyenas and lions, which is hypothesised to facilitate coexistence (Cozzi et al. 2012). In Matusadona National Park, Zimbabwe, lions and hyenas again displayed dietary partitioning (Purchase 2004). A previous literature analysis reveals that hyenas generally have a larger dietary niche breadth than lions, which is unsurprising, considering their opportunistic feeding strategy (Periquet et al. 2014).

10

1.4 Carnivores in small reserves

The behavioural ecology of large carnivores with limited ranging ability living in proximity to humans is poorly understood (Yirga et al. 2012). The studies listed above have taken place in large, unfenced areas and provide information on carnivore ecology within more pristine but do not address the pressure that humans are applying to wildlife today. Given the trend of human population growth and habitat destruction, humans and carnivores are increasingly coming into conflict (Treves and Karanth 2003). Small reserve management with respect to large carnivores is therefore moving to the forefront of wildlife conservation. By understanding the behaviour of carnivores in such areas, policies can be implemented to curb conflict and allow for improved coexistence in the future as human populations expand. Such policies can include metapopulation management, culling, reintroductions, and translocations (Akçakaya et al. 2006).

1.5 LWC and BC

Lewa Wildlife Conservancy (LWC) and Borana Conservancy (BC) in Kenya serve as exemplars of small, fenced, managed protected areas. At LWC and BC, lions and spotted hyenas are the two dominant predators in terms of numerical abundance and behavioural interactions. Research on lions in LBL has been ongoing for several years. However, there is a major gap in knowledge on the influence that hyenas have on prey and competitors alike in the study area. Hyenas display remarkable behavioural plasticity over much of their range; they can adopt a variety of feeding strategies that range from almost exclusive scavenging, as in peri-urban areas in northern (Yirga et al. 2012) to hunting up to 95% of the time, as in the Ngorongoro Crater, (Kruuk 1972). Furthermore, while the presence of the fission-fusion clan system is ubiquitous across their range in sub-Saharan Africa, clan size can vary from under 10 in the Kalahari (Mills 1989) to over 70 in the prey-rich grasslands of East Africa (Kruuk 1972). Territoriality shows a similar pattern: loosely defined territories in the Kalahari and Namib deserts average over 1,000 km2 while territory sizes in the Ngorongoro Crater are generally below 40 km2 (Gittleman 1989). This flexibility has contributed to their success – hyenas are the most abundant large carnivores in Africa today despite being heavily persecuted and lacking the charisma, scientific research interest, and public appeal of canids and felids (Glickman 1995).

Preliminary estimates have shown that the hyena population is between 38-78 individuals at LWC, while the lion population is known to be 17. Lions are therefore under carrying capacity 11

(28) through all four abundance measures used in a recent demographic study (Preston 2015) whereas hyenas may be past the estimated threshold (50). In neighbouring BC, carrying capacities for lions and hyenas are estimated at 22 and 36, respectively, while the lion population is 17 and estimated at 31-57 for hyenas (Preston 2015). Ongoing lion monitoring at the conservancies has revealed that the lion population between the two conservancies is 34, but the actual number of hyenas remains unknown. Hyena dietary and spatiotemporal tendencies are also unknown.

This study addresses the lack of knowledge on large carnivore behavioural ecology and resource partitioning in small, fenced reserves. It also gathers foundational information on hyena demography and space use in LBL. Finally, it will provide management recommendations for large carnivores in small, fenced reserves in close proximity to human communities and evaluate the translatability of such recommendations to a wider audience. In turn, these recommendations will allow for improved policy making and mitigation of human-carnivore conflict in East Africa and beyond.

12

2. Aims and objectives

The aim of this project is to better understand the behavioural ecology of hyenas and lions in the Lewa-Borana Landscape (LBL) and to determine the role of small reserves in large carnivore conservation. The major outcome is to inform management decisions and policies at LWC and BC that develop new solutions to conservation challenges in small, fenced reserves. This overarching aim is met through an investigation of resource partitioning between hyenas and lions at dietary, spatial, and temporal levels and data collection on their demography.

At the dietary level, the objective is to determine prey selection in hyenas and lions, dietary overlap between the species, and niche breadth for both species. A compilation of dietary studies across several countries revealed a 58.6% actual prey species overlap and a 68.8% preferred prey species overlap between the two (Hayward 2006). This is exceptionally high. It is unusual that two apex predators can occupy such similar niches and still coexist. This study will explore whether there is similar extensive dietary overlap or some degree of partitioning.

At the spatial level, the objective is to investigate geographic segregation between hyenas and lions in the study area at the home range and core levels and determine how lion presence affects denning behaviour in spotted hyenas. Denning behaviour in hyenas is affected by lion proximity and presence in Kenya and Zimbabwe (Kolowski and Holekamp 2009, Periquet et al. 2014).

Temporally, both species are considered nocturnal (Kiffner et al. 2008). This project will determine if there is partitioning in the activity times of spotted hyenas and lions. If there is, it may explain their ability to coexist while competing for limited resources (Hayward and Hayward, 2006). Previous studies have demonstrated temporal avoidance amongst large carnivores, especially subordinate species within the large carnivore guild such as African wild dogs and cheetahs (Cozzi et al. 2012).

Foundational knowledge on hyena and lion demography within the study area will be gathered. This information will be gleaned as part of the above three objectives. Furthermore, this information will set the groundwork for future studies on this species in the area and further demographic research on the large carnivore guild.

13

3. Materials and methods

3.1 Study area

3.1.1 Location

This study was carried out in Lewa Wildlife Conservancy (LWC) and Borana Conservancy (BC) in central Kenya, straddling the border of Meru and Laikipia counties (00 06′ to 00 17′ North, 370 21′ to 370 32′ East) (Figure 3.1). The total area of this site is 375 km2 (LWC = 250 km2; BC = 125 km2). The conservancies have fences along their external boundaries, but feature migratory gaps to allow for movement in and out of the conservancies (Dupuis- Desormeaux et al. 2015).

3.1.2 Climate and biodiversity

The climate in LBL is classified as a tropical savanna climate (Aw) under the Koppen Climate Classification system (Peel et al. 2007). LBL experiences two wet seasons, the long rains in March to May, and the short rains from October to December, although periods of drought are relatively common (Lewa Wildlife Conservancy 2014).

Four main habitat types occur in LWC and BC. Based on the dominant plant species, they include forest (dominated by Juniperus-Olea forest), plains (dominated by Pennisetum grasses and Acacia trees), hills and rocky outcrops (dominated by Acacia, and Grewia), and riverine habitat (Mwololo 2011). LBL is situated between montane forests to the south that give rise to Mount Kenya and lowland grasslands to the north.

Despite its small geographic area, from a conservation perspective, LBL is home to strategically important populations of threatened species such as Grevy’s ( grevyi), black (Diceros bicornis), and (Ceratotherium simum) (Low et al. 2009). Apart from dense herbivore populations, LWC and BC contain the full suite of species from the East African large carnivore guild: lions (Panthera leo), spotted hyenas (Crocuta crocuta), leopards (Panthera pardus), cheetahs (Acinonyx jubatus), striped hyenas (Hyaena hyaena), and African wild dogs (Lycaon pictus).

14

3.1.3 Communities and human influence

Human communities reside permanently both in and around LBL (Figure 3.2) and therefore come into close contact with wildlife. Intensive agriculture and pastoralism are primary sources of income for local communities (Nyaligu and Weeks 2013). Herds of cattle, sheep, and goat are grazed in controlled plots within the conservancies and graze freely within community lands. This controlled grazing within the conservancies is carried out to control low-nutrition grass species such as Pennisetum stramineum and P. mezianum (Lewa Wildlife Conservancy 2006) and in turn benefit grazing herbivores with improved grassland biodiversity. Several photographic tourism lodges also operate within LWC and BC and are a primary source of employment and economic activity in the area (Gaitho 2014).

15

Figure 3.1 – Location of LBL within Kenya.

16

Figure 3.2 – Map of LBL with points of interest.

17

3.2 Data collection

The bulk of data collection took place between February and May 2016, but scat data and prey count data from 2014, 2015, and 2016 were used to assess long-term prey selection. Dietary separation was determined using scat analysis based on a reference collection of hairs (Mukherjee et al. 1994). Dietary partitioning focused on prey species relative contribution, prey selection, dietary overlap, and niche breadth.

Spatial separation was determined through point and area overlap of core and home ranges for the two species as well as den site selection by hyenas (Kamler et al. 2013). Data from collars were downloaded as CSV (Comma-Separated Values) files and exported to ArcMap 10.4 for analysis.

Temporal separation was determined through a comparison of activity budgets and analysis of concurrent spatial avoidance between sympatric and non-sympatric pairs (Kitchen et al. 1999, Doncaster 1990, Minta 1992).

Demographic data were collected through camera trapping at known hyena den sites, baited traps, and opportunistic photographing of encountered individuals.

The confidence interval considered for analyses was 95%; therefore, p<0.05 demonstrated significance. All statistical analyses were performed in R statistical software (R Core Team 2014) and ArcMap 10.4 (ESRI 2014).

3.3 Capture and collaring

3.3.1 Hyenas

A total of seven adult hyenas were collared with GPS-GSM collars within the study area, representing five distinct clans (Table 3.1). Collars were manufactured by Savannah Tracking, a Kenya-based telemetry and radio-tracking company.

Adults were targeted because hyenas’ necks grow throughout their lives; collaring a young hyena can put the animal in discomfort as it grows. These collars were deployed between 23/02/2016 and 27/03/2016 (Appendix C). 18

Figure 3.3 – Leg-hold snare trap for hyenas.

Free darting of hyenas at LBL was not possible due to their skittish behaviour. Hyenas were captured using leg snares (n=1; Figure 3.3) and cage traps (n=6; Figure 3.4) with bait. Cages were built by LWC’s workshop in advance of the trapping effort and were made of steel mesh, measuring 3 m x 1.5 m x 1 m. Once baited and set, cage floors were padded with plywood boards and thick, dead grass. Cages were also kept under bushes or trees to provide shade in the event of a hyena being successfully trapped. Bait was purchased from local markets or acquired opportunistically from carcasses. Prey species used for bait included cattle, goat, camel, eland, and sheep. Bait was dragged roughly 100 meters from intended capture sites to form scent trails. As hyenas are olfactory foragers (Woodmansee et al. 1991), this technique proved to be essential. Trapping sites were either within 500 m of communal dens or in areas where rangers frequently sighted hyenas. As hyenas were collared, the collaring effort was expanded outwards to increase the probability of collaring conspecifics from different clans (Appendix A).

19

Figure 3.4 – CH01 recovers after immobilisation and collaring following capture with a baited cage trap in LWC.

In accordance with Kenya Wildlife Service (KWS) policy, veterinarians were present to dart and treat the study animals. Hyenas were immobilized with a combination of ketamine hydrochloride and medetomidine hydrochloride, collared, and checked for body condition and dimensions (Appendix B). They were then reversed with atipamezole. Hyenas were recumbent in an average of 12.1 minutes (range: 8.0-30.0) and recovered in an average of 17.5 minutes (range: 6.0-30.0).

One mortality occurred during the study period. SH01’s collar was found along with her almost entirely consumed carcass near the western boundary of her clan’s territory. The thoroughness of her consumption and the presence of three large hyena scats and hyena spoor in the vicinity of her carcass make it likely that she was killed and cannibalised by conspecifics from the neighbouring Borana clan. As she had previously been re-collared, a significant amount of potential data was lost, resulting in there being less than 14 days’ worth of data from her collar. Her collar was re-deployed on SH02, a clan member whose data was used in this study. Therefore, SH01’s data have been excluded from all analyses in this report.

20

3.3.2 Lions

Long-term information on lion ecology has been collected by the research department at LWC for several years and lions have been the focal species for previous Master’s-level projects at the field site. They are also a primary tourist target at LWC and BC. Accordingly, lions in the area are considered well-habituated. Free-darting was therefore used.

A total of three GPS-GSM collars were deployed on adult lions within the study area, representing three distinct prides (Table 3.2). These collars were deployed between 26/02/2016 and 29/04/2016 (Appendix C).

Lions were recumbent in an average of 8.7 minutes (range: 11.0-15.0) and recovered in an average of 40.3 minutes (range: 12.0–66.0). No mortalities occurred for collared lions during the study period. However, one lion collar, deployed on SL01, stopped working on 03/05/3016 and a new collar was deployed on the same lioness on 18/05/2016. Analysis of her data includes information before and after her re-collaring but excludes the hiatus period.

Figure 3.5 – DL01, the male lion in this study, with unique right and left whisker spot patterns.

21

Table 3.1 – Collared hyenas by sex and clan membership. Hyena ID Sex Clan BH01 Male Borana CH01 Male Charlie NH01 Female Nala NH02 Male Nala SH02 Female Shamba UH01 Male Utalii SH01 Female Shamba

Table 3.2 – Collared lions by sex and pride membership. Lion ID Sex Pride DL01 Male Dalma SL01 Female Sarah WL01 Female Western

3.4 Dietary partitioning

Determination of dietary separation was based on scat analysis (Mukherjee et al. 1994). Scats were identified (Figure 3.6) and collected in the field at kill sites, rest sites, hyena latrines, and opportunistically. One scat was collected per site to prevent pseudo-replication from occurring. Scat was placed in a plastic bag, dried, and treated with a mixture of 150 ml 75% ethanol solution and 350 ml boiled water. 20 hairs were extracted per scat, mounted, and dried overnight. Scats were then examined under microscope using analysis of hair roots. Each hair was compared to reference collections present at LWC for species identification. Frequencies of different prey items were recorded and tabulated for each scat sample. 118 hyena scats and 87 lion scats were used for this analysis.

22

Figure 3.6 – Examples of hyena (L) and lion (R) scats identified in the field. The white coloration, chalky texture, and bulky size of hyena scat is diagnostic.

Raw and relative (# F.O. species n / # occurrences for all species) contributions were calculated for further analysis of selection, overlap, and breadth. Relative contribution of prey species standardises the data to add up to 100% and is useful for determining how much a given prey item contributes to a predator’s diet (Lyngdoh et al. 2014).

3.4.1 Prey selection

Jacobs’ selection index (Jacobs 1974) was used to assess prey selection:

r is the proportion of kills or scat samples of a given prey species; p is the proportional availability of the given species.

The value can range from -1.0 to 1.0 depending on the degree of avoidance or selection, respectively. This method uses abundance data on annual game counts from the LWC Research Department from years 2014, 2015, and 2016 for scats from these respective years. Livestock counts were not available so these have been excluded from the analysis.

3.4.2 Dietary overlap

Pianka’s index (Pianka 1974) was used to assess dietary overlap:

23

Pij is the relative proportion of prey species i in predator species j scat samples;

Pij is the relative proportion of prey species i in predator species k scat samples.

This method results in a standardized output from 0.0 (no overlap) to 1.0 (complete overlap).

3.4.3 Niche breadth

Levins’ measure of niche breadth (MacArthur and Levins 1967, Wallace Jr 1981) was used to determine the diversity of their diets:

Y is the total number of individuals sampled;

Nj is the number of individuals using prey species j.

The raw Levins’ measure was then further standardised to a 0.0-1.0 scale:

n is the number of possible prey species.

3.5 Spatial partitioning

Analysis of spatial segregation included multiple spatial and mapping techniques to compare geographic behaviour in the species. All analyses were completed in ArcMap 10.4. Areas of

24 interest, such as hyena dens, were marked via a GPS device and uploaded to ArcMap base layer maps.

Collars were set to collect one fix every hour over a 24-hour period and report every six hours at 03:00, 09:00, 15:00, and 21:00. A total of 15,912 hyena fixes (range: 2,460–3,025) and 7,041 lion fixes (range: 1,619–3,227) were collected and used for analysis during the study period. Overall, deployed GPS-GSM collars reported fixes 92.1% of the time (range: 79.0%-97.0%). Hyena collars averaged 90.4% report success whereas lion collars averaged 96.2% report success.

3.5.1 Overlap

Home ranges and cores were mapped out for the different species. Cores use the 50% Kernel Density Estimate (KDE) method and home ranges use the 95% KDE. Home ranges and cores were determined on ArcMap using the Home Range Tools extension (Rodgers et al. 2015) and sizes (km2) were compared for the two species using the Mann-Whitney U test. They were then analysed as follows:

(i) Core point and area overlap percentage; (ii) Home range point and area overlap percentage; (iii) Core-home range area overlap percentage.

These overlap percentages for interspecific competitors were further compared to intraspecific overlap between hyena pairs and lion pairs from different social groups – five hyena clans and three lion prides using the Mann-Whitney U test:

n1 is sample size 1;

n2 is sample size 2;

R1 is the rank of the sample size.

3.5.2 Den activity

25

Active vs. inactive minimum den distances were compared to one another using the Mann- Whitney U test for lion cores and home ranges, inter-clan hyena cores and home ranges, and intra-clan hyena cores and home ranges. This was done in order to determine if den sites chosen by hyenas during the study period were influenced by proximity to intraspecific or interspecific competitors.

3.5.3 Den buffers

500 m, 1000 m, and 2000 m ring buffers were placed around each hyena den to further examine spatial relationships between den site selection and intra-clan home ranges and cores, inter-clan home ranges and cores, and lion home ranges and cores. Comparisons were drawn between the six active and 10 inactive hyena dens to determine if interspecific or conspecific pressure play a role in den site choice. A total of 16 den sites were included in this analysis, six of which were confirmed to be active during the field season and 10 of which were inactive.

3.5.4 Community land use

Use of non-protected community land by hyenas and lions was compared in order to gauge spatial avoidance of human habitation. Mean percentages of home range and core areas overlapping with communities were compared between the species using the Mann-Whitney U test.

3.6 Temporal partitioning

3.6.1 Activity budgets

Activity budgets were developed based on data from the GPS-GSM collars, which give location fixes every hour. GPS fixes were uploaded to ArcMap and distances measured between each successive fix. Data from all six hyena collars and all three lion collars were used in this analysis. Mean distance travelled (meters/hour) was calculated for both species for every one- hour window.

Mean distance travelled per hour was then consolidated into time windows: dawn (05:00-09:00), day (09:00-17:00), evening (17:00-21:00), and night (21:00-05:00) in order to determine if there

26 were differences in diurnal, nocturnal, or crepuscular activity patterns. These distances were compared between species using Welch’s two-sample t test (Kitchen et al. 1999).

3.6.2 Dynamic analysis

Spatiotemporal avoidance of lions by hyenas was analysed by comparing the concurrent mean distance between hyena-lion sympatric pairs to the distance expected by chance using dynamic analysis of mutual interaction (Doncaster 1990, Minta 1992). The proportion of simultaneous fixes that were below a critical distance of 200 m and a time window of 30 minutes was compared to that which would be expected based on the distribution of distances between all fixes using the widlifeDI package in R. This produced a contingency table of paired and non- paired fixes with an associated p-value from a Chi-squared test (Long 2014).

Of the 18 possible hyena-lion pairs, 10 were sympatric and therefore used in this analysis. The eight pairs with 0% overlap were excluded from the analysis, as their movements with respect to one another were assumed to be random. Pairs were classified into two groups: those with significant attraction based on Doncaster’s non-parametric test of interaction (p<0.05) and those without (p>0.05). Home range overlap percentages between the two groups were then analysed to determine if the group which displayed significant spatiotemporal attraction had a greater mean overlap than the group with random movements. This was done through a Mann-Whitney U test.

Of the 36 possible hyena pairs, five had any degree of home range overlap and could be used for this analysis, four of which were pairs from different clans. Clan mates NH01 and NH02 were expected to display attraction within their home ranges whereas inter-clan hyenas were expected to display avoidance behaviour.

Of the three possible lion pairs, all three were sympatric. However, while WL01 is sympatric with both SL01 and DL01, there were zero paired events, whether temporally matched or not. Analysis was only run between SL01 and DL01, as they are sympatric and featured matched fixes. They were accordingly expected to display avoidance behaviour. Therefore, SL01-DL01 was the only instraspecific lion analysis possible with Doncaster’s test.

Minta’s test for spatial and temporal interaction was also used for dynamic analysis, as it accounts for differences between shared or overlapping home range areas and non-overlapping areas with respect to space use (Long 2014, Minta 1992, Darnell 2014). Home ranges for a 27 sympatric pair are divided into areas used by individual A (LAA), individual B (LBB), and both. The number of fixes in each area are tested against expected values based on the probability of finding the animal in a given area based on real overlap percentages (Long 2014). LAA or LBB values greater than 0 indicate attraction to the shared area while negative values indicate avoidance. The Lixn statistic indicates simultaneity of use; a positive value indicates mutual attraction to the shared area while a negative value indicates avoidance.

In Minta’s test, the test statistics LAA and LBB are calculated from a 2x2 contingency table, from which a chi-squared value can be calculated for significance. Minta’s test was run for all 10 hyena-lion sympatric pairs, all five sympatric hyena pairs and the three sympatric lion pairs. This technique used rgeos, wildlifeDI, and adehabitatHR packages in R.

3.7 Demographics

3.7.1 Hyenas

Prior to the field season, informal interviews with local rangers and scouts provided information on known hyena dens. Den sites were visited on foot, marked with a GPS device, and camera traps were placed to monitor activity. A photographic database of individual hyenas based on left and right side spot patterns was developed from photos at dens and bait sites. If hyenas were sighted opportunistically, they were photographed and identified.

Hyenas were assigned to a specific clan based on the following criteria:

(i) If the hyena was observed at the clan’s communal den; (ii) If the hyena was observed with a hyena that has been assigned to a given clan; (iii) If the hyena was observed within the core area of a collared hyena’s territory.

The database is split into five sections based on clan membership. Age classes were classified into adult, sub-adult, and cub based on body size and pelage.

28

Figure 3.7 – Example of unique left and right spot patterns on BH01.

3.7.2 Lions

As part of its ongoing predator monitoring work, LWC already has an identification database of the 34 lions within the LBL population. Lions were identified based on whisker spot patterns and the photos within the database were updated when individuals were sighted.

29

4. Results

4.1 Dietary partitioning

4.1.1 Raw proportions

For both species, pains zebra was the most frequently counted prey item in scats (n=112, 70 for hyenas and lions, respectively). For hyenas, impala (n=99) and Grevy’s zebra (n=73) were the second and third most frequently counted prey species. For lions, Grevy’s zebra (n=60) and eland (n=50) were second and third most frequent, respectively (Table 4.1). Zero lion scats contained livestock (cattle, sheep, and goat) hairs while they were counted in 93 hyena scats (78.8% of hyena scats).

Table 4.1 – Raw frequencies of prey hairs in hyena and lion scats. Hyena Lion Species # Scats (n=118) Proportion # Scats (n=87) Proportion Plains zebra 112 0.95 70 0.81 Grevy's zebra 73 0.62 60 0.69 Warthog 10 0.09 22 0.25 Buffalo 53 0.45 25 0.29 Impala 99 0.84 36 0.41 Kudu 0 0.00 4 0.05 57 0.48 34 0.39 Oryx 13 0.11 4 0.05 Waterbuck 45 0.38 6 0.07 Eland 39 0.33 50 0.58 Cattle 36 0.31 0 0.00 Sheep 29 0.25 0 0.00 Goat 28 0.24 0 0.00

4.1.2 Relative proportional contribution

Based on relative frequency of occurrence, the hyenas’ diet consists primarily of plains zebra (18.9%), impala (16.7%), and Grevy’s zebra (12.3%) while the lions’ diet consists primarily of plains zebra (22.5%), Grevy’s zebra (19.3%), and eland (16.1%). Livestock species altogether contribute to 15.7% of the hyenas’ diet and 0.0% to the lions’ diet (Table 4.2, Figure 4.1). 30

Table 4.2 – Relative proportional contribution. Prey Species Hyena Lion Plains zebra 0.19 0.23

Grevy's zebra 0.12 0.19

Warthog 0.02 0.07 Buffalo 0.09 0.08 Impala 0.17 0.12 Kudu 0.00 0.01 Giraffe 0.10 0.11 Oryx 0.02 0.01

Waterbuck 0.08 0.02

Eland 0.07 0.16 Cattle 0.06 0.00 Sheep 0.05 0.00 Goat 0.05 0.00

0.25

0.20

0.15

Proportion 0.10

0.05

0.00

Hyena Lion

Figure 4.1 – Relative proportional contribution of prey species to hyena and lion diets.

31

4.1.3 Jacobs’ index

Jacobs’ index values reveal selection of five prey species each for both hyenas and lions. Hyenas select for waterbuck, giraffe, Grevy’s zebra, eland, and warthog, and against kudu, buffalo, plains zebra, oryx, and impala. Lions select for warthog, eland, Grevy’s zebra, giraffe, and kudu and against buffalo, oryx, impala, waterbuck, and plains zebra (Table 4.3, Figure 4.2). Of the 10 prey species, hyenas and lions shared selection in eight cases; the only differences were in kudu (hyena negative, lion positive) and waterbuck (hyena positive, lion negative).

Table 4.3 – Jacobs’ index and standard error values.

Species Hyena Lion Plains zebra -0.19 ±.02 -0.19 ±.07 Grevy's zebra 0.38 ±.01 0.51 ±.09 Warthog 0.02 ±.06 0.59 ±.17 Buffalo -0.43 ±.07 -0.55 ±.11 Impala -0.03 ±.01 -0.34 ±.02 Kudu -1.00 0.37 ±.18 Giraffe 0.40 ±.10 0.38 ±.08 Oryx -0.07 ±.09 -0.41 ±.04 Waterbuck 0.48 ±.15 -0.27 ±.03 Eland 0.19 ±.07 0.54 ±.14

32

1

0.8

0.6

0.4

0.2

0

-0.2

-0.4

-0.6

-0.8

-1

Hyena Lion

Figure 4.2 – Jacobs’ index values based on game count data from 2014-2016. 4.1.4 Pianka’s index

Pianka’s index of niche overlap was found to be 0.8852 on the standardised scale, indicating high dietary overlap between hyenas and lions at LBL and therefore similar dietary niches. This outcome will be interpreted and compared to findings from other studies in the dietary partitioning discussion (Section 5.2).

4.1.5 Levins’ index

Levins’ index of niche breadth reveals that hyenas have a broader diet than lions in LBL (Table 4.4), owing to their wider prey base and more uniform distribution of prey selection. Species consumed by hyenas that were not consumed by lions include cattle, sheep, and goat – all domesticated livestock. The only species consumed by lions that hyenas did not consume was kudu, which contributed a small (1%) amount to the lions’ diet.

Table 4.4 – Levins’ index of niche breadth, raw and standardised. Species # Spp. Consumed Levins’ index Std. Levins’ index Hyena 12 8.72 0.59 Lion 10 6.61 0.43

33

4.2 Spatial partitioning

4.2.1 Home range and core sizes

Lions had larger home ranges (U = 0, p<0.05) and cores (U = 0, p<0.05) than hyenas in all cases. Home ranges for hyenas averaged 44.58 km2 (S.E.: ±5.81 km2, range: 37.00 km2–68.39 km2, n=6) and for lions averaged 122.63 km2 (S.E.: ±24.46 km2, range: 76.34 km2–159.46 km2, n=3). Cores for hyenas averaged only 9.87 km2 (S.E.: ±1.27 km2, range: 6.61 km2–14.83 km2, n=6) and for lions averaged 40.64 km2 (S.E.: ±14.54 km2, range: 20.95 km2–69.02 km2, n=3) (Figure 4.3).

150

130

110 )

2 90

70

50 Mean Mean (km area

30

10

-10 Core Home Range

Hyena Lion

Figure 4.3 – Average home range and core sizes for hyenas and lions at LBL.

34

Figure 4.4 – Cores for collared hyenas and lions at LBL.

35

Figure 4.5 – Home ranges for collared hyenas and lions at LBL.

36

4.2.2 Overlap

Point and area percent overlap of hyena (HH, n=28), hyena-lion (HL, n=18), lion (LL, n=3), and lion-hyena (LH, n=18) cores and home ranges were calculated and averaged to reveal greater intraspecific avoidance than interspecific avoidance, particularly with respect to hyenas (Table 4.5, Table 4.6). Core-home range area overlap confirmed this trend. Intraspecific hyena relationships revealed 0.00% overlap of cores and cores-home ranges, indicating strong inter- clan avoidance. The NH01-NH02 relationship was excluded from this analysis, as they were in the same clan.

Table 4.5 – Mean percent overlap of difference intraspecific and interspecific relationships based on point-core (PC), point-home range (PHR), area-core (AC), area-home range (AHR), and core-home range (CHR) overlap. Relationship PC PHR AC AHR CHR

HH 0.03 ± 0.00 0.89 ± 0.09 0.00 ± 0.00 2.04 ± 0.25 0.00 ± 0.00

LL 3.67 ± 1.09 33.18 ± 10.88 0.26 ± 0.00 14.65 ± 10.58 14.39 ± 11.27

HL 15.06 ± 2.75 80.82 ± 8.83 11.38 ± 0.28 70.27 ± 6.59 79.23 ± 9.26

LH 8.17 ± 0.48 67.90 ± 3.47 5.53 ± 0.53 51.09 ± 2.18 56.79 ± 4.11

37

Hyenas displayed significant intraspecific avoidance in comparison to interspecific avoidance across all five spatial relationships (PC: U=170.5, p<0.05; PHR: U=150.5, p<0.05; AC: U=182, p<0.01; AHR: U=144, p<0.01; CHR: U=140, p<0.01) (Table 4.6). Lions did not display differences in overlap across any metric. Hyena-lion and lion-hyena overlap also did not reveal a significant difference in interspecific avoidance from either species’ perspective.

Table 4.6 – Results of Mann-Whitney U Test. U p-value PC HH-HL 170.5 0.01 LL-LH 58.5 0.77 HL-LH 155.5 0.83 PHR HH-HL 150.5 0.02 LL-LH 61 0.66 HL-LH 189 0.38 AC HH-HL 182 <0.01 LL-LH 57 0.84 HL-LH 157.5 0.87 AHR HH-HL 144 <0.01 LL-LH 70 0.29 HL-LH 202 0.19 CHR HH-HL 140 <0.01 LL-LH 58.5 0.77 HL-LH 183 0.47

38

Figure 4.6 - Home ranges and cores of collared individual hyenas, differentiated by clan association.

39

Figure 4.7 - Home ranges and cores of collared individual lions, differentiated by pride association.

40

4.2.3 Den activity

Average distance to active (n=6) vs. inactive (n=10) hyena dens was not significantly different across any relationship (Table 4.7). Active dens were predictably closer to intra-clan cores (HC) and home ranges (HHR), but surprisingly, also inter-clan intraspecific cores (OHC) and home ranges (OHHR).

Table 4.7 – Average minimum distances (m) of active dens vs. inactive dens from intra-clan cores and home ranges, lion cores and home ranges, and inter-clan cores and home ranges. Active Inactive U p-value HC 705.5 ± 313.7 1514.5 ± 596.8 37 0.80 HHR 164.1 ± 153.5 297.6 ± 150.8 34.5 0.34 LC 1866.3 ± 771.9 564.9 ± 295.8 46.5 0.08

LHR 210.5 ± 174.5 142.0 ± 134.7 36 0.38

OHC 5471.5 ± 692.3 7363.6 ± 664.6 17 0.18 OHHR 3891.2 ± 652.9 4683.4 ± 585.3 20 0.31

9000

8000

7000

6000

5000

4000 Distance (m) 3000

2000

1000

0 HC HHR LC LHR OHC OHHR

Active Inactive

Figure 4.8 - Average minimum distances (m) of active dens vs. inactive dens from intra- clan cores and home ranges, lion cores and home ranges, and inter-clan cores and home ranges.

41

4.2.4 Den buffers

Placing 500 m, 1000 m, and 2000 m buffers around the hyena dens revealed strong inter-clan avoidance, confirming the territoriality of hyenas at LBL (Table 4.8, Figure 4.9). Only one confirmed den site was located within 2000 m of an inter-clan hyena’s home range; under all other measures, there was no overlap between inter-clan cores or home ranges and hyena dens, whether active or inactive.

Predictably, intra-clan dens were located within cores and home ranges of their respective collared hyenas, regardless of den status (active or inactive). All 10 inactive dens were also within the 2000 m buffer of an intra-clan hyena’s core, indicating the possibility of little seasonal change in home ranges and relatively consistent spatial behaviour.

Given the large home ranges and cores of lions (Figure 4.10), it is also unsurprising that the majority of hyena dens are located within lion home ranges. However, only 2/6 active hyena dens were located within 500 m of lion cores, indicating possible lion avoidance.

Table 4.8 – Frequencies of occurrence of hyena dens within 500 m, 1000 m, and 2000 m buffers to intraspecific and interspecific cores and home ranges.

500 m 1000 m 2000 m

Active Inactive Active Inactive Active Inactive HC 4/6 6/10 5/6 6/10 6/6 8/10 HHR 5/6 9/10 6/6 10/10 6/6 10/10 LC 2/6 7/10 2/6 8/10 4/6 9/10

LHR 5/6 9/10 5/6 9/10 6/6 10/10

OHC 0/6 0/10 0/6 0/10 0/6 0/10

OHHR 0/6 0/10 0/6 0/10 1/6 0/10

42

Figure 4.9 – Hyena dens with buffers overlaid with hyena cores and home ranges. 43

Figure 4.10 – Hyena dens with buffers overlaid with lion cores and home ranges.

44

4.2.5 Community land use

Mean home range-community overlap for hyenas was 15.9% (S.E.: ±1.6%, range: 9.2% - 20.4%, n=6) and for lions was 28.3% (S.E.: ±8.9%, range: 7.3% - 44.4%, n=3). Mean core-community overlap for hyenas was 4.9% (S.E.: ±3.7%, range: 0.0% - 25.0%, n=6) and for lions was 6.5% (S.E.: ±5.3%, range: 0.0% - 19.4%, n=3). Lions and hyenas did not differ significantly with regards to home range (U=12, p=0.55) or core (U=10, p=0.90) overlap with communities lands, indicating that both species utilise the human communities similarly.

0.4

0.35

0.3

0.25

0.2 Proportion 0.15

0.1

0.05

0 Home range Core

Hyena Lion

Figure 4.11 - Mean proportion of home range and core areas located within human communities for hyenas and lions.

45

4.3 Temporal partitioning

4.3.1 Activity budgets

Hyenas and lions displayed similar activity patterns, though hyenas travelled significantly further at night (t=11.37, p<0.01, df=9.47) (Table 4.9). Both species displayed nocturnal tendencies, with night (hyena: 39%, lion: 41%) contributing the greatest proportion to mean daily distance travelled followed by evening (hyena: 31%, lion 28%), dawn (hyena: 27%, lion 26%), and day (hyena: 3%, lion 5%) (Figure 4.12; Table 4.9). Hyenas reach maximum activity between 20:00 – 21:00 (1,020.5 m/hr) whereas lions peak at 03:00 – 04:00 (659.1 m/hr) (Figure 4.13).

Table 4.9 – Mean distances travelled (m/hr) at five separate time windows by hyenas and lions at LBL.

Window Hyena Lion T p-value df

Dawn 582.8 ± 155.3 346.8 ± 102.0 0.81 0.46 5.18

Day 54.8 ± 8.7 68.7 ± 8.7 -1.13 0.28 13.99

Evening 677.4 ± 201.8 373.7 ± 105.8 1.29 0.26 4.53

Night 857.5 ± 10.4 555.8 ± 24.4 11.37 <0.01 9.47

Dawn Day Evening Night Dawn Day Evening Night

Figure 4.12 – Activity budgets of hyenas (L) and lions (R) at LBL based on distance travelled at separate time windows as a proportion of total 24h distance travelled.

46

1200

1000

) 800

600

Distance travelled Distance travelled (m/hr 400

200

0 0-1 1-2 2-3 3-4 4-5 5-6 6-7 7-8 8-9 9-10 10-11 11-12 12-13 13-14 14-15 15-16 16-17 17-18 18-19 19-20 20-21 21-22 22-23 23-0

Hyena Lion

Figure 4.13 - Average distance travelled each hour by hyenas and lions at LBL, with standard errors.

47

4.3.2 Dynamic analysis

Dynamic analysis of mutual interactions using Doncaster’s test on 10 hyena-lion pairs revealed attraction in four cases and indifference in six cases (p-values = <0.01–1.00; Table 4.10). Total paired fixes between sympatric hyena-lion pairs ranged from 1,537–2,868, with paired fixes that fell within the 200 m threshold ranging from 0–14 and those above the 200 m threshold ranging from 1,561–2,855.

Three of the four inter-clan hyena pairs showed no attraction (all p-values = 1.00). Surprisingly, NH01-SH02 showed attraction, though this result was based on just one paired event, which is attributable to a territorial dispute or kill site. To test the validity of these results, the interactions between NH01 and NH02 were also analysed, revealing strong attraction between the two (p<0.01) and underscoring the difference between intra-clan and inter-clan relationships (Table 4.11). The only lion pair analysed using Doncaster’s test revealed no attraction (p=0.43; Table 4.12).

Home range overlap between hyena-lion pairs that expressed attraction and those that expressed indifference did not differ (U=15.5, p=0.51), indicating no relationship between overlap and attraction.

48

Table 4.10 – Table of contingency values for all 10 hyena-lion pairs, including Chi-squared p values. Pair Below Above Totals Paired 3 1561 1564 Non-Paired 1277 2443255 2444532 BH01-WL01 Totals 1280 2444816 2446096 p-value 0.06 Paired 0 2428 2428 Non-Paired 144 5892612 5892756 CH01-DL01 Totals 144 5895040 5895184 p-value 1.00 Paired 0 1330 1330 Non-Paired 141 1767429 1767570 CH01-WL01 Totals 141 1768759 1768900 p-value 1.00 Paired 14 2651 2665 Non-Paired 5242 7094318 7099560 NH01-DL01 Totals 5256 7096969 7102225 p-value <0.01 Paired 5 2591 2596 Non-Paired 5492 6731128 6736620 NH02-DL01 Totals 5497 6733719 6739216 p-value 0.10 Paired 3 2380 2383 Non-Paired 1177 5675129 5676306 SH02-DL01 Totals 1180 5677509 5678689 p-value <0.01 Paired 14 1897 1911 Non-Paired 1114 3648896 3650010 SH02-SL01 Totals 1128 3650793 3651921 p-value <0.01 Paired 0 1537 1537 Non-Paired 77 2360755 2360832 SH02-WL01 Totals 77 2362292 2362369 p-value 1.00 Paired 13 2855 2868 Non-Paired 9407 8213149 8222556 UH01-DL01 Totals 9420 8216004 8225424 p-value <0.01 Paired 2 2264 2266 Non-Paired 1026 5131464 5132490 UH01-SL01 Totals 1028 5133728 5134756 p-value 0.12

49

Table 4.11 – Table of contingency values for all five hyena pairs, including Chi-squared p values.

Pair Below Above Totals Paired 0 2268 2268 Non-Paired 143 5141413 5141556 BH01-SH02 Totals 143 5143681 5143824 p-value 1.00 Paired 0 2121 2121 Non-Paired 15 4496505 4496520 CH01-NH02 Totals 15 4498626 4498641 p-value 1.00 Paired 0 2210 2210 Non-Paired 17 4881873 4881890 CH01-NH01 Totals 17 4884083 4884100 p-value 1.00 Paired 98 2344 2442 Non-Paired 61819 5899103 5960922 NH01-NH02 Totals 61917 5901447 5963364 p-value <0.01 Paired 1 2238 2239 Non-Paired 33 5010849 5010882 NH01-SH02 Totals 34 5013087 5013121 p-value <0.01

Table 4.12 – Table of contingency values for DL01-SL01.

Pair Below Above Totals

Paired 3 2480 2483

Non-Paired 3774 6159032 6162806 DL01-SL01 Totals 3777 6161512 6165289

p-value 0.43

50

Minta’s test revealed hyenas were attracted to shared spaces with lions 40% of the time (LAA), of which 75% were significant, and shared spaces with other inter-clan hyenas 0% of the time. The intra-clan pair NH01-NH02 revealed significant attraction, as expected. Lions were attracted to shared spaces with hyenas 40% of the time (LBB), of which 50% were significant, and shared spaces with inter-pride lions 33% of the time, of which 50% were significant. In terms of simultaneity of use (Lixn), hyena-lion pairs displayed attraction 70% of the time, of which 85.7% were significant, inter-clan hyena pairs displayed attraction 0% of the time, and inter-pride lion pairs displayed attraction 67% of the time, none of which were significant (Table 4.13).

Table 4.13 – Test statistics and p-values for Minta’s spatial interaction test for hyena-lion, hyena-hyena, and lion-lion pairs.

Pair LAA pLAA LBB pLBB Lixn pLixn Hyena-lion BH01-WL01 1.85 <0.01 0.32 <0.01 0.51 <0.05 CH01-DL01 -0.42 <0.01 -1.06 <0.01 0.94 <0.01 CH01-WL01 -0.81 <0.01 0.25 0.21 -0.60 0.18 NH01-DL01 0.89 <0.01 -0.25 <0.01 -0.03 0.74 NH02-DL01 0.17 0.41 -0.29 <0.01 0.37 <0.01 SH02-DL01 -0.08 <0.01 -0.32 <0.01 -0.13 0.16 SH02-SL01 -0.12 <0.01 0.05 0.80 0.14 <0.01 SH02-WL01 -1.06 <0.01 -0.44 <0.01 0.05 0.65 UH01-DL01 1.04 <0.01 0.83 <0.01 0.31 <0.01 UH01-SL01 -0.44 <0.01 -0.48 <0.01 0.29 <0.01 Hyena-hyena BH01-SH02 -1.61 <0.01 -1.43 <0.01 -0.98 <0.01 CH01-NH01 -∞ <0.01 -∞ <0.01 -∞ 0.84 CH01-NH02 -∞ 0.36 -∞ 0.44 -∞ 0.99 NH01-NH02 0.71 <0.01 0.34 <0.01 0.13 0.09 NH01-SH02 -1.23 <0.01 -0.07 0.70 -2.68 <0.01 Lion-lion DL01-WL01 0.03 0.96 -∞ <0.01 -0.03 0.99 DL01-SL01 0.25 <0.05 -0.76 <0.01 0.06 0.27 SL01-WL01 -1.65 <0.01 -0.68 0.02 0.35 0.75

51

4.4 Demographics

4.4.1 Identified individuals

A total of 89 hyenas were identified during the course of this study, with a range of 3-35 between the clans (Table 4.14A). This number is undoubtedly incomplete as it relied on opportunistic encounters with hyenas and camera trap photos from the dens. For the Shamba and Borana clans, photographs were lost due to camera trap malfunctioning and partially recovered.

However, this result provides a minimum density of 0.24 hyenas/km2 and falls within the estimates provided by a demographic study in 2015 (Preston 2015). Continued monitoring of dens and development of the photographic database will complete the picture. However, this effort has provided a baseline knowledge of hyena clan structure and membership within LBL.

Sex ratios were difficult to determine for the hyenas at LBL, leaving 67 of the 89 individuals unidentified. Given the morphological similarities between male and female hyenas (East 2001), females could only be positively identified if they were visibly lactating suckling cubs. Sexes of collared individuals were determined while hyenas were immobilised.

Identification based on pride association of lions (Table 4.14B) confirmed that there are 34 lions in LBL, providing a density of 0.09 lions/km2. Male coalitions were grouped with their resident prides for the purposes of this study. As of July 2016, cubs were only present within one of prides, Western, and none of the cubs had been identified yet by sex. The hyena to lion ratio based on identified individuals is therefore, at minimum, 2.62:1.

52

Table 4.14 - Individually identified hyenas (A) and lions (B) at LBL based on clan and pride membership, respectively.

Clan Adults Sub-adults Cubs Male / Female / Unknown Totals Borana 3 0 0 1 / 0 / 2 3 Charlie 16 3 5 1 / 3 / 20 24 Nala 21 9 5 1 / 8 / 26 35 Shamba 3 2 4 0 / 2 / 7 9 Utalii 10 2 6 1 / 5 / 12 18 Totals 53 16 20 4 / 18 / 67 89 A

Pride Adults Sub-adults Cubs Male / Female / Unknown Totals Dalma 3 0 0 2 / 1 / 0 3 Sarah 3 2 0 0 / 5 / 0 5 Western 10 0 8 5 / 5 / 8 18 Mole 0 6 0 4 / 2 / 0 6 Linda 2 0 0 0 / 2 / 0 2 Totals 18 8 8 11 / 15 / 8 34 B

53

5. Discussion

5.1 Collaring

While this study focused on interactions between hyenas and lions, it did not explore sex differences in spatiotemporal patterns due to small sample size. Hyenas are matriarchal and females tend to spend more time at den sites than males, particularly if they have cubs (Boydston et al. 2003). Furthermore, each member of a hyena clan has a social rank. High ranking females have smaller home ranges and cores and a higher proportion of fixes at the communal dens than do low ranking females (Boydston et al. 2003). They also breed more frequently and display higher fitness (Holekamp et al. 1996, Watts and Holekamp 2009). Finally, high-ranking females display bolder behaviour with lions than do low-ranking females and males, including in disputes over food (Shaw 2012). These factors affect spatiotemporal interactions with lions.

Because this study included the first collaring effort of hyenas and the first study to focus on this species in LBL, hyenas were nervous. This made direct observations impossible, resulting in the inability to differentiate individuals based on rank or sex – the latter only being possible with collared individuals or lactating females. Therefore, differences may have occurred in spatiotemporal patterns since four of the six hyenas in this study are male, and social ranks are unknown for all individuals.

For lions, males tend to have larger home ranges and unless they are resident pride males, display highly variable movement patterns when dispersing (Loveridge et al. 2009). Furthermore, home range size varies across different environments and also according to pride size and environmental variables such as mean annual precipitation and prey density (Schaller 1972, Funston and Mills 2006). DL01, the male lion in this study, is considered resident within his pride and therefore represents a third distinct social group. However, being male, his data may exhibit differences from WL01’s or SL01’s due to sex, both of whom are females.

For the two hyena clan mates NH01 and NH02, analysis was conducted for both separately and they were treated as independent when looking at interspecific and intraspecific spatiotemporal trends. NH01 is a female, and as expected, her core area was smaller than NH02’s, a male (6.61 km2 versus 14.83 km2). Accordingly, despite these factors that may affect spatiotemporal behaviour in these species, the similarities and differences that exist must be looked at through a species-specific lens as

54 opposed to a sex-based one. Ideally, a study can be conducted where all hyenas share the same social rank and are the same sex and all lions are from the same-sized prides and are the same sex, along with other demographic factors such as age. However, this is impossible in a previously completely unstudied population.

5.2 Dietary partitioning

This study revealed high dietary overlap (Ô = 0.8852) between the study species, which corroborates the findings of a meta-analysis of dietary partitioning between hyenas and lions (Periquet et al. 2014). This overlap is indicative both of their similar niches and also a possible scavenging or kleptoparasitic relationship between the two. In the Ngorongoro Crater, lions are noted to scavenge from hyenas more frequently than hyenas do from lions (Kruuk 1972) and lions do not surrender carcasses to hyenas until they have eaten their fill (Höner et al. 2002). Hyenas are also much less successful at supplanting lions from kills when a male is around (Kissui and Packer 2004). Given the small size of LBL and the high spatial overlap between hyenas and lions, there is undoubtedly direct interaction between the species for kills.

As hyenas in this population have a broader diet (퐵̂ = 0.59 versus 0.43), they are utilising a prey base outside of the lions’. One clear difference is that hyenas feed on livestock species in LBL, while no livestock hairs were found in lion scat. This may be a mechanism for the hyenas to avoid or mitigate competition for wild prey species. Therefore, it is possible that hyenas are utilizing livestock species as a small but important alternative to wild prey (15.7% of the hyenas’ diet). This has strong implications from a human-wildlife conflict perspective, as local pastoralists rely on livestock for their livelihoods.

Jacobs’ index reveals that although plains zebra and impala are important contributors to both species’ diets, they do not necessarily selectively feed on them. This suggests an abundance- related tendency to focus on plains zebra and impala – they may simply hunt them because there are more of them. The negative values for both plains zebra should not be regarded as blatant avoidance but rather opportunistic, random feeding. Hyenas and lions both select against buffalo, which supports previous dietary findings for lions in LBL (Pratt 2014). It is also surprising, as the buffalo tends to be among the most favoured prey item for lions across populations (Hayward and Kerley 2005, Davidson et al. 2013). However, different lion populations are known to develop different cultures with prey selection; over time, the tendency to avoid buffaloes at LBL may change (Power and Compion 2009).

55

5.3 Spatial partitioning

There are spatial parallels between the LBL hyena population and that of the Ngorongoro Crater, Tanzania, which has been studied since the early 1970s. Hyenas display strong inter-clan spatial exclusivity, reminiscent of the Ngorongoro population (Kruuk 1972). The Crater floor is 260 km2 and the ecosystem also features human communities both in and around the conservation area. Average home range size at LBL (44.58 km2) is small when compared to other populations of Crocuta, but larger than in areas with extremely high hyena density (≥1.00 individuals/km2) such as the Ngorongoro Crater and Amboseli National Park, where home ranges are closely defended and are generally 20–30 km2 (Holekamp and Dloniak 2011). The death of SH01 may point towards such aggressive territoriality; in populations with tightly packed home ranges, conspecific intruders are sometimes viciously attacked and killed (Smale et al. 1997). This sort of territorial behaviour is in stark contrast to hyena populations in the (Hofer and East 1995), where females travel far outside their core areas for several days due to the wildebeest migration before returning to suckle cubs.

Lion home ranges are also noted to vary across landscapes (Gittleman and Harvey 1982). In , Kenya, lion home ranges are between 25–51 km2, while in the lower density , , home ranges can be over 2,000 km2 (Stander 1991). The home ranges at LBL can therefore be described as moderate in size, despite the small area of the study site and intense human pressure. Lions do not display the same degree of territoriality that hyenas do in this population, perhaps due to a lower population density and therefore less intraspecific competition for prey. Lions are also inherently different from hyenas in that they do not have communal dens; once cubs are mobile, they will accompany the pride as it moves (McComb et al. 1993). Adult females, unless recently whelped, can freely move with their pride members and cubs without needing to return to a den site for suckling and social rituals as adult female hyenas do.

Despite the limited space available in LBL for these predators to coexist, there does appear to be avoidance in core areas between intraspecific and interspecific competitors alike. Hyenas may prefer core areas that minimise their chances of encountering lions, but still may use them as a food source, therefore allowing for some degree of home range overlap. Lions may in turn also utilise hyenas as a food source and are noted to respond favourably to recorded calls of hyenas squabbling over kills (Kiffner 2008).

56

Two of the hyenas dens found during this study were located outside the conservancy boundaries, and one was confirmed as being active. This den was attributed to the Borana clan, as BH01’s fixes have indicated visits to this den site. While the majority (84.1%) of hyena home range area in this study were located within LWC and BC, there is still clearly use of community lands, as reflected in the scat data. Hyena clans are also noted, anecdotally, to reside permanently within the ranches and community lands surrounding LBL, further confirming their ability to reside in close proximity to humans (Yirga et al. 2011, Yirga et al. 2012, Yirga et al. 2013). While lions also display utilisation of human settlements, no livestock hairs were found in their scats. This appears to be a contradictory finding. This discrepancy may be attributed to the chance that livestock hairs may not have been counted in the lion scats; livestock hairs may be counted in the future with increased sample size. There are anecdotal reports of lions preying on cattle in LWC in 2016 (Z. Davidson, Conservation Biologist, Marwell Wildlife, pers. comm.). Hyenas that are based in the conservancies may be limited in how far they can venture into the communities by other clans, as evidenced by the low spatial overlap between clans in this population. Lions do not face such an intraspecific barrier.

5.4 Temporal partitioning

Both hyenas and lions at LBL have similar activity patterns, apart from hyenas being significantly more active at night. As hyenas prey on domestic livestock from the communities, it is possible that this heavily nocturnal behaviour is a response to human presence. Analysis of hyena temporal trends in the Masai Mara Game Reserve (GR) reveals that hyenas in close contact to human habitation tend to be more nocturnal than hyenas towards the centre of the GR (Pangle and Holekamp 2010). All the hyena clans in LBL have home range boundaries that overlap with human communities, which directly puts them at risk of retaliatory persecution due to livestock losses. Overall, both species appear to be largely nocturnal and crepuscular, which is to be expected, especially given the high human impact in LBL as a compounding factor.

Because 40% of the interspecific cases of spatiotemporal avoidance actually revealed attraction by hyenas and lions using both Doncaster’s and Minta’s tests, a likely factor playing a role in these relationships is prey. It is plausible that hyenas are scavenging from lions at LBL, or vice versa. In Minta’s test, mutual simultaneity of use revealed 70% of the hyena-lion cases also shows a surprisingly high level of direct interactions. All hyena pairs revealed avoidance apart from the NH01-NH02 and the NH01-SH02 pairings. The former is unsurprising due to the pair belonging to the Nala clan; the latter is based on just one paired encounter which could be attributed to a dispute over a carcass or territorial conflict. Both NH01 and SH02 are large, adult 57 females, which makes them likely participants in inter-clan conflicts for space or prey (Shaw 2012). Lions showed less intraspecific avoidance than hyenas, with 67% of lion pairs showing simultaneous use and 33% showing attraction. This points towards the exclusivity of hyena clans in this landscape, and to a lesser degree, lion prides. It also reveals the high degree of interspecific sharing of space and time, for the acquisition of prey.

Direct encounters between hyenas and lions often result in hyena mortality. Considering the frequency of intraguild predation by lions on hyenas and the limitation of hyena fitness by lions, the benefit of food is a strong driver for such bold behaviour around lions by hyenas (Watts and Holekamp 2008). In the Masai Mara, prior to human expansion towards the GR in the 1990s, lions were the leading cause of adult hyena mortality but have since been supplanted by humans for hyenas with home ranges bordering human communities (Pangle and Holekamp 2010). Coming into close proximity with lions is a potentially lethal decision for hyenas to make. Therefore, the potential of acquiring food through such encounters must outweigh the risk of mortality; the same can be said for hyena-human contact.

5.5 Demographics

The 89 hyenas identified so far in the LBL population is an incomplete figure, but represents a density of 0.24 hyenas/km2, which is higher than areas such as the (Cozzi et al. 2013), , Etosha National Park, and Central Kalahari Game Reserve (Holekamp and Dloniak 2011). Only three hyenas were identified from the Borana clan, which is a definite underestimate. Anecdotal reports from researchers who have worked in LBL as well as rangers suggest that hyenas are frequently seen and heard in BC (S. Gilisho, Predator Monitoring Officer, LWC, pers. comm.). Furthermore, sightings of groups of 47, 41, 20, 16, and 10 adult hyenas together at different locations within LBL since 2015 (S. Gilisho pers. comm.) point toward a robust population. It is likely that the LBL hyena population is growing, given the identification of 36 individual hyenas below breeding age in this study. Furthermore, the presence of permanent hyena clans outside of LWC and BC is a sign that there is adequate gene flow in and out of the conservancies, given the migratory gaps. The Mount Kenya elephant corridor also provides for landscape connectivity between LBL and Mount Kenya National Park (Nyaligu and Weeks 2013), and hyenas are known to use it. There is limited information, however, on hyena demography in central Kenya. In Laikipia, an estimated 71 hyenas are killed annually by humans on ranches, not counting road kill (Frank 1998), and hyenas are declining outside of protected areas across Africa (Woodroffe 2000). Therefore, while there are some

58 encouraging signs on the stability and growth potential of the hyena population within LBL, there are still gaps in knowledge that need to be filled.

The lion population described in this study, as it stands at 34, is considered accurate. A density of .09 lions/km2 is moderate and healthy in large ecosystems such as Selous in Tanzania (Creel and Creel 1997). However, unlike hyenas, lions are not known to reside permanently in the community areas immediately surrounding LBL and there is limited gene flow in and out of this population. Inbreeding between siblings has been confirmed in at least one case (M. Mwololo, Research Officer 1, LWC, pers. comm.) and is of concern. The lion population in the Ngorongoro Crater has also been noted to have high rates of inbreeding due to the closed nature and small area of the Crater, which in turn has costly fitness consequences and places such populations at risk of local extinction (Riggio et al. 2013). Small populations are inherently susceptible to stochastic variations (Lande 1993). Just eight cubs were identified this year, from only the Western pride. One encouraging sign is that lions are known to reside in ranches to the west of BC in Laikipia (Frank et al. 2005), which may allow for appropriate management measures to be taken to encourage dispersal. There may be potential for dispersal to and from ranches further west in Laikipia, but the presence of inbreeding within this population points to limited connectivity.

5.6 Conclusions and management recommendations

Overall, hyenas display stronger intraspecific than interspecific avoidance in spatiotemporal trends. Both species share similar diets and activity patterns, though lions have larger home ranges and cores. Nonetheless, there are some differences. Hyenas travel further at night than do lions. Hyenas feed on livestock whereas lions do not. Hyenas have small, exclusive home ranges, whereas lions have large, slightly overlapping ones. There are known hyena clans within the communities, but no lion prides. The hyena population overall appears to be healthy, with a high number of cubs and sub-adults, but the lion population may be suffering from inbreeding, a limited ability to disperse, and low cub recruitment.

A human-hyena conflict study would be useful in determining local attitudes towards hyenas, the impact that hyenas are having on livestock, and human economic loss to livestock predation. This can provide for the development of solutions to mitigate losses, such as compensation schemes or the development of corrals to protect livestock. Estimates of livestock abundance in comparison to wild ungulate abundance can provide a clearer picture of whether hyenas are preferentially feeding on domesticated prey. A comparison of hyena clans inside versus outside

59 the conservancies would shed light on the impact of human communities on hyena spatial and dietary ecology, as well as differences in stress hormones such as cortisol. This information can provide for policies that allow for humans and carnivores to coexist. Deploying GPS collars on hyenas from as many clans as possible both inside and outside the conservancies would be ideal in this scenario to maximise sample sizes.

Expanding monitoring efforts to other members of the large carnivore guild will provide a more complete picture of predator ecology in the region. African wild dogs, leopards, cheetahs, and striped hyenas are all present in LBL. Leopards and cheetahs were both observed in chance encounters during the field season, and the cheetahs were well-habituated. A leopard and a striped hyena were also seen on camera trap at hyena baits. It would be useful to gather data on the complete demography, spatiotemporal behaviour, and dietary ecology of these species not just in LBL but in surrounding ranches as well. It is also advised that direct observations be used in this effort. Lions and cheetahs in LBL are already well-habituated to vehicles, but hyenas, leopards, and wild dogs are not. Habituating these species to vehicles for observational research can provide accurate hunting versus scavenging data. Ideally, this research can be done long term to examine seasonal differences in diet or activity and longer term spatial trends.

Culling of predators is strongly advised against based on the findings in this study. Lions and hyenas exist in a natural balance; lion control leads to massive rebounds in hyena populations and disrupts this balance. In Amboseli National Park, hyena populations exploded from a density of 0.13 hyenas/km2 to 1.65 hyenas/km2 between the mid-1990’s and mid-2000’s due to local lion extinction (Watts and Holekamp 2008). Although they demonstrate the ability to exist in human- dominated landscapes, hyenas are vulnerable once locally extirpated due to female philopatry and conservative recolonisation of unclaimed territories, even with intact source populations nearby (Frank 1998). Furthermore, no rhino hairs were found in any of the hyena or lion scats analysed in this study. While hyenas have been noted to attack rhino calves in Aberdare National Park, there were no confirmed successful cases of predation (Sillero-Zubiri and Gottelli, 1991). The lion population is already low and putting external control on them can have strong negative outcomes, including the possibility of extinction. Eight lions were poisoned in LBL since 2014 as retaliation for livestock attacks (Z. Davidson pers. comm.). For a small population, this is a major loss. Lions are classified as Vulnerable by the International Union for Conservation of Nature (IUCN) (Bauer et al. 2008); preserving this population can be a small step towards ensuring their continued persistence in Kenya.

Landscape connectivity and dispersal are important for both hyenas and lions. Male lion dispersal ability is strongly correlated to population connectivity and East African lions now

60 exist in population fragments due to anthropogenic pressure (Dolrenry et al. 2014). While hyenas in this area appear to enjoy the presence of multiple clans both in and around the conservancies, the lion population is more isolated. It may be necessary to provide more wildlife corridors within the study area to allow for genetic exchange between lion populations in surrounding areas such as Laikipia, Samburu, Meru, and Mount Kenya. An alternative to this is metapopulation management, which relies on hands-on removal and introduction of individuals between disconnected areas. Metapopulation management has demonstrated success in South Africa with African wild dogs (Gusset et al. 2006) and has been suggested as a management strategy for lions (Dolrenry et al. 2014), cheetahs (Johnson et al. 2010), and leopards (Balme et al. 2010). Expanding the current predator monitoring effort inside and outside of LBL, facilitating carnivore dispersal, and determining carnivore impacts on human communities will allow for management decisions to be made that allow for the coexistence of predators and humans alike in LBL and beyond.

61

Acknowledgments

I owe many people thanks for their help and support during the course of this project. Many thanks to Dr. Zeke Davidson and Prof. C. Patrick Doncaster for their helpful guidance and feedback during this project. Dr. Davidson opened his home to me and was always helpful with fieldwork-related aspects and the initiation of the project. Prof. Doncaster provided me with considerable support regarding statistical analysis and the writing of this report. Many thanks to both for rigorously reviewing my proposal, dissertation drafts, and providing me with detailed feedback.

I am grateful to Dr. Marc Dupuis-Desormeaux and Prof. Suzanne MacDonald for taking an interest in LBL’s hyenas and getting this project started. They provided funding for the GPS- GSM collars and their guidance at the start of the field season as we prepared for the trapping endeavour was incredibly helpful. I must also give thanks to LWC’s and BC’s management teams and Northern Rangelands Trust (NRT) for allowing this project to take place – Ian Craig and Mike Watson took a keen interest in this project and I appreciate their support and encouragement.

I also would like to thank Dr. Alayne Cotterill for the extremely informative tutorial on trapping hyenas (and for getting our first hyena) as well as Dr. Laurence Frank for generously allowing the team to borrow his equipment for leg-hold snares. This was my first ever crash course in collaring spotted hyenas and it would have been impossible without their support.

Special thanks must also be extended to the research departments at LWC and BC, in particular Geoffrey Chege, Mary Mwololo, Edwin Kisio, and of course, Saibala Gilisho. Chege, Mary, and Edwin were exceptionally organised with vehicles, bait, and reporting hyena and lion sightings. Saibala spearheaded all the work prior to my arrival and very effectively set the foundation for this research, which made the process much smoother than it otherwise would have been. He worked extremely long hours with me that went above and beyond my expectations.

I want to thank the drivers as well – Logeme, Sumbere, and Karmushu – for providing me with transport in challenging field conditions and always assisting during the collaring efforts with a smile.

62

To all the casuals whose names I never learned or could not remember – I express my gratitude here. Studying large carnivores requires a lot of manual labour and they were always keen to lend a helping hand when I had to move traps or baits.

I also greatly appreciate the help provided by Dr. Bernard Rono and Dr. Matthew Mutinda, the two veterinarians I worked with during the collaring effort. Monitoring the spatiotemporal aspects of these predators would have been impossible without collaring them, and collaring them would have been impossible without their immobilisation. Dr. Rono and Dr. Mutinda made themselves available and kept the study animals safe even in extremely logistically challenging situations and on short notice, and for that I am very grateful.

Finally, more than anything, I owe thanks to my parents and siblings. Without their support and encouragement over the years to pursue this field, I would not be where I am today. I understand why it can be challenging at times to have a son (or brother) who wants nothing more than to spend time in the remote wilderness with large carnivores. However, they accept and embrace it with open arms and have taken a genuine interest in my career, and for that I am eternally grateful.

63

References

Akçakaya, H.R., Mills, G. and Doncaster, C.P. (2007). The role of metapopulations in conservation. In Key Topics in Conservation Biology. Oxford, UK, Blackwell Publishing, 64-84.

Balme, G. A., Hunter, L. T., Goodman, P., Ferguson, H., Craigie, J., and Slotow, R. (2010). An adaptive management approach to trophy hunting of leopards Panthera pardus: a case study from KwaZulu-Natal, South Africa. Biology and Conservation of Wild Felids. Oxford University Press, Oxford, 341-352.

Bauer, H., Nowell, K., and Packer, C. (2008). Panthera leo. Red List of Threatened Species. Version 2010.2. International Union for Conservation of Nature.

Boydston, E., Kapheim, K., Szykman, M. and Holekamp, K. (2003). Individual Variation in Space Use by Female Spotted Hyenas. Journal of Mammalogy, 84(3), 1006-1018.

Brooks, T.M., Mittermeier, R.A., Mittermeier, C.G., da Fonseca, G.A.B., Rylands, A.B., Konstant, W.R., Flick, P., Pilgrim, J., Oldfield, S., Magin, G., and Hilton-Taylor, C. (2002). Habitat loss and extinction in the hotspots of biodiversity. Conservation Biology, 16, 909–923.

Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P., Narwani, A., Mace, G.M., Tilman, D., Wardle, D.A., Kinzig, A., Daily, G.C., Loreau, M., Grace, J.B., Larigauderie, A., Srivastava, D., and Naeem, S., (2012). Biodiversity loss and its impact on humanity. Nature, 486(7401), 59-67.

Cozzi, G., Broekhuis, F., McNutt, J., Turnbull, L., Macdonald, D. and Schmid, B. (2012). Fear of the dark or dinner by moonlight? Reduced temporal partitioning among Africa's large carnivores. Ecology, 93(12), 2590-2599.

Cozzi, G., Broekhuis, F., McNutt, J. W., and Schmid, B. (2013). Density and habitat use of lions and spotted hyenas in northern Botswana and the influence of survey and ecological variables on call-in survey estimation. Biodiversity and Conservation, 22(12), 2937-2956.

64

Creel, S., and Creel, N. (1997). Lion density and population structure in the : evaluation of hunting quotas and offtake. African Journal of Ecology, 35(2), 83- 93.

Dalerum, F., Somers, M., Kunkel, K. and Cameron, E. (2008). The potential for large carnivores to act as biodiversity surrogates in southern Africa. Biodiversity and Conservation, 17(12), 2939-2949.

Darnell, A.M., Graf, J.A., Somers, M.J., Slotow, R., and Szykman, G.M. (2014) Space Use of African Wild Dogs in Relation to Other Large Carnivores. PLOS ONE, 9(6): e98846.

Davidson, Z., Valeix, M., Van Kesteren, F., Loveridge, A., Hunt, J., Murindagomo, F. and Macdonald, D. (2013). Seasonal Diet and Prey Preference of the African Lion in a Waterhole-Driven Semi-Arid Savanna. PLOS ONE, 8(2), e55182.

Davies-Mostert, H., Mills, M. and Macdonald, D. (2015). The Demography and Dynamics of an Expanding, Managed Metapopulation. African Journal of Wildlife Research, 45(2), 258-273.

Dirzo, R., and Raven, P. H. (2003). Global state of biodiversity and loss. Annual Review of Environment and Resources, 28(1), 137-167.

Dolrenry, S., Stenglein, J., Hazzah, L., Lutz, R. S., and Frank, L. (2014). A metapopulation approach to African lion (Panthera leo) conservation. PLOS ONE, 9(2), e88081.

Doncaster, C. P. (1990). Non-parametric estimates of interaction from radio-tracking data. Journal of Theoretical Biology, 143(4), 431-443.

Dupuis-Desormeaux, M., Davidson, Z., Mwololo, M., Kisio, E., Taylor, S. and MacDonald, S. (2015). Testing the Prey-Trap Hypothesis at Two Wildlife Conservancies in Kenya. PLOS ONE, 10(10), e0139537.

East, M. (2001). Male spotted hyenas (Crocuta crocuta) queue for status in social groups dominated by females. Behavioral Ecology, 12(5), 558-568.

65

ESRI, 2014. ArcGIS Desktop: Release 10.4. Environmental Systems Research Institute. Redlands, CA.

Frank, L.G. (1998). Living with lions: Carnivore conservation and livestock in Laikipia District, Kenya. Unpublished Report, Development Alternatives, Inc.

Frank, L.G., Woodroffe, R., and Ogada, M.O. (2005). People and predators in Laikipia district, Kenya. Conservation Biology Series-Cambridge, 9, 286.

Funston, P. J., and Mills, M.G.L. (2006). The influence of lion predation on the population dynamics of common large in the Kruger National Park. South African Journal of Wildlife Research, 36(1), 9-22.

Gaitho, V.G. (2014). Impact of community based ecotourism on households’ livelihoods and environmental management in Il Ngwesi and Lekurruki group ranches, Laikipia county, Kenya. Unpublished PhD thesis. Kenyatta University.

Gittleman, J. (1989). Carnivore Behavior, Ecology, and Evolution. Ithaca: Comstock Pub. Associates.

Gittleman, J., and Harvey, P. H. (1982). Carnivore home-range size, metabolic needs and ecology. Behavioral Ecology and Sociobiology, 10(1), 57-63.

Glickman, S.E.. (1995). The from Aristotle to the Lion King: Reputation is Everything. Social Research, 62(3), 501–537.

Gusset, M., Slotow, R., and Somers, M. J. (2006). Divided we fail: the importance of social integration for the re‐introduction of endangered African wild dogs (Lycaon pictus). Journal of Zoology, 270(3), 502-511.

Hayward, M. (2006). Prey preferences of the spotted hyaena (Crocuta crocuta) and degree of dietary overlap with the lion (Panthera leo). Journal of Zoology, 270(4), 606-614.

Hayward, M., and Kerley, G. (2005). Prey preferences of the lion (Panthera leo). Journal of Zoology, 267(3), 309-322.

66

Hayward, M. and Hayward, G. (2006). Activity patterns of reintroduced lion Panthera leo and spotted hyaena Crocuta crocuta in the Addo Elephant National Park, South Africa. African Journal of Ecology, 45(2), 135-141.

Hofer, H., and East, M. (1995). Population dynamics, population size, and the commuting system of Serengeti spotted hyenas. Serengeti II: dynamics, management, and conservation of an ecosystem, 2, 332.

Holekamp, K., Smale, L., and Szykman, M. (1996). Rank and reproduction in the female spotted hyaena. Journal of Reproduction and Fertility, 108(2), 229-237.

Holekamp, K., and Dloniak, S. M. (2010). Intraspecific variation in the behavioral ecology of a tropical carnivore, the spotted hyena. Advances in the Study of Behavior, 42, 189-229.

Höner, O.P., Wachter, B., East, M.L., and Hofer, H. (2002). The response of spotted hyaenas to long‐term changes in prey populations: functional response and interspecific kleptoparasitism. Journal of Animal Ecology, 71(2), 236-246.

Jacobs, J. (1974). Quantitative measurement of food selection. Oecologia, 14(4), 413-417.

Johnson, S., Mengersen, K., de Waal, A., Marnewick, K., Cilliers, D., Houser, A. M., and Boast, L. (2010). Modelling relocation success in southern Africa using an Iterative Bayesian Network Development Cycle. Ecological Modelling, 221(4), 641-651.

Kamler, J., Stenkewitz, U., Klare, U., Jacobsen, N. and Macdonald, D. (2012). Resource partitioning among cape foxes, bat-eared foxes, and black-backed jackals in South Africa. The Journal of Wildlife Management, 76(6), 1241-1253.

Kamler, J., Stenkewitz, U. and Macdonald, D. (2013). Lethal and sublethal effects of black- backed jackals on cape foxes and bat-eared foxes. Journal of Mammalogy, 94(2), 295- 306.

Kamler, J., Stenkewitz, U., Sliwa, A., Wilson, B., Lamberski, N., Herrick, J. and Macdonald, D. (2015). Ecological relationships of black-footed cats (Felis nigripes) and sympatric

67

canids in South Africa. Mammalian Biology - Zeitschrift für Säugetierkunde, 80(2), 122- 127.

Kiffner, C., Waltert, M., Meyer, B. and Mühlenberg, M. (2008). Response of lions (Panthera leo) and spotted hyaenas (Crocuta crocuta) to sound playbacks. African Journal of Ecology, 46(2), 223-226.

Kissui, B., and Packer, C. (2004). Top–down population regulation of a top predator: lions in the Ngorongoro Crater. Proceedings of the Royal Society of London B: Biological Sciences, 271(1550), 1867-1874.

Kitchen, A.M., Gese, E.M., and Schauster, E.R. (1999). Resource partitioning between coyotes and swift foxes: space, time, and diet. Canadian Journal of Zoology, 77(10), 1645-1656.

Kolowski, J. and Holekamp, K. (2009). Ecological and anthropogenic influences on space use by spotted hyaenas. Journal of Zoology, 277(1), 23-36.

Kruuk, H. (1972). The Spotted Hyena. Chicago: University of Chicago Press.

Lande, R. (1993). Risks of population extinction from demographic and environmental stochasticity and random catastrophes. American Naturalist, 911-927.

Lewa Wildlife Conservancy, 2006. Research and Monitoring Annual Report. Unpublished report. Lewa Downs, Kenya.

Lewa Wildlife Conservancy, 2014. Research and Monitoring Annual Report. Unpublished report. Lewa Downs, Kenya.

Lindsey, P.A., Alexander, R., Frank, L.G., Mathieson, A., and Romanach, S.S. (2006). Potential of trophy hunting to create incentives for wildlife conservation in Africa where alternative wildlife‐based land uses may not be viable. Animal Conservation, 9(3), 283- 291.

68

Linnell, J.D., Swenson, J.E., and Andersen, R. (2000). Conservation of biodiversity in Scandinavian boreal forests: large carnivores as flagships, umbrellas, indicators, or keystones?. Biodiversity and Conservation, 9(7), 857-868.

Long, J.A. (2014). wildlifeDI–A Suite of R Tools for Exploring Dynamic Interaction Patterns in Wildlife Telemetry Data Version 0.1.

Loveridge, A.J., Valeix, M., Davidson, Z., Murindagomo, F., Fritz, H., and Macdonald, D.W. (2009). Changes in home range size of African lions in relation to pride size and prey biomass in a semi-arid savanna. Ecography, 32(6), 953-962.

Low, B., Sundaresan, S., Fischhoff, I. and Rubenstein, D. (2009). Partnering with local communities to identify conservation priorities for endangered Grevy’s zebra. Biological Conservation, 142(7), 1548-1555.

Lyngdoh, S., Shrotriya, S., Goyal, S. P., Clements, H., Hayward, M. W., and Habib, B. (2014). Prey preferences of the snow leopard (Panthera uncia): regional diet specificity holds global significance for conservation. PLOS ONE, 9(2), e88349.

MacArthur, R.H. and Levins, R. (1967). The limiting similarity, convergence, and divergence of coexisting species. American Naturalist, 101, 377-385.

McComb, K., Pusey, A., Packer, C., and Grinnell, J. (1993). Female lions can identify potentially infanticidal males from their roars. Proceedings of the Royal Society of London B: Biological Sciences, 252(1333), 59-64.

Miller, B., Dugelby, B., Foreman, D., del Rio, C. M., Noss, R., Phillips, M., and Willcox, L. (2001). The importance of large carnivores to healthy ecosystems. Endangered Species Update, 18(5), 202-210.

Mills, M.G.L. (1989). The comparative behavioral ecology of hyenas: the importance of diet and food dispersion. In Carnivore Behavior, Ecology, and Evolution, 125-142. Springer US.

Minta, S.C. (1992) Tests of spatial and temporal interaction among animals. Ecological Applications, 2, 178-188.

69

Mora, C., and Sale, P. F. (2011). Ongoing global biodiversity loss and the need to move beyond protected areas: a review of the technical and practical shortcomings of protected areas on land and sea. Marine ecology progress series, 434, 251-266.

Mukherjee, S., Goyal, S. and Chellam, R. (1994). Refined techniques for the analysis of Asiatic lion Panthera leo persica scats. Acta Theriologica, 39, 425-430.

Mwololo, M. (2011). The impact of predation on Grevy’s zebra (Equus grevyi) population on Lewa Wildlife Conservancy. Unpublished Master’s thesis. University of Nairobi.

Nyaligu, M. O., and Weeks, S. (2013). An elephant corridor in a fragmented conservation landscape: preventing the isolation of Mount Kenya National Park and National Reserve. Parks, 19(1), 91-101.

Packer, C., Loveridge, A., Canney, S., Caro, T., Garnett, S., Pfeifer, M., Zander, K., Swanson, A., MacNulty, D., Balme, G., Bauer, H., Begg, C., Begg, K., Bhalla, S., Bissett, C., Bodasing, T., Brink, H., Burger, A., Burton, A., Clegg, B., Dell, S., Delsink, A., Dickerson, T., Dloniak, S., Druce, D., Frank, L., Funston, P., Gichohi, N., Groom, R., Hanekom, C., Heath, B., Hunter, L., DeIongh, H., Joubert, C., Kasiki, S., Kissui, B., Knocker, W., Leathem, B., Lindsey, P., Maclennan, S., McNutt, J., Miller, S., Naylor, S., Nel, P., Ng'weno, C., Nicholls, K., Ogutu, J., Okot-Omoya, E., Patterson, B., Plumptre, A., Salerno, J., Skinner, K., Slotow, R., Sogbohossou, E., Stratford, K., Winterbach, C., Winterbach, H. and Polasky, S. (2013). Conserving large carnivores: dollars and fence. Ecology Letters, 16(5), 635-641.

Pangle, W.M., and Holekamp, K. (2010). Lethal and nonlethal anthropogenic effects on spotted hyenas in the Masai Mara National Reserve. Journal of Mammalogy, 91(1), 154–164.

Peel, M.C., Finlayson, B.L., and McMahon, T.A. (2007). Updated world map of the Köppen- Geiger climate classification, Hydrology and Earth System Sciences, 11, 1633-1644

Periquet, S., Fritz, H. and Revilla, E. (2014). The Lion King and the Hyaena Queen: large carnivore interactions and coexistence. Biological Reviews, 90(4), 1197-1214.

70

Periquet, S., Mapendere, C., Revilla, E., Banda, J., Macdonald, D., Loveridge, A. and Fritz, H. (2015). A potential role for interference competition with lions in den selection and attendance by spotted hyaenas. Mammalian Biology - Zeitschrift für Säugetierkunde, 81(3), 227-234

Pianka, E.R. (1974). Niche overlap and diffuse competition. Proceedings of the National Academy of Sciences, 71(5), 2141-2145.

Power, R.J. and Compion, R.X.S. (2009) Lion predation on elephants in the Savuti, Chobe National Park, Botswana. African Zoology, 44(1), 36-44.

Pratt, L. (2014). Co-conserving Threatened and Endangered species: the Paradoxical Kenyan Fence. Unpublished Master’s thesis. University of Southampton.

Preston, B. (2015). Multi-Species Management in Small Fenced Reserves: Consequences for Conservation and Management. Unpublished Master’s thesis. University of Southampton.

Purchase, G. (2004). Factors Affecting the Ratio of Lions (Panthera leo) to Spotted Hyaenas (Crocuta crocuta) in Protected Areas of Africa: Competition and/or Available Prey. Unpublished PhD thesis. University of Aberdeen.

R Core Team, 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Available at: http://www.r-project.org/.

Riggio, J., Jacobson, A., Dollar, L., Bauer, H., Becker, M., Dickman, A., Funston, P., Groom, R., Henschel, P., de Iongh, H., Lichtenfeld, L., and Pimm, S. (2013). The size of savannah Africa: a lion’s (Panthera leo) view. Biodiversity and Conservation, 22(1), 17-35.

Rodgers, A.R., Kie, J.G., Wright, D., Beyer, H.L., and Carr., A.P. (2015). HRT: Home Range Tools for ArcGIS. Version 2.0. Ontario Ministry of Natural Resources and Forestry, Centre for Northern Forest Ecosystem Research, Thunder Bay, Ontario, Canada.

71

Schmitz, O., Hambäck, P. and Beckerman, A. (2000). Trophic Cascades in Terrestrial Systems: A Review of the Effects of Carnivore Removals on Plants. The American Naturalist, 155(2), 141-153.

Schuette, P., Wagner, A., Wagner, M. and Creel, S. (2013). Occupancy patterns and niche partitioning within a diverse carnivore community exposed to anthropogenic pressures. Biological Conservation, 158, 301-312.

Shaw, K.C., Van Meter, P.E., and Holekamp, K.. (2012). Variation among free-living spotted hyenas in three personality traits. Unpublished PhD thesis. Michigan State University.

Sillero-Zubiri, C., and Gottelli, D. (1991). Threats to Aberdare rhinos: predation versus poaching. Pachyderm, 14, 37-38.

Smale, L., Nunes, S., and Holekamp, K. (1997). Sexually dimorphic dispersal in : patterns, causes, and consequences. Advances in the study of behaviour, 26, 181-251.

Stander, P. E. (1991). Demography of lions in the Etosha national park, Namibia. Madoqua, 18(1), 1-9.

Treves, A. and Karanth, K. (2003). Human-Carnivore Conflict and Perspectives on Carnivore Management Worldwide. Conservation Biology, 17(6), 1491-1499.

Vieira, E. M. and Port, D. (2007), Niche overlap and resource partitioning between two sympatric fox species in southern Brazil. Journal of Zoology, 272: 57–63.

Wackernagel, M., Schulz, N. B., Deumling, D., Linares, A. C., Jenkins, M., Kapos, V., Monfreda, C., Loh, J., Myers, N., Norgaard, R., and Randers, J. (2002). Tracking the ecological overshoot of the human economy. Proceedings of the National Academy of Sciences, 99(14), 9266-9271.

Wallace Jr, R. K. (1981). An assessment of diet-overlap indexes. Transactions of the American Fisheries Society, 110(1), 72-76.

72

Watts, H., and Holekamp, K. (2008). Interspecific competition influences reproduction in spotted hyenas. Journal of Zoology, 276(4), 402-410.

Watts, H. and Holekamp, K. (2009). Ecological Determinants of Survival and Reproduction in the Spotted Hyena. Journal of Mammalogy, 90(2), 461-471.

Weber, W. and Rabinowitz, A. (1996). A Global Perspective on Large Carnivore Conservation. Conservation Biology, 10(4), 1046-1054.

Woodmansee, K.B., Zabel, C.J., Glickman, S.E., Frank, L.G., and Keppel, G. (1991). Scent marking (pasting) in a colony of immature spotted hyenas (Crocuta crocuta): A developmental study. Journal of Comparative Psychology, 105(1), 10.

Woodroffe, R. (2000). Predators and people: using human densities to interpret declines of large carnivores. Animal Conservation, 3(2), 165-173.

Woodroffe, R., and Ginsberg, J.R. (1998). Edge effects and the extinction of populations inside protected areas. Science, 280(5372), 2126-2128.

Yirga, G.., Bauer, H., Gebrihiwot, K., and Deckers, J. (2011). Peri-urban spotted hyena (Crocuta crocuta) in northern Ethiopia: diet, economic impact, and abundance. European Journal of Wildlife Research, 57(4), 759-765.

Yirga, G., De Iongh, H., Leirs, H., Gebrihiwot, K., Deckers, J. and Bauer, H. (2012). Adaptability of large carnivores to changing anthropogenic food sources: diet change of spotted hyena (Crocuta crocuta) during Christian fasting period in northern Ethiopia. Journal of Animal Ecology, 81(5), 1052-1055.

Yirga, G., Ersino, W., De Iongh, H. H., Leirs, H., Gebrehiwot, K., Deckers, J., and Bauer, H. (2013). Spotted hyena (Crocuta crocuta) coexisting at high density with people in Wukro district, northern Ethiopia. Mammalian Biology-Zeitschrift für Säugetierkunde, 78(3), 193-197.

73

Appendix A

Figure A.1 – Collaring sites for hyenas and lions between February – April 2016.

74

Appendix B

Capture Data Sheet Date Sex Time Age (approx.) ID Collar Freq. Condition (1-5) Area/ Grid ref.

Skull length Weight Skull width Mane lower jaw Nose – tail root (HB) Mane cheek Tail Mane at crest skull Chest girth(half x2) Mane b/w shoulders

Shoulder height Mane mid chest Canine upper length curve Comments:- Canine upper base width Time dart in: Canine length straight Time to Ataxa: Canine lower length Time to Recumbency: Canine lower base width Reversal time: Wear on teeth Recovery time: Front paw length Front paw width Drugs: Rear leg (knee - paw) Immobilisation (Mg/Kg): Base of tail circum. Reversal (Mg/Kg):

Table B.1 – Sample collaringSmooth data sheet Immobilisation? template.

Top-up Administered?

Smooth Recovery? Condition mammae: Blood sample: Nasal swab: Hair sample: Other:

75 Other:

Appendix C

Figure C.1 – Progression of collar deployment and collar data windows during field season.

76