POPULATION GENETICS OF THE SCRUB- AT KENNEDY

SPACE CENTER

by

CORY JAMES SPERN

B.S., University of Central Florida

A thesis submitted to the Department of Biological Sciences at Florida Institute of Technology in partial fulfillment of the requirements for the degree of

MASTERS OF SCIENCE in ECOLOGY

Melbourne, Florida May 2017

POPULATION GENETICS OF THE -JAY AT KENNEDY SPACE CENTER

A THESIS

by

CORY JAMES SPERN

Approved as to style and content by:

______Christin L. Pruett, Ph.D., Chairperson Assistant Professor Department of Biological Sciences

______Robert van Woesik, Ph.D., Member Professor Department of Biological Sciences

______Munevver Subasi, Ph.D., Member Assistant Professor Department of Mathematical Sciences

______David R. Breininger, Ph.D., Member Ecologist Kennedy Space Center

______Richard B. Aronson, Ph.D. Professor and Head Department of Biological Sciences

May 2017

ABSTRACT

POPULATION GENETICS OF THE FLORIDA SCRUB-JAY AT KENNEDY SPACE CENTER

by Cory James Spern, B.S., University of Central Florida

Chairperson of Advisory Committee: Christin L. Pruett, Ph.D.

The Florida Scrub-Jay ( coerulescens) has suffered ~90% reduction in population size due to degradation of . This loss of habitat has led to the fragmentation of many subpopulations and it is unknown how this change in connectivity has affected their population genetics. A greater understanding of how populations have changed through time can be achieved using genetics. Past population genetic studies have primarily focused at the state-wide level, but more information is needed about how groups interact at smaller spatial scales. Using nine microsatellite loci, the genetic structure of scrub-jay was investigated at the

Kennedy Space Center (KSC) Brevard County, Florida. This study examined subpopulations at a spatial scale of ~35 km, compared with coarse grain studies, that were at larger spatial scales (>75 km). Genetic diversity, subpopulation connectivity, and structure were examined among four subpopulations. Results indicate that scrub-jays on KSC have similar levels of genetic diversity in comparison with previous county and state-wide studies, and that scrub-jays at

KSC exhibit limited subpopulation structure, indicating levels of gene flow high iii

enough to limit divergence among sites. Results indicate that connectivity among

Scrub Reserve Units (SRU) is currently not limited by fragmentation, most likely due to corridors facilitating scrub-jay movement among subpopulations and immigration from patches of scrub habitat where genetic samples of scrub-jay were not taken. However, if these sites become isolated, it is possible for genetic diversity to be lost and inbreeding to occur.

Isolated populations are at a greater risk of due to potential inbreeding and loss of genetic diversity. Inbreeding of closely related individuals can increase the chance of receiving identical alleles (homozygosity), possibly causing inbreeding depression through the exposure of deleterious recessive alleles.

Two long-term pedigree datasets (N = 253 pairings) were collected for ~25 years at

KSC from two of the subpopulations. Pedigrees were constructed based on parentage data to calculate levels of inbreeding for individuals. In spite of several state-wide disease epidemics over the past 30 years, causing declines in the population at KSC, only one instance was found with an inbreeding coefficient of

0.25 due to a parent-offspring pairing. Therefore, inbreeding among close relatives is uncommon in scrub-jay at two of four sites and is likely to be avoided by this . Subpopulation-wide inbreeding values (FIS) were calculated from microsatellite data from all four subpopulations with values ranging from -0.02 to

0.12 showing a range of inbreeding estimates. These results suggest that in some

iv

subpopulations, inbreeding has occurred possibly due to small population size and the inability to avoid low-level inbreeding.

In small populations, low effective population size (Ne) can increase the risk of extinction. Small populations are more susceptible to stochastic events, losses in heterozygosity, and genetic drift. Determining Ne can give more information on the state of the population than census size alone, given some individuals in the population do not breed. Several different methods were used based on genetic and demographic data to calculate current and historical effective population size.

Current Ne, using genetic methods, ranged from 482–865 and based on a demographic method, was 547 with an Ne/N ratio of 0.66. Historical levels of Ne were roughly 30 times greater (Ne = 29,550) than current size, before scrub-jay began to decline ~7,500 YBP. In the population on KSC, current Ne has been markedly reduced compared to historical numbers, though it appears this was due to climatic changes thousands of years ago before humans had any significant influence on scrub-jay populations and habitat. Ultimately, the remaining populations require continued habitat management to maintain genetic diversity and to maximize long-term evolutionary potential given ongoing threats to scrub- jay populations caused by habitat loss and alteration.

v

TABLE OF CONTENTS

ABSTRACT ...... iii TABLE OF CONTENTS ...... vi LIST OF TABLES ...... viii LIST OF FIGURES ...... ix ACKNOWLEDGEMENTS ...... xi

CHAPTER I: FLORIDA SCRUB-JAY ON KENNEDY SPACE CENTER ...... 1 INTRODUCTION ...... 1 AND HABITAT PREFERENCES ...... 1 COOPERATIVE BREEDING, CONNECTIVITY, AND DISPERSAL ...... 5 PREVIOUS GENETIC STUDIES ...... 11

CHAPTER II: GENETIC DIVERSITY AND POPULATION STRUCTURE OF FLORIDA SCRUB-JAY ON KENNEDY SPACE CENTER ...... 13 INTRODUCTION ...... 13 HABITAT, DISPERSAL AND COOPERATIVE BREEDING ...... 14 FLORIDA SCRUB-JAY CORE POPULATIONS ...... 16 POPULATION GENETICS ...... 18 RESEARCH QUESTIONS ...... 22 MATERIALS AND METHODS ...... 23 SAMPLING, GENOTYPING, AND RELATEDNESS ...... 23 GENETIC DIVERSITY AND POPULATION BOTTLENECKES ...... 25 POPULATION STRUCTURE AND GENEFLOW ...... 26 RESULTS ...... 30 SAMPLING, GENOTYPING, AND RELATEDNESS ...... 30 GENETIC DIVERSITY AND POPULATION BOTTLENECKS ..... 30 POPULATION STRUCTURE AND GENEFLOW ...... 30 DISCUSSION ...... 35 GENETIC DIVERSITY AND RECENT BOTTLENECKS ...... 35 POPULATION STRUCTURE AND GENEFLOW ...... 37

CHAPTER III: INBREEDING AND EFFECTIVE POPULATION SIZE OF FLORIDA SCRUB-JAY ON KENNEDY SPACE CENTER ...... 40 INTRODUCTION ...... 40 INBREEDING IN WILD POPULATIONS ...... 40 EFFECTIVE POPULATION SIZE ...... 44 FLORIDA SCRUB-JAY AND SCRUB HABITAT IN DECLINE .... 46 RESEARCH QUESTIONS ...... 50 MATERIALS AND METHODS ...... 51 vi

GENETIC DATA...... 51 INBREEDING ...... 51 EFFECTIVE POPULATION SIZE ...... 53 RESULTS ...... 58 INBREEDING ...... 58 EFFECTIVE POPULATION SIZE ...... 58 DISCUSSION ...... 65 INBREEDING ...... 65 EFFECTIVE POPULATION SIZE ...... 68 GUIDELINES FOR LAND MANAGERS ...... 73

REFERENCES ...... 75

Appendix ...... 86

vii

LIST OF TABLES

Table II.1. List of microsatellite primers, number of alleles, allele sizes, and annealing temperature for Florida Scrub-Jay (Stenzler and Fitzpatrick 2002) ...... 24

Table II.2. Genetic diversity of Florida Scrub-Jay populations at the Kennedy Space Center, Khodadad (2008), and Coulon et al. (2008). Number of individuals (N), observed heterozygosity (HO), expected heterozygosity (HE), and allelic richness (AR) ...... 31

Table III.1. Parameters used in the minimal method (Nunney and Elam 1994) for estimating demographic effective population size ...... 57

Table III.2. Inbreeding coefficient (FIS) of four Florida Scrub-Jay subpopulations at Kennedy Space Center, FL. FIS is a measurement of relatedness by descent within a group...... 60

Table III.3. Effective population size estimates using three different methods and historical start of population decline using MSVAR. 95% CI – Bayesian credibility intervals...... 64

viii

LIST OF FIGURES Figure I.1. Aerial and satellite imagery showing habitat alterations over 70 years. Top image (1943) shows more open habitat and less anthropogenic development. Bottom image shows anthropogenic disturbance (2013) ...... 7

Figure I.2. Transitional stages of scrub habitat a) short, b) optimal, c) tall- mixed and, d) tall. Based on natural fire cycles, scrub patches will exist in various growth stages known as a patchy mosaic ...... 8

Figure I.3. Florida Scrub-Jay in optimal scrub habitat...... 8

Figure II.1. Four Florida Scrub-Jay subpopulations and the number of individuals sampled at Kennedy Space Center on Merritt Island: Shiloh, Happy Creek, Schwartz Road, and Tel-4. Distances between sites are shown as arrows: Shiloh-Happy Creek ~20 km, Happy Creek-Schwartz Road ~9 km, and Schwartz Road-Tel4 ~8 km ...... 21 Figure II.2. Graph of the likelihood [-ln(X|K)] of the number of genetic clusters (K) based on STRUCTURE output of Florida Scrub-Jay at Kennedy Space Center ...... 32 Figure II.3. Results of Principal Coordinates Analysis (PCoA) at four Scrub Reserve Units at the Kennedy Space Center: Shiloh (SH), Happy Creek (HC), Schwartz Road (SR), and Tel-4 (T4). Lack of subdivision indicates panmixia ...... 33 Figure II.4. Map showing assignment from Geneclass2. Red arrows indicate individuals assigned to Happy Creek but sampled in other locations. Green arrow indicates individuals assigned to Schwartz Road that are currently found at Tel-4 ...... 34 Figure III.1. Pedigree from census data showing inbreeding in Florida Scrub-Jay at the Tel-4 location (Fig. II.1, p. 21). A parent sibling pairing of mother (S-LW) with son (LO-SL) producing offspring (SL-). Squares and circles around individuals represent males and females respectively...... 60 Figure III.2. MSVAR estimates of current and historical effective population size (Ne) for Florida Scrub-Jay at Kennedy Space Center based on kernel density estimates ...... 61

ix

Figure III.3. MSVAR output showing kernel density estimate of time (in log years) since effective population size decline in the Florida Scrub- Jay at the Kennedy Space Center, approximately 7,500 years (dashed line) ...... 61 Figure III.4. Comparison of maturation time (M) and effective population size (Ne) based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate ...... 62

Figure III.5. Comparison of standardized variance in lifespan (Iai) and effective population size to census size ratio (Ne/N) based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate...... 62 Figure III.6. Proportion of females fledging one or more offspring (p) and effective population size to census size ratio (Ne/N) based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate...... 63

x

ACKNOWLEDGEMENTS

I would like to thank my advisor, Dr. Christin Pruett for her guidance and advice throughout my time at Florida Institute of Technology. I am grateful to

David Breininger and Geoff Carter for their expertise on scrub-jays, answering all of my questions, and collecting samples. I am thankful for my committee for their advice and help in sculpting my research questions. I would like to thank Angie

Ricono for her help in PCR and running gels. Of course, I am grateful to Deirdre

Van Horn for every extra form or task I needed to get done, and always being willing to talk. And finally, I want to thank my parents for their never-ending support.

Furthermore, none of this would have been possible if not for Kennedy

Space Center for providing funding to perform all of the genetic work. If it were not for the Ecology Program at Kennedy Space Center, I would not have had samples to analyze. So, I thank Stephanie Legare for all the hours and effort collecting samples. I am grateful to Carlton Hall for general advice and allowing me to intern and work at the Kennedy Space Center, Angela Czupta for helping me with all the badging issues, and Eric Stolon for his advice on habitat and movement modeling.

xi

1

CHAPTER I

FLORIDA SCRUB-JAY ON KENNEDY SPACE CENTER

INTRODUCTION

Based on the fossil record, current extinction rates on Earth have been estimated at ~ 1,000 times higher than natural background levels (de Vos et al.

2014). Humankind’s negative interactions with the environment are closely tied to climate change (Benito-Garzon et al. 2013) and habitat alteration (Woolfenden and

Fitzpatrick 1984). Anthropogenic actions have led to the reduction and isolation of many species throughout their ranges. These influences are compounded for species that are endemic, habitat specialists (Woolfenden and Fitzpatrick 1984,

Clavel et al. 2011).

CONSERVATION STATUS AND HABITAT PREFERENCES

Florida Scrub-Jay (Aphelocoma coerulescen) is the only species endemic to Florida (Woolfenden and Fitzpatrick 1984) and is a species of conservation concern that was listed by the U.S. Fish and Wildlife Service

(USFWS) as Threatened in 1987 (Mumme et al. 2000). The International Union for

Conservation of Nature (Birdlife International 2012) lists the Florida Scrub-Jay as

Vulnerable. The reason for these listings is that populations have declined to 7,000-

10,000 individuals (~90% loss compared with pre-European colonization,

2

Boughton and Bowman 2011), in roughly 4,000 family groups across the state

(Mumme et al. 2000). Declines are linked to anthropogenic exploitation of unique types of habitat (Woolfenden and Fitzpatrick 1984).

As their name suggests, scrub-jays are found in scrub habitat. Scrub was common in Florida during the late Pleistocene (44,000-10,000 YBP; Myers and

Ewel 1990). During the Wisconsin glaciation, Florida became much drier; this was most likely caused by a decrease in sea level, reduction in the water table, a decrease in precipitation, or a combination of these factors (Watts 1980). During the Holocene, however, scrub became rare as the climate shifted to wetter conditions, in part due to rising sea levels and subsequent increases in the Florida water table (Myers and Ewel 1990, van Soelen et al. 2012). Scrub is now restricted to scattered pockets throughout Florida on ridge systems (Root 1998, Breininger

1989).

Along with changes in historical climatic patterns, more recent anthropogenic land use has altered scrub habitat within Florida. Native Americans began changing the landscape of Florida approximately 800 years ago through the beginnings of agriculture and use of fire (Myers and Ewel 1990). However, when

European colonists arrived in the 1700s, alteration of the landscape occurred on a greater scale (Myers and Ewel 1990). A good illustration of anthropogenic land use can be seen on Merritt Island, specifically around the Kennedy Space Center

3

(KSC), which is considered one of the three core populations of Florida Scrub-Jay

(Stith et al. 1996).

Initially, sugarcane farming altered scrub in the late 1700s on

Merritt Island, where KSC is now located (U.S. Fish and Wildlife Service

(USFWS) 2014). The climate in Florida is also favorable for the cultivation of citrus, which requires sandy soils and minimal exposure to frosty temperatures

(Myers and Ewel 1990). Consequently, citrus groves were planted in scrub and caused a substantial loss of habitat for scrub-jays throughout their distribution.

Further alterations in habitat have occurred due to human population expansion and development (Schmalzer et al. 2002). In 1962, the National Aeronautics and Space

Administration (NASA) purchased north Merritt Island to further the burgeoning space program, and it was during this time that the Merritt Island National Wildlife

Refuge (MINWR) was established on KSC property (Schmalzer et al. 1994). The refuge was responsible for managing much of the ~57,000 ha of land and lagoon not utilized by KSC (Schmalzer et al. 1994). In the 1990s, the Fish and Wildlife Service (USFWS) instituted a compensation plan in which NASA was required to mitigate scrub habitat lost due to construction projects around KSC

(Schmalzer et al. 1994). This could be accomplished either through habitat restoration or the creation of new scrub habitat (i.e. from old agricultural sites;

Schmalzer et al. 1994).

4

Along with anthropogenic habitat changes, fire suppression was also implemented by KSC about 50 years ago, further degrading scrub habitat (Fig. I.1;

Myers and Ewel 1990, Breininger et al. 1996b). Fire is a crucial element in maintaining scrub habitat (Myers and Ewel 1990). Historically, scrub was maintained by fire at a frequency of 5-20 years per burn (Main and Menges 1977).

When fire suppression occurs, scrub habitat can become overgrown, Florida Scrub-

Jay populations decline, and family groups can be outcompeted by Blue Jays

( cristata) that prefer denser, forested habitats (Woolfenden and

Fitzpatrick 1984). Fragmentation of scrub habitat can also limit the spread of fire, preventing unconnected patches of scrub from burning (Fitzpatrick et al. 1991).

Predation is also more common in areas with overgrown scrub patches or tree canopies. Canopy height can hinder effective sentinel behavior in scrub-jays, while increasing the number of hunting perches for predators such as hawks and owls

(Fitzpatrick et al. 1991). Nest depredation by is also more common in overgrown areas (Carter et al. 2007). Currently, habitat is maintained and restored using prescribed fire (Schmalzer and Hinkle 1992). Land that requires burning has been divided up into fire management units (FMUs) across KSC (Schmalzer et al.

1994). Breininger et al. (2014) quantified the influence that habitat quality, breeder experience, and helpers had on recruitment in Florida Scrub-Jay, based on data from 22 years at 36 scrub sites along the Atlantic Coast of central Florida. Scrub habitat can be classified into our transitional states, based on time since fire (Fig.

I.2). These four states are short ( with a height < 1.3 m), optimal (oak height of

5

1.3-1.7 m), tall-mix, and tall (oak height > 1.7 m). Breininger et al. (2014) determined that oak scrub of 1.3-1.7 m tall with sandy openings was the optimal habitat for Florida Scrub-Jay (Fig. I.3; Root 1998). Territories in these non-optimal states have low recruitment of scrub-jay (Breininger et al. 2014).

Several of the sites used for the scrub mitigation plan were also proposed to be part of a network of primary scrub habitat for Florida Scrub-Jay conservation around KSC (Breininger et al. 1996b). These sites, known as scrub reserve units

(SRUs), included Shiloh, Happy Creek, Schwartz Road, and Southern Pinelands

(Tel-4 area), south of the space center industrial area (Fig. II.1, p. 7). The Merritt

Island National Wildlife Refuge Comprehensive Conservation Plan has set a goal to increase the number of scrub-jay families to 500-650 on KSC over the next 8 years, as this was a 15 year plan (USFWS 2008). Currently, there are ~275 breeding pairs at KSC (David Breininger, personal comm.).With proper management, the USFWS (2008) believes that it is possible to double the number of breeding pairs.

COOPERATIVE BREEDING, CONNECTIVITY, AND DISPERSAL

The historical patchiness of scrub habitat likely forced the population density of scrub-jays to reach a saturation point so that all available habitat was occupied by breeding pairs (Woolfenden and Fitzpatrick 1984). Saturation of habitat is believed to be one of the main drivers of the of cooperative breeding in this species (Breininger et al. 2010). Cooperative breeding typically

6 involves a breeding pair and several helpers (usually offspring from previous years), who to varying degrees, help rear young, watch for predators, and defend territory (Breininger et al. 2006). Studies have indicated that even one-year-old males are capable of pairing and producing offspring (Woolfenden and Fitzpatrick

1984). However, offspring forego breeding for several seasons until there is a breeding vacancy or a dispersal event occurs (Woolfenden and Fitzpatrick 1984), providing support that a lack of habitat is driving this behavior. The relationship between family size and territory size is rather complex (Woolfenden and

Fitzpatrick 1984). Territory size is greatly influenced by quality of habitat, which has a direct effect on reproductive success and family size (Woolfenden and

Fitzpatrick 1984, Fitzpatrick et al. 1991, Breininger and Oddy 2004). Higher quality scrub gives breeding pairs a greater chance of successfully producing offspring and increasing territory size (Woolfenden and Fitzpatrick 1984,

Breininger et al. 2009).

7

Figure I.1. Aerial and satellite imagery showing habitat alterations over 70 years. Top image (1943) shows more open habitat and less anthropogenic development. Bottom image shows anthropogenic disturbance (2013).

8

Figure I.2. Transitional stages of scrub habitat a) short, b) optimal, c) tall-mixed and, d) tall. Based on natural fire cycles, scrub patches will exist in various growth stages known as a patchy mosaic.

Figure I.3. Florida Scrub-Jay in optimal scrub habitat.

9

In other species of jays in the genus Aphelocoma, multiple degrees of sociality have evolved (Peterson and Burt 1992). According to Peterson and Burt

(1992), Aphelocoma jays ancestrally bred as plural cooperative breeding groups, in which multiple females would breed in the same territory. As Aphelocoma jays occupied new habitats, sociality began to vary (Peterson and Burt 1992). Peterson and Burt (1992) hypothesized that southern Mexico and Florida scrub-jays began singular cooperative breeding before speciation events occurred, while scrub-jays in the western U.S. began breeding as pairs without any helpers. Burt and Peterson

(1993) studied cooperative breeding of A. californica in Mexico and concluded that habitat saturation and interspecific competition were not the reasons for the evolution of cooperative breeding, as it was in Florida (Woolfenden and Fitzpatrick

1984, Burt and Peterson 1993). Burt and Peterson (1993) proposed several reasons for these differences. First, Western Scrub-Jays of southern Mexico use multiple habitat types; not limited to oak-scrub habitat like Florida Scrub-Jay. Second, anthropogenic habitat use has opened up previously unusable habitat types, allowing scrub-jays of southern Mexico to utilize new areas, unlike scrub-jays in

Florida where anthropogenic land use has limited their range. And third, as these habitats vary in quality (and type), A. californica helpers have higher potential fitness by waiting for a higher quality territory to become available, instead of dispersing to an unknown area, as is the case with scrub-jays in Florida. Ultimately, it appears that Western Scrub-Jays in southern Mexico have likely been able to

10 adapt to changing habitats more effectively than Florida Scrub-Jays, which are restricted to very limited optimal scrub habitat (Burt and Peterson 1993).

Florida Scrub-Jay typically do not breed until 2-3 years of age (Woolfenden and Fitzpatrick 1984). By delaying reproduction and lengthening generation time, populations of cooperative breeding are likely to have slower responses to changes in population size than non-cooperative breeders, because helpers are reliant on vacancies in territories to begin breeding (Woolfenden and Fitzpatrick

1984). Consequently, stochastic events (i.e. hurricanes and epidemics) can have a profound effect on population sustainability (Breininger et al. 1996a). These events, along with poor habitat, greatly increase the risk of extinction, as large numbers of individuals can be lost during these times, and recovery can be slow (Breininger et al. 1996a). In addition, Florida Scrub-Jay are short-distance dispersers (~2.8 km for females and 1.6 km for males in fragmented habitats; Breininger et al. 2006) and thus connectivity among habitat fragments is limited. Males may spend their entire lives in one territory, if they are successful enough to “bud” off from their parents and establish a territory. However, females may traverse several territories to become a helper or (eventual) breeder once a vacancy opens (Woolfenden &

Fitzpatrick 1984). It is important to note that secondary habitats, such as flat woods, are used as corridors to enable jays to reach fragments of scrub and thus jays can move through suboptimal habitats (Woolfenden and Fitzpatrick 1984,

Breininger et al. 1991).

11

Habitat connectivity was evaluated by Stith et al. (1996), who performed a statewide survey of Florida Scrub-Jay populations. One hundred ninety-one isolated groups were identified, however many of these groups had less than 10 breeding pairs (Stith et al. 1996). Based on the limited dispersal range of Florida

Scrub-Jay, Stith et al. (1996) created a 12-km buffer zone, the average furthest distance a scrub-jay will disperse, and included known barriers to movement, such as water and dense forested areas. Using this buffer zone of 12 km, Stith et al.

(1996) identified 42 metapopulations. Of these 42 metapopulations, Merritt Island

National Wildlife Refuge (on Kennedy Space Center property) was designated as a

“core metapopulation”, a population that has a very low chance of extinction (Stith et al. 1996).

PREVIOUS GENETIC STUDIES

McDonald et al. (1999) studied differences between the Western Scrub-Jay and Florida Scrub-Jay using ten microsatellite markers. Based on historical differences in climate and habitat between the two species, McDonald et al. (1999) hypothesized that Florida Scrub-Jay would have comparatively lower population size and therefore be prone to genetic drift and loss of genetic diversity. As expected, Florida Scrub-Jay had lower genetic diversity (HE = 0.148) than Western

Scrub-Jay (HE = 0.343). A state-wide population genetic survey was conducted to identify genetic units of Florida Scrub-Jay (Coulon et al. 2008). Ten to twelve genetic groups were identified across Florida (Coulon et al. 2008). The scrub-jay of

12

Kennedy Space Center, already known as one of the largest metapopulations in

Florida (Stith et al. 1996), were grouped together with scrub-jays in Volusia and

Flagler Country into a single genetic cluster (“C”; Coulon et al. 2008). Another coarse grain study, performed by Khodadad (2008), sampled scrub-jay groups across Kennedy Space Center and Brevard County, Florida. Khodadad (2008) found that smaller isolated groups had lower observed heterozygosity. Studies by

Coulon (2008) and Khodadad (2008) have been useful in understanding coarse grain dynamics of population genetics in Florida Scrub-Jay, however, small scale changes can be overlooked. Therefore, the scrub network at the Kennedy Space

Center is an ideal area to study fine grain population dynamics and the effect of dynamics on population genetics.

13

CHAPTER II

GENETIC DIVERSITY AND STRUCTURE OF THE FLORIDA SCRUB-JAY ON KENNDY SPACE CENTER

INTRODUCTION

The global increase in human population size has triggered reductions in natural lands around the planet at an extraordinary rate (UNEP 2007). Loss of habitat is considered one of the major threats to species at a global scale (UNEP

2007). A study by McKee et al. (2013) found that human population density was positively correlated with declines in mammal and bird species. Contemporary mass (in the Anthropocene) have occurred at such an alarming rate that these extinctions have been compared with the five previous mass extinctions, which have occurred throughout Earth’s history (Dirzo et al. 2014). Locally, populations are also going extinct, which is the precursor to species-level extinctions (Ceballos et al. 2015). Therefore, understanding risks at the population- level will be beneficial in preventing further losses at the species-level.

Habitat fragmentation leads to a loss in population size by decreasing the amount of available habitat, and by forming habitat islands throughout historically connected lands (Frankham et al. 2010). Fragmentation can ultimately lead to a loss of genetic diversity and an increased risk of extinction (Frankham et al. 2010).

Determining the connectivity among habitat islands is necessary in understanding if

14 these groups function as a single metapopulation, or as isolated populations

(Coulon et al. 2008). An understanding of metapopulation dynamics has been especially useful in identifying risks to the Florida Scrub-Jay (Aphelocoma coerulescens). Florida Scrub-Jay are an important study species for understanding the effects of habitat fragmentation on genetic diversity and connectivity, as scrub- jays have had a 90% loss in population size over the past century (compared with pre-European settlement numbers; Boughton and Bowman 2011). This loss is associated with scrub-jays unique relationship to scrub habitat. (Woolfenden and

Fitzpatrick 1984, Breininger 1989, Myers and Ewel 1990). Studies on Florida

Scrub-Jay have defined “subpopulations” as habitat patches being separated by ≤

3.5 km, which is considered the longest average dispersal distance (Stith et al.

1996). While “metapopulations” are patches separated by 12 km, which is considered the maximum distance a scrub-jay can disperse (Stith et al. 1996).

Genetics studies have shown, however, that it is possible for jays to move further distances than behavioral studies indicate, through the use of corridors (Breininger

1992a, Coulon et al. 2008). Therefore, for the purposes of this study,

“subpopulations” will be defined by scrub-jay’s ability to move between patches rather than distance alone.

HABITAT, DISPERSAL, AND COOPERATIVE BREEDING

The relationship between Florida Scrub-Jay and habitat is complex

(Woolfenden and Fitzpatrick 1984). Scrub is a relict habitat that is rare, due to dramatic historical climatic changes (Myers and Ewel 1990). During these climatic

15 changes, scrub-jay most likely became isolated and developed increased habitat specialization. Much of the recent decline in scrub-jay populations is primarily due to anthropogenic removal of habitat (Schmalzer et al. 2002) and fire suppression of natural lands (Breininger et al. 1996b). Fire suppression can have a long-term effect on scrub because oak, a primary component of scrub, becomes fire resistant after reaching a large size and can prevent the spread of fire to other patches of scrub

(Schmalzer et al. 1994). If habitat becomes degraded, scrub habitat can transition into unfavorable states which reduce survival of Florida Scrub-Jay (Breininger et al. 2009; Fig. I.1, p.7; Fig. I.2, p.8). Degraded scrub may also require mechanical treatments, as well as fire, to return it to an optimal state. Scrub-jays prefer scrub height of 1.3-1.7 m with sandy openings and minimal pine tree canopy on well- drained soil (Breininger 1992a, Root 1998). Secondary habitat, which consists of scrub and pine on poorly drained soil, is typically unfavorable for breeding territories, but useful as corridors (Breininger 1992a). Corridors are necessary when dispersal is restricted by barriers such as roads, buildings, forested habitats, and bodies of water (Stith et al. 1996, Harrison and Bruna 1999).

Compounded by specific habitat requirements, Florida Scrub-Jay are also short distance dispersers (Stith et al. 1996, Breininger et al. 2006). The evolution of short dispersal distance has been attributed to the fitness costs associated with dispersing into new areas (Coulon et al. 2010, Bonte et al. 2012) and the weak flight ability of Florida Scrub-Jay (Breininger 1992a). Therefore, scrub-jays rarely disperse more than a few kilometers (~2.8 km for females and 1.6 km for males in

16 fragmented habitats) from their natal territory (Breininger et al. 2006). If scrub patches become isolated, then limited gene flow and reduction in genetic diversity can reduce adaptability and increase the risk of extinction (Frankham et al. 2010).

The Florida Scrub-Jay is a cooperatively breeding bird that is non-migratory and defends the same territory year-round (Woolfenden and Fitzpatrick 1984).

Each family territory is large, averaging 10 ha (Carter et al. 2006). Helpers assist breeders in territory defense, mobbing predators, and aid in rearing young

(Breininger et al. 2006). Woolfenden and Fitzpatrick (1984) found that breeders with helpers have higher reproductive success than breeders without helpers. Also, experienced breeders produce more offspring than inexperienced breeders

(Breininger et al. 2003). Females disperse further distances than males to find a breeding vacancy (Breininger et al. 2006); while males typically have the opportunity of inheriting their natal territory or taking a section of their natal territory, known as territorial budding. Most scrub-jays do not have the opportunity to breed within the first year, as competition for territories is quite high. Consequently, birds do not breed until ~3 years of age (Woolfenden 1978,

Woolfeden and Fitzpatrick 1984). Thus, population recovery, after a reduction in size, can be slow due to the time needed to accumulate experienced breeders and helpers over successive years (Breininger et al. 2003).

FLORIDA SCRUB-JAY CORE POPULATIONS

During the mid-1990s, efforts were made to identify remaining scrub-jay groups and assess risks of extinction throughout Florida. Stith et al. (1996)

17 surveyed Florida Scrub-Jay populations across the state. A total of 191 groups were recognized, though many had less than 10 breeding pairs (Stith et al. 1996).

Metapopulations were then designated based on a maximum dispersal distance between scrub patches of 12 km. Stith et al. (1996) also identified three metapopulations as having greater than 100 breeding pairs, which are considered core populations. One of these core populations encompassed federal lands of

Kennedy Space Center (KSC) on Merritt Island.

The core population at KSC consists of four focal areas of scrub (scrub reserve units; SRUs) used as long-term study sites, Happy Creek, Tel-4, Schwartz

Road, and Shiloh (Fig. II.1). It should be noted that these SRUs have expanded in area since the initial studies began due to declines in overall population size of scrub-jay (Breininger et al. 1996a, David Breininger, personal communication). In

1988, Happy Creek had 23 breeding pairs; over time, it has increased to 43 breeding pairs (210 potential breeding pairs; Geoff Carter and David Breininger, personal communication). The Happy Creek unit required both prescribed fire and mechanical treatments to improve habitat quality (Schmalzer et al. 1994). Tel-4 is different from Happy Creek as it has more of a pine tree canopy (Breininger et al.

1996b). The study at Tel-4 began in 1989 with 15 breeding pairs and now has 74 breeding pairs (156 potential breeding pairs; Geoff Carter and David Breininger, personal communication). Studies at Schwartz Road began in 2000 with 12 breeding pairs and now has 20 breeding pairs (122 potential breeding pairs; Geoff

Carter, personal communication). Shiloh was proposed as the northern most SRU

18 in the early 1990s, however it required prescribed fire, mechanical treatments, and scrub restoration from old citrus groves (Schmalzer et al. 1994). A long-term study at Shiloh was initiated in 2010 when 16 breeding pairs were found, and currently, in 2017, there are 33 breeding pairs (114 potential breeding pairs; Geoff Carter, personal communication). Overall, there are ~275 breeding pairs at KSC, based on recent census estimates (David Breininger, personal communication). Due to fragmentation, the four study sites are geographically isolated from each other, however corridors between study sites do exist, as suboptimal secondary habitat

(Breininger 1992a, Breininger et al. 1996b; Fig. II.1). Thus, connectivity and maintenance of genetic diversity in the smallest sites is possible through dispersal events.

POPULATION GENETICS

Previous studies of the genetic diversity of Florida Scrub-Jay have focused on comparisons with closely related species, or within species comparisons at broad spatial scales. McDonald et al. (1999) sought to determine the influence of habitat and isolation between the Western Scrub-Jay (Aphelocoma californica) and

Florida Scrub-Jay. Unlike Florida Scrub-Jay, which have very limited, patchy habitat and short dispersal distances, Western Scrub-Jay have continuous habitat throughout the western U.S. and can disperse greater than 50 km (compared with 2-

3 km for Florida Scrub-Jay; Carmen 2004, Breininger et al. 2006). Based on these differences, McDonald et al. (1999) predicted that Florida Scrub-Jay would have greater genetic subdivision due to habitat isolation and shorter dispersal distances.

19

Also, McDonald et al. (1999) hypothesized that isolation of scrub patches on ridges within Florida most likely caused genetic drift in small populations, reducing genetic diversity. As predicted, McDonald et al. (1999) discovered that the genetic distance (GST) for the Florida Scrub-Jay populations was 0.048; while GST was

0.015 for Western Scrub-Jay populations. Also, Western Scrub-Jay had almost twice as many alleles per locus than (4.4 to 2.4, respectively), averaged across ten microsatellite loci.

Coulon et al. (2008) performed a state-wide study of Florida Scrub-Jay population genetics using microsatellite markers. They concluded that 10-12 genetic groups existed within Florida. Coulon et al. (2012) also discovered that isolation by distance was not as crucial as gap size (and types of barriers) between individuals in adjacent territories for limiting connectivity, indicating that scrub-jay use corridors during dispersal events (Breininger 1992a). One unpublished study of

Florida Scrub-Jay population genetics was conducted at Kennedy Space Center and surrounding mainland areas in Brevard County, between 2004-2007 (Khodadad

2008). Khodadad (2008) examined the relationship between habitat quality and genetic diversity in subpopulations. She defined habitat as optimal (primary scrub habitat for scrub-jay nesting territory) or suboptimal (overgrown scrub that is unfavorable; Khodadad 2008). She found a loss of heterozygosity in subpopulations with poor habitat quality, and that sites in suboptimal habitat had

FIS values above zero (suggesting inbreeding) based on microsatellite data. This

20 data is of great benefit in understanding the relationship that habitat plays in genetic diversity; however, no analyses were performed to assess connectivity among habitat fragments or overall population structure. The studies by Coulon et al.

(2008) and Khodadad (2008) are useful for understanding population genetics at a coarse grain scale. However, information about connectivity and population dynamics can be lost at such a broad scale. The study sites at Kennedy Space

Center were established as a network of scrub habitat (Breininger et al. 1996b); therefore to understand how and if these sites function as a network, a genetic study is needed at a finer spatial scale.

21

Figure II.1. Four Florida Scrub-Jay Subpopulations and the number of individuals sampled at Kennedy Space Center on Merritt Island: Shiloh, Happy Creek, Schwartz Road, and Tel-4. Distances between sites are shown as arrows: Shiloh-Happy Creek ~20 km, Happy Creek-Schwartz Road ~9 km, and Schwartz Road-Tel4 ~8 km (arrows not to scale).

22

RESEARCH QUESTIONS

The following questions were examined using microsatellite loci and population genetic analyses: 1) what is the current level of genetic diversity of

Florida Scrub-Jay at Kennedy Space Center within each Scrub Reserve Unit? 2)

Have any recent genetic bottlenecks occurred at Kennedy Space Center? And, 3) what kind of genetic structure and connectivity is found among Scrub Reserve

Units at Kennedy Space Center?

23

MATERIALS AND METHODS

SAMPLING, GENOTYPING, AND RELATEDNESS

Feathers were taken from nine to 24 Florida Scrub-Jay individuals at each SRU

(Fig. II.1). Samples were collected between 2013 and 2015. Birds were captured using baited potter traps and drop traps (Breininger et al. 2009). Several contour (or one instance in which a flight fell off the captured bird) were plucked and stored in paper envelopes to keep them dry until DNA could be extracted. The use of feathers for the purpose of DNA analyses has been found to be only moderately invasive when compared with venipuncture, especially when considering handling threatened or endangered species (Bush et al. 2005). The calamus was clipped from at least three feathers for each individual and then prepared following QIAGEN's tissue extraction method (Qiagen Inc., Velancia,

California, USA), along with the addition of 30 µL 1M dithiothreitol (DTT) as an aid in keratin digestion during the initial incubation step (De Volo et al. 2008).

After extractions were completed, PCR was used to amplify 9 microsatellite loci

(Table II.1; Stenzler and Fitzpatrick 2002). Six of the nine loci were chosen for comparison with earlier studies, as they were also used by Khodadad (2008). PCR amplifications were performed using a Bio-Rad MyCycler Thermocycler.

Thermocycler settings were optimized by Coulon et al. (2008) and are as follows: one cycle at 95.0 ºC for 2 min, 35 cycles at 95.0 ºC for 50s, 57-58 ºC for 1 min

(based on locus specific recommendations or temperature gradient trials; Table

24

II.1), and 72 ºC for 1 min; these cycles were followed by a final extension of 72 ºC for 30 min. Samples were then multiplexed, sent to the University of Florida genomics facility (ICBR), and run on an ABI 3130XL sequencer (Applied

Biosystems, Foster City, CA). Alleles were scored manually using Genemapper software (ABI, Foster City, CA). The software MICRO-CHECKER was run on genotyped microsatellites to detect null alleles, stutter bands, and large-allele drop out (Van Oosterhout et al. 2004, Barson et al. 2009).

Table II.1. List of microsatellite primers, number of alleles, allele sizes, and annealing temperature for Florida Scrub-Jay (Stenzler and Fitzpatrick 2002).

Allele Locus # Allele Annealing Temperature Size

ApCo 2* 12 217-254 58˚

ApCo 15 9 228-261 58˚

ApCo 18* 6 146-174 58˚

ApCo 22* 7 157-170 58˚

ApCo 23* 5 152-160 58˚

ApCo 29 13 96-145 58˚

ApCo 30* 11 203-245 57˚

ApCo 37 15 125-186 58˚

ApCo 41* 11 205-239 58˚

* indicates locus used by Khodadad (2008)

25

The program ML-RELATE (Kalinowski et al. 2006) was used to identify relatedness (r) of individuals within the same territory. This was accomplished by calculating the maximum likelihood of relatedness between pairs of individuals.

ML-RELATE calculates the likelihood of relatedness by identifying shared alleles through descent. This program assumes there is no inbreeding and a closed population (no immigration; Kalinowski et al. 2006). If two individuals had an r value of 0.5 or greater they could have a parent-offspring or sibling-sibling relationship (Blouin 2003). If individuals were found to be closely related, one of the individuals was removed from analysis to prevent bias. Exceptions were made for individuals that were several SRUs apart and shared almost completely fixed alleles in the subpopulation, as these individuals would appear related by chance

(Frankham et al. 2010).

GENETIC DIVERSITY AND POPULATION BOTTLENECKS

ARLEQUIN 3.5 software (Excoffier and Lischer 2010) was used to test for

Hardy-Weinberg equilibrium and linkage disequilibrium and to measure genetic diversity including expected heterozygosity (HE) and observed heterozygosity (HO).

Due to unequal sample sizes (Shiloh and Schwartz Road compared to Happy Creek and Tel-4; Table II.2), allelic richness was calculated using the software FSTAT

(Goudet 1995). Bonferroni corrections were made for multiple tests (Sokal and

Rohlf 1995).

26

During an initial decline in population size (a bottleneck), alleles are lost

(Frankham et al. 2010). However, during this period, there is a greater level of average heterozygosity, than would be predicted from the number of alleles alone

(Cornuet and Luikart 1996). This heterozygosity excess has been useful in identifying recent bottlenecks. Bottleneck version 1.2.02 (Piry et al. 1999) was used to determine if any subpopulation had experienced a recent bottleneck (within 4Ne generations, Cornuet and Luikart 1996). Following recommendations by Piry et al.

(1999), the two-phase model of mutation was used (10,000 replications) with 30% multistep mutations and tests of significance were performed using Wilcoxon signed-rank tests.

POPULATION STRUCTURE AND GENE FLOW

Subpopulation structure among SRUs was assessed using the programs

STRUCTURE (Pritchard et al. 2000, Hubisz et al. 2009) and Geneland (Guillot et al. 2005, Guillot et al. 2008). STRUCTURE assigns individuals to “genetic clusters” that are in Hardy-Weinberg and linkage equilibrium; no a priori geographical information is used in the analysis (Pritchard et al. 2000).

STRUCTURE uses a Markov Chain Monte Carlo (MCMC) approach to assign individuals to genetic groups. The admixture model was used, as scrub habitat was historically contiguous at KSC (Breininger et al. 1996b). Three independent runs were performed for each possible number of genetic clusters (K = 1 – 6) with

100,000 initial burn-in and 1,000,000 subsequent iterations. STRUCTURE outputs

27 were run through StructureHarvester (Earl and von Holdt 2012) to obtain graphical representations of the likelihood for each K value. These outputs were loaded into

DISTRUCT to graphically display population structure (Rosenberg 2004).

Geneland uses a similar method to that in STRUCTURE, but includes preloaded priors of geographical locations to aid in clustering populations.

Geneland was run 6 times with K of 1 to 6 clusters. Various combinations of initial burn-in (500; 1,000; 2,000), MCMC replicates (100,000; 1,000,000; 2,000,000), and delta coordinate (0.001, 0.0001, 0.00001) were used to find convergence.

Geographical coordinates were input as the centroid of each territory. This was done as jays are typically caught in an opening near a boundary where traps are easier to bait and use. Therefore, location of capture would not be a good representation of actual residence within a territory as there are several instances of individuals residing on the same territory (such as both members of a breeding pair). To account for this in Geneland, the delta coordinate was set in decimal degrees to adjust for individual overlap in each territory (Guillot et al. 2005). The delta coordinate is adjusted when there is a level of uncertainty as to the exact location of each individual (Guillot et al. 2005).

Arlequin was also used to compute subpopulation pairwise differentiation

(FST). Permutations (10,100) were performed to determine if FST values were significantly different from zero. Bonferroni correction was performed to correct for multiple tests (Sokal and Rohlf 1995). GenAlEx 6 was used to construct a

28 principal coordinates analysis graph (Peakall and Smouse 2006) based on individual genetic relatedness.

To test for recent migration events between subpopulations, GeneClass2 was used to perform assignment tests. GeneClass2 uses an MCMC method to identify genotype likelihoods for populations and then determines the likelihood an individual would belong to each population examined (Piry et al. 2004). The program simulated populations of 10,000 individuals to create probability thresholds for each study site. To identify first generation migrants (Piry et al.

2004, Paetkau al.2004), scrub-jays were ranked using the test statistic L=L-home

(Paetkau et al. 2004). This test statistic is used when there is uncertainty that all locations have been sampled (Breininger and Carter 2003, Paetkau et al. 2004).

Type I error, the chance that an individual would be classified as a migrant when it was not, was set to 0.01 (Paetaku et al. 2004). GeneClass2 was also used to determine how well individuals assigned to their respective subpopulations by using methods of Rannala and Mountain (1997) and Paetkau et al. (2004). If the probability of assignment was below 95% (P or α ≤ 0.05), that individual would be considered to have come from another location. Individuals not assigned to their population of origin, would be loosely assigned to another population based on the lowest mean Bayesian probability of assignment for all individuals in a subpopulation (P = 0.19). If an individual did not meet the minimum probability for any of the tested subpopulations, it was assumed to come from an unsampled

29 location, as there is scrub habitat and jays outside the study areas (Breininger and

Carter 2003).

30

RESULTS

SAMPLING, GENOTYPING, AND RELATEDNESS

DNA was successfully extracted from all birds. Seven individuals were removed from initial analysis based on relatedness. A total of 64 individuals were used in all analyses (Table II.2.)

GENETIC DIVERSITY AND POPULATION BOTTLENECKS

All loci were in Hardy-Weinberg equilibrium as well as in linkage equilibrium after Bonferroni correction for multiple tests. Observed heterozygosity

(HO), expected heterozygosity (HE), and allelic richness (AR) were similar for all subpopulations (Table II.2). Genetic diversity estimates from this study are comparable with other coarse grain studies of scrub-jays (Khodadad 2008, Coulon et al. 2008; Table II.2). Pairwise FST tests showed that no subpopulations were significantly different from zero after Bonferroni correction, indicating that there is very little divergence among subpopulations. Bottleneck version 1.2.02 did not indicate any evidence of a recent bottleneck (p > 0.05) for any subpopulation.

POPULATION STRUCTURE AND GENEFLOW

The clustering program STRUCTURE identified K = 1 as the most probable number of genetic clusters in the dataset (Fig. II.2). STRUCTURE indicated that all individuals are admixed; this suggests that subpopulations are panmictic and act as a single population. Six independent Geneland runs were unable to converge, even

31

with changing parameters such as burn-in lengths, number of MCMC replicates,

and the delta coordinate. This indicates that there was a lack of genetic information

to correctly identify clusters at nine loci or that there is only one genetic cluster in

the dataset (Guillot et al. 2005). PCoA confirms that there is little or no genetic

structuring among subpopulations based on individual relatedness (Fig. II.3.)

Table II.2. Genetic Diversity of Florida Scrub-Jay populations at the Kennedy Space Center, Khodadad (2008), and Coulon et al. (2008). Number of individuals (N), mean observed heterozygosity (HO), mean expected heterozygosity (HE), and allelic richness (AR).Standard deviations (SD) provided in parentheses after means. SD is unavailable for Coulon et al. (2008).

Location N HO HE AR

Shiloh 9 0.65 (0.29) 0.71 (0.20) 5.44

Happy Creek 20 0.73 (0.24) 0.72 (0.13) 5.69

Schwartz Road 11 0.68 (0.28) 0.77 (0.15) 5.45

Tel-4 24 0.68 (0.23) 0.72 (0.17) 5.56

Khodadad (2008) 48 0.63 (0.21) ------– Happy Creek

Khodadad (2008) 72 0.57 (0.23) ------– Tel-4

Coulon et al. (2008) – “Cluster 59 0.71 0.70 --- C”

32

Results from GeneClass2 indicate that three individuals were first generation migrants (P < 0.01). One individual from Shiloh and one individual from Schwartz Road were assigned to Happy Creek, while one individual from

Schwartz Road most likely originated from Tel-4. Results from the assignment tests indicate that five individuals could not be assigned to any subpopulation and most likely originated from an unsampled area (Fig. II.4).

Figure II.2. Graph of the likelihood [-ln(X|K)] of population clusters (K) based on STRUCTURE output of Florida Scrub-Jay at Kennedy Space Center.

33

1.000

0.800

0.600

0.400

SH 0.200

HC Coord. 2 Coord. SR 0.000 T4 -0.200

-0.400

-0.600 -0.600 -0.400 -0.200 0.000 0.200 0.400 0.600 0.800 Coord. 1

Figure II.3. Results of Principal Coordinates Analysis (PCoA) at four Scrub Reserve Units at the Kennedy Space Center: Shiloh (SH), Happy Creek (HC), Schwartz Road (SR), and Tel-4 (T4). Lack of subdivision indicates panmixia.

34

Figure II.4. Map showing assignment from Geneclass2. Red arrows indicate individuals assigned to Happy Creek but sampled in other locations. Green arrow indicates individuals assigned to Schwartz Road that are currently found at Tel-4. Yellow arrows indicate individuals from unsampled areas currently found in one of the scrub reserve units.

35

DISCUSSION

GENETIC DIVERSITY AND RECENT BOTTLENECKS

Kennedy Space Center, on Merritt Island, contains one of the largest core populations of Florida Scrub-Jay. Four main tracts of scrub habitat have been maintained and have the best chance of minimizing the risk of extinction. Distance between each study site, however, is greater than the average dispersal distance

(Fig. II.1; Stith et al. 1996). Thus, a lack of gene flow could cause small subpopulations to lose genetic diversity.

Based on the results from microsatellite data, each subpopulation had similar observed heterozygosity (HO; 0.65 – 0.73) and expected heterozygosity (HE;

0.71 – 0.77) values (Table II.2). The results of this study were also compared to two larger scale studies done in the mid-2000s. Khodadad (2008) found HO values were approximately 0.63 and 0.57 for Happy Creek and Tel-4, respectively. The statewide genetic survey by Coulon et al. (2008) clustered Kennedy Space Center in with Volusia and Flagler counties as cluster “C”. During the study, Coulon et al.

(2008) found HO and HE to be ~ 0.7 for cluster “C”. In comparison, another Florida

Scrub-Jay metapopulation, on the Lake Wales Ridge in the center of Florida

(cluster “B”), had an HE of ~0.69 (Coulon et al. 2008). These studies indicate that genetic diversity has not changed to a large extent in the past 7-8 years at Kennedy

Space Center. Given that scrub-jays have a generation time of approximately 7 years (Woolfenden and Fitzpatrick 1984), this is not a surprising result because

36 unless a population remains at a very small size (<10 individuals) over many generations, there is a minimal effect on heterozygosity (Frankham et al. 2010). A much larger effect is found in comparisons of the number of alleles that are lost in a single generation; however, comparisons with Coulon et al. (2008) or Khodadad

(2008) are not possible given the differences in sample size between studies as well as the greater geographical ranges in these studies (Table II.2). Khodadad (2008) also did not indicate that individuals were removed based on relatedness.

Therefore, it is possible that some of the randomly selected individuals used in her study could be relatives causing a reduction in Khodadad’s (2008) heterozygosity estimates. This could certainly be the case given Florida Scrub-Jays short dispersal distance (Stith et al. 1996). Another possibility for the lower heterozygosities in the

Khodadad (2008) study are the substantial population declines at Tel-4 during

2005-2007. Subsequent gene flow could have caused an increase in present heterozygosity measurements (Table II.2).

The similarity among subpopulations within the core at KSC, for both heterozygosity and allelic richness, suggests that there is gene flow among sites

(Table II.2). Thus, sites are acting as a network of interconnected scrub units and it is crucial that corridors are maintained to enable continued movement between these sites. Ultimately, for any management action to be successful, genetic diversity must be maintained to prevent extinction.

37

At least two mosquito-borne arbovirus epidemics caused statewide die-offs of scrub-jays in 1979 and 1997 (Breininger et al. 2009). Predation can also negatively influence scrub-jay population size (Breininger et al. 1996a). At Shiloh, the breeding population suffered substantial declines (Table. II.2), due to predation by migratory hawks in 2014-2015 (Geoff Carter, personal communication).

However, no recent bottlenecks were detected, even in areas with known small subpopulation size, such as at Shiloh and Schwartz Road. This further supports the idea that management and restoration efforts have successfully maintained connectivity among sites.

POPULATION STRUCTURE AND GENEFLOW

Florida Scrub-Jay are short distance dispersers (Breininger et al. 2006) and typically become sedentary after establishing territories (Woolfenden and

Fitzpatrick 1984). The core population at KSC consists of several fragmented

Scrub Reserve Units that are geographically isolated (Fig. II.1) with small pockets of scrub and secondary habitat (used as corridors; Breininger 1992a). In spite of this fragmentation, the program STRUCTURE identified a single genetic cluster

(Fig. II.2), indicating that there has been recent gene flow among subpopulations.

Lack of population divergence is further supported by the Geneland analysis that did not converge on the optimal likelihood, even when adjusting multiple parameters. Pair-wise FST values also confirm the high levels of gene flow as none of the values were significantly different from zero, after Bonferonni correction.

38

Gene flow among subpopulations and from unsampled locations has likely enabled the maintenance of genetic diversity despite losses in size of some subpopulations

(e.g., Tel-4 and Shiloh).

In a study of an endangered fish species, Lippe et al. (2006) found that despite a declining population, genetic diversity was still high due to a longer generation time. Florida Scrub-Jays have a mean generation time ~7 years

(Woolfenden and Fitzpatrick 1984). Maturation time, in conjunction with helpers awaiting breeding vacancies (Woolfenden and Fitzpatrick 1984), could act as a genetic buffer minimizing losses in genetic diversity to the population as a whole.

Scrub-jays, in less than favorable habitats, may also be immigrating into the four

Scrub Reserve Unit as there are individuals outside of the areas sampled

(Breininger and Carter 2003).

The program Geneclass2 was used to identify first generation migrants and assign individuals to subpopulations (Fig. II.4). Three individuals were determined to be first generation migrants. Two individuals were assigned to Happy Creek and one to Schwartz Road. Although subpopulations are geographically isolated beyond average dispersal distances (Fig. II.1; Stith et al. 1996), there is likely to be some connectivity through corridors. This is encouraging, considering that it is possible for a small number of individuals to restore much of the genetic diversity lost due to small population size (Ingvarsson 2001). Five individuals were not assigned to any area that was sampled (Fig. II.4). This can indicate high rates of

39 gene flow, or lack of genetic information from the nine microsatellites used in this study. Higher resolution from extra microsatellite markers could illuminate scrub- jay dispersal patterns at a finer spatial scale.

In spite of continual population decline, ongoing management efforts have successfully maintained a genetically diverse core population at Kennedy Space

Center. Extensive efforts have been made to utilize and implement management plans at the refuge since the early 1990s (Schmalzer et al. 1994). However, groups of Florida Scrub-Jay are still in decline (Birdlife International 2012). Kennedy

Space Center is one of the largest core metapopulations of Florida Scrub-Jay (Stith et al. 1996). The KSC scrub-jay population appears to currently have high levels of gene flow, though only time will tell whether land management will remain consistent and continue the same prescribed fire regime in the years to come. In the midst of global climate change (Frankham et al. 2010), it is difficult to know how scrub-jay populations will respond. Merritt Island is at most 3 m above sea level

(Breininger et al. 1996a), which could potentially reduce available habit as sea levels increase. Corridors must be maintained and created, especially around barriers such as industrial complexes and bodies of water. These corridors will ultimately increase gene flow between small subpopulations that may be impacted by predation or poor habitat. Ultimately land managers must focus on increasing scrub habitat, corridors, and fire regimes across the state of Florida if the scrub-jay has any chance of persisting.

40

Chapter IIII

INBREEDING AND EFFECTIVE POPULATION SIZE OF THE FLORIDA SCRUB-JAY ON KENNEDY SPACE CENTER

INTRODUCTION

INBREEDING IN WILD POPULATIONS Inbreeding is defined as the pairing of individuals who share one or more common ancestors (Frankham et al. 2010). Studies on inbreeding in laboratory settings (Frankham 2005) have been numerous; however until recently, inbreeding was thought to be rare in wild preferentially outbreeding populations. Due to the lack of genetic studies in wild populations, inbreeding was believed to have a minor role in extinction risk to small populations (Frankham et al. 2010). As genetic techniques became more accessible, many studies have been published on inbreeding in both plants and in the wild (Keller and Waller 2002). A number of these studies have shown that inbreeding can have detrimental effects to small populations (Frankham et al. 2010).

Inbreeding depression, the cumulative loss of fitness due to inbreeding, has been documented in a variety of taxa (Frankham et al. 2010). This can occur by unmasking of deleterious recessive mutations or the disruption of beneficial gene complexes (Frankham et al. 2010). Out of 154 species in 34 taxa that were inbred,

Crnokrak and Roff (1999) found that 90% of those tested had significant levels of

41 inbreeding depression. Traits associated with inbreeding depression include small body size, deformities, low reproductive success compared with non-inbred individuals, and low survival rates, which ultimately increase the risk of extinction

(Frankham et al. 2010). It is possible however, for a few immigrants, from an unrelated population, to have a large impact on genetically depauperate areas

(Ingvarsson 2001). This phenomenon, referred to as “genetic rescue”, has been used as a conservation measure to increase genetic diversity in a number of invertebrate, vertebrate, and plant species (Frankham et al. 2010, Johnson et al.

2010, Bateson et al. 2014, Frankham 2015, Harrison et al. 2016). However,

Hedrick and Fredrickson (2010) caution that increases in genetic diversity may only be temporary in small and isolated populations due to genetic drift. Without an increase in population size and suitable habitat, genetic rescue may only be a short- term solution.

The Florida Scrub-Jay is a cooperatively breeding bird (Woolfenden and

Fitzpatrick 1984). This breeding system consists of a breeding pair and several helpers that defend a territory year-round (Breininger et al. 2006a). Optimal territory size is ~ 10 ha, however this is dependent on quality of habitat in the territory and population density (Woolfenden and Fitzpatrick 1984). Male offspring typically stay in the same territory while waiting to inherit their natal territory, or obtain a small piece of a larger territory through the process known as “territory budding”. Alternatively, females may disperse several territories away as helpers

42 until a breeding vacancy opens (Woolfenden and Fitzpatrick 1984). These behaviors have been suggested as a way of limiting chances of inbreeding

(Woolfenden and Fitzpatrick 1984).

Inbreeding is believed to be rare in Florida Scrub-Jays because of “incest taboo”, which discourages individuals from mating within a group and is maintained by differences in dispersal distances between sexes (Woolfenden and

Fitzpatrick 1984). At Archbold Biological Station, two exceptions were noted by

Woolfenden and Fitzpatrick (1984): full-sibling pairings and a mother-son pairing.

The full-sib pairings occurred in a territory by a breeding pair that had a high rate of reproductive output and produced siblings over successive generations. Seven siblings formed pair-bonds and apparently produced fertile . The mother-son inbreeding occurred when a male breeder died and the son took his place, producing fertile eggs. The authors, however, did not discuss the fate of these chicks after hatching. Woolfenden and Fitzpatrick (1984) also suggested that once an individual leaves a family group, it most likely will lose the ability to recognize kin. Therefore, if two individuals are related, but come from different family groups or different generations, there is a chance they would pair and breed.

Coulon et al. (2008) used 19 microsatellites and determined that the population-level inbreeding coefficient (FIS) ranged between –0.1 and 0.06.

Considering an FIS value of 0.0625 is a first cousin pairing (Frankham et al. 2010), some of the groups do show signals of inbreeding. At Kennedy Space Center, FIS

43 was found to be 0.005 (Coulon et al. 2008). However, it should be noted that this was a coarse grain study with 59 individuals from three different counties (Brevard

– Kennedy Space Center, Volusia and Flagler; Coulon et al. 2008). Townsend et al.

(2011) examined extra pair paternity and inbreeding in habitats at various states of quality, which included natural scrub at Archbold Biological Station, a suburban site, and a highly-fragmented site. Townsend et al. (2011) found that ~17% of the offspring at the fragmented site had an inbreeding coefficient above zero, whereas only ~3% - 7% of offspring at the natural scrub and suburban sites had inbreeding coefficients above zero. This indicates that habitat quality can influence inbreeding potential. As subpopulation sizes continue to decline the likelihood of inbreeding increases; if inbreeding avoidance limits reproductive opportunities, then extinction is highly probable (Frankham et al. 2010).

Chen (2014) examined pedigree records and genetics of Florida Scrub-Jay at Archbold Biological Station from a ~30-year period. Two types of breeding pairs were designated “resident” pairs, in which both individuals were born in the study site, and “immigrant” pairs, when at least one of the individuals was born outside of the study site (Chen 2014). She then demonstrated that even if migrants came from genetically depauperate areas, genetic diversity was still being introduced, while resident breeding pairs had higher rates of inbreeding and consequently higher rates of hatch failure. Based on this information, smaller, more isolated

44 groups at KSC should have higher rates of inbreeding; these higher rates should have occurred before the initiation of habitat restoration in the 1990s.

EFFECTIVE POPULATION SIZE

Effective population size (Ne) is defined as “… an idealized population that would lose genetic diversity (become inbred or drift) at the same rate as the actual population” (Frankham et al. 2010). In many populations, Ne can be much lower than the census size, and thus the population will undergo genetic drift and loss of diversity at a much higher rate than predicted from census size alone (Frankham et al. 2010, Pruett et al. 2011). Ne can be influenced by several ecological factors, which include fluctuations in size of the population, variance in family size, and unequal sex ratios (Frankham et al. 2010). Estimates have shown that Ne varies between 0.1-0.5 of actual population sizes, based on the method of analysis (Mace and Lande 1991, Frankham et al. 2010). Several recommendations have been made on minimum Ne to reduce the risks associated with loss of genetic diversity and extinction. Recommendations vary between 50 to 500 for immediate prevention of inbreeding depression, and 500 to 5,000 for maintenance of long-term evolutionary potential (Mace and Lande 1991, Franklin and Frankham 1998, Lynch and Lande

1998, Frankham et al. 2010). It is important to understand that these numbers are simply guidelines to help in the development of species management plans.

In Florida Scrub-Jay, populations can fluctuate due to changes in habitat quality (Woolfenden and Fitzpatrick 1984, Frankham et al. 2010), epidemics

45

(Woolfenden and Fitzpatrick 1984), and predation (Breininger et al. 2009). Family size can also vary as quality of scrub habitat can be heterogenous within territories, which can have a large influence on recruitment (Woolfenden and Fitzpatrick 1984,

Breininger and Carter 2003, Breininger et al. 2014). Experienced breeders also typically have higher rates of breeding success (Woolfenden and Fitzpatrick 1984).

Sex ratios in jays have been found to be roughly 1:1 (Khodadad 2008). Florida

Scrub-jay typically delay breeding until three years of age (Woolfenden and

Fitzpatrick 1984).

Woolfenden and Fitzpatrick (1984) estimated an Ne of ~290 by modeling dispersal distance, adult longevity, and adult movement patterns. Although this is a good initial estimate, Woolfenden and Fitzpatrick (1984) only used demographic information from a single population in central Florida (Archbold Biological

Station), and at the time, no genetic information was available. This estimate is only valid for the specific population under study and does not take into account the demographic history of the population, as population fluctuations can severely depress Ne estimates (Nabholz et al. 2013). Delaney and Wayne (2005) studied Ne of the Western Scrub-Jay (Aphelocoma californica) and the Island Scrub-Jay

(Aphelocoma insularis) and found an Ne of 12,082 for A. californica, while A. insularis had an Ne of 1,603 (Delaney and Wayne 2005). As Florida Scrub-Jay are typically isolated on their own scrub ridge “islands” due to habitat specialization

46 and habitat loss, Florida Scrub-Jay effective size should be more similar to the

Island Scrub-Jay than the Western Scrub-Jay.

FLORIDA SCRUB-JAY AND SCRUB HABITAT IN DECLINE

Florida Scrub-Jay (Aphelocoma coerulescens) populations have drastically declined over the past century (Woolfenden and Fitzpatrick 1996), with a ~90% loss compared to pre-European settlement number (Boughton and Bowman 2011).

In the past two decades, surveys have shown that the current population size has declined by at least 35-40% (Boughton and Bowman 2011), and continue to decline, primarily due to habitat degradation and habitat loss (Birdlife 2012). As habitat specialists with short dispersal distances and a complex breeding system, the Florida Scrub-Jay requires specific habitat management needs for continued persistence. Scrub habitat, for which scrub-jay are named, has suffered extensive losses due to anthropogenic land-use change and degradation (Schmalzer et al.

2002). Scrub is comprised of oak (Quercus spp.) in a patchy mosaic with saw palmetto (Serenoa repens), herbaceous shrubs, and grasses (Woolfenden and

Fitzpatrick 1984, Breininger et al. 1996b). Much of the degradation associated with scrub has been linked to fire suppression on oak patches, which consequently become dense and woody (Fitzpatrick et al. 1991). Fire suppression in Florida has persisted for ~50 years in some areas (Breininger et al. 1996a). In addition, anthropogenic land use, primarily from agriculture and real estate development, has had a pronounced influence on how fire spreads through scrub after fire

47 suppression (Schmalzer et al. 1994). Scrub in a natural state, burns at a frequency of 5-20 years (Main and Menges 1977). Saw palmetto is more pyrogenic than oak

(Schmalzer and Hinkle 1992). Therefore, fragmentation between ignition areas can prevent oak from burning, creating dense woody stands that is sub-optimal for scrub-jays (Schmalzer and Hinkle 1992, Schmalzer et al. 1994). Four transition states within scrub have been identified (Fig. I.2, p.8) based on the time since a fire has occurred: short (oak with a height < 1.3 m), optimal (oak height of 1.3-1.7 m), tall-mix (a mix of tall and short ), and tall (oak height > 1.7 m). Breininger et al. (2009) determined that scrub oak, with a height of 1.3-1.7 m tall, intermixed with sandy openings and minimal pine canopy would be optimal for scrub-jays.

Predation is also a great concern for scrub-jays in sub-optimal transition habitat states (Woolfenden and Fitzpatrick 1984, Breininger 1996a). By day, hawks prey on scrub-jay using pine canopies as hunting perches (Woolfenden and Fitzpatrick

1984, Breininger et al. 1996), especially during raptor migration (Geoff Carter, person communication). By night, females and eggs or chicks are susceptible to nest depredation by snakes (Carter et al. 2007). Poor habitat quality, limited dispersal, and delayed reproduction can exacerbate the effects of mortality leading to slow recovery rates following a population reduction (Breininger et al. 1996b).

Ultimately, suitable habitat must be maintained and restored to sustain large populations and buffer losses due to predation and stochastic events.

48

A statewide survey of Florida Scrub-Jay was performed during 1992 and

1993 (Fitzpatrick et al. 1994). Stith et al. (1996) sought to designate metapopulations and ultimately evaluate extinction risks across the state. The researchers calculated the dispersal distances for scrub-jay and found that ~3.5 km was the average longest distance a scrub-jay would travel between patches, whereas

12 km was the maximum total distance a scrub-jay would disperse (Stith et al.

1996). Subpopulations were grouped into connected metapopulations based on this maximum distance (beyond 12 km; Stith et al. 1996). This information was then applied to a population viability analysis (PVA) model, which also included other demographic parameters and stochastic events such as hurricanes and epidemics

(e.g. a mosquito-borne arborvirus; Breininger et al. 2009). From the results, Stith et al. (1996) determined that if a population had less than 10 breeding pairs, it would have ~50% probability of extinction within 100 years, while a population with 100 breeding pairs would have only a 2-3% chance of extinction. They identified 42 metapopulations (based on the 12 km metapopulation buffer) across the state (Stith et al. 1996). This information was useful for incorporating land management practices into discrete habitat patches and for initiating management goals for each metapopulation (Stith et al. 1996). Three of these sites (which included Ocala National Forest, Lake Wales Ridge, and Merritt Island, Kennedy

Space Center property) were designated as “core populations”, each contained >

100 breeding pairs (Fitzpatrick et al. 1994). Over the past ~30 years, the Kennedy

Space Center has instituted a management regime to sustain focal areas of scrub

49 habitat and maintain a stable population of Florida Scrub-Jays. The Kennedy Space

Center contains four focal areas of scrub managed for Florida Scrub-Jay

(Breininger 1996b). The first focal area, Happy Creek, was established in 1988 and a year later, Tel-4 was added. In the 1990s, Shiloh and Schwartz Road were also added (Breininger et al. 1996b); these two locations are smaller and more fragmented than Happy Creek or Tel-4 (Geoff Carter, personal communication;

Fig. II.1, p. 21). Although the vast majority of scrub-jays are found in these four areas, some birds are found outside the study sites in small scrub patches

(Breininger and Carter 2003).

The space center was under fire suppression from 1963 to 1975 (Schmalzer et al. 1994). After 1975, minimal prescribed burning measures were taken; however, it was not until 1981 that a three-year prescribed fire cycle was implemented (Schmalzer et al. 1994). Tel-4 was burned at least one time during the

1960s-1970s due to , causing more openings at this site than the three other sites (Schmalzer et al. 1994). During the 1990s, habitat restoration began in small increments to increase the population size of Florida Scrub-Jay around

Kennedy Space Center (Schmalzer et al.1994), although the population is still declining (David Breininger, personal communication) .

50

RESEARCH QUESTIONS

Based on Florida Scrub-Jays habitat fragmentation and declining populations, the following questions were asked. 1) What levels of inbreeding have occurred within Florida Scrub-Jay populations over ~30 year time-frame scale based on pedigrees of birds at Kennedy Space Center? 2) Do inbreeding levels differ between pedigree data and population-based estimates from genetic data? 3)

What is the current effective population size (Ne) at Kennedy Space Center and does this size differ from historical estimates over the last century?

51

MATERIALS AND METHODS

GENETIC DATA

Three to four contour feathers were collected from Florida Scrub-Jay individuals (N = 64) from four locations at the Kennedy Space Center, FL (Fig.

II.1, p. 21). DNA from the feather sample tissues were purified and isolated following QIAGEN’s DNeasy tissue extraction kit (Qiagen, Inc., Velancia,

California, USA), and 30 µL 1M dithiothreitol (DTT) was added to break down keratin and release DNA during the initial incubation step (Volo et al. 2008).

Polymerase chain reaction (PCR) was then used to amplify 9 fluorescent-dye labeled microsatellite loci (Stenzler and Fitzpatrick 2002). Amplification products were genotyped using an ABI 3130 sequencer (Applied Biosystems, Foster City,

CA). Genemapper software (ABI, Foster City, CA) was then used to manually score allele fragment sizes.

INBREEDING

To estimate the inbreeding coefficient (F) of individuals, a pedigree (N =

253 total pairings) was created from two long-term study sites, Happy Creek (26 years; David Breininger and Geoff Carter, unpub. data) and Tel-4 (25 years; David

Breininger and Donna Oddy, unpub. data). Despite persistent habitat restoration, the population at KSC is declining (David Breininger, personal communication).

The census size of these two study sites increased over the 26 years of sampling,

52 due to expansions in the area of the study sites (Breininger et al. 1996a, David

Breininger, personal communication). Data were gathered by color-banding individual jays and monitoring family groups throughout the year to identify immigration/emigration at each site, as well as mortality (Breininger and Carter

2003). Pedigree information from Happy Creek (1988-2014) and Tel-4 (1989-

2014) was analyzed using ENDOG software to provide inbreeding coefficients

(Gutierrez and Goyache 2004). Several assumptions have been made based on the nature of collecting wild pedigree data. The sex of birds was determined by behavior, a method deemed reliable by Woolfenden and Fitzpatrick (1984).

Juveniles and “First Years” (FYs) were chosen for the analysis based on the assumption that juveniles and FYs are the offspring of breeders on each territory site, because adoption is rare (Woolfenden and Fitzpatrick 1984, Geoff Carter, personal communication). Parentage data was used to identify parent/offspring relationships and to construct a long-term pedigree. Individuals were removed that were either never banded or shared a band combination with another individual (for instance, from a nearby, but still separated location). Individuals with unknown sex were also removed from the pedigree. As age was required by ENDOG software, any founders or migrants with unknown age were given a birthdate of the year before they began breeding as it is possible (though not typical) for Florida Scrub-

Jay to begin breeding in the first year (Woolfenden and Fitzpatrick 1984).

53

Population-level inbreeding was estimated using the inbreeding coefficient

(FIS) based on microsatellite genotypes. This statistic is derived from the equation

FIS = (1 – HI) / HS, where HI is the averaged observed heterozygosity across all

Scrub Reserve Units, and HS is the averaged expected heterozygosity across all

Scrub Reserve units (Frankham et al. 2010). FIS was calculated using ARLEQUIN

3.5 (Excoffier and Lischer 2010). This method utilizes bootstrapping, which resamples the data to test whether it is significantly different from zero (Manly

2007).

EFFECTIVE POPULATION SIZE

Estimates of Ne were calculated using two genetic methods and a demographic method based on life-history characteristics. In large, randomly mating populations, loci are typically unlinked, meaning there is independent sorting (Frankham et al. 2010). However, drift in small populations can accumulate linked loci causing linkage disequilibrium. Linkage disequilibrium can be calculated between each set of loci (Waples and Do 2008). The program LDNE utilizes these linkage disequilibrium distances to estimate effective population size

(Waples and Do 2008). Effective size is also estimated based on the number of individuals (S) and the mating system (i.e. monogamy). There are three benefits of using this method. First, it includes a simple calculation that does not require random mating to be an assumption (Waples 2006). Second, it is useful for highly polymorphic loci (such as microsatellites; Waples and Do 2010). Last, it has

54 increased precision by correcting for bias associated with low frequency alleles

(Waples and Do 2008). Alleles were excluded if they fell below a frequency of 0.01 and the jackknife method was used to estimate 95% confidence intervals (Waples and Do 2008). Initially, each subpopulation was tested separately. However, due to lack of genetic structure (Fig. II.2, p. 32; Fig.II.3, p. 33) all individuals were grouped into a single population.

Current and historical Ne were estimated using the program MSVAR

(version 1.3), which uses a Bayesian coalescent-based approach (Beaumont 1999,

Storz and Beaumont 2002). Using Markov Chain Monte Carlo (MCMC), MSVAR infers the most likely current and historical Ne estimates using current microsatellite data (Beaumont 1999). Three independent runs were performed with different priors as suggested by Beaumont (1999). Markov chains were computed for 2 X 109 steps with values taken every 105 steps for a total of 20,000 draws from each run. If each run had similar posterior distribution, the last 50% were combined for the final analysis. Kernel density estimates were then performed on the MCMC outputs for the current effective population size, historical effective population size, and time (in years) when the population began to increase or decrease. The program

R (R Core Team 2015) was used for kernel density estimation. The mode of the posterior distribution for each MSVAR output (current Ne, historical Ne, and time) was then calculated in R using a function found in Appendix A

55

(http://www.sumsar.net), while the package coda (Plummer et al. 2006), was used to calculate 95% confidence intervals.

A demographic estimation of Ne was performed using the minimal method of Nunney and Elam (1994):

NeD/Nc = 4r(1-r)T / [rAf (1 + IAf ) + (1 – r) Am (1 + IAm ) + (1 – r) Ibm + r Ibf]

A list of parameters used in the calculation are found in Table III.1. Two parameters, x, number of offspring per female, and p, proportion of females fledgling one or more offspring, needed to calculate Ibf, were estimated using parentage data from scrub-jays at Happy Creek and Tel-4. Parameter x, was calculated by dividing the number of total offspring by the number of total females producing offspring. While p was calculated by taking the ratio of females producing offspring over the number of females in a breeding pair. In Florida

Scrub-Jays, sex ratio between males and females are equal, on average

(Woolfenden and Fitzpatrick 1984, Khodadad 2008). In monogamous species, male and female variance in breeding success (Ibm and Ibf, respectively) are equal

(Nunney and Elam 1994). It should be noted that the minimal method does not account for variation in family size, which is one of the key factors influencing Ne

(Frankham et al. 2010). As the demographic method gives an Ne/N ratio, multiplying this number by the census size gives an estimate of Ne for comparison with genetic methods. The current estimate at KSC is ~275 breeding pairs (Dave

Breininger, personal communication). With a mean family size of ~3 individuals

56

(Breininger 1989, Breininger and Oddy 2004), there would currently be ~825 individuals at the four Kennedy Space Center study sites. It should be stated however, that most researchers and land mangers working with Florida Scrub-Jay gauge population sizes based on breeding pairs and territories and not on the total number of individuals in a population. This can be problematic, as most population genetics analyses are focused on the total number of individuals. As a way to bridge these two areas of research, an effective size ratio is shown for the total population and the number of breeding pairs. After the Ne/N ratio was determined , five parameters: number of offspring per female (x), lifespan in years (Ai), maturation time (M), standardized variance in lifespan (Iai), and proportion of females fledging one or more offspring (p) were tested to identify how each influenced effective population size. However, only M, Iai, and p are reported, as x and Ai had minimal impact on Ne/N ratios.

57

Table III.1. Parameters used in the minimal method (Nunney and Elam 1994) for estimating demographic effective population size of Florida Scrub-Jays from KSC.

Parameter Represents Value Source r Sex ratio 0.5 Khodadad (2008) T Generation time (in years) 7 T = M – 1 + A

Ai Adult life span (in years) 5 Wilcoxen et al. 2010 Standardized variance in Iai 0.83 Vi = Iai lifespan Variance in male Ibm 1.39 Equal to Ibf if monogamous reproduction Variance in female Ibf 1.39 Iβf=(1-βf)/βf reproduction Woolfenden and Fitzpatrick M Maturation Time 3 1984

αf Female breeding success 0.42 αf = xp/βf Number of offspring per Parentage Data from Happy x 1.87 female Creek and Tel-4 Proportion fledging one or Parentage Data from Happy p 0.32 more offspring Creek and Tel-4

Age independent Vi 0.83 Vi = Ai / (1+Ai) survivorship

58

RESULTS

INBREEDING

One inbreeding event was detected in the pedigree at Tel-4 (1989-2014

(Fig. III.1). A pair, LO-SO (♂) and S-LW (♀), produced a male offspring LO-SL

(♂). By 2006 the male breeder, LO-SO (♂), disappeared and a new breeder filled the vacancy, LO-SG (♂). In 2007 and 2008, the female breeder, S-LW (♀), mated with her son, LO-SL (♂), and produced five offspring. After 2009, neither the mother, S-LW (♀), the son, LO-SL (♂), or their inbred offspring were seen and so either they moved outside the Tel-4 study site or there was a mortality and the family disbanded. There were no inbreeding events detected at Happy Creek.

Three of the study sites revealed FIS levels significantly higher than zero

(Table III.2). This suggests that inbreeding has occurred at the subpopulation-level for Shiloh, Schwarz Road, and Tel-4 but not for Happy Creek. Average across all subpopulations was found to be 0.08, slightly higher than an inbreeding event between first cousins (F = 0.0625).

EFFECTIVE POPULATION SIZE

Effective population size was analyzed at the population level (grouping subpopulations). The linkage disequilibrium method (LDNE) estimated Ne of 482

(95% CI: 240- 4,667; Table III.3). Based on MSVAR, the current Ne is 856 individuals (95% CI: 178- 4,141) and a historical Ne of 29,550 individuals (95%

59

CI: 7,187- 119,146 years ago; Fig. III.2). Time since population decline was estimated at 7,516 ybp (95% CI: 1,087-38,022 ybp; Fig. III.3). Based on MSVAR estimates, the Florida Scrub-Jay at KSC have experienced a 97% decline in Ne over the past ~7,500 years, which is about the time Merritt Island is believed to have formed due to receding glaciers and dramatic changes in climate (Mayhew 2000).

The demographic estimation of Ne//N was ~0.67. To compare this ratio with the genetic Ne estimates, this number was multiplied by a population estimate of 825. This yielded an Ne of ~547, which is between the MSVAR (Ne = 865) and

LDNE (Ne = 482) estimates (Table III.3). While the demographic Ne based on breeding pair size alone, is understandably lower at 365. Five parameters : (1) lifespan (Ai), (2) maturation time (M), (3) standard variance in lifespan (Iai), (4) number of offspring (x), (5) proportion fledging one or more offspring (p), were each manipulated to examine how these values affect Ne/N. Only three parameters, maturation time (M), standardized variance in lifespan (Iai), and proportion of females fledging one or more offspring (P) are shown here as they had the greatest impact on the effective size ratio. Comparisons between maturation time and Ne/N reveal that longer amounts of time spent as a helper, will increase effective size ratio (Fig. III.4). A comparison between standardized variance in lifespan (Iai) and

Ne/N ratio (Fig. III. 5) shows a negative correlation. Conversely, there is positive relationship between proportion of females fledging one or more offspring and the effective population size/census size ratio (Fig. III.6).

60

LS-GP LS-BO

LO-SO S-LW

LO-SL SL-

Figure III. 1. Pedigree from census data showing inbreeding in Florida Scrub-Jay at the Tel-4 location (Fig. II.1, p. 21). A parent sibling pairing of mother (S-LW) with son (LO-SL) producing offspring (SL-). Squares and circles around individuals represent males and females respectively.

Table III.2. Inbreeding coefficient (FIS) of four Florida Scrub-Jay subpopulations at Kennedy Space Center, FL. FIS is a measurement of relatedness by descent within a group. P values are based on 1023 permutations of the data.

Subpopulation FIS P value Shiloh 0.12 0.040 Happy Creek -0.02 0.730 Schwartz Road 0.14 0.002 Tel-4 0.08 0.020 Mean of subpopulations 0.08 ---

61

Figure III.2. MSVAR estimates of current and historical effective population size (Ne) for Florida Scrub-Jay at Kennedy Space Center based on kernel density estimates.

Figure III.3. MSVAR output showing kernel density estimate of time (in log years) since effective population size decline in the Florida Scrub-Jay at the Kennedy Space Center, approximately 7,500 years (dashed line).

62

1.2

1

0.8

0.6

Ne/N Ratio Ne/N 0.4

0.2

0 1 2 3 4 5 6 7 Maturation time (in years)

Figure III.4. Comparison of maturation time (M) and effective population size to census size ratio (Ne/N) based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate.

1.2

1

0.8

0.6

Ne/N Ratio Ne/N 0.4

0.2

0 0 0.15 0.3 0.45 0.6 0.75 0.9

lai

Figure III.5. Comparison of standardized variance in lifespan (Iai) and effective population size to census size ratio (Ne/N) based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate.

63

0.9 0.8 0.7 0.6 0.5 0.4

Ne/N Ratio Ne/N 0.3 0.2 0.1 0 0.1 0.25 0.4 0.55 0.7 0.85 1 Proportion fledging one or more offspring (p)

Figure III.6. Proportion of females fledging one or more offspring (p) and effective population size to census size ratio (Ne/N)based on the minimal method of Nunney and Elam (1994). Dashed line indicates current Florida Scrub-Jay estimate.

64

Table III.3. Effective population size estimates using three different methods and historical start of population decline using MSVAR. 95% CI – Bayesian credibility intervals.

Method Ne estimate 95% CI

Demographic Ne/N 0.67 ------

Demographic – Breeding Pairs 365 ------

Demographic – Population size 547 ------

LDNE 482 240 4,667

MSVAR- Current 865 178 4,141

MSVAR- Ancestral 29,550 7,187 119,146

Time (years) 95% CI

MSVAR- time of decline in Ne 7,516 1,087 38,022

65

DISCUSSION

In conservation biology, small populations are of a great concern due to a higher risk of extinction from increased genetic drift, lower genetic diversity, and inbreeding (Frankham et al. 2010). Tools, such as pedigrees and genetic analyses, can be useful in identifying historical and current impacts to species of conservation concern. Florida Scrub-Jay have suffered a ~90% loss in total population size over the past century (Breininger 1989, Boughton and Bowman

2011) and are considered vulnerable by the IUCN (Birdlife 2012). The population of Florida Scrub-Jay at KSC is considered to be a core population that must be conserved to ensure long-term species persistence (Stith et al. 1996). However, information about levels of inbreeding and current and historical effective population size at small spatial scales are lacking.

INBREEDING

Studies on small populations in the wild have shown that inbreeding occurs in both plants and animals (Keller and Waller 2002). Inbreeding can accumulate deleterious alleles that increase the risk of extinction through defects and lower chances of survival (Frankham et al. 2010). Birds in disturbed habitats had increased inbreeding and strains on development (Lens et al. 2001). A study on an

Aphelocoma jay, the Aphelocoma ultramarina, was observed to have lower nestling survival from inbred pairs (Brown and Brown1998). As the Florida

Scrub-Jay has a short dispersal distance and is limited to scrub habitat, there would

66 be more opportunity to mate with a close relative (Woolfenden and Fitzpatrick

1984, Keller and Waller 2002).

At Kennedy Space Center, the long-term pedigree showed that inbreeding is rare among close relatives, as between two sites and 253 pairing events over almost

30 years, only one case was found (Fig. III.1). This is similar to one of the few inbreeding events observed by Woolfenden and Fitzpatrick (1984) at Archbold

Biological Station in which, the father became injured and eventually died, which led the mother and son to pair. In central Florida, a study by Townsend et al. (2011) found that a pedigree of Florida Scrub-Jay in fragmented areas had an inbreeding coefficient of 0.02 (Townsend et al. 2011). Though, inbreeding is rare in scrub-jay, the study shows that habitat fragmentation can lead to increased levels of inbreeding. Based on genetics, FIS estimates indicate inbreeding is occurring in three of the four study sites at KSC (Table III.2.). Shiloh and Schwartz Road had the highest FIS values of 0.12 and 0.14 respectively, which is at, or slightly higher than, breeding between half-siblings (F = 0.125; Frankham et al. 2010). Thus, individuals on average are as closely related as half-siblings in these two subpopulations. Without long-term pedigrees at these two sites, however, it is difficult to determine how long inbreeding has occurred. Considering that these sites are smaller and more fragmented than Happy Creek and Tel-4, it is reasonable to expect that these sites would also have higher FIS values. Tel-4 had an FIS value of 0.08. Even though this subpopulation has >100 individuals, it has been through

67 several fluctuations in subpopulation size, due to fires (David Breininger, personal communication), which can increase the inbreeding coefficient over time

(Frankham 1995a). Happy Creek, with an FIS value of -0.02, was not significantly different from zero, therefore showed no evidence of inbreeding. However, it should be noted, that the study sites have expanded to include more families as there have been declines to overall population size.

In comparison with the state-wide genetic survey of Florida Scrub-Jay by

Coulon et al. (2008; FIS = -0.1 to 0.06), 3 of the 4 scrub-jay subpopulations at KSC had higher FIS values (Table III.1). In the state-wide study, birds from Kennedy

Space Center, which were grouped together with jays from Volusia and Flagler county, had an FIS value of 0.005 (Coulon et al. 2008). However, given that Coulon et al. (2008) did not sample the exact same locations, and those individuals were collected over multiple counties (> 100 km apart), this might be an artifact of sampling bias. Therefore, the findings of Coulon et al. (2008) may not be comparable to the study at KSC. Chen (2014) found that chicks of residents were more inbred and had higher rates of mortality than those of immigrants. Therefore, more connectivity is required, especially at Shiloh and Schwartz Road to facilitate immigration, which is necessary to prevent further rises in inbreeding rates at these locations (Ingvarsson 2001).

68

EFFECTIVE POPULATION SIZE

Effective population size estimates (Ne) are useful in identifying how much drift or loss of adaptive potential a population would be expected to lose in a population of that size (Frankham et al. 2010). In small populations, the ratio of

Ne/N can be much less than 1 (Mace and Lande 1991, Frankham et al. 2010), indicating the need to restore adaptive potential to genetically impoverished populations. Several factors have been identified as the most influential in reducing

Ne including fluctuations in population size, variance in family size and unequal sex ratios (Frankham et al. 2010). Florida Scrub-Jays at KSC have experienced such fluctuations. Two mosquito-borne arbovirus epidemics caused statewide die-offs in

1979 and 1997 (Breininger et al. 2009). Though no bottleneck was detected in

Chapter 2, these die-offs would certainly decrease Ne (Frankham et al. 2010).

Unrelated to the arbovirus epidemics, Tel-4 had some drastic declines in census size from 2009-2013, due to fires (David Breininger, personal communication).

Variation in family size follows fluctuations in population size as having the second largest effect on reducing Ne below N (Frankham et al. 2010).Variation in family size is also one of the major influences of Ne that is not accounted for in the minimal method (Nunney and Elam 1994). This can explain some of the differences in the demographic and genetic based approaches. As both habitat quality and breeder experience can influence family size in Florida Scrub-Jay

(Woolfenden and Fitzpatrick 1984), family size variation has a large impact on Ne.

69

Un-even sex ratios can also influence Ne greatly (Frankham 1995b). However,

Khodadad (2008) demonstrated through genetics that the sex ratio is ~1:1. Thus, fluctuations in population size and variation in family size are the most likely drivers of differences impacting effective population size (Frankham et al. 2010); the effects of both of these factors can be examined using genetic methods.

The Florida Scrub-Jay has declined by ~90% in the past century when compared with pre-European settlement numbers (Boughton and Bowman 2011).

More recent declines have been influenced by fire suppression, agriculture and human population expansion (Fitzpatrick et al. 1991, Schmalzer et al. 1994,

Schmalzer et al. 2002). The historical MSVAR estimate (Ne = 29,550) did not pick up any signals of recent decline due to European settlement (Table III.3). There is minimal overlap between the current and ancestral effective sizes (Fig. III.2), which indicates that there was a substantial decline in the population over time (Sharma et al. 2012). The historical decline is consistent with a reduction in glacier size and an increase in sea level that occurred around ~7,500 YBP (Mayhew 2000). As the climate would have begun to shift towards wetter conditions, a reduction in scrub is also believed to have occurred (Myers and Ewel 1990), which could account for such a drastic decline from the current Ne estimate of 865. This decline most likely reduced or prevented any gene flow between mainland Florida and Merritt Island

(where Kennedy Space Center resides) as it is rare for Florida Scrub-Jay to fly long distances, especially over large bodies of water (Stith et al. 1996).

70

Using the minimal method by Nunney and Elam (1994), another estimate of

Ne was performed that utilizes demographic parameters. This method is useful in identifying what parameters have the greatest influence on Ne currently, while the genetic methods take into account historical fluctuations in size. The initial Ne estimate of 547, is between the LDNE (Ne = 482) and MSVAR (current Ne = 865) values (Table III.2) suggesting that both the demographic and genetic methods are likely to provide reliable estimates of current effective size. Three parameters, maturation time, standardized variance in lifespan, and proportion of females fledging one or more offspring, were examined to identify how changes in each parameter may influence effective population size. Maturation time (or the age of first reproduction) is positively associated with Ne/N (Fig.III.4). Most Florida

Scrub-Jay delay breeding until ~3 years of age (Woolfenden and Fitzpatrick 1984).

However, in urban or degraded habitats, higher rates of breeder mortality give the opportunity for young jays to breed in their first or second year, which could lead to a reduction in Ne/N (Breininger 1999). The delayed breeding of scrub-jays on

KSC has likely led to the high Ne/N ratio found in this study, given that most monogamous birds have an Ne/N ratio of 0.5 or less based on demographic methods (Nunney and Elam 1994, Frankham 1995b). Thus, the longer maturation and generation time has likely buffered this population from the effects of small population size.

71

Variance in lifespan was found to have a large influence on the Ne/N ratio

(Fig.III.5); a higher variance led to a lower ratio. As scrub patches are quite heterogenous (Woolfenden and Fitzpatrick 1984, Breininger and Carter 2003), differences in habitat quality can greatly influence survival in a specialist species like the Florida Scrub-Jay, thus causing a large variance in lifespan. A reduction in pine tree canopy can remove hunting perches of raptors (Fitzpatrick et al. 1991) and thinning overgrown scrub patches can limit nest depredation by snakes (Carter et al. 2007). Stochastic events, such as hurricanes and epidemics can also increase the variance in lifespan (Breininger et al. 1996a). Increasing the number of territories with quality scrub habitat will give scrub-jays the best chance at minimizing variance in lifespan (Breininger et al. 2009).

Tests were also performed on how changes in the proportion of females fledging one or more offspring (p) influenced the Ne/N ratio. Results indicate that if more females were fledging young, effective population size would increase to an extent (Fig. III.6). When p values higher than the current estimate for scrub-jays

(0.32) were examined the Ne/N ratio only increased from 0.67 to 0.75 suggesting that the current proportion of females fledging offspring will maintain the effective size. However, a reduction in the proportion fledging offspring would have a large effect (Fig. III. 6). Ultimately, these tests reveal the importance of maintaining and expanding the population, which can only be achieved through proper habitat restoration.

72

Delaney and Wayne (2005) performed a study on two sister taxa, the

Western Scrub-Jay (Aphelocoma californica) and the Island Scrub-Jay (A. insularis). It was discovered that Western Scrub-Jay (Ne = 12,082) had almost a tenfold larger effective population size than the Island Scrub-Jay (Ne = 1,603). As expected, effective population size at Kennedy Space Center (482-865) is more similar to the isolated Island Scrub-Jay, considering the large distribution of

Western Scrub-Jay throughout the western USA (Delaney and Wayne 2005).

Woolfenden and Fitzpatrick (1984) estimated Ne at Archbold Biological Station to be ~290, compared with KSC’s lowest Ne estimate of 482. Archbold is an inland population that is smaller than that at KSC. However, this estimate at Archbold was performed in the early 1980s (Woolfenden and Fitzpatrick 1984). Over the past ~30 years, substantial efforts have been made to restore habitat and connectivity at

Archbold (Chen 2014), therefore if Ne were assessed again using demographic methods, it may be closer to that found in this study.

As a guideline, an effective size of 500 adults is needed to prevent the immediate effects of inbreeding and increase adaptive potential (Mace and Lande

1991, Franklin and Frankham 1998). This indicates that the population of scrub-jay at KSC are at or above the guideline based on current Ne estimates (482-865). As researchers and land mangers working with Florida Scrub-Jay utilize the number of breeding pairs as a measure of population success, this number was found to be

365. Inbreeding was found to be between 0.12-0.14 at the two smallest and

73 fragmented subpopulations (Shiloh and Schwartz Road). Therefore, connectivity might need to be increased at these two locations to prevent further inbreeding and inbreeding depression. All four study sites are geographically isolated from each other (Fig.II.1, p. 21), and corridors are necessary to increase movement between each site.

GUIDELINES FOR LAND MANAGERS

As this study was only an initial step in understanding the population genetics of Florida Scrub-Jay at Kennedy Space Center, there is still much to be done through the use of additional loci to increase the resolution of population structure and computer modelling. However, several recommendations are noteworthy based on current understanding of Florida Scrub-Jay biology and the tests performed in this study. First, more helpers are needed to improve breeding success and increase the total population size (Breininger et al. 2014). By manipulating the demographic method of effective population size, it was found that as more females are producing offspring, Ne/N increases (Fig.III.6). This would indeed benefit recruitment in the population as well as increase the genetic potential over time. Second, collect genetic samples from areas outside of the known study sites to identify where individuals are dispersing from and to better understand how they are dispersing, especially through unfavorable habitat. Third, maintain habitat through mechanical treatment and prescribed burning to mimic natural fires, which is found in a patchy mosaic of various transition states

74

(Woolfenden and Fitzpatrick 1984, Breininger et al. 2006, Kent and Kindell 2010).

This will not only increase recruitment, but should reduce variance in lifespan

(Fig.III.5). As raptors can use pine trees as hunting perches and snakes use dense scrub patches to depredate nests (Carter et al. 2007), habitat restoration will be beneficial for the Florida Scrub-Jay, overall. Ultimately, the genetic diversity of scrub-jay at KSC is quite high, in spite of continual population decline. This is encouraging and steps should be taken to prevent further losses to both total population size and losses to genetic diversity.

75

REFERENCES

Bååth, R. URL http://www.sumsar.net/blog/2014/10/probable-points-and-credible- intervals-part-one/. Barson NJ, Cable J, Van Oosterhout C. 2009. Population genetic analysis of microsatellite variation of guppies (Poecilia reticulata) in Trinidad and Tobago: Evidence for a dynamic source-sink metapopulation structure, founder events and population bottlenecks. J. Evolutionary Biology. 22:485–497. Bateson ZW, Dunn PO, Hull SD, Henschen AE, Johnson JA, Wittingham LA. 2014. Genetic restoration of a threatened population of greater prairie- chickens. Biological Conservation 174: 12-19. Beaumont MA. 1999. Detecting population expansion and decline using microsatellites. Genetics 153:2013–2029. Benito-Garzon M, Leadley PW, Fernandez-Manjarres JF. 2013. Assessing global biome exposure to climate change through the Holocene-Anthropocene transition. Global Ecology and Biogeography 23: 235-244. BirdLife International. 2012. Aphelocoma coerulescens. The IUCN Red List of Threatened Species. Version 2014.3. . Downloaded on 1 December 2014. Blouin MS. 2003. DNA-based methods for pedigree reconstruction and kinship analysis in natural populations. TRENDS in Ecology and Evolution. 18 10:503-511. Bonte D, Van Dyck H, Bullock JM, Coulon A, Delgado M, Gibbs M, Lehouck V, Matthysen E, Mustin K, Saastamoinen M, Schtickzelle N, Stevens VM, Vandewoestijne S, Baguette M, Barton K, Benton TG, Chaput-Bardy A, Clobert J, Dytham C, Hovestadt T, Meier CM, Palmer SCF, Turlure C, Travis JMJ. 2012. Costs of dispersal. Biological Reviews 87: 290-312. Boughton RK, Bowman R. 2011. State Wide assessment of Florida Scrub-Jays on managed areas: A comparison of current populations to the results of the 1992-1993 survey. Technical report, submitted to the USFWS. Breininger DR. 1989. A new population estimate for the Florida scrub jay on Merritt Island National Wildlife Refuge. Florida Field. Naturalist. 17:25– 31.

76

Breininger DR. 1992a. Habitat Model for the Florida Scrub Jay on John F. Kennedy Space Center. NASA Technical Memorandum 103496. Kennedy Space Center (FL): The Bionetics Corporation. P. 92. Breininger DR. 1992b. Birds of Swale Marches on John F. Kennedy Space Center. Florida Field Naturalist 20: 36-41. Breininger DR. 1999. Florida Scrub-Jay demography and dispersal in a fragmented landscape. Auk 116:520–527. Breininger DR, Carter G. 2003. Territory quality transitions and source-sink dynamics in a Florida Scrub-Jay population. Ecological Applications 13:516–529. Breininger DR, Larson V, Duncan B, Smith R, Oddy D, Goodchild M. 1995. Landscape patterns of Florida scrub jay habitat use and demographic success. Conservation Biology 9:1442–1453. Breininger D, Larson V, Oddy D, Barkaszi M. 1996a. Florida scrub-jay demography in different landscapes. Auk 113:617–625. Breininger D, Larson V, Schaub R, Duncan B et al. 1996b. A conservation strategy for the Florida scrub-jay on John F. Kennedy Space Center/Merritt Island National Wildlife Refuge: An initial scientific basis for Recovery. NASA Technical Memorandum 111676. Kennedy Space Center (FL): Dynaman Corporation. P. 134. Breininger DR, Morris JG. 1982. Habitat Preference of the Florida Scrub Jay (Aphelocoma-coerulescens-coerulescens) at Merrit-Island-National- Wildlife-Refuge, Florida. American Zoologist. 22:906. Breininger DR, Nichols JD, Carter GM, Oddy DM. 2009. Habitat-specific breeder survival of Florida Scrub-Jays: inferences from multistate models. Ecology 90:3180–9. Breininger DR, Nichols JD, Duncan BW, Stolen ED, Carter GM, Hunt DK, Drese JH. 2010b. Multistate modeling of habitat dynamics: factors affecting Florida scrub transition probabilities. Ecology 91:3354–64. Breininger DR, Oddy D. 2004a. Do habitat potential, population density, and fires influence Scrub-Jay source-sink dynamics? Ecological Applications 14:1079–1089. Breininger DR, Provancha M, Smith R. 1991. Mapping Florida scrub jay habitat for purposes of land-use management. Photogrammetric Engineering Remote Sensing 57:1467–1474

77

Breininger DR, Stolen ED, Carter GM, Oddy DM, Legare SA. 2014. Quantifying how territory quality and sociobiology affect recruitment to inform fire management. Conservation 17:72-79. Breininger DR, Stolen ED, Nichols JD, Carter GM, Oddy DM,Hunt DK. 2010a. A Model-Selection Approach to Predicting Whether Florida Scrub-Jays Delay Breeding. The Condor 112:378-389. Breininger DR, Schmalzer PA. 1990. Effects of Fire and Disturbance on Plants and Birds in a Florida Oak/ Palmetto Scrub. American Midland Naturalist 123:64–74. Breininger DR, Toland B, Oddy D et al. 2003. Biological criteria for the recovery of Florida scrub-jay population on public lands in Brevard County and Indian River County. Final Report to Endanger Species Office, US Fish and Wildlife Service, Jacksonville, Florida. Dynamac Corporation, DYN-2, Kennedy Space Center, Florida. Breininger DR, Toland B, Oddy DM, Legare ML. 2006. Landcover characterizations and Florida scrub-jay (Aphelocoma coerulescens) population dynamics. Biological Conservation 128:169–181. Brooks TM, Mittermeier RA, Mittermeier CG, da Fonseca GAB, Rylands AB, Konstant WR, Flick P, Pilgrim J, Oldfield S, Magin G, Hilton-Taylors C. 2002. Habitat Loss and Extinction in Hotspots of Biodiversity. Conservation Biology 16: 909-923. Brown JL, Brown ER. 1998. Are inbred offspring less fit? Survival in a natural population of Mexican jays. Behavioral Ecology 9:60-63. Burt D, Peterson A. 1993. Biology of cooperative-breeding scrub jays (Aphelocoma coerulescens) of Oaxaca, Mexico. Auk 110:207–214. Bush KL, Vinsky MD, Aldridge CL, Paszkowski CA. 2005. A comparison of sample types varying in invasiveness for use in DNA sex determination in an endangered population of greater Sage- (Centrocercus uropihasianus). Conservation Genetics 6:867-870. Carmen WF. 2004. Non-cooperative breeding in the California scrub-jay. Studies in Avian Biology 28:1-100. Carter GM, Breininger DR, Stolen ED, Oddy DM. 2011. Determinants of Nest Survival in a Managed Florida Scrub-Jay Population. The Condor 113: 629- 636.

78

Carter GM, Legare ML, Breininger DR, Oddy DM. 2007. Nocturnal nest predation: a potential obstacle to recovery of a Florida Scrub-Jay population. Journal of Field 78:390-394. Carter GM, Stolen ED, Breininger DR. 2006. A rapid approach to modeling species-habitat relationships. Biological Conservation 127: 237-244. Ceballos G, Ehrlich PR, Barnosky AD, Garcia A, Pringle RM, Palmer TM. 2015. Accelerated modern human-induced species losses: Entering the sixth mass extinction. Sciences Advances 1: e1400253 Chen N. 2014. Genomics of Population Decline in the Florida Scrub-Jay [dissertation]. Ithaca (NY): Cornell University. Clavel J, Julliard R, Devictor V. 2011. Worldwide decline of specialist species: toward a global functional homogenization? Frontiers in Ecology and the Environment 9: 222-228. Cornuet JM, Luikart G. 1996. Description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. Genetics 144:2001-2014. Coulon A, Fitzpatrick JW, Bowman R, Lovette IJ. 2010. Effects of Habitat Fragmentation on Effective Dispersal of Florida Scrub-Jays. Conservation Biology 24:1080-1088. Coulon A, Fitzpatrick JW, Bowman R, Lovette IJ. 2012. Mind the gap: genetic distance increases with habitat gap size in Florida scrub jays. Biology Letters 8:582–585. Coulon A, Fitzpatrick JW, Bowman R, Stith BM, Makarewich CA, Stenzler LM, Lovette IJ. 2008. Congruent population structure inferred from dispersal behaviour and intensive genetic surveys of the threatened Florida scrub-jay (Aphelocoma coerulescens). Molecular Ecology 17:1685–701. Crnokrak P, Roff DA. 1999. Inbreeding depression in the wild. Heredity 83:260- 270. Delaney KS, Wayne RK. 2005. Adaptive Units for Conservation: Population Distinction and Historic Extinctions in the Island Scrub-Jay. Conservation Biology 19:523–533. De Volo SB, Reynolds RT, Douglas MR, Antolin MF. 2008. An Improved Extraction Method To Increase Dna Yield From Molted Feathers. Condor 110:762–766.

79

De Vos JM, Joppa LN, Gittleman JL, Stephens PR, Pimm SL. 2014. Estimating the normal background rate of species extinction. Conservation Biology 29:452-462. Di Rienzo A, Peterson AC, Garza JC, Valdes AM, Slatkin M, Freimer NB. 1994. Mutational processes of simple-sequence repeat loci in human populations. Proceedings of the National Academy of Sciences 91: 3166-3170. Dirzo R, Young HS, Galetti M, Ceballos G, Isaac NJB, Collen B. 2014. Defaunation in the Anthropocene. Science 345:401-406. Earl DA, von Holdt BM. 2012. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetics Resources 4: 359-361. Excoffier L, Lischer HEL. 2010. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Resources 10:564–567. Fitzpatrick J, Pranty B, Stith B. 1994. Florida scrub jay statewide map 1992-1993. Archbold Biological Station. Lake Placid, Florida. Fitzpatrick J, Woolfenden G, Kopeny M. 1991. Ecology and development-related habitat requirements of the Florida Scrub Jay (Aphelocoma coerulescens coerulescens). Frankham R. 1995a. Conservation Genetics. Annual Review of Genetics 29:305- 327. Frankham R. 1995b. Effective population size/adult population size ratios in wildlife: a review. Genetical Research 66:95-107. Frankham R. 2005. Genetics and extinction. Bioological Conservation 126:131- 140. Frankham R. 2015. Genetic rescue of small inbred populations: meta-analysis reveals large and consistent benefits of gene flow. Molecular Ecology 24:2610-2618. Frankham R, Ballou JD, Briscoe DA. 2010. Introduction to Conservation Genetics. Cambridge University Press. Franklin IR, Frankham R. 1998. How large must populations be to retain evolutionary potential? Animal Conservation 1:69-73.

80

Goudet J. 1995. FSTAT (Version 1.2): A Computer Program to Calculate. Journal of Heredity 86:485–486. Guillot G, Mortier F, Estoup A. 2005. Geneland: a computer package for landscape genetics. Molecular Ecology Notes 5:712–715. Guillot G, Santos F, Estoup A. 2008. Analysing georeferenced population genetics data with Geneland: a new algorithm to deal with null alleles and a friendly graphical user interface. Bioinformatics 24:1406–7. Gutierrez JP, Goyache F. 2005. A note on ENDOG: a computer program for analyzing pedigree information. Journal of Animal Breeding and Genetics 122:172-176. Harrison, S. 1991. Local extinction in a metapopulation context: an empirical evaluation. Biological Journal of the Linnean Society 41:73-88. Harrison S, Bruna E. 1999. Habitat Fragmentation and Large-Scale Conservation: What Do We Know for Sure? Ecography 22: 225-232. Harrisson KA, Pavlova A, Da Silva AG, Rose R, Bull JK, Lancaster ML, Murray N, Quin B, Menkhorst P, Magrath MJL, Sunnucks P. 2016. Scope for genetic rescue of an endangered subspecies though re-establishing nautral gene flow with another subspecies. Molecular Ecology 6:1242-1258. Hedrick PW. Fredickson R. Genetic rescue guidelines with examples from Mexican wolves and Florida panthers. Conservation Genetics 2:615-626. Hubisz MJ, Falush D, Stephens M, Pritchard JK. 2009. Inferring weak population structure with the assistance of sample group information. Molecular Ecology Resources 9:1322–32. IHA Envrionemtanl Services Branch. 2014. Kennedy Space Center Florida Scrub- Jay Compensation Plan. Prepared for: Environmental Management Brach TA-A4C. Ingvarsson PK. 2001. Restoration of genetic variation lost – the genetic rescue hypothesis. TRENDS in Ecology & Evolution 16:62-63. Jansson M, Ståhl I, Laikre L. 2013. mPed: A computer program for converting pedigree data to a format used by the PMx-software for conservation genetic analysis. Conservation Genetic Resources 5:651–653. Johnson FA, Breininger DR, Duncan BW, Nichols JD, Runge MC, Williams BK. 2011. A Markov Decision Process for Managing Habitat for Florida Scrub- Jays. Journal of Fish Wildlife Management 2:234–246.

81

Johnson WE, Onorato DP, Roelke ME, Land ED, Cunningham M, Belden RC, McBride R, Jansen D, Lotz M, Shindle D, Howard J, Wildt DE, Penfold LM, Hostetler JA, Oli MK, O’Brien SJ. 2010. Genetic Restoration of the Florida Panther. Science 329:1541-1645. Kalinowski ST, WagnerAP, Taper ML. 2006. ML-Related: a computer program for maximum likelihood estimation of relatedness and relationship. Molecular Ecology Notes 6:576-579. Keller LF, Waller DM. 2002. Inbreeding effects in wild populations. TRENDS in Ecology & Evolution 17:230-241. Kent A, Kindell C. 2010. Scrub management guidelines for peninsular Florida: using the scrub-jay as an umbrella species. Florida Fish and Wildlife Conservation Commision, Tallahassee, Florida. Khodadad CLM. 2008. Genetic Variation, Population Substructure and Sex Ratio of the Florida Scrub-jay, Aphelocoma Coerulescens Coerulescens [dissertation]. Melbourne (FL): Florida Institute of Technology. Lande R. 1995. Mutation and Conservation. Conservation Biology 9:782-791. Lens L, van Dongen S, Galbusera P, Schenck T, Matthysen E, Van de Casteel T. 2000. Developmetanl instability and inbreeding in natural bird populations exposed to different levels of habitat disturbance. Journal of Evolutionary Biology 13:889-896. Lynch M, Lande R. 1998. The critical effective size for a genetically secure population. Animal Conservation 1:70-72. Lippe C, Dumont P, Bernatchez L. 2006. High genetic diverisy and no inbreeding in the endangered copper redhorse, Moxostoma hubbsi (Catostomidae, Piesces): the positive sides of a long generation time. Molecular Ecology 15: 1769-1780. Mace GM, Lande R. 1991. Assessing extinction threat: Towards a reevaluation of IUCN threatened species categories. Conservation Biology 5:148-157. Main K, Menges E. 1997. Archbold Biological Station Fire Management Plan. Land Management Publication 97-1. Lake Placid, FL., US. Manicacci D, Olivieri I, Perrot V, Atlan A, Gouyon PH, Propseri JM, Couvet D. 1992. Landscape ecology: Population genetics at the metapopulation level. Landscape Ecology 6:147-159.

82

Manly, BFJ. 2007. Randomization Boostrap and Monte Carlo Methods in Biology, 3rd edition. Chapman and Hall, London. Maurer G, Beck N, Double MC. 2010. A “feather-trap” for collecting DNA samples from birds. Molecular Ecology Resources 10:129–134. Mayhew TA. 2000. Late Holocene geologic evolution of a barrier island complex, east-central Florida [thesis]. Melbourne (FL): Florida Institute of Technology. McDonald D, Potts W, Fitzpatrick J, Woolfenden G. 1999. Contrasting genetic structures in sister species of North American scrub-jays. Proceedings of the Royal Society B 266:1117–1125. McKee J, Chambers E, Guseman J. 2013. Human Population Density and Growth Validated as Extinction Threats to Mammal and Bird Species. Human Ecology 41:773-778. Mumme R, Schoech S, Woolfenden GE, Fitzpatrick JW. 2000. Life and Death in the Fast Lane: Demographic Consequences of Road Mortality in the Florida Scrub‐Jay. Conservation Biology 14:501–512. Myers RL, Ewel JJ. 1990. Ecosystems of Florida. University of Central Florida Press. Nabholz B, Uwimana N, Lartillot N. 2013. Reconstructing the phylogenetic history of long-term effective population size and life-history traits using patterns of amino acid replacement in mitochondrial genomes of mammals and birds. Genome Biology and Evolution 5:1273–1290. Nunney L, Elam DR. 1994. Estimating the effective population size of conserved populations. Conservation Biology 8:175-184. Van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P. 2004. Micro-Checker: Software for Identifying and Correcting Genotyping Errors in Microsatellite Data. Molecular Ecology Notes 4:535–538. Paetkau D, Slade R, Burdens M, Estoup A. 2004. Genetic assignment for the direct, real-time estimation of migration rate: a simulation-based exploration of accuracy and power. Molecular Ecology 13:55-65. Peakall R, Smouse PE. 2006. GenAlEx 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes 6:288- 295.

83

Peterson AT, Burt DB. 1992. Phylogenetic history of social evolution and habitat use in the Aphelocoma jays. Animal Behavior 44:859–866. Piry S, Luikart G, Cornuet JM. 1999. BOTTLENECK: a computer program for detecting recent reductions in effective population size using allele frequency data. Journal of Heredity 90: 502-503 Piry S, Alapetite A, Cornuet J-M, Paetkau D, Baudouin L, Estoup A. 2004. GeneClass2: a software for genetic assignment and first-gernation migrant detection. Journal of Heredity 95:536-539. Plummer M, Best N, Cowles K, Vines K. 2006. CODA: Convergence Diagnosis an d Output Analysis for MCMC, R News, vol 6, 7-11.

Pritchard J, Stephens M, Donnelly P. 2000. Inference of population structure using multilocus genotype data. Genetics 155: 945-959. Pruett CL, Johnson JA., Larsson LC, Wolfe DH, Patten MA. 2011. Low effective population size and survivorship in a grassland grouse. Conservation Genetics 12:1205–1214. R Core Team. 2015. R: A language and environment for statistical computing. R F oundation for Statistical Computing, Vienna, Austria. URL https://www.R- project.org/.

Rannala B, Mountain JL. 1997. Detecting immigration by using multilocus genotypes. Proceedings of the National Academy of Sciences 94:9197- 9201. Ricklefs RE. 1968. Patters of growth in birds. Ibis (Lond. 1859). 110:419–451. Root K. 1998. Evaluating the effects of habitat quality, connectivity, and catastrophes on a threatened species. Ecological Applications 8:854–865. Rosenberg NA. 2004. DISTRUCT: a program for the graphical display of population structure. Molecular Ecology Notes 4:137-138 Schmalzer PA, Breininger DR, Adrian FW, Schaub R, Duncan BW. 1994. Development and implementation of a scrub habitat compensation plan for Kennedy Space Center. Schmalzer PA, Hinkle CR. 1992. Recovery of Oak-Saw Palmetto Scrub after Fire Recovery of Oak-Saw Palmetto Scrub after Fire. Castanea 57:158–173.

84

Schmalzer PA, Shannon R. Turek, Foster TE, Dunlevy C a., Adrian FW. 2002. Reestablishing Florida scrub in a former agricultural site: survival and growth of planted species and changes in community composition. Castanea 67:146–160. Sharma R, Arora N, Goossens B, Nater A, Morf N, Salmona J, Bruford MW, Van Schaik CP, Krutzen M, Chikhi Lounes. 2012. Effective Population Size Dynamics and the Demographic Collapse of Bornean Orang-Utans. PLoS ONE 7(11): e49429. Sokal RR, Rohlf FJ. 1995. Biometry, 3rd edn. WH Freeman Co, New York. 887p.

Stenzler L, Fitzpatrick J. 2002. Isolation of microsatellite loci in the Florida Scrub‐ Jay Aphelocoma coerulescens. Molecular Ecology Notes 2:547–550. Stith B, Fitzpatrick J, Woolfenden G, Pranty B. 1996. Classification and Conservation of Metapopulations: A Case study of the Florida scrub jay. In: McCullough D, editor. Metapopulations and Wildlife Conservation. Washington DC: Island Press. p. 187–215. Storz JF, Beaumont M a. 2002. Testing for genetic evidence of population expansion and contraction: an empirical analysis of microsatellite DNA variation using a hierarchical Bayesian model. Evolution 56:154–166. Tallmon DA, Koyuk A, Luikart G, Beaumont M a. 2008. COMPUTER PROGRAMS: onesamp: a program to estimate effective population size using approximate Bayesian computation. Molecular Ecology Resources 8:299–301. Townsend AK, Bowman R, Fitzpatrick JW, Dent M, Lovette IJ. 2011. Genetic monogamy across variable demographic landscapes in cooperatively breeding Florida scrub-jays. Behavioral Ecology 22:464–470. UNEP. 2007. The Global Environment Outlook 4 (GEO-4). Nairobi, Kenya: UNEP. U.S. Fish and Wildlife Service. 2008. Merritt Island National Wildlife Refuge Comprehensive Conservation Plan. Titusville, Florida. U.S. Fish and Wildlife Service. 2014. History of the Shiloh Area Merritt Island National Wildlife Refuge Refuge Management North of Haulover Canal. van Soelen E, Brooks G, Larson R, Sinninghe Damste J, Reichart G. 2012. Mid- to late-Holocene coastal environmental changes in southwest Florida, USA. The Holocene 22:929–938.

85

Waples RS. 2005. Genetic estimates of contemporary effective population size: to what time periods do the estimates apply? Molecular Ecology 14: 3335- 3352. Waples RS, Do C. 2008. LDNE: a program for estimating effective population size from data on linkage disequilibrium. Molecular Ecology Resources 8:753- 756.

Waples RS, Do C. 2010. Linkage disequilibrium estimates of contemporary Ne using highly variable genetic markers: a largely untapped resource for applied conservation evolution. Evolutionary Applications 3: 244-262. Watts WA. 1980. The Late Quaternary Vegetation History of the Southeastern United States. Annual Review of Ecology and Systematics 11:387–409. Wilcoxen TE, Boughton RK, Schoech SJ. 2010. Older can be better: physiological costs of paternal investment in the Florida scrub-jay. Behavioral Ecology and Sociobiology 64: 1527-1535 Woolfenden G. 1978. Growth and survival of young Florida scrub jays. Wilson Bulletin. 90:1–18. Woolfenden GE, Fitzpatrick JW. 1978. The Inheritance of Territory in Group- Breeding Birds. Bioscience 28:104-108. Woolfenden GE, Fitzpatrick JW. 1984. The Florida Scrub Jay: Demography of a Cooperative-breeding Bird. Princeton University Press (Monographs in population biology). 426 p. Woolfenden GE, Fitzpatrick JW. 1996. Florida Scrub-Jay (Aphelocoma coerulescens). The Birds of North America online, Volume 228. Cornell Lab of Ornithology. https://birdsna.org/Species- Account/bna/species/flsjay/introduction Zimmerman M, Oddy D, Stolen E, Breininger D, Pruett CL. 2015. Microspatial sampling reveals cryptic influences on gene flow in a threatened mammal. Conservation Genetics 16:1403-1414.

86

APPENDIX setwd("C:\\CorysR") estimate_mode <- function(s) { d <- density(s) d$x[which.max(d$y)] }

#Reading the files #Run Master 7 yr hpars.run.master7 <- read.table("hpars run masters 7 yr GT.txt", header = TRUE) curr.size01<-(hpars.run.master7$curr.size) anc.size01<-(hpars.run.master7$anc.size) mean.time01<-(hpars.run.master7$mean.time) fit<-density(curr.size01) fit2<-density(anc.size01) #Plot Current and Ancestral Effective Population Size plot(fit, xlim=c(0,6), ylim=c(0,1.5), main = "MSVAR Effective Population Size") lines(fit2, lty=2) legend("topleft",c("Current","Ancestral"), lty=1:2) mean(curr.size01)10^2.936825 mean(anc.size01) 10^4.470558 estimate_mode(mean.time01) 10^3.875994

87

#Current Population Upper and Lower Bound 95% CI library(coda) HPDinterval(mcmc(curr.size01), 0.95) #Lower bound 10^2.250944 #Upper bound 10^3.617082 #Ancestral Population Upper and Lower Bound 95% CI HPDinterval(mcmc(anc.size01), 0.95) #Lower bound 10^3.856522 #Upper bound 10^5.076079 #Plot Time since Decline of Ancestral Population fit<-density(mean.time01) plot(fit, main = "Estimated Time since Decline in Ne") estimate_mode(mean.time01) 10^3.875994 #Upper and Lower Bound 95% CI HPDinterval(mcmc(mean.time01), 0.95) #Lower bound 10^3.036252 #Upper bound 10^4.58003