The University of Southern Mississippi The Aquila Digital Community

Dissertations

Fall 12-2007 THE BLACK PINESNAKE: SPATIAL ECOLOGY, PREY DYNAMICS, PHYLOGENETIC ASSESSMENT AND HABITAT MODELING Danna Leah Baxley University of Southern Mississippi

Follow this and additional works at: https://aquila.usm.edu/dissertations Part of the Sciences Commons, Biodiversity Commons, Biology Commons, and the Terrestrial and Aquatic Ecology Commons

Recommended Citation Baxley, Danna Leah, "THE BLACK PINESNAKE: SPATIAL ECOLOGY, PREY DYNAMICS, PHYLOGENETIC ASSESSMENT AND HABITAT MODELING" (2007). Dissertations. 1298. https://aquila.usm.edu/dissertations/1298

This Dissertation is brought to you for free and open access by The Aquila Digital Community. It has been accepted for inclusion in Dissertations by an authorized administrator of The Aquila Digital Community. For more information, please contact [email protected]. The University of Southern Mississippi

THE BLACK PINESNAKE: SPATIAL ECOLOGY, PREY DYNAMICS,

PHYLOGENETIC ASSESSMENT AND HABITAT MODELING

by

Danna Leah Baxley

A Dissertation Submitted to the Graduate Studies Office of The University of Southern Mississippi in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

Approved:

December 2007

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. COPYRIGHT BY

DANNA LEAH BAXLEY

2007

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. The University of Southern Mississippi

SPATIAL ECOLOGY, PREY DYNAMICS, HABITAT MODELING, RESOURCE

SELECTION, AND PHYLOGENETIC ASSESSMENT OF THE BLACK PINESNAKE

by

Danna Leah Baxley

Abstract of a Dissertation Submitted to the Graduate Studies Office of The University of Southern Mississippi in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

December 2007

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ABSTRACT

SPATIAL ECOLOGY, PREY DYNAMICS, HABITAT MODELING, RESOURCE

SELECTION, AND PHYLOGENETIC ASSESSMENT OF THE BLACK PINESNAKE

by Danna Leah Baxley

December 2007

As human-caused losses approach unprecedented levels, informed

biological studies at the landscape, community, organismal, and genetic levels become

increasingly important. The six chapters of this dissertation target the black Pinesnake,

Pituophis melanoleucus lodingi, an imperiled specialist of the ecosystem, to

answer biological questions on all four of these levels. At the landscape level, we sought

to identify areas where black Pinesnake populations persist as well as the landscape-level

variables driving presence of this secretive taxon. At the community level, we assessed

habitat associations of and communities within the longleaf pine

ecosystem, and identified habitat parameters that appear to drive community

composition. On a finer scale, we conducted a radio-telemetry study to examine spatial

ecology of black Pinesnakes, and to test several hypotheses regarding prey dynamics and

resource selection of these . Lastly, we used molecular markers (mitochondrial

DNA, i.e. mtDNA) to assess the intraclade relationships of eastern , which

includes P. m. lodingi, P. m. mugitus, and P. m. melanoleucus, in an attempt to add to

existing phylogenetic knowledge of these eastern congeners.

ii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Results of our landscape and community-level studies demonstrate the importance

of large forested tracts of longleaf pine as well as the importance of frequent fire and low

stand densities (low basal areas) to black Pinesnake persistence and restoration. Our

radio-telemetry study indicates that these snakes travel great distances, inhabit large

home ranges, and frequently cross and suffer mortality from roads. Core home ranges of

P. m. lodingi contained significantly more hispid cotton rats than peripheral areas,

indicating a correlation between prey availability and movement for these snakes.

Results of molecular analysis show a lack of sub-species monophyly within eastern

Pituophis, and little geographic structure; however, since there are high levels of

morphological and ecological distinction between these three sub-species, future work

using alternative molecular markers may clarify these within-clade relationships. By

using multiple scales and divergent methods of study, we hope to provide land managers

and biologists with the information necessary to conserve and restore P. m. lodingi

populations.

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ACKNOWLEDGMENTS

Numerous people have been invaluable to the completion of the research involved

in this dissertation. My deepest thoughts of appreciation go to Drs. Carl Qualls, Brian

Kreiser, Jake Schaefer, and Michael Davis, members of my graduate committee, and to

Linda LaClaire of U.S. Fish and Wildlife Service. Each member of my committee has

played a major part in the progress and completion of my research. Dr. Jake Schaefer

provided statistical guidance, conceptual ideas, help with manuscript revisions, and

access to multivariate software. Dr. Micheal Davis assisted in plant identification, and

facilitated thought in plant-animal interactions, which resulted in my second dissertation

chapter. Dr. Brian Kreiser provided manuscript reviews, constructive criticism, and

countless hours of instruction and assistance in the genetics laboratory. I am grateful to

Dr. Carl Qualls, my major professor, for diligently seeking and attaining project funding,

for editing and contributing to countless manuscripts, abstracts, and papers, and for his

availability and openness as a mentor.

This dissertation would not have been possible without funding from the

Mississippi Department of Wildlife, Fisheries, and Parks, U. S. Forest Service, Gopher

Tortoise Council, The Nature Conservancy, The Greater Cincinnati Herpetological

Society, and The University of Sothem Mississippi Biological Sciences Department.

Special thanks goes to Sandra Kilpatrick of the U.S. Forest Service for her interest and

involvement, to Dr. Thomas Ricks who conducted all transmitter implantation surgeries,

and to Mike Sisson, Krista Noel, Josh Ennen, Sam Holcomb, Shea Hammond, Wyatt

iv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Rochelle, and Will Selman for their steadfast assistance in the field and lab. The GIS

habitat modeling portion of this dissertation would not have been possible without the

assistance, patience, and commitment of Greg Lipps. I would also like to express

appreciation to Tom Mann and Dr. Robert Jones of the Mississippi Museum of Natural

Science for providing scientific collection permits as well as their opinions and advice.

Numerous biologists contributed tissue samples to our genetics project and, although I

will not list them all here, I would like to thank them all for their gracious donations of

time and tissue to our project. Lastly, I would like to thank my parents, Jody and Danny

Baxley, my sister Tracy Baxley, and Mike Sisson for their unwavering support and

encouragement throughout my graduate career.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. TABLE OF CONTENTS

ABSTRACT...... ii

ACKNOWLEDGMENTS...... iv

LIST OF TABLES...... viii

LIST OF ILLUSTRATIONS...... x

CHAPTER

I. GENERAL INTRODUCTION...... 1

II. HABITAT ASSOCIATIONS OF REPTILE AND AMPHIBIAN COMMUNITIES IN LONGLEAF PINE HABITATS OF SOUTH MISSISSIPPI...... 7

Abstract Introduction Materials and Methods Results Discussion

III. THE BLACK PINESNAKE { LODINGI): SPATIAL ECOLOGY AND ASSOCIATIONS BETWEEN HABITAT USE AND PREY DYNAMICS...... 29

Abstract Introduction Materials and Methods Results Discussion

vi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. IV. ASSESSING SECOND AND THIRD ORDER HABITAT SELECTION BY BLACK PINESNAKES {PITUOPHIS MELANOLEUCUS LODINGI): A BOOLEAN MODEL USING GEOGRAPHIC INFORMATION SYSTEMS...... 48

Abstract Introduction Materials and Methods Results Discussion

V. RAPID POPULATION EXPANSION AND RECENT DIVERGENCE OF EASTERN PITUOPHIS: P. M. LODINGI, P. M. MUGITUS, AND P. M. MELANOLEUCUS...... 69

Abstract Introduction Materials and Methods Results Discussion

VI. GENERAL CONCLUSION AND RECCOMENDATIONS...... 82

APPENDIXES...... 85

LITERATURE CITED...... 96

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF TABLES

Table

2.1 Property type, number of years surveyed, species richness, and habitat type for sites included in habitat analysis ...... 13

2.2 and surveyed using drift fences and active searching and percentage of sites where each species was surveyed from 2004-2006 ...... 15

2.3 Summary of habitat variables quantified at 20 sites surveyed for reptiles and amphibians in south Mississippi...... 19

2.4 PCA loadings of habitat variables from 20 sites surveyed in 2004 ...... 21

2.5 Canonical correspondence analysis (CCA) final scores for 34 reptile and amphibian species...... 23

3.1 Tracking and home range data for five Black Pinesnakes radio-tracked during 2005-2006...... 38

3.2 Means, standard deviations, test statistics, and p-values for habitat variables collected from mammal transects within core versus outer home ranges of Black Pinesnakes ...... 43

4.1 Mean percent composition, standard deviation, test statistic, and significance level for each variable tested for compositional differences between real and simulated home ranges ...... 59

4.2 Mean, test statistic, and p-values of Wilcoxon Sign-Rank tests comparing minimum distances of radio-tracked Black Pinesnakes and minimum distances of random simulated points to environmental variables ...... 60

4.3 Input parameters for Boolean model ...... 61

4.4 Summary of habitat variables quantified (mean + standard deviation) at 131 sites of recent (post-1987) and historic P. m. lodingi occurrence ...... 64

5.1 Mitochondrial NADH sequences of Pituophis and Lampropeltis obtained from GenBank ...... 74

viii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5.2 Number and location of Pituophis samples used in this study ...... 75

5.3 Haplotype diversity (h), nucleotide diversity, range of pair-wise distance values (d), number of samples (s), and number of unique haplotypes (u) for Pituophis using ND4 and combined ND2 and ND4 ...... 78

ix

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF ILLUSTRATIONS

Figure

2.1 Sites surveyed for reptiles and amphibians 2004-2006 ...... 10

2.2 Linear regression between reptile diversity and canopy cover for all sites surveyed 2004-2006...... 18

2.3 Principal components analysis (PCA) of habitat variables for all sites surveyed in 2004...... 22

3.1 Locations of radio-tracked Black Pinesnakes in South Mississippi ...... 33

3.2 Radio-telemetric locations of eight Black Pinesnakes tracked 2005-2006 ...... 39

3.3 Mean monthly movement distances of Black Pinesnakes in South Mississippi ...40

3.4 Species and number of small mammal individuals captured in core versus outer home ranges of Black Pinesnakes ...... 42

4.1 Boolean model representing areas of optimal suitability (Boolean score = 6), for Black Pinesnakes ...... 62

4.2 Percent occurrence records plotted against percent reduction in land area for Boolean model ...... 63

5.1 Distribution of eastern Pituophis, adapted from Conant and Collins, 1991 ...... 72

5.2 Strict consensus of 15 most parsimonious trees from ND4 data set. Maximum Parsimony, Maximum Likelihood, and Bayesian support values are indicated along branches ...... 77

5.3 Mismatch distribution of pair-wise sequence differences using NADH dehydrogenase subunit 4 (ND4) ...... 79

x

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 1

CHAPTERI

GENERAL INTRODUCTION

The federal Endangered Species Act (ESA), established in 1973, is the current

legislative effort to safeguard and protect the biodiversity within the United States of

America. Section 2 of this Act states, “Congress finds and declares that various species

of fish, wildlife, and plants in the United States have been rendered extinct as a

consequence of economic growth and development untempered by adequate concern and

conservation.” As a result, “the United States has pledged itself as a sovereign state in

the international community to conserve to the extent practicable the various species of

fish or wildlife and plants facing extinction.” Despite efforts of the ESA, trends of

accelerated loss and declines of species have been recognized for nearly all taxonomic

groups, including reptiles and amphibians both within the United States and globally

(Dirzo and Raven, 2003). Assessing population trends of reptiles and amphibians proves

problematic for several reasons. Long-term population monitoring for herpetofauna is

resource intensive, requires significant amounts of time, and is rarely implemented

(Gibbons et al., 2000). Furthermore, it is difficult to separate natural population

fluctuations from population declines, especially for amphibians (Pechmann et ah, 1991;

Blaustein et ah, 1994; Green, 2003). As a result of these factors, reptiles and amphibians

often invoke conservation concern and attain candidate or protected ESA status only after

populations suffer reductions so severe that recovery is difficult or improbable. For

example, the Mississippi Gopher Frog (Rana sevosa) was reduced to one breeding

population before attaining ESA protection (Richter and Seigel, 2002) and several reptile

species, such as the Eastern Indigo (Drymarchon corais couperi) and the Southern

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2

Hog-nosed Snake (Heterodon simus), quietly disappeared from Mississippi before

biologists noticed or acted upon declines (Tuberville et al., 2000).

Small, isolated populations, especially those that require unique or sensitive

habitats, are particularly susceptible to population declines and extinction. The longleaf

pine ecosystem, a landscape that once almost entirely characterized the southeastern

United States and has been severely diminished, is home to approximately 72 reptile and

amphibian species. Of these, about one third are classified as specialists and are found

only within the historic range of longleaf pine forests (Guyer and Bailey, 1993).

Countless factors have contributed to the decline and degradation of the longleaf pine

ecosystem. The main causes for habitat loss include fire suppression, removal of pine

stumps for the naval stores industry, invasion of non-native species, and the conversion

of longleaf pine forests to urban, silvicultural, or agricultural areas (Frost, 1993; Outcalt,

1997; Varner and Kush, 2004).

Even in conserved areas, such as DeSoto National Forest, Mississippi, longleaf

pine forests suffer ongoing quality reduction because prescribed bums predominately

occur during the winter to ameliorate issues of liability and smoke control. Historically,

lightning-caused fires burned longleaf pine forests during the growing season. These

types of fires bum at higher temperatures and are more effective at eradicating competing

midstory and shrub vegetation (Glitzenstein et al., 1995). Winter bums have also been

implicated in amphibian declines, though research on this topic has been highly

controversial (Schurbon and Fauth, 2003; Means et al., 2004).

In addition to creating logistical problems with prescribed burning, habitat

fragmentation is emerging as a serious threat to remaining tracts of longleaf pine forest.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3

The size, growth, and resource demands of the human population are the most important

agents of change in spatial patterns of biodiversity (Vitousek et al., 1997; Sala et al.,

2000). Although global estimates of human population predict declines in U.S.

population by 2050 (Cohen, 2003), U.S. trends of migration to warmer and more

temperate areas of the country will probably continue. As human population growth and

resulting development in the southeast increases, many isolated tracts of longleaf pine

forest will become functionally extinct because prescribed fire will no longer be feasible.

Without natural or prescribed fire, the longleaf ecosystem quickly loses the successional

battle and hardwood species become dominant. In the absence of nearby longleaf

habitats containing metapopulations, to which amphibians and reptiles are capable of

dispersing, the reptile and amphibian populations within these fragments may become

extinct (McCullough, 1996; Marsh and Trenham, 2001).

The efforts of the ESA to manage habitat use on both public and private lands and

to mitigate effects of human development and urbanization upon threatened and

endangered taxa have been widely criticized (Tear et al., 1993; Schemske et al., 1994;

Foin et al., 1998). This act has been criticized for exhibiting a taxonomic bias in favor of

over plants, vertebrates over invertebrates, and and mammals over fish and

herpetofauna (Tear et ah, 1995), although the main criticism of the ESA is the lack of

encouragement within this statute for proactive actions that may preclude the need to list

species as threatened or endangered (Noss et ah, 1997). In an attempt to amend some of

the problems associated with this reactive approach to conservation, over 275 U.S.

species are listed as “candidates” for federal ESA listing; this means not enough

information is currently available to make informed decisions about listing these species,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. but population declines are suspected. Among these federal candidate species is the

black Pinesnake ( Pituophis melanoleucus lodingi), a longleaf pine specialist that is also

state-listed in Mississippi with the status of “endangered.”

Until recently, very little was known concerning the natural history, spatial

ecology and current distribution of the Black Pinesnake; much of the available literature

concerns captive specimens or laboratory trials (Clibum, 1957; Clibum, 1962; Clibum,

1976). Although the historic range of P. m. lodingi spanned from western Florida to

extreme eastern (Crain and Clibum, 1971; Conant and Collins, 1991), this sub­

species is thought to be extirpated from Louisiana, and has not been reported west of the

Pearl River for nearly forty years. Approximately 61% of historical Black Pinesnake

populations are thought to be extirpated or are in decline (Duran, 1998b). Aside from

one unpublished radio-telemetry study conducted within optimal habitat (Duran, 1998a)

and comparison of prey handling techniques between P. m. lodingi and P. ruthveni

(Rudolph et al., 2002), there is a paucity of P. m. lodingi research by which to inform

candidate ESA status.

In addition to the general lack of ecological data for Black Pinesnakes, there is

also very little available data concerning the genetic relationships among eastern

Pituophis. Currently one of fifteen recognized sub-species of Pituophis, this snake is

geographically isolated from all other members of the genus; consequently, some

researchers support elevation of this sub-species to full species status (Duran, 1998b).

Debates regarding the relationships among eastern and western Pituophis are

characterized by several taxonomic hypotheses (see Rodriguez-Robles and Jesus-

Escobar, 2000 for a complete review). Most recently, Burbrink and Lawson (2007)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5

suggested inclusion into the Pituophis genus of all members of the clade formed by the

most recent common ancestor of Pantherophis guttatus. Another recent systematic

analysis of new world gopher, bull, and Pinesnakes indicates that the eastern Pinesnake

clade (P. m. lodingi , P. m. mugitus, and P. m. melanoleucus) and the western Pinesnake

clade (P. c. affinis, P. c. annectens, P. c. bimaris, P. c. catenifer , P. c. deserticola, P. c.

sayi, and P. c. vertebralis) represent divergent lineages and should be recognized as two

distinct species (Rodriguez-Robles and Jesus-Escobar, 2000). Although Rodriguez-

Robles and Jesus-Escobar (2000) found monophyly of the eastern Pinesnake clade to be

well supported, within-clade relationships for this group were characterized by low

bootstrap values, and were relatively unresolved. Aside from the Rodriguez-Robles and

Jesus-Escobar study, there is little information concerning molecular systematics of the

eastern Pinesnake clade as a whole, and the Black Pinesnake in particular. For instance,

total molecular systematic work on the Black Pinesnake consists of mitochondrial DNA

analysis from two black Pinesnakes obtained via the pet trade, where snakes are often

hybridized or back-crossed resulting in uncertain lineages.

Informed biological studies at the landscape, community, organismal, and genetic

levels become increasingly important as P. m. lodingi populations continue to decline.

The objective of this dissertation was to use a variety of research scales to expand the

base of available knowledge regarding P. m. lodingi to provide land managers and

biologists with decision-making tools. At the landscape level, we sought to identify areas

where black Pinesnake populations persist as well as the landscape-level variables

driving presence of this secretive taxon. At the community level, we assessed habitat

associations of reptile and amphibian communities within the longleaf pine ecosystem,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 6

and used multivariate analysis to identify habitat parameters that appear to drive

community composition. On a finer scale, we conducted a radio-telemetry study to

examine spatial ecology of Black Pinesnakes, and to test several hypotheses regarding

prey dynamics and spatial ecology of these snakes (e.g. quantify 3rd order resource

selection within home ranges). Lastly, we used molecular markers (mtDNA) to assess

the intraclade relationships of the eastern Pituophis clade, which includes P. m. lodingi,

P. m. mugitus, and P. m. melanoleucus, in an attempt to add to existing phylogenetic

knowledge of these eastern congeners. By using multiple scales and divergent methods of

study, we hope to provide land managers and biologists with the information necessary to

conserve and restore P. m. lodingi populations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7

CHAPTER II

HABITAT ASSOCIATIONS OF REPTILE AND AMPHIBIAN COMMUNITIES IN

LONGLEAF PINE HABITATS OF SOUTH MISSISSIPPI

Abstract.—Habitat associations of reptiles and amphibians native to longleaf pine forests

in southern Mississippi are scantly studied and poorly understood. From 2004 to 2006,

we surveyed the herpetofauna of twenty-four longleaf pine communities in twelve

counties in south Mississippi. For each site, we quantified herpetofaunal diversity and

relative abundance and a suite of habitat variables to address the following objectives: (1)

determine what levels of habitat heterogeneity exist in longleaf pine forests in south

Mississippi; (2) determine if there are differences in reptile and amphibian community

composition among these sites; and (3) if habitat-faunal differences exist among sites,

identify what habitat variables are driving these community differences. Multivariate

analysis identified three distinct longleaf pine habitat types, differing primarily in soil

composition and percent canopy cover of trees. Canonical correspondence analysis

indicated that canopy cover, basal area, percent grass in the understory, and soil

composition (percent sand, silt, and clay) were the predominant variables explaining

community composition at these sites. Many species exhibited associations with some or

all of these habitat variables. The significant influence of these habitat variables,

especially basal area and canopy cover, upon herpetofaunal communities in south

Mississippi indicates the importance of incorporating decreased stand density into

management practices for longleaf pine habitat.

Key Words.—amphibian; canopy cover; community; habitat associations; herpetofauna;

longleaf pine, Mississippi; reptile;

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 8

Introduction

The longleaf pine (Pinus palustris) ecosystem, a landscape that once almost entirely

characterized the southeastern coastal plain of the United States, has been severely

diminished. The reduction of longleaf pine forest to less than five percent of its historic

range is largely due to fire suppression and the conversion of longleaf pine forests to

agricultural, silvicultural, or human use areas (Frost 1993; Outcalt 1997; Varner and

Kush 2004). Due to the elusive, fossorial, and often cryptic nature of many reptile and

amphibian species, the community assemblages and habitat associations of herpetofauna

native to longleaf pine forests in Mississippi are poorly known. Guyer and Bailey (1993)

described the unusually diverse herpetofauna of longleaf forests and estimated that

approximately 72 reptile and amphibian species are associated with this habitat; while

one third of these species are found only within the historic longleaf pine range and are

classified as specialists. Although studies describing reptile and amphibian communities

in upland longleaf pine habitats are relatively common (Tuberville et al., 2005; Smith et

al., 2006), most studies addressing habitat-faunal relationships within the longleaf pine

ecosystem have targeted Florida and South Carolina (Dodd and Franz, 1995; Yale et al.,

1999; Litt et al., 2001).

Southern Mississippi contains a relatively under-studied sector of longleaf pine

forest and, to our knowledge, there are no published studies concerning the habitat

associations of herpetofaunal communities within this area. In an attempt to better define

the faunal-habitat relationships of these communities, we sought to address the following

three objectives: (1) determine what levels of habitat heterogeneity exist in longleaf pine

forests in south Mississippi; (2) determine if there are differences in reptile and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 9

amphibian community composition among these sites; and (3) if habitat-faunal

differences exist among sites, identify what habitat variables are driving these

differences. As proposed by Dodd and Franz (1993), we also sought to add to the base of

information concerning entire herpetofaunal communities, including common species,

such that monitoring protocols may ultimately be developed and implemented in South

Mississippi. By better defining the faunal-habitat relationships of these communities, the

natural history and ecology of many longleaf pine resident and specialist species may be

better understood.

M a t e r ia l s a n d M e t h o d s

Study Sites.—The search for suitable longleaf pine habitat was primarily restricted

to the southern, coastal plain region of Mississippi. Habitat targeted in this study was

characterized by sandy soils, an overstory of longleaf pine trees, a fire-maintained,

herbaceous, grassy understory, and a fire-suppressed mid-story. In total, twenty-four

sites with suitable habitat were identified and surveyed in twelve counties (Figure 2.1).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 10

Sites 1 i

17,18 v,- -

Figure 2.1 Sites surveyed for reptiles and amphibians 2004-2006. Shaded areas represent public land (National Forest or Wildlife Management Area).

Of these sites, twenty-two were on public land and two were on private land. Seventeen

of the public sites were within the DeSoto National Forest (six within the Chickasawhay

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 11

district and eleven within the DeSoto District), we surveyed one site on the Mississippi

Sandhill Crane National Wildlife Refuge, three sites within the Marion County Wildlife

Management Area, and one site owned by the Hancock County School Board.

Additionally, several private landowners in both Hancock County and George County

allowed their lands to be surveyed.

Survey techniques.—To address habitat variation as well as habitat/community

associations of reptiles and amphibians in Mississippi, our project was divided into two

parts. In 2004, all identified sites were surveyed using standard herpetological sampling

techniques using a combination of drift fences with pitfall and funnel traps (Gibbons and

Semlitsch, 1981; Campbell and Christman, 1982; Fitch, 1987), manual searching, cover

boards, and road surveys. Due to time and budget constraints, a subset of the sites

surveyed in 2004 were selected for further surveys in 2005 and 2006, and three sites were

surveyed only in 2005 (see Table 2.1 for number of survey years for each site). In 2004

each drift fence (100 ft long by 3 ft high) was equipped with four funnel traps, one at

each end on each side, and four pitfall traps, one near each end on each side. To sample

multiple sites within a local region simultaneously, twelve or more drift fence/trap arrays

were deployed at all times. As a result of low capture rates for species of interest with

cover boards and pit-fall traps in 2004, we did not employ these techniques in 2005 or

2006. The number of drift fences erected at each site was determined by the area of

suitable habitat present at each site. Each year, sites were surveyed for at least 14

consecutive days in the spring (March 15- May 27) and 14 consecutive days in the

summer (May 28-July 29).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 12

All arrays were checked at least once per day, and were checked twice per day

during hot conditions. All captured animals were removed from traps and released at the

site of capture each time traps were checked. For each animal captured, the location,

date, and species were recorded. During the first week of July 2004, 2005, and 2006,

habitat data was collected at each site where sampling occurred. The following

characteristics of each site were quantified and recorded: soil composition, dominant

vegetation species, percent canopy cover of trees, percent mid-story coverage, percent

shrub coverage, basal area of trees, percentage cover of grasses, forbs, and shrubs in

understory, ground-cover percentage of litter and bare soil, and recency of fire. A

spherical densiometer (Forest Densiometers model-C) was used to measure canopy

cover. At each drift fence array, four densiometer readings were taken to estimate

percent canopy cover facing north, south, east, and west of the array; these readings were

averaged to estimate overall cover. A Cruz-All basiometer (model 59793) was used to

measure factor-10 basal area. Measurements for both canopy cover and basal area were

taken from the middle of each 100’ drift fence array. We used a one-meter quadrat (four

replicates per drift fence) to estimate average percentage of grasses, forbs, shrubs, litter

and bare soil comprising the groundcover, and we visually estimated the percentage of

mid-story coverage and percentage of shrubs in the mid-story. To increase measurement

consistency, all habitat parameters were measured by one individual (D. B.). Lastly, we

used the soil hydrometer method (Sheldrick and Wang, 1993) to determine percentage

content of sand, silt, and clay for all soil samples.

Data Analysis .— To assess degree of habitat heterogeneity among survey sites,

we used principal components analysis (PCA) coupled with analysis of similarity

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 13

(ANOSIM) using the Bray Curtis similarity index. For this analysis, we included all sites

surveyed in 2004 (twenty sites). To explicitly assess habitat associations of reptiles and

amphibians, we employed canonical correspondence analysis (CCA) using all sites (and

all years) surveyed.

Table 2.1. Site numbers corresponding to map (Figure 2.1), property type, number of years surveyed, species richness, and habitat type for sites included in the PCA habitat analysis (O = open canopied group, M = moderate canopy closure group, and C = closed canopy group).

Site # County Property Type Years Surveyed So. Rich Hab

1 Jones DeSoto National Forest 3 20 C 2 Wayne DeSoto National Forest 2 16 - 3 Wayne DeSoto National Forest 3 20 C 4 Wayne DeSoto National Forest 3 24 C 5 Wayne DeSoto National Forest 3 29 c 6 Wayne DeSoto National Forest 1 8 o 7 Greene DeSoto National Forest 1 17 M

8 Marion Marion Co. Wildlife Management Area 1 10 -

9 Marion Marion Co. Wildlife Management Area 1 21 -

10 Marion Marion Co. Wildlife Management Area 1 15 - 11 Forrest DeSoto National Forest 1 11 c 12 Forrest DeSoto National Forest 1 14 M 13 Forrest DeSoto National Forest 1 15 O 14 Perry DeSoto National Forest 32 M 15 Forrest DeSoto National Forest 1 7 C 16 Pearl River DeSoto National Forest 1 10 C 17 Perry DeSoto National Forest 1 13 M 18 Perry DeSoto National Forest 1 15 O 19 George Private 1 11 M 20 Harrison DeSoto National Forest 1 17 M 21 Harrison DeSoto National Forest 1 8 M 22 Hancock Private 1 19 O 23 Hancock Private 1 16 M 24 Jackson Sandhill Crane National Wildlife Refuge 1 21 O

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Since habitat parameters for each site did not remain constant across years (substantial

changes in habitat resulted from prescribed fire and Hurricane Katrina), we considered

each survey year for a given site a different replicate. All community data were

standardized to adjust for small disparities in number of trap days for each site (resulting

from differences in suitable habitat and number of drift fence arrays per site). We did not

include arboreal snakes (Opheodrys aestivus and Elaphe obsoleta), frogs o f the genus

Hyla, or extremely rare reptiles and amphibians (total number of captures less than five

individuals) in our community analyses. We excluded arboreal snakes and tree frogs

from our analysis because these groups are not effectively sampled using drift fences

with funnel and pitfall traps (Dodd, 1991), and we sought to minimize sampling bias to

the greatest extent possible. Scaling options for CCA were as follows: axis scores

centered and standardized to unit variance, axes scaled to optimize representation of sites,

and we chose to use Monte Carlo permutation tests (Manly, 1991) with 9,999 iterations to

test two null hypotheses: that there is no relationship between the species and habitat

matrices, and that there are no species-environment correlations. To test the null

hypothesis that slope of reptile diversity plotted against percent canopy cover is equal to

zero, we conducted a one-way ANOVA using JMP (Version 6. SAS Institute., Cary, NC,

1989-2005); alpha was 0.05 for all tests. All multivariate analyses were conducted using

PRIMER 5.2.9 (Primer-E Ltd, Roborough, Plymouth, United Kingdom) and PcOrd 5

(MJM Software Design, Gleneden Beach, Oregon, United States).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 15

R e s u l t s

Over three field seasons, 3,464 trap days (one drift fence array open for one 24-

hour period = one trap day) resulted in 2,192 captures (Table 2.2), representing 54

species (20 amphibian and 34 reptile).

Table 2.2. Reptiles and amphibians surveyed using drift fences and active searching and percentage of sites where each species was surveyed from 2004-2006. Abbreviations in parentheses are used in Table 2.5.

Scientific Name Common Name # Surveyed % o f Sites

Turtles Gopherus polyphemus (Gophpoly) Gopher Tortoise 68 66.7 Terrapene Carolina (Terracar) Eastern Box Turtle 13 33.3

Lizards Anolis carolinensis (Anolcar) Green Anole 37 66.7 Aspidocelis sexlineatus (Aspsex) Six-lined Racerunner 31 37.5 Eumeces fasciatus (Eumefasc) Common Five-lined Skink 11 33.3 Eumeces inexpectatus (Eumeinex) Southeastern Five-lined Skink 53 66.7 Eumeces laticeps Broad-headed Skink 2 8.3 Ophisaurus attenuatus Slender Glass Lizard 1 4.2 Ophisaurus ventralis Eastern Glass Lizard 2 8.3 Sceloporus undulatus (Scelopun) Eastern Fence Lizard 160 88 Scincella lateralis (Scinclat) Ground Skink 362 92

Snakes Agkistrodon contortrix (Agkiscon) Copperhead 7 29.2 Agkistrodon piscivorus (Agkistpi) Cottonmouth 5 20.8 Cemophora coccinea (Cemcocc) Scarlet Snake 63 58.3 Coluber c. priapus (Coicon) Black Racer 158 87.5 Crotalus adamanteus (Crotadam) Eastern Diamondback Rattlesnake 6 25.0 Diadophis punctatus Ring-necked Snake 3 12.5 Elaphe guttata (Elaphgut) Com Snake 9 25.0 Elaphe obsoleta Gray Rat Snake 10 16.6 Heterodon platirhinos (Hetplati) Eastern Hog-nosed Snake 19 37.5 Lampropeltis getula (Lampgetu) Eastern King Snake 18 33.3 Lampropeltis triangulum (Lamptria) Scarlet King Snake 24 33.3 Masticophis flagellum (Mastflag) Coachwhip 52 45.8

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 16

Table 2.2. (cont.) Reptiles and amphibians surveyed using drift fences and active searching and percentage of sites where each species was surveyed from 2004-2006.

Scientific Name Common Name # Surveyed % of ______Sites

Snakes Micrurus fulvius Coral Snake 3 8.3 Nerodia erythrogasterflavigaster Yellow-bellied Water Snake 2 8.3 Opheodrys aestivus Rough Green Snake 3 12.5 Pituophis melanoleucus (Pitmel) Pine Snake 10 20.8 Regina rigida Crayfish Snake 3 12.5 Sistrurus miliarius (Sistrmil) Pygmy Rattlesnake 8 29.2 Storeria dekayi (Stordeka) Dekay’s Snake 8 16.7 Storeria occipitomaculata Red-bellied Snake 1 4.2 Tantilla coronata (Tantcoro) Southeastern Crowned Snake 13 54.2 Thamnophis sauritus (Thamsaur) Eastern Ribbon Snake 16 33.3 Thamnophis sirtalis (Thamsirt) Common Garter Snake 19 79.2

Salamanders Eurycea guttolineata (Eurycgut) Three-Lined Salamander 8 25.0 Eurycea quadridigitata Dwarf Salamander 2 8.3 Notopthalmus viridescens Eastern Newt 1 4.2 Plethodon Mississippi (Plethmis) Mississippi Slimy Salamander 22 37.5

Frogs Acris gryllus (Accgryl) Southern Cricket Frog 22 37.5 Rufo terrestris (Bufter) Southern Toad 244 87.5 Bufo quercicus (Bufquerc) Oak Toad 16 4.2 Gastrophryne carolinensis (Gascar) Eastern Narrow-mouthed Toad 557 91.7 Hyla chrysoscelis/versicolor Gray Treefrog 2 8.3 Hyla cinerea Green Treefrog 2 8.3 Hyla femoralis Pinewoods Treefrog 20 29.2 Hyla gratiosa Barking Treefrog 8 12.5 Hyla squirella Squirrel Treefrog 2 8.3 Pseudacris crucifer Spring Peeper 2 8.3 Rana catesbeiana (Rancat) Bull Frog 6 20.8 Rana c. clamitans (Ranclam) Bronze Frog 51 41.7 Rana grylio Pig Frog 1 4.2 Rana palustris Pickerel Frog 1 4.2 Rana sphenocephala (Ransphen) Southern Leopard Frog 7 29.2 Scaphiopus holbrooki (Scaphhol) Eastern Spade foot Toad 154 25.0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Average trap mortality was 3.3% and red imported fire ants ( Solenopsis invicta ) were

present in 71% of these mortality events. There were significant yearly differences in

reptile and amphibian communities sampled (ANOSEM Global R = 0.316, P = 0.008);

lowest numbers of both species and individuals were sampled in the severe drought year

of 2006, though yearly differences were not significant (relative abundance x2 = 4.82, df

= 2, P = 0.089; diversity (S) x2 = 4.07, df = 2, P = 0.13). It is of note that we surveyed

few true longleaf pine specialists; furthermore, we failed to detect Heterodon simus,

Drymarchon covais couperi, Rana sevosa or Pseudacris ornata. A one-way ANOVA

test (F-ratio = 4.92, df = 1,33, P = 0.03) indicated a significant negative linear

relationship between percent canopy cover and reptile diversity (Figure 2.2). Relative

abundances of reptiles and amphibians surveyed are listed in Table 2.2. Specific locality

data and further survey data for sites surveyed in 2004 and 2005 are catalogued with the

Mississippi Museum of Natural Science (Smith and Qualls, 2004; Baxley and Qualls,

2005).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 18

Q_

O r~ i Q Average % Canopy Cover

Figure 2.2 Linear regression indicating a significant negative linear relationship between reptile diversity and canopy cover for all sites surveyed 2004-2006.

For our habitat analysis of twenty sites in eleven counties, the first three PCA

axes explained 71.5% of the habitat variation between sites. Percent canopy coverage

and soil composition (percent sand silt and clay in the soil) were the factors with highest

loadings on the first two axes, explaining most of the variation in the model, while basal

area, and percentage of shrub growth and bare soil in the understory explained most of

the variation on the third axis (Table 2.2). Three distinct habitat-group clusters resulted

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 19

from the PCA (Figure 2.3). We identified three habitat-group clusters from our PCA, and

subsequently defined each group as separate factors within ANOSIM. Monte Carlo tests

indicated that these three groups represent significantly different site types (Global R =

0.452, P = 0.001). Means and standard deviations for habitat variables for these three

groups are displayed in Table 2.4.

Table 2.3. Summary of habitat variables quantified (mean ± standard deviation) at 20 sites surveyed for reptiles and amphibians in south Mississippi. Groups represent different site types identified by PCA and ANOVA (see Fig. 2.3).

Habitat variable Group 1 Group 2 Group 3 Closed Canopy Moderate Canopy Open Canopy

% Sand 61.7+11.9 83.3 ±12.6 80.2 ± 11.6

% Clay 6.2+ 2.2 3.4 ±2.2 3.9 ±2.4

% Silt 32.1 ± 10.9 13.2 ±10.5 15.9 ±9.8

% Forb 3.9 ±3.9 1.8 ±1.1 3.8 ±4.3

% Shrub 22.3 + 14.0 33.1 ± 15.3 25.8 ±25.8

% Litter 66.7 ±26.7 78.2 ±16.3 57.0 ± 15.5

% Bare ground 16.4 ± 17.1 12.1 ± 12.9 17.9 ± 14.9

Basal area m2 110 ±27.7 69.0 ±9.9 24.0 ± 8.9

% Canopy Cover 67.1 ± 12.1 45.0 ±5.9 23.6 ±8.8

Bum recency (months) 17.6 ± 18.0 28.0 ±8.8 28.8 ± 15.5

% Grass 15.9 ±11.0 11.7 ± 12.8 21.7 ± 15.6

% Midstory coverage 17.0^30.0 18.0 ± 12.2 28.2 ±25.8

% Shrubs in midstory 48.2 ±22.1 68.0 ± 16.2 48.9 ±27.1

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 20

The first three axes of the CCA explained 22.9% of the variance within the site

and species matrices. Soil composition (percent sand and silt in the soil) accounted for

most of the variation explained on the first axis, percent grass in the understory and

percent sand accounted for most of the species-habitat variation explained by axis two,

and canopy cover and basal area explained most of the variation on the third axis. Monte

Carlo permutation tests revealed significant structure (i.e. P = proportion of randomized

runs with eigenvalues greater than or equal to the observed eigenvalue) in our habitat and

community data for the second and third axes, but not for axis one (axis one P = 0.12,

axis two P = 0.02, axis three P = 0.0006). Although species-environment correlations

were statistically significant only for axis three (axis one P = 0.22, axis two P = 0.49, and

axis three P = 0.002), these correlations were extremely high for both observed (real data

mean correlation = 0.90) and simulated data (Monte Carlo mean correlation = 0.85).

When sites are plotted in ordination space according to habitat parameters, the majority

of sites cluster near the middle of the habitat ordination, although there are several sites

that are apparent outliers, characterized by habitat dissimilar to most of the other sites.

When species are plotted in ordination space according to habitat parameters, several

species are correlated with one or more habitat parameters (See Table 2.4 for loadings).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 21

Table 2.4. PCA loadings of habitat variables from 20 sites surveyed in 2004. The first three axes of the PCA explain 71.5% of habitat variance. The major axis loading for each habitat variable is displayed in bold.

Variable PCI PC2 PC3

% midstory -0.25 -0.20 -0.12

% shrub coverage -0.14 0.15 0.46

% sand -0.41 0.25 -0.09

% clay 0.40 -0.12 0.07

% silt 0.40 -0.27 0.08

Ave. forb 0.28 0.16 0.34

Ave. Litter -0.30 -0.33 0.21

Ave Bare Soil 0.12 0.26 0.58

Ave Shrub -0.10 0.02 0.59

Basal Area 0.12 0.22 -0.49

% Canopy Cover 0.02 -0.54 -0.07

Bum recency -0.34 -0.08 0.01

Ave Grass 0.36 0.17 0.09

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 22

I

% Canopy Cover

Basal Area

cn O CL

-2 --

-3 -■

-4 -■

3 2 0 1 2 4 f> 4 Percent Silt + Percent Sand Percent Clay

PC1

Figure 2.3. Principal components analysis (PCA) of habitat variables for all sites surveyed in 2004. Triangles represent closed canopied sites, inverse triangles represent sites with moderate canopy coverage, and squares represent sites with relatively open canopies. Numbers correspond to site numbers used in both Figure 2.1 and Table 2.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table 2.5. Canonical correspondence analysis (CCA) final scores for 34 reptiles and amphibian species. See Ter Braak (1986) for detailed explanation of CCA output. Species abbreviations are given in Table 2.2. The major axis loading for each species displayed in bold.

Species Axis 1 Axis 2 Axis 3

Accgryl 0.14 0.62 -0.94 Agkiscon 2.47 -0.10 0.29 Agkistpi -1.36 2.23 1.93 Anolcar 0.83 0.45 -0.52 Aspsex -2.63 -1.08 0.23 Bufter -0.39 0.04 -0.12 Bufquerc -5.23 5.35 6.80 Cemcocc 0.50 -0.31 0.75 Coicon -0.03 0.56 -0.01 Crotadam 1.32 0.53 0.84 Elaphgut -0.92 0.88 0.98 Eumefasc -0.98 1.99 0.28 Eumeinex 0.08 -0.10 0.97 Eurycgut 0.12 -1.00 -1.96 Gascar -0.60 0.29 -0.28 Gophpoly 0.18 -2.33 0.42 Hetplati 0.61 -0.26 -1.17 Lampgetu -1.00 -1.82 -0.24 Lamptria 1.88 0.01 0.13 Mastflag -0.27 -2.18 1.04 Pitmel 0.95 -1.54 0.63 Plethmis -0.92 1.03 -2.98 Rancat -0.77 1.58 -0.92 Ranclam -0.28 0.53 -1.90 Ransphen -1.12 0.21 0.30 Scaphhol -2.60 -1.78 -1.27 Scelopun -0.10 0.32 0.47 Scinclat 0.96 0.40 -0.50 Sistrmil -0.68 -0.02 0.44 Stordeka 1.34 -0.53 3.98 Tantcoro 0.48 1.44 -0.25 Terracar 1.69 -0.16 2.16 Thamsirt 1.35 -0.06 0.65 Thamsaur 2.11 1.05 0.01

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 24

For example, the Oak Toad ( Bufo quercicus) is positively correlated with high

percent grass in the understory and extremely low percent canopy cover. Several species,

such as the Six-lined Racerunner (Aspidoscelis sexlineatus), Spadefoot Toad ( Scaphiopus

holbrooki), and Eastern Coachwhip (Masticophis flagellum) are positively associated

with sandy soils and open canopies. Conversely, several species seem to be associated

with sites characterized by more closed canopies and less sand in the soil: Eastern Ribbon

Snake ( Thamnophis sauritus), Copperhead (Agkistrodon contortrix), Scarlet Kingsnake

(Lampropeltis triangulum), Garter Snake {Thamnophis sirtalis ), and Dekay’s Snake

{Storeria dekayi). As expected, many common/generalist species plot near the center of

the ordination, and are neither positively or negatively correlated with specific habitat

parameters: Southern Toad {Bufo ter res tr is), Fence Lizard {Sceloporus undulatus),

Narrowmouth Toad {Gastrophryne carolinensis), Black Racer {Coluber constrictor

priapus), Southern Cricket Frog {Acris gryllus), and Bronze Frog {Rana clamitans

clamitans).

D is c u s s io n

Defining and understanding regional habitat associations of reptiles and

amphibians is a multi-dimensional and complicated facet of community ecology. With

so much stochasticity inherent to natural systems, questions at the community level are

difficult to address, and results of such studies can be equally difficult to interpret. When

interpreting the results of this study, we assumed that any given species present in upland

habitat in south Mississippi has an equal capture probability across sites; therefore, the

likelihood of capturing this species over a standardized amount of trap days should be

equal for any given site. Variation in weather patterns and landscape variables surely

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 25

affected our study; despite these factors, we believe that the duration and intensity of our

sampling allows us to compare relative abundances of reptiles and amphibians across

sites, and correlate habitat variables to these relative abundances.

The first objective of this study was to determine levels of heterogeneity within

historical upland longleaf pine habitat in south Mississippi, and it is apparent from our

ANOSIM that substantial variation exists. In contrast to the densely planted silvicultural

forests typifying much of the gulf coast plain, old-growth longleaf pine forests are

characterized by relatively open canopies and low basal areas, and were described by

Bartram (1791) as “savannah” habitat. This frequently burned, open-canopied longleaf

habitat described by Bartram is characterized by high percentages of grasses and forbs in

the groundcover, whcih this translates into some of the richest plant biodiversity in North

America: up to thirty species per square meter (Frost et al., 1986). As most forest stands

in south Mississippi are in some stage of forest regeneration, it follows that basal area and

percent canopy cover are two factors explaining much of the variation between surveyed

stands.

Our second objective was to correlate reptile and amphibian community

differences with the observed differences in longleaf pine habitat. Although we did see

significant structure in our CCA, indicating that our measured habitat variables explained

some of the variation in observed herpetofuanal communities, most of the community

variation observed in our study was not explained in the CCA model. The missing

variation in our model is likely due to both historical effects and landscape level effects

(habitat fragmentation, proximity to roads, previous forestry practices such as intensive

site preparation, etc.).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 26

The highly significant structure (P = 0.0006) and species-environment correlation

(P = 0.0020) on axis three of the CCA (where most of the variation is explained by

canopy cover and basal area) indicate that percent canopy cover and basal area are two

factors contributing greatly to faunal-community relationships in longleaf pine forests of

South Mississippi. Canopy cover is widely recognized as a key factor determining forest

microhabitat, plant growth and survival, and wildlife habitat within the forest (Jennings et

al., 1999). Furthermore, fire is a key factor in maintaining the open canopies of longleaf

pine savannahs in the southeast (Komarek, 1965). Without fire, hardwood encroachment

ensues which results in canopy closure and, as less sunlight filters to the forest floor, loss

of herbaceous grandcover (Waldrop et ah, 1992). This herbaceous groundcover is

essential for growth, development, and reproduction of the gopher tortoise (Mushinsky

and McCoy, 1994), which has been portrayed as an ecosystem engineer (Jones et ah,

1994) of the longleaf pine ecosystem. Although recommended percent canopy cover for

gopher tortoises is between thirty and fifty percent (Wilson et al, 1997), our CCA did not

positively associate gopher tortoises with open canopied forests. This anomaly is likely

because most active gopher tortoise burrows within our study sites were encountered

along roads or powerlines. These edge areas are characterized by more open canopies

and grassier understories than the interior areas containing our drift fences, where we

collected habitat parameters. To further stress the importance of open canopies to gopher

tortoises, this animal has been shown to abandon its burrow primarily in association with

changes in overstory structure (increased canopy cover) which result in shading of active

burrows (Aresco and Guyer, 1999).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Our research supports the findings of several previous studies regarding the

importance of canopy cover to herpetofaunal community structure in the southeast. For

example, Meshaka and Layne (2002) described openings with sparse or no shrub cover as

critical factors allowing the persistence of xeric specialists on a long unbumed site in

Florida. In addition, a study of herpetological diversity within young growth forests of

South Carolina found a negative correlation between reptile diversity and upland canopy

closure (Russell et al., 2002). Lastly, Dodd and Franz (1995) found snake communities

of high pine habitats in north-central Florida to be more diverse and evenly distributed

than snake communities sampled in closed xeric hammock habitat.

Many previous habitat association studies targeting herpetofaunal communities

specifically address community differences in response to management practices; for

example, Litt et. al. (2001) found higher capture rates of both Aspidoscelis sexlineatus,

and Sceloporus undulatus in burned habitat plots when compared to control or herbicide

treated plots. Although we did not find notable habitat associations for either of these

species when examining recency of prescribed bum, we did find Aspidoscelis sexlineatus

to be associated with open habitats. This is not surprising given that teiid lizards are

known to have the highest rates of oxidative metabolism among reptiles (Asplund, 1974),

and may require open habitats to achieve their preferred high body temperatures.

The results of this study and previous studies indicating the significant influence

of basal area and canopy cover upon longleaf pine communities have important

management and conservation implications. In addition to the gopher tortoise, many

other longleaf pine residents and specialists have undergone similar evolutionary

pressures within this ecosystem; consequently, lowering stand densities and basal areas

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 28

will likely benefit many members of this imperiled community. In the aftermath of

Hurricane Katrina, development pressures have increased, and continue to increase, in

coastal Mississippi. As remaining tracts of longleaf pine forests become more and more

fragmented, and forested land continues to be sub-divided and developed, the long-term

persistence of many native longleaf pine species relies heavily on management practices

on public lands. To conserve the wealth of biodiversity represented within the longleaf

pine ecosystem, it is important to recognize the benefits to herpetofauna of lowered stand

densities, and the marked effect these management decisions have on community

structure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 29

CHAPTER III

THE BLACK PINESNAKE (PITUOPHIS MELANOLEUCUS LODINGI): SPATIAL

ECOLOGY AND ASSOCIATIONS BETWEEN HABITAT USE AND PREY

DYNAMICS

A b s t r a c t .— As remaining longleaf pine forests become increasingly

fragmented, wildlife management becomes a daunting task, especially for rare and

relatively unstudied taxa such as the Black Pinesnake (P. m. lodingi). Appropriate

management for P. m. lodingi depends on our understanding of the ecology of this sub­

species; consequently, there are two main objectives of this study: to address questions of

spatial ecology for Black Pinesnakes on a range-wide scale, and to employ small

mammal trapping within home ranges of telemetered snakes to investigate the

relationship between prey density and spatial ecology of these snakes. P. m. lodingi

exhibited large home ranges (Minimum Convex Polygon (MCP) home ranges were 92 -

396 ha), frequently crossed roads, and average (non-zero) movement distance for these

snakes was 338 meters per location event. Core home ranges of telemetered pinesnakes

were characterized by significantly greater abundances of hispid cotton rats ( Sigmodon

hispidus), and significantly higher total captures of hispid cotton rats and cotton mice

(Peromyscus gossypinus). By better defining the relationship between spatial ecology

and prey dynamics for P. m. lodingi, land managers will be better equipped to manage

and conserve remaining Black Pinesnake populations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 30

The longleaf pine savanna ecosystem once stretched along the coastal plain

landscape, nearly unbroken, from southwestern Virginia to eastern Texas (Bartram,

1791). As of 2000, approximately 841,800 hectares of naturally regenerated longleaf

pine stands remain; this figure represents less than 3% of the historic range (Ware et al.,

1993; Frost, 2006). A myriad of factors have contributed to the decline of this fire-

dependent ecosystem including: decimation from logging and the naval stores industry,

conversion into agricultural or silvicultural areas, fire-suppression, invasive species, and,

more recently, habitat fragmentation and development (Outcalt, 1997; Varner and Kush,

2004; Frost, 2006). The historic and continued decline of longleaf pine forests poses a

threat to approximately 72 resident reptile and amphibian species and 30 specialist

species: reptiles and amphibians whose ranges fall completely within the historic range of

longleaf pine forests (Guyer and Bailey, 1993; Means, 2006). The Southern Hognose

Snake (Heterodon simus), Flatwoods Salamander ( Ambystoma cingulatum), and Gopher

Tortoise ( Gopherus polyphemus) are just a few examples of highly imperiled longleaf

pine specialists characterized by declining populations (McCoy and Mushinsky, 1992;

Means et al., 1996; Gibbons et al., 2000; Tuberville et al., 2000). Although it is difficult

to distinguish natural population fluctuations from population declines, many populations

of longleaf pine taxa, such as the Black Pinesnake ( Pituophis melanoleucus lodingi) and

the Mimic Glass Lizard (Ophisaurus mimicus) are suspected to be in decline, but

insufficient historical and natural history data exists to determine geographic extent and

severity of these declines. With the precipitous decline in area of this habitat type,

concern has arisen regarding the dearth of natural history knowledge available for many

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 31

longleaf pine specialists that are apparently declining, especially those that are cryptic or

occur at naturally low densities.

Until recently, very little was known concerning the natural history and spatial

ecology of the Black Pinesnake, a large colubrid longleaf pine specialist whose historic

range spanned from far western Florida to extreme eastern Louisiana (Conant and

Collins, 1991; Paul Moler pers. comm.). Currently, this sub-species is thought to be

extirpated from Louisiana and has not been reported west of the Pearl River for nearly

forty years. C. M. Duran’s (unpubl. data) radio-telemetry study during the late 1990’s,

conducted within a managed gopher tortoise reserve on Camp Shelby (an Army National

Guard training facility, DeSoto National Forest), elucidated habitat preferences and

movement patterns of this taxon and dispelled the belief that Black Pinesnakes primarily

use gopher tortoise burrows as refugia. The average minimum convex polygon (MCP)

home range in Duran’s study (unpubl. data) was 47.5 hectares, and he found that Black

Pinesnakes use pine stumpholes almost exclusively for refugia. Since the only field

research addressing Black Pinesnake ecology and natural history occurred on a managed

gopher tortoise preserve, on habitat far superior to most habitat remaining within the

range of Black Pinesnakes, our first objective was to address questions of spatial ecology

on a range-wide scale, especially in areas characterized by fragmented or sub-optimal

habitat.

As remaining longleaf pine forests become increasingly fragmented, wildlife

management becomes a daunting task, especially for relatively unstudied taxa such as the

Black Pinesnake. Since Black Pinesnakes move long distances (C. M. Duran unpubl.

data) and often fall victim to road mortality (Olsen and Lee, 2007; Baxley and Qualls,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 32

this study), our second objective was to employ small mammal trapping within home

ranges of telemetered Black Pinesnakes to investigate the relationship between prey

density and spatial ecology of these snakes. We hypothesized that core home ranges of

Black Pinesnakes would have greater abundances of small mammals and greater total

mammal biomass than areas within home ranges that were seldom used. By examining

these relationships, we sought to indirectly test one of the basic tenets of ideal free

distribution theory (Fretwell and Lucas, 1970; Fretwell, 1972) which states that predator

distribution should be proportionally tied to prey distributions. If our small mammal

community experiment supports this theory, we should find that black pine snakes spend

more time in areas characterized by “better” (i.e. higher abundance, greater biomass)

small mammal communities. By better defining the spatial ecology of Black Pinesnakes

and the relationship between spatial ecology, small mammal communities, and habitat

use, we will be better able to manage and conserve this unique longleaf pine specialist.

M a t e r ia l s a n d M e t h o d s

Study Area.—Between April 2005 and December 2006, we conducted radio­

telemetry within six counties in South Mississippi: George, Harrison, Jones, Marion,

Perry, and Wayne (Figure 3.1). Four of the six study sites were within DeSoto National

Forest; the remaining two sites, Marion County and George County, are within the

Marion County Wildlife Management Area and on private land respectively. Overstory

vegetation at all sites was dominated by either longleaf pine (Pinus palustris), or a

combination of longleaf pine and slash pine {Pinus elliotti) or loblolly pine ().

All sites have a history of prescribed fire and have been burned at least once within the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 33

past five years. Level of fragmentation varied among sites; for example, the two sites

within the Chickasawhay District of DeSoto National Forest (Jones and Wayne Counties)

were fragmented only by forest service roads, while all of the remaining sites were

bisected by both paved roads and private in-holdings.

Figure 3.1. Locations of radio-tracked Black Pinesnakes in South Mississippi.

Survey and Radio-telemetry Techniques.—Black Pinesnakes were captured using

common herpetological sampling techniques including drift fences with funnel traps

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 34

(Gibbons and Semlitsch, 1981; Campbell and Christman, 1982; Fitch, 1987), manual

searching, and road surveys. Each drift fence (30.5 m long by 1.2 m high) was equipped

with four funnel traps, one at each end on each side. To sample multiple sites within a

local region simultaneously, twelve or more drift fence/trap arrays were deployed at all

times between 15 March and 1 July 2004, 2005 and 2006.

All Black Pinesnakes captured were transported to the University of Southern

Mississippi where we assessed feasibility of radio-transmitter implantation for each

individual (transmitter < 5% of snake mass). Temperature-sensitive radio transmitters

(Holohil Systems, Ltd. SI-2 and AI-2 transmitters) were surgically implanted within the

coelomic cavity of suitably large Black Pinesnakes following the procedures of Reinert

and Cundall (1982); furthermore, we did not perform surgery on snakes captured between

August and February, to avoid late-season surgery-related mortalities (Rudolph et al.,

1998).

In an attempt to increase post-surgury survival, all snakes were released at the

point of capture after being held in the laboratory for a maximum of five days (Hardy and

Greene, 1999). After release, snakes were located every three days during the spring and

summer, and once per week during the fall and winter using a Telonics telemetry receiver

and hand-held antenna. Microhabitat occupied, activity status, body temperature, road

crossing occurrences, and behavioral observations were recorded for each location

episode. Tracked individuals were located at different times of day, to assess the full

range of activity patterns, and care was taken not to disturb the animals when locating

them during active periods (which may bias microhabitat, activity, or behavioral

observations). All snakes were tracked until mortality ensued or until transmitters began

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 35

to fail. When transmitters began to fail, we captured the snake, removed the transmitter,

and released the individual at the point of capture.

Small Mammal Trapping .—From 2 June through 10 July 2006 small mammal

trapping, using Sherman traps, was conducted within the home ranges of four telemetered

Black Pinesnakes. ArcGIS was used to randomly select point locations within the core

home range of each snake (inner 10% useage area) as well as the outer home range of

each snake (outer 10% usage area) to serve as transect center-points. Within each home

range, two sets of two parallel transects (40 total Sherman traps) were simultaneously

employed within the core and outer home range of each snake; transect direction was

east-west for all replicates. We aimed to reduce temporal bias in our analysis of mammal

communities by using this paired survey technique. Transects were used to sample small

mammals because transect arrangements have been shown to yield more total captures,

more individual captures, and more species than grid arrangements (Pearson and

Ruggiero, 2003). Sherman traps were place 10 meters apart, and large (7.62cm x 8.89cm

x 22.86cm) and small (5.08cm x 6.35cm x 16.51cm) Sherman traps were equally

distributed between transects and placed such that no two small traps and no two large

traps were adjacent. We chose to space traps ten meters apart because this is the most

frequent distance for small mammal trap spacing reported in the literature (Bowman et

al., 2001). Traps were closed during the heat of the day, opened and baited with oatmeal

at dusk, and checked shortly after dawn on the following day. Species, total length, and

mass were recorded for each captured mammal, and all individuals were marked via a

unique spot-shaving pattern. Processing time varied from five to fifteen minutes;

subsequently, mammals were released at the point of capture.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 36

Habitat Data.— During the first week of July 2006, habitat data was collected at

the center point for each small mammal transect. The following characteristics were

quantified and recorded for each habitat sampling locality: dominant overstory plant

species, dominant understory plant species, percent canopy cover of trees, percent mid­

story coverage, percent over-story shrub coverage, basal area of trees, percentage cover

of grasses, forbs, and shrubs in understory, ground-cover percentage of litter and bare

soil, and recency of fire. A spherical densiometer (Forest Densiometers model-C) was

used to measure canopy cover. At the center-point of each transect, four densiometer

readings were taken to estimate percent canopy cover facing north, south, east, and west;

these readings were averaged to estimate overall cover. A Cruz-All basiometer (model

59793) was used to measure factor-10 basal area. We used a one-meter quadrat (four

replicates per transect) to estimate average percentage of grasses, forbs, shrubs, litter and

bare soil comprising the groundcover, and we visually estimated the percentage of mid­

story coverage and percentage of shrubs in the mid-story. To increase measurement

consistency, one individual (D. B.) measured all habitat parameters.

Analysis .—We used Hawth’s Analysis Tools for ArcGIS (Beyer, 2004) to

calculate minimum convex polygon (MCP) and fixed kernel home ranges for telemetered

Black Pinesnakes. For fixed kernel home ranges, we adjusted the smoothing factor until

the area of the 95% kernel equaled the area of the MCP, as recommended by Row and

Blouin-Demers (2006). As a result of low sample sizes and high mortality of Black

Pinesnakes, we did not statistically compare differences in home ranges and movement

activity between male and female individuals. Although given the paucity of published

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 37

information concerning this taxon, we feel that a descriptive report of these values is

warranted.

Since we hypothesized that biomass and abundance of small mammals would be

greater in the inner versus the outer home ranges of telemetered Black Pinesnakes, we

used one-tailed Wilcoxon signed-rank tests (alpha = 0.05) to assess these differences.

For our purposes, abundance was defined as mean number of captures per species per

day, and we used only non-recaptured mammals in our analysis (i.e., each mammal was

counted only once, even if re-captured several times). We also used a non-parametric

Wilcoxon test (two-tailed) to assess differences in total mammal captures between large

and small Sherman traps

To assess differences in habitat between core and outer small mammal transects,

we conducted non-parametric Wilcoxon signed-rank tests or parametric matched-pairs

tests (depending upon whether habitat parameters met assumptions of normality and

equal variances). To decrease the probability of Type I error, we applied a Bonferroni

correction (Miller, 1981) resulting in an alpha level of 0.005. We used JMP version 5.1

(2003) for all statistical analyses.

R e s u l t s

Survey results.—Between 2004 and 2006, we acquired twenty-three new Black

Pinesnake records. Using drift fences and funnel traps, we captured six Black Pinesnakes

over the course of 4,088 trap days (1 per 688 trap days). Five Black Pinesnakes were

hand-captured, and road cruising resulted in twelve new records (five live and seven dead

on road, DOR, Black Pinesnakes). For all methods and all years, May and June were the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 38

two most productive months for detecting Black Pinesnakes, presumably as a result of

breeding activity.

Radiotelemetry.—After implanting radio-transmitters into eight suitably large

adults, we tracked snakes between April 2005 and December 2006 for a total of 463

location events (Table 3.1).

Table 3.1. Tracking and home range data for five Black Pinesnakes (Pituophis melanoleucus lodingi) radio-tracked during 2005 and 2006.

#Days # Location Kernel (ha) Snake ID Sex Tracked Events MCP(ha) 50% 90%

1 F 401 101 196.99 27.9 142.4

2 M 136 29 182.61 45.2 145.9

3 M 381 85 395.98 62.1 289.2

4 M 561 114 91.68 18.4 64.3

5 F 487 118 171.72 24.2 116.4

Roads did not pose a barrier to movement; 18.6% of the time, Black Pinesnakes

crossed roads between location events, and two of eight Black Pinesnakes fitted with

radio-transmitters were found dead on paved roads. Black Pinesnakes predominantly

used rotted pine stump holes for refuge, and seldomly frequented abandoned gopher

tortoise burrows and other types of refugia (Figure 3.2). Telemetered individuals were

found above ground only 29.6% of the time. In our study, snakes were seen fleeing into

refugia 5.8% of the time, and snakes exhibited a decrease in body temperature during

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 39

13.8% of all location attempts, indicating a change in body position within the

environment (e.g. snake fled from observer). When all location events are included in

analysis, average movement distance per location event is 214.46 m.

Blackberry thicket Non-stump hole Unidentifiable hole c o Brush Pile oC3 j Armadillo Burrow

Open

# of Location Events

Figure 3.2. Radio-telemetric locations of 8 Black Pinesnakes, Pituophis melanoleucus lodingi, tracked during 2005 and 2006.

When only non-zero movement distances are included in analysis, average movement

distance per location event jumps to 337.98 meters; snakes exhibited greatest movement

activity during the month of June (Figure 3.3). It is of note that snakes moved shorter

distances during the severe drought of 2006 when compared to 2005 (Figure 3.3).

Black Pinesnakes did not appear to exhibit true hibernation behavior and were

seen above-ground, or found to have moved to different refugia, on warm days

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 40

throughout all months of the year. The longest movement over a three day period by any

snake was 2,050 meters. Snakes appeared to exhibit some degree of fidelity to stump-

holes; 6.5% of the time snakes moved to a previously occupied stump-hole. Five snakes

tracked for more than 135 days exhibited large minimum convex polygon (MCP) home

2 c 400

= 300

S 200

IE 100

loco coidlo inm cn iococo cococo coco to cococo co oo ooo ooo ooo ooo ooo ooo o rS® OXDCD

Month

Figure 3.3. Mean monthly movement distances (meters) of Black Pinesnakes, Pituophis melanoleucus lodingi, in South Mississippi.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 41

ranges; estimates ranged from 92 to 396 hectares, much larger than the 47.5 hectare

(average MCP home range) previously reported for Black Pinesnakes on Camp Shelby

(C. M. Duran unpubl. data).

Black Pinesnakes in this study exhibited extremely high mortality rates; of eight

snakes implanted with transmitters, only two survived to the end of the study. Two

snakes were found dead on paved roads, one snake was apparently predated, one snake

died from transmitter migration, and the remaining two snakes were found dead,

unpredated, above-ground, during the height of the severe drought of 2006. Both snakes

found dead during the drought were emaciated, and necropsy of these two individuals

revealed no transmitter complications. One of these snakes, a female, crawled to a

seasonally flooded bottomland area (which was dry at the time of the drought) and

remained there for nearly a month before she eventually died.

Small Mammal Trapping .—Over the course of 1440 trap days, small mammal

trapping resulted in 179 captures of 64 individuals representing six species (Figure 3.4).

Wilcoxon signed rank tests (one-tailed) indicated a greater abundance of hispid cotton

rats (Paired T-ratio = 2.25, p = 0.022) in core home ranges as well as higher total captures

of hispid cotton rats and cotton mice in core home ranges of telemetered Pinesnakes

(Paired T-ratio= -2.76, p = 0.005). Total biomass was higher in the core home range,

though this difference was not statistically significant (Paired T-ratio = -1.77,/) = 0.059).

Furthermore, significantly more mammals were captured in large Sherman traps (x2 =

7.38, p = 0.007) when compared to small Sherman traps.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 42

Peromyxus gossypnus

Neotoma floridana

vi Sigmodon hispid us

znI Ochrotomjs nuttalli

Didelphis marsupialis

Blarina bre vicauda 1------1------l------1------1------| Core Home Range 0 5 10 15 2 0 2 5 3 0 Number of Individuals H f Outer H ome Range

Figure 3.4. Species and number of small mammal individuals captured in core versus outer home ranges of Black Pinesnakes.

Habitat Analysis .—Small mammal transects within core home ranges of

telemetered Black Pinesnakes were characterized by significantly less canopy cover, and

less forbs, litter, and bare soil in the groundcover (Table 3.2). Transects within core

home ranges had higher percentages of grass comprising the groundcover, had been

burned more recently, and had lower percentages of midstory and shrub coverage, though

these differences were not statistically significant.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 43

Table 3.2. Means, standard deviations, test statistics, and p-values for habitat variables collected from mammal transects within core (I) versus outer (O) home ranges of black Pinesnakes (matched pair analysis). Habitat variables followed by an asterisk did not meet parametric assumptions and were analyzed using Wilcoxon Matched Pairs Signed Rank tests.

Mean SD Variable I 0 I O Test Statistic P

% Canopy Cover 41.0 67.2 17.2 19.5 4.27 0.0007

Bum recency (mo) 22.5 40.5 9.9 13.3 1.88 0.16

% Shrubs in midstory 48.8 70.0 22.1 14.1 3.83 0.03

% Midstory coverage 14.3 40.0 9.4 32.4 1.83 0.16

% Forbs* 5.5 0.1 5.5 0.3 -45.5 0.0001

% Grass* 11.8 4.8 15.3 7.1 -21.5 0.14

% Litter* 88.8 98.1 8.5 2.5 43.5 0.001

% Bare soil* 4.7 0.6 5.2 1.4 -39.0 0.0001

% Shrubs 29.8 45.4 19.1 23.9 1.97 0.07

DISCUSSION

The results of our study highlight many similarities and a few striking differences

between the Black Pinesnake and its eastern and central congeners, the Northern

Pinesnake ( Pituophis melanoleucus melanoleucus), Louisiana Pinesnake {Pituophis

ruthveni) and Florida Pinesnake {Pituophis melanoleucus mugitus). Average movement

distances and minimum convex polygon home ranges of Black Pinesnakes in this study

are larger than, though still comparable to, those previously reported for eastern and

central members of the genus Pituophis (Franz, 1986; Himes and Hardy, 2002; Gerald et

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 44

al., 2006); all of these snakes are known to inhabit large home ranges and move long

distances. Percent of location events where snakes returned to a previously occupied

retreat for Black Pinesnakes (6.5%) was similar to values recorded for Northern

Pinesnakes (7.2%) (Gerald et al., 2006); this potentially reflects both the importance of

these retreat sites, and existence of some level of fidelity to refugia for both of these sub­

species.

Perhaps one of the most notable ecological difference between the Black

Pinesnake and its eastern and central congeners is that the black pine is not known to

excavate nests, hibemacula, or pocket gopher burrows (Burger et al., 1988; Burger and

Zappalorti, 1991; Himes, 2000; Franz, 2001). Furthermore, the Black Pinesnake inhabits

a unique ecological niche because, unlike the Louisiana Pinesnake and the Florida

Pinesnake, the Black Pinesnake is not sympatric with pocket gophers (Geomys breviceps

shares its range with the Louisiana Pinesnake and Geomys pinetis shares its range with

the Florida Pinesnake) (Himes, 2000; Franz, 2001). Given the absence of pocket gophers

within the range of the Black Pinesnake, alternative food sources, hibemacula, and

refugia become increasingly important. As most old-growth longleaf pine stumps were

removed and utilized in the Naval Stores industry, and gopher tortoise burrows are no

longer prominent features of the landscape, the continued removal of longleaf pine

stumps during intensive site preparation and salvage logging operations poses a serious

threat to this taxon. It is very possible that presence and distribution of pine stump holes

may be a feature that limits within-range abundance of Black Pinesnakes.

The disturbingly high mortality of Black Pinesnakes in this study and Louisiana

Pinesnakes in a previous study (19 mortalities of 38 telemetered snakes, Himes et al.,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2002) suggests that population declines may be due in part to adult mortality in excess of

annual recruitment. Given the high frequency with which Black Pinesnakes cross, and

suffer mortality from, roads, the boom in post-Hurricane Katrina development of the

Mississippi Gulf Coast does not bode well for already-fragmented Black Pinesnake

populations. Furthermore, the slow-moving nature of Pinesnakes increases their

susceptibility to vehicular mortality, especially in areas where intentional killing prevails

because of the negative attitudes toward all snake species. Two of the snakes tracked in

this study died during the severe drought of 2006. Although we cannot conclusively say

that these mortality events are a result of drought-related stress, we can recognize the

potential for drought-related population fluctuations, a phenomenon that is completely

unstudied in upland snakes.

The results of our paired small mammal experiment fit within the context of ideal

free distribution theory, and hold important management and conservation implications.

Corroborating our hypothesis, core home ranges of four telemetered Black Pinesnakes

had higher mammal abundance, especially of hispid cotton rats, and greater mammal

biomass than little-used areas along the periphery of these snake’s home ranges.

Although there have been no previous studies using telemetry and small mammal

trapping to assess potentially prey-driven habitat use within the genus Pituophis, a recent

study examining the relationship between edge habitat use and small mammal abundance

in Elaphe obsoleta and Coluber constrictor found no evidence that the preferential use of

edge habitats by both snakes was a consequence of small mammal abundance (Carfagno

et al., 2006). Carfagno et al. (2006) further postulated that prey distribution is less

important than thermoregulatory reasons for generalist predators, especially in higher

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 46

latitudes. This hypothesis is supported by findings that Elaphe in Ontario, Canada prefer

edge habitats regardless of the fact that these habitats do not contain higher abundances

of small mammals (Blouin-Demers and Weatherhead, 2001). In contrast, Madsen and

Shine (1996) and Heard et al. (2004) found evidence for prey-driven habitat use of

Australian pythons ( Liasis fuscus and Morelia spilota metcalfei).

It is impossible to separate causation from correlation regarding the preferential

use of open, herbaceous habitats, which also contain higher abundances of hispid cotton

rats, by Black Pinesnakes. Both taxa have been previously shown to be denizens of open,

grassy areas. Our findings of higher percentages of forbs within core home range

mammal transects (P = 0.0001) is not surprising given the heavy reliance of Hispid

cotton rats on herbaceous ground-cover for refugia and food (Goertz, 1964; Kincaid and

Camereon, 1982; Duran, 1994). Further, bluestem grass ( Schizachyrium spp.), the

historically dominant grass of western longleaf pine ecosystems, is thought to be a major

food item for this taxon (Miller, 1999). Duran (1998) found that Black Pinesnakes

significantly avoided areas of high canopy closure, and were found more often in areas

with high percentages of herbaceous vegetation. Regardless of whether prey abundance

drives (or is correlated with) habitat use by these snakes, our findings hold important

management implications, and indicate that habitat preferences of P. m. lodingi and

Sigmodon hispidus tend strongly toward the historic longleaf pine ecosystem, and away

from the current status quo of densely planted silvicultural land. Although efforts to

increase the use of growing season prescribed burning on public land have increased over

the past few years, dormant season burning of longleaf pine forests in Mississippi is still

the norm. In the aftermath of Hurricane Katrina, problems characteristic of growing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 47

season bums (liability, smoke, etc.) have been exacerbated due to the resulting fuel loads

from the storm. By prioritizing management resulting in open, grassy, forb-rich, longleaf

pine ecosystems (growing season bums, lower planting densities) populations of Black

Pinesnakes, and other longleaf pine specialists will likely benefit.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 48

CHAPTER IV

ASSESSING SECOND AND THIRD ORDER HABITAT SELECTION BY BLACK

PINESNAKES {PITUOPHIS MELANOLEUCUS

LODINGI): A BOOLEAN MODEL USING GEOGRAPHIC INFORMATION

SYSTEMS

Abstract.—With the historic and continued fragmentation and degradation of longleaf

pine forests, an increasing need emerges to identify remaining Black Pinesnake

populations and understand species-environment correlations such that conservation

efforts for this taxon are both directed and appropriate. This study has three main

objectives: 1) To use radio-telemetry data in conjunction with GIS modeling to assess

second and third order resource selection (habitat suitability model) for Black Pinesnakes

west of the Mobile River; 2) to use occurrence data (historic and recent) to validate our

habitat suitability model; and 3) to quantify on-the-ground habitat characteristics of Black

Pinesnake locality records and evaluate differences in environmental variables between

historic and recent occurrence records, potentially as a barometer of habitat loss. On the

landscape level, Black Pinesnakes require large tracts of evergreen forests (historically

Pinus palustris) in upland areas lacking cultivated crops, pasture/hay fields, development,

and roads. Within home ranges, Black Pinesnakes require some degree of patchiness and

habitat heterogeneity; snakes were found closer to scrub/shrub habitat, open areas, and

deciduous forests than expected. Our Boolean model reduced the area deemed suitable

by 84.3%, while accurately categorizing 75% of recent Black Pinesnake occurrences

(post-1987), demonstrating the ability to develop robust models with very small sample

sizes. We found significant habitat differences between sites of recent and historic P. m.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 49

lodingi occurrence, which corroborate the findings of our Boolean model. Recent sites of

occurrence were, on average, 597 m from areas with a model score of “optimal,” while

sites with no recent records were, on average, 1582 m from areas with an “optimal”

score; furthermore, recent sites were characterized by significantly less canopy cover,

lower basal area, less midstory cover, greater percentages of grass, bare soil, and forbs in

the groundcover, less shrubs and litter in the groundcover and a more recent bum history

than sites with no recent records. These data suggest that areas containing historic Black

Pinesnake populations are simultaneously becoming unsuitable on both the landscape and

microhabitat levels. The Boolean model approach we utilized is easy to understand and

update, and can be used to develop testable hypotheses of changes to Black Pinesnake

populations resulting from changes in the landscape.

INRODUCTION

Recent advances in Geographic Information Systems (GIS) technology have

provided conservation biologists with much-needed tools to assess habitat selection and

clarify range distributions of vertebrates. Prior to the widespread adoption of GIS,

vertebrate distribution maps and indices of habitat selection were most often constructed

using historic locality records, which are inherently biased and unreliable (Freitag et al.,

1998; Reddy and Davalos, 2003). The problems of biased locality data and false

absences are exacerbated for snakes and other taxa that are found in inaccessible habitats,

are difficult to detect, or are characterized by cryptic behaviors (Rodda, 1993; Kery,

2002). Over the past two decades, GIS has been widely used to predict occurrences of or

generate species-environment suitability models for birds and large mammals (e.g. Dubuc

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 50

et al., 1990; Clark et al., 1993; Knick and Dyer, 1997; Corsi et al., 1999). In contrast,

herpetofauna are under-represented in the GIS-modeling literature, presumably because

reptiles and amphibians typically move less and inhabit smaller home ranges than large

mammals and birds, thus increasing the difficulty of landscape-level habitat modeling.

Despite the potential difficulty of applying these techniques to herpetofauna, species-

specific modeling for reptiles and amphibians is becoming more common (Lipps, 2005;

Browning et al., 2005; Dayton and Fitzgerald, 2006; Rouse, 2006); however, most of

these studies address questions at fine scales (e.g. timber rattlesnake hibemacula within

Madison County Wildlife Management Area; terrestrial salamander abundance within

Hoosier National Forest).

Herpetological modeling studies on a broad scale are feasible only for species that

have large home ranges, move long distances, or have extremely specialized species-

environment associations. For example, Baskaran et al. (2005) utilized the known

association of Gopher Tortoises ( Gopherus polyphemus) with specific soil and land cover

types to create a GIS model which successfully predicted the occurrence of Gopherus

polyphemus burrows over a five-county region in Georgia. Although most North

American snake species do not occupy sufficiently large home ranges on which to focus

broad-scale GIS modeling efforts, one notable exception is the Black Pinesnake

(.Pituophis melanoleucus lodingi). P. m. lodingi move long distances, inhabit large (up to

396 ha) home ranges (Duran, 1998; Baxley and Qualls, in prep), and is considered one of

the most imperiled snakes in North America.

Currently a federal candidate for Endangered Species Act listing, the Black

Pinesnake once ranged from western Florida to eastern Louisiana along the Gulf Coast

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 51

Plain, but is now extirpated from Louisiana and exhibiting trends of decline in

Mississippi and Alabama. Guyer and Bailey (1993) list the Black Pinesnake as a

specialist of the longleaf pine ecosystem, an ecosystem that has been reduced to less than

five percent of its historic range due to fire suppression and conversion of these forests to

agricultural, silvicultural, or human use areas (Frost 1993; Outcalt 1997; Varner and

Kush 2004). With the historic and continued fragmentation and degradation of longleaf

pine forests, an increasing need emerges to identify remaining Black Pinesnake

populations and understand species-environment correlations such that conservation

efforts for this taxon are both directed and appropriate. This study has three main

objectives: 1) To use radio-telemetry data in conjunction with GIS modeling to assess

second and third order resource selection, and create a habitat suitability model for Black

Pinesnakes west of the Mobile River; 2) to use occurrence data (historic and recent) to

validate our habitat suitability model; and 3) to quantify on-the-ground habitat

characteristics of Black Pinesnake locality records and evaluate differences in

environmental variables between historic and recent occurrence records, potentially as a

barometer of habitat loss. In this study, we follow the selection hierarchy presented by

Johnson (1980) that defines second order selection as the home range of an individual

and third order selection as the usage of different habitat components within the home

range of an individual. With the overall efforts to conserve species in the United States

viewed as opportunistic and uncoordinated (Ando et al., 1998), we hope the results of this

study will provide tools to organize and implement both directed and appropriate

conservation actions aimed at restoring Black Pinesnake populations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 52

M a t e r ia l s a n d M e t h o d s

Radio-telemetry

Between March and July 2004-2006, we surveyed for Black Pinesnakes in South

Mississippi using standard herpetological techniques including drift fences with funnel

traps (Gibbons and Semlitsch, 1981; Campbell and Christman, 1982; Fitch 1987), manual

searching, and road surveys. We implanted radio-transmitters into and subsequently

radio-tracked eight individuals (see Baxley and Qualls, in prep for further methods).

From April, 2005 through December, 2006, snakes were located every three days during

the spring and summer, and once per week during the fall and winter. Home ranges of

radio-tracked individuals fell within six counties in South Mississippi: George, Harrison,

Jones, Marion, Perry, and Wayne. Due to high mortality, only five of the eight radio­

tracked Black Pinesnakes were tracked for sufficient periods (> 130 days) to warrant

inclusion in the habitat model.

Selection o f Environmental Variables

Several different approaches have been used to determine the species-

environment requirements of vertebrates for the purpose of discriminating between

environmental variables for use in habitat models. Deductive approaches use a-priori

species requirements and expert opinion to select habitat parameters of importance, while

inductive approaches derive an “ecological signature” from the characterization and

statistical testing of animal location data (Corsi et al., 2000). As model performance may

be reduced by the inclusion of expert opinion (Clevenger et al., 2002; Seone, et al.,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 53

2005) we chose an inductive approach, using data collected in the field to drive the model

development, instead of relying on a-priori assumptions of suitable habitat.

To assess second order resource selection by Black Pinesnakes, we first generated

random home ranges in ArcGIS (Version 9. Redlands, CA: ESRI, Inc; 2004) using

Hawth’s Tools (Beyer, H. L. 2004. Hawth’s Analysis Tools for ArcGIS. Available at

http://www.spatialecology.com/htools) such that each random home range was equal in

area to the average minimum convex polygon (MCP) home range of telemetered Black

Pinesnakes (radius = 812.3 m). We chose to use randomly generated home ranges to

represent habitat availability because landscape proportions are not equal to proportions

of habitats in random home ranges; consequently, inclusion of the whole landscape as

“available” results in high type I error rates (Katnik and Wielgus, 2005).

After generating random home ranges in ArcGIS 9, we used land use / land cover

(LULC) layers (30 meter pixels, USGS National Land Cover Dataset, 2001, Homer, et al.

2004), of fifteen environmental variables as well as a roads layer obtained from the U. S.

Census Bureau (2000; Tiger/line file) as the starting point for environmental variable

selection (Table 4.1). After using ArcGIS 9 to calculate percent composition, within each

real and simulated home range, for each variable, we employed Analysis of Similarity

(ANOSIM, a Monte Carlo re-sampling technique) to test the null hypothesis that the

compositional proportion of each variable is equal across simulated and real home

ranges. The Black Pinesnake is an upland longleaf pine specialist, and there is no doubt

that well drained soils are important to Black Pinesnake distribution. Surprisingly, soil

analysis of sites for which recent (2004-2007) black Pinesnake records exist revealed

large variances in soil composition (range of percent sand 5-95%, range of percent silt 1-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 54

58%, range of percent clay 0.5-37%; soil hydrometer method Sheldrick and Wang, 1993).

This high variance in soil composition leads us to believe that habitat structure may

trump soils (as the limiting factor for Black Pinesnakes) in many fragmented or marginal

areas whereP. m. lodingi populations have managed to persist. In order to err on the side

of suitable habitat overestimation instead of potentially eliminating areas of suitable

habitat characterized by marginal soils, we chose not to incorporate a soils layer into our

model. Furthermore, unlike its congeners, P. m. lodingi is not known to excavate

burrows for nesting or hibemacula (like P.ruthveni, and P. m. melanoleucus; Burger et al,

1988; Burger and Zappalorti, 1991; Himes, 2000), nor does it feed on Geomys sp., pocket

gophers, (like P. ruthveni and P. m. mugitus; Himes, 2000; Franz, 2001); these life

history traits serve to reduce the probability of strict “priority” (McDearman, 2005) soil

association for P. m. lodingi.

To identify variables of importance for third order habitat selection, we used

Euclidean distance analysis following the methods of Conner et al. (2003). Within each

home range, we generated 1000 random points. For each of these random points, and for

each radio-telemetry point, we measured mimiumum distances to each land use/land

cover feature. Habitat variables that were not present in any home range, and were, on

average, greater than 800 m from telemetry points (e.g. open water made up 0% of total

home range area for all snakes; mean distance from telemetry points was 1883 m) were

eliminated from the distance analysis, because selection at this level is considered 2nd

order selection. We used a Hotelling’s T-test to test the null hypothesis that the resulting

vector of mean distances did not differ from a vector of 1 ’s; if the vector of mean

distances differs from a vector of l ’s, habitat selection exists (see Conner et al., 2003 for

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 55

a complete review of this method). After using MANOVA to establish overall habitat

selection, we tested each of the remaining variables for statistical significance using

Wilcoxon Matched-Pairs Signed Rank tests. By using matched-pairs tests, and removing

among-home range variation, we were confident that we were assessing within-home

range differences in resource selection/use by snakes.

Habitat Modeling and Validation

After identifying the environmental associations of Black Pinesnakes on the

landscape level, we constructed a Boolean overlay model incorporating the six habitat

variables found to be significant in our 2nd order analysis. To establish model parameters

for each significant habitat variable, we used the minimum and maximum values

observed for each variable in home ranges of radio-tracked Black Pinesnakes. Using the

focal statistics function of ArcGIS, each 30 m raster cell was assigned a predicted

suitability score based on the number of model parameters satisfied within a radius of

812.3 m. Areas that fulfilled all six model input parameters (Table 4.3) were given a

rank of “6,” areas fulfilling five of six parameters were given a score of “5” and so forth.

Methods of Boolean overlay are widely used in GIS-based approaches of multi-

criteria-decision-making (Hopkins, 1977; Tomlin, 1990; Berry, 1993; Malczewski,

1996), and have proven to be simple yet effective tools for questions at this scale. One

major criticism of habitat suitability models is the general absence of efforts to validate

and ground-truth models after completion. To determine the quality of our model, we

compiled all known locality data (154 records) for P. m. lodingi, and overlaid these

records on our Boolean model. As our model output is essentially a prediction of the

suitability of a raster cell as the center of a hypothetical circular home range, occurrence

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 56

records were considered correctly classified by the model when they were <813 m from

areas of predicted suitable habitat (Boolean score of “6”). We also assessed model

performance by constructing a plot of percent occurrence records versus percent

reduction in land area to determine what Boolean ranking should serve as a threshold for

“suitability” (i.e. the point where number of occurrence records is maximized while

percent land area is minimized). We then visited all locality records west of the Mobile

River (barring those that were inaccessible due to military road closures or gated, private

property) to evaluate on-the-ground habitat and examine microhabitat differences

between recent and historic records. By assessing microhabitat variables at these P. m.

lodingi occurrence sites, we sought to explain some of the missing variation in our model

(i.e. investigate if low quality forests are present at sites where historic, but not recent,

records exist for Black Pinesnakes). When validating our model, we used twenty years

(1987) as a cut-off for “recent” records. Although we realize there are likely some under­

sampled areas where Black Pinesnakes persist, but have not been seen post-1987, we felt

that twenty years was a conservative (albeit subjective) cut-off point by which we could

be relatively confident in the existence of present Black Pinesnake populations.

To assess habitat on a small-scale for each P. m. lodingi occurrence, we quantified

and recorded the following characteristics of each site: dominant vegetation species,

percent canopy cover of trees, percent mid-story coverage, percent shrub coverage, basal

area of trees, percentage cover of grasses, forbs, and shrubs in understory, ground-cover

percentage of litter and bare soil, and recency of fire. We used a spherical densiometer

(Forest Densiometers model-C) to measure canopy cover. At each site, four densiometer

readings were taken to estimate percent canopy cover facing north, south, east, and west

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 57

of the locality point; these readings were averaged to estimate overall cover. A Cruz-All

basiometer (model 59793) was used to measure factor-10 basal area. We used a one-

meter quadrat (four replicates per site) to estimate average percentage of grasses, forbs,

shrubs, litter and bare soil comprising the groundcover, and we visually estimated the

percentage of mid-story coverage and percentage of shrubs in the mid-story. To increase

measurement consistency, one individual (D. B.) measured all habitat parameters. We

used non-parametric Wilcoxon tests (alpha = 0.05) to examine differences between

habitat variables at recent (post-1987) and historic sites of P. m. lodingi occurrence. Sites

with no viable habitat (i.e. 15 sites that were completely urban) were not included in

analysis. We used JMP (Version 6. SAS Institute, Cary, NC, 1989-2005) and PRIMER

5.2.9 (Primer-E Ltd, Roborough, Plymouth, United Kingdom) for statistical analyses.

R e s u l t s

2nd and 3rd Order Selection

In our second order analysis, ANOSIM revealed six habitat variables that differed

significantly between real and simulated home ranges on the landscape level: percent

evergreen forest, percent pasture/hay, percent crops, percent open water, percent

developed, and density of “A3” (secondary) roads (Table 4.1). Home ranges of Black

Pinesnakes were characterized by less open water, pasture/hay, cropland, and roads, and

more evergreen forest than random home ranges. For third order analysis, MANOVA

indicated highly significant overall selection in all four tests (P<0.0001 for Wilk’s

Lambda, Pillai’s Trace, Hotelling Lawley, and Roy’s Max Root tests). Distance analysis

to evaluate 3rd order habitat selection by Black Pinesnakes resulted in significant

selection or avoidance of seven variables (Table 4.2); we eliminated open water,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 58

developed (medium), developed (low), developed (high), barren, roads, and cultivated

crops from analysis since these variables represented 0% total area of all home ranges. P.

m. lodingi exhibited selection “for” shrub/scrub, deciduous forest, and developed (open)

areas within their home ranges and selection “against” woody wetland, pasture/hay,

mixed forest, and emergent wetland. Although 3rd order selection for evergreen forests

was not statistically significant, radio-tracked Black Pinesnakes were found closer to this

habitat class than any other (mean distance = 13.1m). The only other LULC where Black

Pinesnakes were found, on average, within 150 meters was scrub/shrub (mean distance =

141.7 m); mean distances of radio-tracked Black Pinesnakes to all other LULC classes

averaged 3,236 meters.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 59

Table 4.1. Mean percent composition, standard deviation, test statistic, and significance level for each variable tested for compositional differences between real (R) and simulated (S) home ranges. Means and standard deviations were calculated from the entire data set, while, statistical values reflect random 2 sub-sampling (ANOSIM). Instead of percent composition, we used density (m/km ) for the roads data.

Variable Mean SD Test Statistic P R S R S (Global R)

Evergreen forest 68.5 31.4 9.5 22.2 0.52 0.008

Open water 0 1.7 0 3.9 0.48 0.048

Developed (open) 2.8 5.2 1.1 5.9 0.35 0.056

Developed (low) 0 1.4 0 4.0 0.54 0.048

Developed (med) 0 0.5 0 1.9 0 1.000

Developed (high) 0 0.3 0 2.6 0 1.000

Barren 0 0.4 0 1.4 0.24 0.167

Deciduous forest 0.6 5.6 1.0 10.5 -0.06 0.444

Mixed forest 9.3 11.1 5.2 10.6 -0.11 0.873

Shrub/scrub 9.1 11.8 8.3 9.1 -0.11 0.738

Grass/herbaceous 1.0 1.3 2.0 3.5 -0.09 0.841

Pasture/hay 0.3 7.5 0.7 10.3 0.38 0.032

Cultivated crops 0 1.9 0 4.0 0.94 0.008

Woody wetland 8.2 19.0 7.6 21.7 -0.084 0.841

Emergent wetland 0.1 1.1 0.3 5.7 -0.06 1.000

Roads (secondary) 0 97.7 0 242.4 0.31 0.050

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 60

4.2. Mean, test statistic, and p-values of Wilcoxon Sign-Rank tests comparing Euclidean distances of radio-tracked Black Pinesnakes (R) and Euclidean distances of random simulated points (S) to land use/land cover variables.

Variable Mean Distance (m) Test Statistic P R S

Evergreen forest 13.1 15.4 -639.5 0.44

Developed (open) 252.7 360.8 -16000 0.0001

Deciduous forest 3520.6 3588.2 -7760.5 0.004

Mixed forest 199 135.9 22443.5 0.0001

Shrub/scrub 141.7 253.8 -19448 0.0001

Grass/herbaceous 1991.6 2002.5 3404 0.207

Pasture/hay 1093.3 1044.2 6315 0.017

Woody wetland 477.2 352.3 -21129 0.0001

Emergent wetland 2222.5 2069.6 16021 0.0001

Boolean Model

Within ArcGIS 9, we constructed model parameters (Table 4.3) using the results

of our 2nd order analysis (landscape-level selection for evergreen forests and against

cultivated crops, pasture/hay, roads, developed areas, and open water). Each raster cell

within the study area was reclassified according to the number of model parameters met

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 61

within a 812.3 m radius, from a score of 0 (meeting none of the model parameters) to 6

(meeting all of the model parameters).

Table 4.3. Input parameters for Boolean model

Variable Minimum % Maximum %

Evergreen forest 57.17 no upper limit

Open water 0.00 0.00

Cultivated crops 0.00 0.00

Pasture/hay 0.00 1.48

Developed areas 1.78 4.14

Roads 0.00 0.00

Validation

We overlaid a total of 154 historic and recent P. m. lodingi locality points on our

Boolean model. Although only 15.7% of the geographic area within our model was <

813 meters from a pixel with the optimal Boolean score of six, 75% of “recent” black

Pinesnake records (84 of 112) were located <813 m from areas with an optimal suitability

score of six. In effect, the model reduced the area deemed suitable by 84.3%, while

accurately

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 62

*v , *Y> 'f'* . < f V, > J\ ■I * Mf\ * * ■ - i . * ■v *.: y s

■ ; ‘.' V ’> -—“ '•’4-•. ■ •*,, ’ -V *.< i'S; *4' > I M*& *£•' . ' r a ,------4 * V % ' *V V r I Ii ■1 s

.*.• •« ?> « *

** ^.-y j • * .. . ^ c / f'U ■> ( r .. . ■ o c. { v,\ -* t '.> ■ '4 j t ',. ° J i' "4 r ) i

0 25 50 100 Kilometers 1 i i i 1 i i i I

Legend H ♦ <813 m eters from predicted suitable m Predicted Suitable Habitat it >813 m eters from predicted suitable

Figure 4.1. Boolean model representing areas of predicted suitable habitat (Boolean score = 6), for Black Pinesnake. Seventy-five percent of occurrence records were located <813 m of suitable habitat (dark diamonds).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 63

100.0 9 0 .0

8 0 .0 - 7 0 .0 6 0 .0 I ■ % Occurrence records!; 5 0 .0 I - A % Reduction in Area ; 4 0 .0 3 0 .0 20.0 10.0 0.0

Boolean score threshold for habitat suitability

Figure 4.2. Percent reduction of total area classified as suitable habitat and percent of occurrence records correctly classified as being on suitable habitat are graphed against the corresponding Boolean score thresholds. A Boolean score of 6 indicates predicted suitable habitat, while a value of zero indicates furthest from suitable. The graph indicates the trade-off that occurs between the reduction in total area classified as suitable and the accuracy in classifying occurrence records.

categorizing 75% of recent test points. Mean distance of “recent” (post-1987) points of

Black Pinesnake occurrence to optimally suitable habitat (Boolean score = 6) was 597 m,

while mean distance of “historic” records areas with a score of six was 1582 m. In

contrast, random points within the landscape had a mean distance of 2710 m from habitat

with a Boolean score of six. While our model best performs when areas <813 m of cells

scoring six are considered suitable habitat, a plot of percent occurrence records versus

percent reduction in land area finds a Boolean rank of “4” is the threshold for suitability

if distance is not taken into account. Only 35.5% of the study area received a Boolean

score >4, while 71.9% of all occurrence records were found in these areas (Figure 4.2);

targeting these areas will optimize the probability of finding new Black Pinesnake

populations, while minimizing area to be searched.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 64

Habitat Data

Sites of historic P. m. lodingi occurrence with no known recent (post-1987)

records differed from those with known recent occurrences. When comparing recent and

historic sites, recent sites were characterized by significantly less canopy cover, lower

basal area, less midstory cover, greater percentages of grass, bare soil, and forbs in the

groundcover, less shrubs and litter in the groundcover and a more recent bum history

(Table 4. 4). The only habitat parameter that did not differ between recent and historic

sites was percent shrubs in the midstory.

Table 4.4. Summary of habitat variables quantified (mean ± standard deviation) at 131 sites of recent (post-1987) and historic Black Pinesnake occurrence.

Variable Mean ± SD Test Statistic P Pre-1987 Post-1987 Chi2

Canopy Cover (%) 63.3 ±27.6 60.0 ±86.0 10.89 0.001

Bum recency (mo) 73.4 ±46.7 45.5 ±44.5 13.76 0.0002

Basal area (10F) 10.0 ±6.1 6.3 ±3.7 13.56 0.0002

% Shrubs in midstory36.8 ±32.8 30.6 ±29.7 1.20 0.27

% Midstory coverage 49.2 ± 37.0 28.4 ±28.7 8.08 0.0045

% Grass 8.2 ±9.0 21.4 ± 21.7 17.3 0.0001

% Forbs 3.7 ±5.5 6.2 ±8.0 4.48 0.0344

% Shrubs 39.8 ±25.8 27.8 ± 16.5 5.65 0.0175

% Litter 92.3 ± 13.1 73.8 ±27.0 21.62 0.0001

% Bare soil 5.2 ± 12.3 17.0 ±20.6 16.83 0.0001

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 65

D is c u s s io n

Habitat structure and resource availability are two of the most important factors

that affect movement patterns in snakes (Gregory et al., 1987). In this study, resource

selection by Black Pinesnakes was evident both within home ranges of snakes (3rd order)

and on the landscape level (2nd order). On the landscape level, Black Pinesnakes require

large tracts of evergreen forests (historically Pinus palustris) in upland areas lacking

cultivated crops, pasture/hay fields, development, and roads. With the size, growth, and

resource demands of the human population acting as the most important agents of change

in spatial patterns of biodiversity (Vitousek et al., 1997; Sala et al., 2000), it is no surprise

that our model predicted P. m. lodingi occurrence in areas where wide-spread human

disturbance is not yet evident on the landscape level. Although global estimates of

human population predict declines in U.S. population by 2050 (Cohen, 2003), U.S. trends

of migration to warmer and more temperate areas of the country will probably continue.

As human population growth and resulting development in the southeast increases, many

isolated tracts of longleaf pine forest will become functionally extinct because prescribed

fire will no longer be feasible. This impending fragmentation coupled with the notion

that large home range size increases extinction vulnerability (Woodroffe and Ginsberg,

1998) adds to the urgency and importance of conserving and managing large tracts of

relatively undisturbed land, especially areas that buffer existing Black Pinesnake

populations.

Within home ranges, Black Pinesnakes require some degree of patchiness and

habitat heterogeneity; snakes were found closer to scrub/shrub habitat, open areas, and

deciduous forests than expected. At first glance, the within-home range preference for

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. deciduous forest patches seems anomalous, given the fidelity of P. m. lodingi to longleaf

pine forests. Upon closer inspection of this finding, we see that mean distance from

radio-telemetry points to deciduous forest is over 3,500 m; consequently, the statistical

preference for this habitat type is probably caused by one snake spending substantial time

in a disturbed, cut-over area characterized by deciduous trees, which are often the only

trees remaining after longleaf pine forests are harvested. Thermal quality has been shown

to influence habitat use in other snakes (Huey et al., 1989; Row and Blouin-Demers,

1996; Shine and Fitzgerald, 1996; Blouin-Demers and Weatherhead, 2002; Row and

Blouin-Demers, 2006b), so it is likely that these open/patchy areas provide

thermoregulatory opportunities for Black Pinesnakes that are not present in more densely

forested, homogenous areas. It is also of note that the Northern Pinesnake (Pituophis

melanoleucus melanoleucus) is positively associated with open areas, both in Tennessee

and in New Jersey (Gerald et al., 2006b; Burger et al., 1988). These open, patchy areas

likely take on greater importance for the Black Pinesnake in the current landscape due to

the artificially high planting density and basal area of present-day longleaf pine forests

in comparison to the historic landscape (Landers et al., 1995).

Areas containing historic Black Pinesnake populations are simultaneously

becoming unsuitable on both the landscape and microhabitat levels. Our Boolean model

results coupled with the results of on-the-ground habitat analysis give congruent support

to the notion that reduced black Pinesnake populations are primarily due to habitat

fragmentation and habitat alteration. When compared to recent Black Pinesnake

occurrence records, historic (pre-1987) sites of Black Pinesnake occurrence are

consistently farther from areas deemed “suitable” by our model; furthermore, these areas

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 67

are typified by qualities indicative of high planting densities, and fire suppression (closed

canopies, high percentages of leaf litter, low percentages of grass etc.). Given the status

of Black Pinesnakes as longleaf pine specialists, the importance of open-canopied forests

and dense, herbaceous ground-cover can not be over-emphasized. Moreover, the

selection “for” open/patchy areas within home ranges is not so surprising, given the

structural patchiness and high stand variability typical of old-growth longleaf pine forests

(Noel et al., 1998). In a recent study, Waldron et al. (2006) postulated that habitat

specificity to the imperiled longleaf pine ecosystem may be a significant contributor to

the decline of the Eastern Diamondback Rattlesnake; we suspect that this is also true for

the Black Pinesnake.

In terms of model performance, we accurately classified 75% of recent black

Pinesnake occurrence points by using independent home range data to construct model

input parameters. Boolean Models often overestimate the amount of suitable habitat

(Type 2 error; Standora, 2002), and this can be difficult to detect when only presence data

is available to evaluate the model results. However, our model appeared to perform well

in this category also, with <16% of the study area predicted suitable in the best

performing model. A paradox of working with endangered and rare species is that

sample sizes large enough to develop robust estimates of habitat composition are rarely

available. With few exceptions, habitat modeling is rarely attempted with sample sizes as

low as those in the present study (five home ranges). The success of this model serves as

an excellent example of the utility of GIS modeling amidst the constraints of studying

rare taxa.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 68

Two other strengths of our modeling approach in terms of its application to

applied conservation and management are deserving of comment. First, our Boolean

approach utilizes an easily comprehended methodology requiring the end user to have no

expertise in GIS or habitat modeling. Roux et al. (2006) argue that unobstructed

knowledge flow between science and management is one of the great challenges for

sustainable ecosystem management and overly complex models may not be integrated

into policy-making if they are perceived by managers as merely an “intellectual

curiosity.” Second, our model can be easily updated as additional data on Black

Pinesnakes or updated digital data such as LULC and roads become available. Even

without new data, hypothetical data could be utilized to make testable hypotheses

concerning the effects of different land management or conservation plans, or to predict

the impact that future landscape changes may have on populations of the Black

Pinesnake. This forecasting of the anthropogenic effects on patterns of biodiversity is

seen as one of the most important uses of habitat models (Guisan and Thuiller, 2005) and

our model lays the framework for additional study in this area. In summary, we believe

the current model provides both land managers and biologists with a much-needed tool to

assist in directed conservation efforts and further our knowledge of the biology of the

Black Pinesnake.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTERV

RAPID POPULATION EXPANSION AND RECENT DIVERGENCE OF EASTERN

PITUOPHIS: P. M. LOD1NGI, P. M. MELANOLEUCUS, AND P. M. MUGITUS

A b s t r a c t .—Systematics of the genus Pituophis have been described by several

phylogenetic hypotheses, and the intraclade relationships of eastern Pituophis: P. m.

melanoleucus, P. m. lodingi, and P. m. mugitus are particularly unstudied and unresolved

This study has three main objectives: 1) to expand upon Rodriguez-Robles and Jesus-

Escobar’s phylogenetic study through additional sampling of wild caught individuals of

the three easternPituophis sub-species from across their ranges; 2) to employ an

additional mitochondrial gene to see if we can obtain better phylogenetic resolution of the

three sub-species; and 3) to use molecular data to make inferences about demographic

history. Phylogenetic analysis using Maximum Parsimony, Maximum Likelihood, and

Bayesian methods did not recover Pituophis sub-species as monophyletic groups.

Analysis of pair-wise sequence differences resulted in a distribution not significantly

different from the rapid expansion model (P = 0.57), and we estimate rapid population

expansion as occurring between 63,125 and 1,518,174 years BP. Given the high degree

of morphological and ecological dissimilarities between P. m. mugitus, P. m. lodingi, and

P. m. melanoleucus, future genetic work using microsatellite loci will likely reveal

increased geographic structure within this clade.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 70

Introduction

Relationships within the transcontinental genus Pituophis have long been debated

and are characterized by several taxonomic hypotheses (see Rodriguez-Robles and Jesus-

Escobar, 2000 for a complete review). Recent systematic analysis of new world gopher,

bull, and Pinesnakes indicated that the eastern Pinesnake clade (P. m. lodingi, P. m.

mugitus, and P. m. melanoleucus) and the western Pinesnake clade (P. c. affinis, P. c.

annectens, P. c. bimaris, P. c. catenifer, P. c. deserticola, P. c. sayi, P. c. vertebralis)

represent divergent lineages and should be recognized as two distinct species (Rodriguez-

Robles and Jesus-Escobar, 2000). In their study, monophyly of the eastern Pinesnake

clade was well supported (MP bootstrap value = 100%); however, within-clade

relationships for this group were characterized by low bootstrap values, and were

relatively unresolved. Particularly unresolved in the Rodriguez-Robles and Jesus-

Escobar study (2000) was the monophyly of the black Pinesnake {Pituophis

melanoleucus lodingi) group (MP Bootstrap value = 56%). Aside from this study, to our

knowledge there is no other information concerning molecular systematics of the eastern

Pinesnake clade as a whole, and the black Pinesnake in particular. Non-molecular work,

a phenetic analysis (using morphological variables) of eight North American subspecies

of Pituophis conducted by Reichling (1995), found that canonical coefficients for eastern

Pinesnakes clustered independently from those of both western and Louisiana

Pinesnakes. Although Reichling did not test differences within the eastern Pinesnake

clade for significance, he divided the eastern Pinesnakes into four Operational

Taxonomic Units (OTUs), and found several morphological variables that, when

averaged for each OTU, differed between groups.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 71

The interest in resolving the phylogenetic relationships within Pituophis is

attributable to several factors: the trans-continental distribution of this genus, the unique

and charismatic natural history of many currently recognized sub-species of the genus,

and the growing conservation concern for several members of the Pituophis clade.

Eastern Pinesnakes (P. m. lodingi , P. m. mugitus, and P. m. melanoleucus) are rare and

secretive taxa differing in range distribution (Fig. 5.1), morphology, and ecological niche

(Franz, 1986; Burger et al., 1988; Burger and Zappalorti, 1991; Conant and Collins,

1991; Himes, 2000; Franz, 2001; Himes and Hardy, 2002; Gerald et al., 2006). P. m.

lodingi is currently a federal candidate for Endangered Species Act listing, and several

states have acknowledged declining eastern Pituophis populations. For example, P. m.

lodingi is listed as imperiled in Mississippi (S2 designation), critically imperiled in

Louisiana (SI), and of high conservation concern in Alabama. Additionally, P. m.

melanoleucus is listed as threatened in New Jersey, and P. m. mugitus is listed as rare in

Florida (S3 designation). As habitat of the eastern Pinesnakes continues to be fragmented

and degraded, populations continue to decline, and an increasing need emerges for

resolution of the phylogenetic relationships among these snakes. Despite ongoing

debates concerning the validity of the sub-species rank (see Burbrink et al., 2000 for a

review) and perpetual discordance among scientists regarding species concepts (Mayden,

1997; Coyne and Orr, 2004), we feel that attempts to clarify and expand upon knowledge

of molecular variation are warranted, especially for taxa of conservation concern.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 72

^ / j - P. raffcvenf

- P. m. lo d in g i

■ P. 771. meianoieucus Area of hybridization P. m. mugitus

W l * a /

Figure 5.1. Distribution of eastern Pituophis, adapted from Conant and Collins, 1991.

This study has three main objectives: 1) to expand upon Rodriguez-Robles and Jesus-

Escobar’s study through additional sampling of wild caught individuals of the three

eastern Pituophis sub-species from across their ranges; 2) to employ an additional

mitochondrial gene to see if we can obtain better phylogenetic resolution of the three sub­

species; and 3) to use molecular data to make inferences about demographic history.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 73

M a t e r ia l s a n d M e t h o d s

Molecular techniques.—Total genomic DNA was extracted from scale clips or

sloughed skins of Pituophis using the Qiagen Dneasy extraction kit (QIAGEN Inc.,

Valencia, CA). The polymerase chain reaction was then used to amplify 855 bp of the

NADH dehydrogenase 4 (ND4) gene using two sets of primers developed specifically for

Pituophis based on sequences from GenBank (5’-CTAAAACTGGGRGGATACGG-3’

and 5 ’-CTATTATCCTTTAGTAGAGGG-3 ’ and 5 ’ -TC AATTTT ACT A A AACT GGG-3 ’

and 5 ’ -GAGTT AGC AGGTCTT ATGGC-3 ’). We also amplified 517 bp of the NADH

dehydrogenase 2 gene (ND2) using primers (H3518 and H5382) from deQueiroz et al.

(2002), one primer (L4437b) from Kumazawa et al (1996) and one primer (16Sb) from

Palumbi (1996). PCR cycling conditions consisted of an initial 1 min denaturing step at

95°C followed by 30 cycles of 1 min at 95°C, 1 min at 50°C and 3 min at 72°C. A final

elongation step of 7 min at 72°C completed the cycle. PCR products were cleaned using

the ExoSAP-IT system (USB Co., Cleveland, OH) and then used as template in a cycle

sequencing reaction with ABI BigDye™ Terminator cycle sequencing kit (Perkin Elmer

Applied Biosystems, Foster City, CA). All sequencing reactions were sephadex cleaned

(Princeton Separations, Adelphia, NJ) prior to gel runs by the Iowa State University DNA

Sequencing and Synthesis Facility.

Phylogenetic Analysis. — We obtained sequence data from Rodriguez-Robles and

Jesus-Escobar’s systematic study from GenBank (Table 5.1). We aligned GenBank

Sequences to our Pituophis sequences using Sequencher v. 4.1 (GeneCodes Co.,

Madison, WI); see Table 5.2 for all samples used in this study. In order to compare our

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 74

sequences to those obtained via GenBank, we truncated ND4 sequences to achieve a

length of 808 base pairs.

Table 5.1. Mitochondrial NADH sequences of Pituophis and Lampropeltis obtained from GenBank. N represents the number of sequences for each species

Species N GenBank Species N GenBank accession # accession #

L. getula 1 AF138759 P. m. melanoleucus 3 AF138770 AF141116- AF141117

P. m. mugitus 2 AF138769- P. m. lodingi 2 AF141114- AF141115 AF141118

P. eaten ifer 1 AF141106

We inferred phylogenetic relationships with PAUP* v.4.0bl0 (Swofford, 2002)

using the methods of maximum parsimony (MP) and maximum likelihood (ML), and a

Bayesian inference of the phylogeny was performed with Mr. Bayes v. 3.1 (Ronquist and

Huelsenbeck, 2003). For ML analyses, we inferred a suitable model of sequence

evolution using MODEL-TEST 3.06 (Posada and Crandall, 1998). Lampropeltis getula

was used as an outgroup in our ND4 phylogenetic analyses because this was one of the

outgroups used by Rodriguez-Robles and Jesus-Escobar, and Lampropeltis has been

previously been shown to be closely related to Pituophis (Rodriguez-Robles and Jesus-

Escobar, 1999). For analysis of our combined ND4 and ND2 data set (which excludes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 75

sequences used in Rodriguez-Robles and Jesus-Escobar, 2000), we used one wild-caught

Pituophis catenifer as an outgroup.

Table 5.2. Number and location of Pituophis samples used in this study.

Taxon # Samples State

P. m. lodingi 26 Mississippi

P. m. mugitus 3 Georgia

P. m. mugitus 2 Florida

P. m. melanoleucus 2 North Carolina

P. m. melanoleucus 9 New Jersey

P. m. melanoleucus 2 Alabama

P. catenifer 2 California

Phylogenetic support was assessed through bootstrapping (Felsenstein, 1985) with 1,000

rounds of resampling for the MP and ML analyses and through posterior probabilities for

the Bayesian analysis. To assess geographic structure, we used 1315 bp of combined

ND2 and ND4 sequences (from 36 individuals) and 808 bp of ND4 sequence (from 43

individuals) to construct mismatch distributions, conduct Analysis of Molecular Variance

(AMOVA), and estimate haplotype diversity (h) and nucleotide diversity (%) using

Arlequin version 2.001 (Schneider et al., 2000). We used models of Rogers and

Harpending (1992) to determine the probability of rapid population expansion in eastern

Pituophis and to estimate the time since expansion. For the mismatch distribution, we

used 3.5 years as the average generation time for Pituophis (Fitch, 1970), and we selected

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 76

1,4%/MY as the mutation rate, a conservative value similar to previously reported

herpetofaunal mutation rates for ND4 and cytB (Carranza et al., 2006; Castoe et al.,

2007).

R e s u l t s

All phylogenetic methods resulted in the recovery of basically the same topology

(Figure 5.2). We found little difference between analysis of ND4 sequences alone and

analysis of the combined ND2 and ND4 data set; consequently, we display only the

results of the ND4 analysis since these are directly comparable to the previous analysis of

Rodriguez-Robles and Jesus-Escobar (2000). Analysis of 808 bp of sequence from the

ND4 gene for 44 individuals of Pituophis melanoleucus resulted in fifteen unique

haplotypes, and maximum divergence between haplotypes was 14 bp, while analysis of

1315 bp of combined ND4 and ND2 sequence resulted in twelve unique haplotypes and

maximum divergence between haplotypes was 7 bp (see Table 5.3 for h and n). While

the eastern Pituophis clade formed a strongly supported monophyletic group, sub-species

within this clade were not recovered as monophyletic; furthermore, one of the two

weakly supported clades within eastern Pituophis included members of all three sub­

species. There are considerable differences between bootstrap values obtained in

Rodriguez and Escobar’s analysis when compared to the re-analysis with added samples

of eastern Pituophis (Fig. 5.2). Although both analyses indicated very high support for

monophyly of the eastern Pinesnake clade (bootstrap value = 100% for both analyses),

coherent subspecies groups deteriorate with the addition of samples.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 77

P. m. lodingi (1)

i P.m. fcdiRgf(l)*

55/ 54/73 Pm . lodingi (Id) P. m. mugitus (1)

P. m. lodingi (1) *

P. m. lodingi (3)

P. m. melanoleucus (8) *

P. m. melanoleucus (2) P. «.lodingi (4)

P.m. mugitus(1)

P. m. melanoleucus (1) *

P. m, melanoleucus (1) * 80/92/78

P. m. mugitus (2) *

100/ 98/100 P. m. mugitus (1) *

P. catenifer (1)

------p. catenifer(I)*

Lampropeltis getula (1) ’ 0.01 substitutions/sito

Figure 5.2. Maximum liklihood tree from the ND4 data set. Support values are indicated along the branches in the following order: MP bootstrap, ML bootstrap, and Bayesian posterior probability. Number of sequences for each taxa are indicated in parentheses. * indicates a group contains sequence(s) from GenBank.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 78

Table 5.3. Haplotype diversity (h), nucleotide diversity, range of pair-wise distance values (d), number of samples (s), and number of unique haplotypes (u) for Pituophis using ND4 and combined ND2 and ND4 datasets.

71 u

Total

ND4 0.80 +0.05 0.003+0.002 2.25+1.26 43 15

ND4 & 0.81±0.06 0.002+ 0.001 2.15+1.23 35 12 ND2

By Subspecies

P. m. lodingi

ND4 0.60+0.10 0.001±0.001 1.06+0.73 26

N D4& 0.62+0.11 0.001+ 0.001 1.57+0.97 23 ND2

P. m. melanoleucus

ND4 0.56+0.15 0.002+ 0.001 1.64+1.04 12

ND4& 0.56+0.17 0.001+ 0.001 1.39+0.94 ND2

P. m. mugitus

ND4 0.90+0.16 0.007+0.005 5.80+3.34

ND4& 1.00+0.27 0.004+0.003 4.67+3.13 ND2

While Rodriguez-Robles and Jesus-Escobar found 82% support for monophyly of all P.

m. melanoleucus, this relationship is completely unresolved when additional eastern

Pituophis haplotypes are included in the analysis. Little geographic structuring is evident

for eastern Pituophis; furthermore, haplotypes from the northern portion of the range

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 79

appear to be embedded within the matrix of southern haplotypes. AMOVA indicated that

sequence variation was much higher within populations (63% ND4 only; 67% using

combined ND2 and ND4 data set) than between populations (37% ND4 only; 33% using

combined data set). Upon analysis of pair-wise sequence differences, we found that our

distribution was not significantly different from the rapid expansion model (P = 0.54 for

combined ND2 and ND4; P = 0.57 for ND4 only), and we estimate rapid population

expansion as occurring between 63,125 and 1,518,174 years BP (Fig. 5.3).

Mismatch Distribution

300

250

200

I------Observedj 150 I______Si m ula ted^ |

100

50

0 0/ 0-

Figure 5.3. Distribution of pair-wise sequence differences (using 808 bp ND4). Estimate of rapid population expansion: 63,125 - 1,135,938 years BP. Estimate of rapid population expansion for combined ND4 and ND2 data set (1315 bp, not shown) 74,729 -1,518,174 years BP.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 80

D is c u s s io n

The first objective of this study was to re-assess Rodriguez-Robles and Jesus-

Escobar’s phylogenetic hypothesis for eastern Pituophis; unexpectedly, expanded

geographic sampling failed to recover monophyletic groups representing the three eastern

Pituophis sub-species. The lack of mitochondrial evidence for geographic structure

among eastern Pituophis coupled with support for the rapid population expansion model

suggests relatively recent divergence of these three taxa. Given our estimates of recent

population expansion (between 63,125 and 1,518,174 years BP), it is likely that the

distribution of Pituophis populations was limited to southern North America during the

last Pleistocene glacial advance, and underwent subsequent rapid population expansion

from the southeastern United States. Many vertebrate taxa have displayed similar

phylogeographic patterns of vicariant separation during glacial periods, followed by

dispersal to contemporary ranges (Jaarola and Tegelstrom, 1995; Bematchez and Wilson,

1998; Taberlet et al., 1998; Hewitt, 1999). An anomalous outcome of this and Rodriguez

and Escobar’s (2000) study, perhaps lending support to the hypothesis of post-Pleistocene

rapid expansion from the Southeastern United states, is the P. m. mugitus individual that

appears to be basal to the eastern Pituophis clade (Fig. 5.2), differing from all other

eastern Pituophis individuals by 11-14 bp. Given the recent divergence of eastern

Pituophis, we must keep in mind potential effects of incomplete lineage sorting (Avise,

2000), which may obscure our understanding of the phylogenetic history of this clade.

Although our data suggest that these three sub-species do not represent

reciprocally monophyletic groups, by no means should this be viewed as evidence to treat

eastern Pituophis as one management unit. There are many ecological and behavioral

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 81

differences among the eastern Pinesnakes; consequently, we feel that these three forms of

Pituophis inhabit distinct ecological and evolutionary niches, and may represent incipient

species. Oftentimes adaptive differences between populations are not partitioned in a

monophyletic manner, and since the process of speciation produces polyphyletic

relationships, which eventually progress to paraphyletic, and ultimately monophyletic

relationships, Crandall et al. (2000) argues that using criteria of reciprocal monophyly to

define management units for conservation is overly stringent. Furthermore, ecological

differentiation may provide better evidence than molecular data for divergence relevant

to population persistence and conservation (Legge et al., 1996; Crandall et al., 2000).

Reciprocal monophyly is unlikely to be found in vertebrate populations

characterized by high levels of gene flow (Crandall et al., 2000; Avise, 2004). Given the

large home ranges of eastern Pituophis, it is possible that hybridization between

subspecies has occurred with sufficient frequency to blur mitochondrial distinction

between these groups; however, range reductions among all three taxa will likely reduce

contact between sub-species thereby making it imperative to conserve and restore

populations of all three eastern Pituophis. In efforts to preserve biodiversity,

conservation biologists must attempt to maximize evolutionary potential through the

protection of populations on separate evolutionary trajectories (O’Brien and Mayr, 1991;

Hey et al., 2003; Vignieri et al., 2006), even if these populations are not reciprocally

monophyletic. Considering the morphological and ecological dissimilarities between P.

m. mugitus, P. m. lodingi, and P. m. melanoleucus, we recommend future genetic work

using microsatellite loci to further investigate geographic structure and levels of gene

flow within this clade.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 82

CHAPTER VI

GENERAL CONCLUSION AND RECOMMENDATIONS

The broad objective of this dissertation was to utilize multiple research tools (field

surveys, radio-telemetry, GIS technology, molecular techniques) to ultimately provide

land managers and biologists with information necessary to conserve and restore P. m.

lodingi populations. Although this project is predominantly species-specific in focus,

experimental biologists have given recent attention to the efficacy of merging single­

species research with ecosystem-oriented studies to achieve conservation goals amidst

limited resources (Lindenmayer et al., 2007). Given the quantity of vertebrate and

invertebrate species of current conservation concern, it is understandable that species-

specific monitoring and management for all of these taxa is cost-prohibitive. Fortunately,

species-specific research often identifies unique requirements of an organism which can

easily be incorporated into ecosystem management plans. For example, recent evidence

indicating the importance of rotting pine stumps to individual longleaf pine specialists

(e.g. Black Pinesnakes, Eastern Diamondback Rattlesnakes) has resulted in new

management trends aimed at preserving pine stumps, thus benefiting multiple taxa.

We hope that the information presented here regarding black Pinesnake habitat

preferences and spatial ecology is taken into consideration during future habitat

management decisions. Our research supports several management practices that we

believe will be beneficial to Black Pinesnakes, and surely other longleaf pine specialists,

including: planting at lowered stand densities to reduce canopy cover (< 60%),

substituting growing-season prescribed fires for the conventionally used winter bums,

and taking precautions during logging activities such that pine stumps remain in-situ. Of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. these recommendations, the use of growing season bums is especially important to Black

Pinesnakes for two main reasons: 1) growing season bums result in increased herbaceous

vegetation and forbs which are required by prey (Sigmodon hispidus), and 2) these fires

typically bum in an uneven manner resulting in low to moderate levels of landscape

patchiness. Our GIS model and distance analysis of third order habitat selection

demonstrates that low levels of patchiness and heterogeneity are “preferred” by Black

Pinesnakes.

Unless major effort is invested into preventing the development and urbanization

of private lands containing P. m. lodingi populations (i.e. via cooperative landowner

incentives), Black Pinesnakes will likely exist only on public lands within the next half-

century. With this in mind, biologists and land managers need to acknowledge the

growing evidence (Bernardino and Dalrymple 1992; Hels and Buchwald, 2001; Smith

and Dodd, 2003; Andrews and Gibbons, 2005; Aresco, 2005; Steen et al., 2006) detailing

the negative impacts of roads on herpetofauna. Since P. m. lodingi inhabit large home

ranges and move long distances, it is no surprise that road mortality (25% in this study) is

high. As developed areas encroach on public lands, and National Forests and Wildlife

Management Areas continue to be un-buffered and fragmented by new, paved roads, it is

possible that, even on protected land, road mortality for Black Pinesnakes may be in

excess of annual recruitment. Consequently, there is an urgent need for the scientific

community to convey the importance of and foster research into new, innovative, cost-

effective underpass designs such that human growth and ecosystem health are not

mutually exclusive.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 84

Although we found no evidence for reciprocal monophyly of eastern Pituophis

using mitochondrial genes, this by no means provides evidence to support treating

eastern Pituophis as a single management unit. All three eastern Pituophis have suffered

historic range reductions, and areas of potential gene flow between sub-species are

shrinking due to habitat fragmentation and local population extinction. With this in

mind, future genetic work employing additional molecular markers (microsatellites) will

better reveal Pituophis genetic diversity and how this diversity is distributed between and

among Pituophis populations. Further analysis of genetic diversity within the eastern

Pinesnake clade will serve to better inform future conservation decisions, especially in an

era of endangered species mitigation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 85

APPENDIX A

Staftsn No. so be Cnsditad Permit femtie. UNITED STATES DEPA -.TVIENT OF THE 1'NTERiCR 43615 04-0 ils FiS'H AND WILDLIFE SERVICE S»JKV*f~V. D ate

Mississippi Sandhill Crane National Wildlife Refuge 03-02-2004 period of Use (inclusive) SPECIAL USE PERMIT From March 15 . 2004

T o J u ly 30 .2 0 0 4 permittee Name I Permitis Danna Smith (PI), Sam Holcomb, Cart Qualls I Department of Biological Sciences University of Southern Mississippi i Hattiesburg, MS 1304/237-1941, danna.*mith®usm.edu

Purpose (specify in detail privilege requested, or units of products involved) To conduct the "Survey to Determine the Conservation Status of toe Mimic Glass Lizard in Mississippi", During a 3-wee!; period, permittees may survey for glass lizards at designated areas on the refuge.A combination of drift fences, pitfall and funnel traps, as wait as manual sea'ching, and road surveys witl be techniques used. Once traps are set up, permittees wish to survey daily for 14-21 consecutiva days. Captured fcards wiil fee removed from tfaps; they may be measured and mproductuve status assessed. One voucher specimen may i Reded Lorn We refuge. Prior to release, each lizard may be marked with dorsal paint mark and PIT tag for iderrtificafkm. Habitat data i he collected.

j Description (specify unit m e n b e ia c jm e lw * « n 4 i»ui«t», w other rSssgrstabfe -Ui-aignWions;-

Amourtt of fee if not a fixed payment, specify rate and unit of charge:

{" Payment Exempt -Justification: * T Foil Payment i p merit -Balance of payments to be made as follows: Reeor f /merits

Special Conditions , ; Permittees will closely coordinate the location and timing of their surveys with refuge biologist to reduce chance of disturbance to rwsSop ! . r i as t nuit c i ltd tim ng of surveys must be OK’d by refuge biologist (During the crane nesting season Mar 01- Jun 30, a I are* i8,u..« j — l - d to disturbance unless arranged]. Permitees will stop in at refuge office to check in at beginning of the 21-day - ri « -»»s wi I sgn out keys for refuge gate at office, Permittees will needa 4WD vehicle (preferably with winch'' to r Of ua-» vv ’"Vm fleas w.ll keep this permit with them or ;n their vehicle white on the refuge. A report on sumey fmd-’-g w u' -'-voted to the refuge biologist.

m p t m l :s tMMKi i f f* » O.S. fak snd vv lr» «- t~ rrovftT: 3t>on£ -fi r®gervatksns. sxprsssod or STipfed h©.w., ^ o «lO*««««,*, y^F / ; ^

s « m u ttse / / 7 3'7" -yy'~-C'9^0? /

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 86

APPENDIX B

Authorization iD: FS-27O!W5s02S9j

Contact ID: gmb no . gsOd-cosz

Expiration Dais:

Use Code: U.S. DEPARTMENT OF AGRICULTURE Forest Service TEMPORARY SPECIAL - USE PERMIT (FSH 2709.11, sec. 54.6} AUTHORITY: Act of June 4,1897

University of Southern Mississippi hereinafter called the Holder, is hereby authorized to use, subject to the terms and conditions of this permit. National Forest System land identified within the unit area and described as 36 sites within T2N.R9W.S3S: T2N.Rf2W.S3: T1N.R3VV.S2: T1N.R10W.S22: TIN R12W.S?3:T1S.R12W.S12: TiS.R11W.S17&20: T1S.R12W.S19: T1S.R14W.S14: T3S.R9W.S20: T4S.R11W.S27&28: T6S R11W.SS. T55 R11W.S14 as shown on the attached Maps. This authorization cavers approximately2 acres.

The holder is authorized to conduct the following activities and/ or install the following temporary improvements on the permitted area. Conduct surveys for the mimic glass lizard, including construction of temporary drift fences with pitfali and fur.nei traps (see attached Exhibit 1 - Research Proposal),

TERMS AND CONDITIONS

I. Use under this permit shall begin on 3/15/2004 and end on 3/14/2005. The permit shall not be extended. 2 The fee for this use is S2_. tt shall be paid in advance and is not refundable 3 The holder shall conduct the authorized activities according to the attached approved plans and specifications, Exhibits 1 & 2 . 4. The holder shall not install any improvements not specifically identified and approved above. 5. No soil, frees, or other vegetation may be destroyed or removed from National Forest System lands without specific prior written permission from the authonzed officer. 6 The holder shall comply with all Federal, State, county, and municipal laws, ordinances, and regulations which are applicable to the area or operations covered by this permit 7 The holder shall maintain the improvements and premises to standards of repair, orderliness, neatness, sanitation, and safety acceptable to the authorized officer. The holder dial! fully repair and bear the expense for alt damage, other than ordinary wear and tear, to National Forest System lands, roads and trails caused by the holder's activities. 8 Tire holder has the responsibility of inspecting the use area and adjoining areas for dangerous trees, hanging limbs, and other evidence of hazardous conditions which would pose a risk, of injury to individuals. After securing permission from the authorized officer, the holder shall remove such hazards. S. The holder shall be liable for any damage suffered by the United States resulting from or related to use of this permit, including damages to National Forest resources and costs of fire suppression. 10, The holder shall hold harmless the United States from any liability from damage to life or property arising from the holder’s occupancy or use of National Forest lands under this permit. I I . The holder agrees to permit the free and unrestricted access to and upon the premises at ail times for al! lawful and proper purposes not inconsistent with the intent of the permit or with the reasonable exercise and enjoyment by the holder of the privileges thereof. 12, This permit is subject to all valid existing rights and claims outstanding in third parties. 13. This penult may be revoked upon breach of any of the conditions herein or at the discretion of the authorized officer, Upon expiration or revocation of this permit, the holder shall immediately remove all improvements except those owned by the United States, and shall restore the site within 15 days, unless otherwise agreed upon in writing. If the holder fails to remove the improvements, they shall become the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 87

property of the Unites States, tut teat will not relieve the holder of liability for the cost of their removal and restoration of the site, 14. This permit Is 8 license for the use of federally owned land. It does not grant any interest in real property This permit Is not transferable. The holder shatl not enter into any agreements with third parties for occupancy of the authorized premises and improvements, 16, Appeal of any provisions of this permit or any requirements thereof shall be subject to tee appeal regulations at 36 OFR, Subpart C. or revisions thereof. 16. This permit is accepted subject to the conditions set forth herein, conditions 1-17 and Exhibits JL&J. attached to and made a part of this permit, 17, The above clauses shall control if they conflict with additional clauses or provisions

I have read and understand the terms and conditions and agree to abide by them,

HOLDER U. S. DEPARTMENT OF AGRICULTURF University of Southern Mississippi Forest Service

By By: •/f- Address:! 118 CoBweDrive #6013 HatBesbura. MS 39406-0001 Name dudith Henry

THie:Dlsthct Render (Authorized Officer) Phone #: (601) 266-6906

Date: -iy~j Date.

is lh«P e c w u fc WsvketfopAciaf TOporm »a«»3vtr^»«»gpiec^cfOR * Mapicy* & wfee 0M8ccwatThAwpt&Cfcre feawinr i;f fWnbrifo i SOfaCStor ie 0588-6082.

nee*n^cjfBw?^fs*J fS*n8rMftsSJsr

S**wn tor«*. t tm s> * m » m , e to ty w m *»T «mP o m * **a, (*ftor«t Fanw!4»i An* f *m* Ad, H im tti Umv*9 Act MMh* 7#ro Panr-fc Act Act Of 3- ‘iNWwtvwa .Act fattooM itom USA* ms 7f*ia Act Aefof NovwnNw 1§, tS73, ArdtHtogkal feaiaspgt* P*otaelion Aest. «r*2 Alass* MMprt^ Hwniwrt iK-xta C »hi^r*t?afT Act P* $*cr«t*y 4 A^rieiture to a«u* $w**erfMf»r* 'or tfw us* *w oocuperey «f ?**¥«■*» fg ^ a Syswrt tank. *>• S*pr*t«y of A ^roitsrrj nojtafiavs St. X O f tUSS’t. Pm Sb43pw1 B, »«st«5A prtj

^2#)*»slm FrwKterB efV»ann*BAc;yus.C.&&aommtHtartis*ii!dtyi$e*prmm6&WQm&mrK»i»mtoflbii Pcracrt Serv«j»

not* fwrnponete swap** Of W5«3^ op»f*i»A aod’of rrv«?w»Mfi» t*Lir>r««gxfie« «#<«?; 8fttf*ntvwoapri ns^fef.Mt* *«*■*»<« me? tefcAteAeHttrgeaa rsppm •^s. <**% *Mi t*r»a»r.flfan. at**** Wn«waSoh. iM or«r saw*# fnfemtfEA oqtMW. 1>& fer mv®wsi *-**.«**». *»«**<$ cwwjne «**

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. a p p e n d ix c

HANCOCK S O t a .

J gave the University of Southern Mississippi lab permission to conduei^i'foctclogtcal surveys for the mimic glass lizard f Qphteaurus minucus) on . my private land or 16 Section land

5 fence Date 8f &-CV y

HANCOCK SCHOOL DISTRICT 17304 HWY. 603 KILN, MS 3SS58

x--"

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 89

APPENDIX D

THE UNIVERSITY OF SOUTHERN MISSISSIPPI

June 15. 2004

Dr, Carl Qualls Uepaifnctu of B iology I nnerMt} of Southern Mississippi Ucttk -.butg. MS 39406

Ri F \( Cc p ro v e d U H00SW

Dear Dr, Quails,

V.>m protocol entit.ed " Ophis-aurns Project” has, been reviewed and approved by the University of southern \!t >Ms»inpi Institutional Animal Care and Use Committee.

The project is approved for the period from March 15. 2004 until September 30. 2007. If there are any substantive changes in the protocol, you are required to submit an addendum to the original proposal, which must be approved by the IACUC.

’1 he 1 \('l U gives assurance that the proposal complies with the public health policy on humane c.itv and use of laboratory animals and all applicable provisions of the Animals Welfare Act.

cd Sheree Watson, Ph.D. Interim Chair. IACUC

DEPARTMENT OF PSYCHOLOGY Sox it CD • Mscuesbuts,itriesburj MS ■ ’9406-Mi T.orse 16041 ioiv4177 • Fax (oOlS 266-551' •w.usaiedu Hattiesburg * L<>«g Beacf * Q a-a n Springs * BilsM * .bhn V. Sct'tinisS[xicc Center

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. APPENDIX E

AfcO ’■« !» fR!> * s n i p

LUCtDALE MISS. February 1!, 2004

* * Oatma Smith 404 M orth 2 151 Avenue Hattiesburg. MS 39401

Dear Danna.

Please find enclosed maps for three different tracts as well as a plat map. The tracts are in Lucedale, Mississippi (George County) and are easily accessible. To find the tracts, all you have to do is drive on Highway 49 South to Highway 9£ East. You will stay or. Highway 98 for about 50 to 60 miles mid all three tracts are located on Highway 98. When you come to a new Chevron cm your right- hand side (south side of 98), you are on the tract labeled "Bush Dirt Pit". Please note where I drew in the Chevron station on the stand and plat maps.

All tracts are under a hunting lease and therefore I need to know in advance when you will be present.

Please call me with any questions.

Sincerely,

fSjUAMM

Sarah Watkins

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 91

APPENDIX F

13 March 2004

:;'::;:,ADX1 X1 STRATIVS SCIENTIFIC ' LL M PERMIT

:-0 WHOM I T WAV ...COKCBSN:

Per-'.seiort is granted, to;

Danna S m ith■ Ur:rvebSityi'df '■ Southern M issi 82 sIpp 1 Dept&vof B iological Science's lI?8,sCol lege Drive # SQiS m t t i e s b u r g , MS 324 On --OOOl,

assisted by Dr. Carl Quails; Krista Noel, Thomas Mohrmar-, and Samuel Holcomb, no bbiltect amphibians, reptiles, and ama.11 mammals while undertaking a survey for :a&h±saurus (glass Lizard) species oh vei1-drained to excessiveiy wall-drained ■ &oi I s in s o u th e r n M i s s i s s i p p i .

Collection hods may include cover boards, drift fences with pitfall traps and funnel traps, ana nand collection during foot surveys. Because- drift fence pitfall trap .’ill typically be placed in areas suitable for tortoises:, cam 'Utt oe taxen to avoid incidental captur f i or sec. Bucket traps should be fitted y>.th a cylinder or wire mesh of m l enough diameter to exclude juven-le e ’ j‘f■:**.& but large- enough to permit r i h - glass Iisards and smaller, lizards. anct snakes. Bucket traps should be tat o prevent over-heat ing o f captives and the buckets should have some sort of cover and moisture source on the bottom, buckets must also have holes in the bottom to allow drainage of rainwater.

This habitat, type also supports the federally and state endangered Mississippi, g o p h e r f r o g {Rana se ro sa ), the state enaanqerea, federal candidate black pine sn a k e (PLtuopkis lodlngi: the. fed* r» 1 . state endangered gopher co r y o iw e iGovhorus polypnea *)Mi , t ) ed, state endangered mdlgo snak e i Drym&jrchGn c o ra ls} . a? * t southern hognose snake 'Hoterodor: si.vus} . It is ur.Iikv I t \ pecies except the tortoise w ill be encountered, and as stresses above, trapping apparatus should bo designed i.e. minimize the likelihood of incidental capture of tortoises. However, should any of these species be 1 ncxderta 3.ly captured within trapping gear or ericounfeied while, conducting surveys on foot, brief handling is authoritea to secure voucher photographs adequate to permit unequivocal species deterrninat,on. Turing such handling permittees are sc be regarded as Agents of the State.

Collection gear left unattended in the field must be properly identi.fie-d.

with the exception of incidental capture as described above, this permit does not. authorize the taking of any threatened or endangered species (state or federal, list attached?. r.'f.-crdi'y S.iiuni: X’i^-moinn

f>:V-^VD: &!«SV? » MC.M O’;. IS i m fM 553 * PHW E f-OI ::f~1C 0; fAX St 1 ^-7 1 2 7 •

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 92

Because most of tac proposed collection nit kg w-‘i be on the Desoto Natinr.jQ Forest, and because of the possibility of incidental capture of federally listed specie a vithir. trap arrays and/or fur dncumtrcation purposes during foot surveys of trapping sites, your collect, i one iroy require a permit from Lbs National Forest Service. The state permit s invalid until such time as you obtain the *J.S. ForesL Secvice Permit I if necessary), and it ie aLtached hereto and a copy forwarded to ths M ississippi xuaeurr. c i Natural Science (ATTH; Scientific Collection Ferm.it Ocaruttee) .

This permit cr a copy of this permit must be in possession of tbs permittee Is whan o r g a n is e s a r e ta k e n o r pm;i;K:;r.;e.d u n d e r t h e c o n d itio n s o f t h e p e r m i t .

A complete report of collections must be sent to the Mlnniasippi Museum of Natural Science, 2148 Riverside Dr., Jackson, MS 19232-l.wi (Attn; Scientific Collection Ferrr.it Review Committee) within 15 days of permit expiration. Collection reports should list Laxa collected, number of individuals of each, exact collection locality and d»?,e of collection. locality information muct include me county of collection, and if. is preferred that precise locality information be provided in I a t.itude/longitude (GPS) or in the township, range, a n d s e c ti o n 1TR8 ; system. If the TKS system is used, precise location within a section Bhould be indicated (e.g.; KV?4 of 5E4 of Sec 11). if ponsiblc.

A single ramhr.r ssccinm of Pphisaurus mimicus (the Tiicmic glaeo liiardi nr.d •other non listed species may be retained from, each population (trapping sitoB ir. close proximity should be considered a single srte for voucher purposes) All other Bpecimens w ill be narked and re.eased alive at. t-.hr point of capture, Voucher specimens w ill be deposited at the M ississippi Museum of Natural Science.

A cany of publications, survey reports, and other printed materials produced as a result of this sol 1 ect.i on shov'-d be 3ont to the Mississippi Kuaeum o£ Natural Science (Att.o: Scientific Collection Permit Review Committee;- ,

r'-!.r r->- ;-nf --rn;- f r -i-“x .; f on 11 M arch 2 DOS

m :tm , e-ouservation hioicgy section

CC Cary Bcrrv. Manager, Rr.g-.on. 2 and Pegian 2 Diane Tyrone, DcSoto National forest Mjae-iui

Enc lo s u r c

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 93

APPENDIX G

United States Forest National Forests 654 W. Frontage RdL, P.O. Box 243 Ip Defiartnictit of Service In Mississippi Wiggins, MS 3957*? Agriculture ...... liy^^?2:rrf'Y60{/02S-2Si0

File Code: 2670 Dale: February 16,20G5 Route To:

Subject: Black Pine Snake Research

T»: University o f Southern Mississippi Department o f Biological Sciences Attn; Danna Smith ! 18 College Dr. #5018 Hattiesburg, MS 39406-0001

This letter is your authorization to conduct research on the conservation biology and ecology of

the Mack pine snake in Mississippi We have burned the site on February 12 and you can install

the fences immediately. Please follow the guidelines in the permit from the Mississippi Museum

o f Natural Science. I look forward to getting a copy of your research results.

'JUDITH HENRY District Ranger De Soto Ranger District

C aring Tar the Land and Serving People Pratted 30 KeeveJrcs Pnoe* o

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 94

APPENDIX H

KOf,T?L?:ri Apr'iNisT'iTivi sctettitk collect ::.;!-!

It; VAt CTHCc.KN: eel is gra-Tsd v.t:

Southern 1\ i to.-; losipp i.Mp v,. e f B jolagisai Sc^ncc.s MS C ciie# Drive # iai'B :Uvt.M o:.; vp MS Mi-M-f ft M

:,-T,;: ;.fi» £■ , C« » QU-Sii"; P lS V i NoML, TOCOO': HC-V. r C? 7: . W i -M O O .. A S fj , .--ro M .s s ls s i.p p .f. A lth ou gh th e fm;..-; o f b ls c k p in e momo rnmo;rah w;.,., v rtixP M'-'- M-:otc; Km tonal Kntosf' In Kayne arm -Miry CoaAften. to r bo Mart on :'.'.:o.o v _•.• W.i. I-.if :: ••• IMnacovonf: Aron tr. Hsrtnn County M ono!ftonal upon rorrmoMo-n of Mot omaqor f sony Birin qe.M, ono on Loco Paoktno Ccopony land Mor-dif i.onn i open pmoum;: on to/ the landov-nerf tn loorcro Cn .snty, b!.ao> pine snakes incidents •:, v u.o nmM msmMare ray be .included wi b m ?: the etody protcool, reopee fcecw t

C o :. o-cti-.-n nafhodn ;;oiy inoitdo oovof boardv, Sheroon live Irene, dvirf. Moon •■ at v.n .funnel fceye, and tend collection durino foot, survey-. lee.nor drift 'Mr:-v farmo.i. trap -mays wl.H fypiceii y oo pCaced ir. areas suitable ter for' o ;.t?.t, ony t reuse b e otvo: to avoid .MoidnnTai oepto.ro of juvenile tns'ocl o n e . f ore-- : ■• r e r o told .f-r noau-M to preves* coe.r-heet..; v.o of captives and onoulo ca ere:.!- J 0 o f ficien t.'. y rrom onb'y m t ioasf once per day, one nore cider, -e tor conn.i.:;on.- • to emut.re M e t c e p f o i e d a n i o n M do .not d e c o d e n e b y n r e r ec. o.r ev e m e a t o t:. •Mcroro; 0 ...e;:- a n d o r. ir. o r r o t on C o : s n a k e m t u a i a v-u.ll b e i r- ae"* t to o Loci vnitk m ;.;;fre.rod '; J-.t- c c:ne r e . o ' e o t p i n e e r e o o r a M I h a v e PIT f a c e t r t p i a n t o i. or:.- Pu.itad.:/ CtiOT irtrii ;-C.d'-o M M r ; f o ;v,.rru .;e 1 ;.y irtplnrced ’with, a f crepe: e ;o; r r - t.onto: : . ■•;•; O l d : ; t r atiortitr . Tvansciitter implaniration should fc© done according to acc^ptoc prococol (contact Dr. Craxg Rudolph: (S36> 56^-7981) by 3 veterinarian or an individual experienced with this procedure. Surgery-.related pmbloms should be immediately reported to Tom Mann or Dr. Bob Jones at the Mississippi Museum cf Science ({501> 354-6367, extensions 116 and 113. r e s p e c t i v e l y ) . n rs 1 ;.,v - t v v t ..:., i f p r: i: t I t i t ,. ' f:t .:.. i ;; * • f ,t -;. '. ri ". -t i / ili-1. fr^M-v t: t -•' f cay bv wc-ti t\ qt:; :;.i, ptii.ttt.r, 1 • -Vi't'tb t .: t t t i 't f tbCiv;Vti y .1 f i V'.:? " t ft'ty ,

Phis h :s;v:. t At. eb -P btpp.t.ttb f P5 fudv.:-;.M / bPP P t to tv;, h-': t q “ r pr'i ";. ss'.Mt:; /ty ., pppiuv- f .Atti s t’.vs.b , ": t; f e-ttrt. 11 y f pr-.at e:;t p, « tb‘.v tr. .M 'pevec yt'pf t- • t t r t - t t .' tot t r u t p.1" .’yyt.f .it's:::; , f.tt M tvrM ;y t f. r. vt tv.-:' ---.A, ?f a:.' r-tti M.".;:; v f : ft' . t / t ; .the;- t i t . , t t a thu :P a:.t tvidunyv-r.'-tc sourV;#-rr. t t y b s t :. .-MMt'tM:. ;. .'•..'. tu-bM 7 f lb b;'.ti y t.haL u:q, o f : oxc^ p t t'.ho t b t t 1.1 .. • vtto Oo t"o.:r.: e.t'-d, and .-s sf ?:-t st^d M tv t, f r-.vp.triq optirat os t- Pt-j.': o to tvs .:. at. I to oirO.-t;-o: tte 1 i to f ; f yoJ o i ,t-.a ..dvoval to:.:-;, a :. o o f f o r t o la a a . Uowa-’O;;, t aa f a aoy to fhto^ sp~o.:,po >>r- top;dootM.iy vapopr-;.o oi't’otrappiA O bear or «ACbo;t*o;roi vobilt to.ittort; ourvoyo f tot , bri-rf h -'tio q is A vtfiiitt;: tc to o.oo: v— too- : : : .t o e r t j t: •; Joqoa' o to ro-rroo aoo y ; ;. vooo ; t - t t o ia s tio f o i r:o o o ' : an . fo r io q o.oo. - -.0 M. op p o o -o .tto o b to'o t o b-; r t p r c t f a s P:;-;.a..a tf t :•<; B ta to .

it-oori oy; Mv-i/ar A'f.'.'.a.vt’p’’?

o i O.-P-.oi Mof * P 6 P .MbM.ob * PH0« ‘ fo POOob FM 69i ^-?52T * o.-iooono;:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 95

Cellection ^ a r ^ eft anattended in the field j o l s be properly identified

With 5.h« exr;yp-iori c-f capture of clack pir.e ftimicns *nd incident* 1 c a p tu re cjI ether listed aperies os described above, this permit. docs not authorize* the taking of *iry threatened or endangered species (.^tate or federal, lis t attached! , most of the proposed collection ail.** will be on the UsSoto Natrona} torest, and became of the possibility cf incidental capture of federally 'isted .‘rpneie? within trap arrays and/or for docur^entat-.ion purposes >: obtain the :j.s. Forest Service Permit (if r.tt^wi'aryi, and it is attached nerel.o and a copy forwarded Lo hi>« M ississippi BCuaeum of Natural Science (ATTN: Scientific Collection i'ermit CoQuniLtooi .

This permit or a copy of this permit must be in possession of the permittee (x when organisms are taken or possessed under Uv- conditions of the permit. A complete report or col' not •ons must be z>~nt to the Mississippi ^4u3eum or Natural Su.ier.cn, 2.14S Riverside Or., JotAscn, MS 3$202-1R53 (Attnt Scientific Collection iermit Review Committee) within '5 days of permit vxpiratior.. Collection reports should list Lax*. collected, number of individuals of each, exact collection locality and date ol' collection. Locality information muxt include the county of collection, and if is preferred that precis* locality information be provided ir. latitL.c.e/3unqitude (GPS) or U\ the township, tvsnqe, and s e c tio n (TR5) system - i f th e TRS system u scd r p r e c is e i:>»:a<;ion w ith in a section fthrnlc be indicated (e.g.: NV4 cf Si4 of Sec ll)j if possible.

Voucher specimens of Aon-ilsl.ijd species may ne ret.*ined from each pcpnl$tior. (trapping sites in close proximity should b« considered a single site for-vcicber purp::s«r;; . All other specimens will be marked (which m^y include insertion of pit tags; and released alive at the poir.i of rapture. Vauihcr specimens will be rlep c situ d a t th e M is s is s ip p i Museum o f n a tu r a l Scior.ee.

A cc-ipy of publications, survay reports, and other printed m aterials produced sk a result of this collection ahouLfi bn r,*nt to the Kissl-s^lppi Xuaeur. of Natural Science IAttns SciecLific Collection Eerrnil. Review Committee:-.

3 Haxch 200*-, a f V

Ulrtcr., zcti.*;c'rvotion biciogy sc rtim

<:c Gary b e rry , [O n ag er, P.r*:j

F.ncl<35ure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LITERATURE CITED

Ando, A., J. Camm, S. Polasky, and A. Solow. 1998. Species distributions, land values,

and efficient conservation. Science 279:2126-2128.

Andrews, K. M., and J. W. Gibbons. 2005. How do highways influence snake

movement? Behavioral responses to roads and vehicles. Copeia 4:772-782.

Aresco, M. J. 2005. The effect of sex-specific terrestrial movements and roads on the sex

ratio of freshwater turtles. Biological Conservation 123:37-44.

Aresco, M. J., and C. Guyer. 1999. Burrow abandonment by gopher tortoises in slash

pine plantations of the Conecuh National Forest. Journal of Wildlife Management

63:26-35.

Asplund, K. K. 1974. Body size and habitat utilization in whiptail lizards

(Cnemidophorus). Copeia 1974:695-703.

Avise, J. C. 2000. Phylogeography: The History and Formation of Species. Harvard

University Press, Cambridge, Massachusetts 464 p.

. 2004. Molecular Markers, Natural History and Evolution. Sinauer Associates

Inc., Sunderland, Massachusetts. 511 p.

Bartram, W. 1791. Travels through North Carolina and South Carolina, Georgia,

East and West Florida. Dover Publications, Inc., New York.

Baskaran, L. M., V. H. Dale, R. A. Efroymson, and W. Birkhead. 2005. Habitat modeling

within a regional context: An example using gopher tortoise. The American

Midland Naturalist 2:335-351.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 97

Baxley, D. and C. Qualls. 2005. Conservation and ecology of the black

Pinesnake ( Pituophis melanoleucus lodingi) in South Mississippi. Report to the

Mississippi Department of Wildlife, Fisheries, and Parks. Jackson, Mississippi,

USA. 38 pp.

Bernardino, Jr., F. S., and G. H. Dalrymple. 1992. Seasonal activity and road mortality of

the Pa-hay-okee wetlands of the Everglades National Park, USA. Biological

Conservation 61:71-75.

Bematchez, L., and C. C. Wilson. 1998. Comparative phylogeography of Nearctic and

Palearctic fishes. Molecular Ecology 7:431-452.

Berry, J. K. 1993. Beyond Mapping-Concepts, Algorithms, and Issues in GIS. GIS World

Books, Fort Collins, Colorado.

Blaustein, A.R., D. B. Wake, and W. P. Sousa. 1994. Amphibian declines: judging

stability, persistence, and susceptibility of populations to local and global

extinctions. Conservation Biology 8:60-71.

Blouin-Demers, G., and P. J. Weatherhead. 2001. Habitat use by black rat snakes ( Elaphe

obsoleta obsoleta) in fragmented forest. Ecology 82:2992-2896.

------. 2002. Habitat-specific behavioral thermoregulation by black rat snakes {Elaphe

obsoleta obsoleta). Oikos 97:59-68.

Bowman, J., C. V. Corkum, and G. J. Forbes. 2001. Spatial scales of trapping in small-

mammal research. Canadian Field Naturalist 115:472-475.

Browning, D. M., S. J. Beaupre, and L. Duncan. 2005. Using partitioned Mahalanobis D2

to formulate a GIS-based model of timber rattlesnake hibemacula. Journal of

Wildlife Management 69:33-44.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 98

Burbrink, F. T., and R. Lawson. 2007. How and when did old world ratsnakes disperse

into the new world? Molecular Phylogenetics and Evolution 43:173-189.

Burbrink, F. T., R. L. Lawson, and J. B. Slowinski. 2000. Mitochondrial DNA

phylogeography of the polytypic North American rat snake ( Elaphe obsoleta): A

critique of the subspecies concept. Evolution 54:2107-2118.

Burger, J., and R. T. Zappalorti. 1991. Nesting behavior of Pinesnakes ( Pituophis

melanoleucus melanoleucus) in the New Jersey Pine Barrens. Journal of

Herpetology 25:152-160.

Burger, J., R. T. Zappalorti, M. Gochfeld, W. I. Boarman, M. Caffrey, V. Doig, S. D.

Garber, B. Lauro, M. Mikovsky, C. Safina, and J. Saliva. 1988. Hibemacula and

summer den sites of Pinesnakes {Pituophis melanoleucus) in the New Jersey Pine

Barrens. Journal of Herpetology 22:425-433.

Campbell, H. W., and S. P. Christman. 1982. Field techniques for herpetofaunal

community analysis. In N. Scott, Jr. (ed.), Herpetological Communities, pp. 193-

200. U.S. Fish and Wildlife Service, Wildlife Research Report 13.

Carfagno, G. L., E. J. Heske, and P. J. Weatherhead. 2006. Does mammalian prey

abundance explain forest-edge use by snakes? Ecoscience 13:293-297.

Carranza, S., E. N. Arnold, and J. M. Pleguezuelos. 2006. Phylogeny, biogeography, and

evolution of two Mediterranean snakes, Malpolon monspessulanus and

Hemorrhois hippocrepis (, ), using mtDNA sequences.

Molecular Phylogenetics and Evolution 40:532-546.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Castoe, T. A., C. L. Spencer, and C. L. Parkinson. 2007. Phylogeographic structure and

historical demography of the western diamondback rattlesnake ( Crotalus atrox):

A perspective on North American desert biogeography. Molecular Phylogenetics

and Evolution 42:193-212.

Clark, J. D., J. E. Dunn, and K. G. Smith. 1993. A multivariate model of female black

bear habitat use for geographic information systems. Journal of Wildlife

Management 57:519-526.

Clevenger, A. P., J. Wierzchowski, B. Chruszcz, and K. Gunson. 2002. GIS-generated,

expert-based models for identifying wildlife habitat linkages and planning

mitigation passages. Conservation Biology 16:503-514.

Clibum, J. W. 1957. Behavior of a captive black Pinesnake. Herpetologica 13:66.

. 1962. Further notes on the behavior of a captive black Pinesnake {Pituophis

melanoleucus lodingi Blanchard). Herpetologica 18:34-37.

------. 1976. Observations of ecdysis in the black Pinesnake Pituophis melanoleucus

lodingi (Reptilia, Serpentes, Colubridae). Journal of Herpetology 10:299-301.

Cohen, J. E. 2003. Human population: The next half century. Science 5648:1172-1175.

Conant, R., and J. T. Collins. 1991. A Field Guide to Amphibians and Reptiles of Eastern

and Central North America. Houghton Mifflin Co. Boston.

Conner, M. L., M. D. Smith and L. W. Burger. 2003. A comparison of distance-based

and classification-based analyses of habitat use. Ecology 84:526-531.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 100

Corsi, F., J. De Leeuw, and A. Skidmore. 2000. Modeling species distribution with GIS.

In L. Boitani and T. K. Fuller (eds.), Research Techniques in Animal Ecology:

Controversies and Consequences, pp. 389-434. Columbia University Press, New

York.

Corsi, F., E. Dupre, and L. Boitani. 1999. A large-scale model of wolf distribution in

Italy for conservation planning. Conservation Biology 13:1-11.

Coyne, J. A., and H. A. Orr. 2004. Speciation. Sinauer Associates, Sunderland,

Massachusetts, USA.

Crain, J. L. and J. W. Clibum. 1971. Pituophis melanoleucus lodingi from the Western

part of its range. The Southwestern Naturalist 15:496-497.

Crandall, K. A., O. R. P. Bininda-Emonds, G. M. Mace, and R. K. Wayne. 2000.

Considering evolutionary processes in conservation biology. Trends in Ecology

and Evolution 15:290-295

Dayton, G. H., and L. A. Fitzgerald. 2006. Habitat suitability models for desert

amphibians. Biological Conservation 132:40-49.

De Queiroz, A., R. Lawson, and J. A. Lemos-Espinal. 2002. Phylogenetic relationships of

North American garter snakes ( Thamnophis ) based on four mitochondrial genes:

How much DNA sequence is enough? Molecular Phylogenetics and Evolution

22:315-329.

Dirzo, R., and P. H. Raven. 2003. Global state of biodiversity and loss. Annual Review of

Energy and the Environment 28:137-167.

Dodd, C. K. 1991. Drift fence-associated sampling bias of amphibians at a Florida

sandhills temporary pond. Journal of Herpetology 25:296-301.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 101

Dodd, C. K. and R. Franz. 1993. The need for status information on common

herpetofaunal species. Herpetological Review 24:47-49.

. 1995. Seasonal abundance and habitat use of selected snakes trapped in xeric

and mesic communities of north-central Florida. Bulletin of the Florida Museum

of Natural Flistory 38:43-67.

Dubuc, L. J., W. B. Krohn, and R. B. Owen Jr. 1990. Predicting occurrence of river otters

by habitat on Mount Desert Island, Maine. Journal of Wildlife Management

54:594-599.

Duran, C. M. 1994. Ecology of the Hispid Cotton Rat (Sigmodon hispidus ) With Comments

on the Effects of Artificial Population Manipulation on an Eastern Texas Small

Mammal Community. Unpubl. M.S. Thesis, Stephen F. Austin State University

Nacogdoches, Texas.

,1998a. Radio-telemetric study of the Black Pinesnake {Pituophis

melanoleucus lodingi) on the Camp Shelby Training Site. Final Report.

Misssissippi Natural Heritage Program and the Mississippi National Guard.

. 1998b. Status of the black Pinesnake (Pituophis melanoleucus lodingi).

Final Report to the United States Fish and Wildlife Service . 32 pp.

Felsenstein, J. 1985. Confidence limits on phylogenies: an approach using the bootstrap.

Evolution 39:783-791.

Fitch, H. S. 1970. Reproductive cycles of lizards and snakes. Miscellaneous Publications

of the Museum of Natural History, University of Kansas 52:1-247.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 102

. 1987. Collecting and life-history techniques. In R. Seigel, J. Collins and S.

Novak (eds.), Snakes: Ecology and Evolutionary Biology, pp. 143-181. McGraw-

Hill, New York.

Foin, T. C., S. P. D. Riley, A. L. Pawley, D. R. Ayres, T. M. Carlsen, P. J. Hodum, and P.

V. Switzer. 1998. Improving recovery planning for threatened and endangered

species. Bioscience 48:177-184.

Franz, R. 1986. The Florida gopher frog and the Florida pinesnake as burrow

associates of the gopher tortoise in northern Florida. In D. R. Jackson and R. J.

Bryant (eds.), The Gopher Tortoise and Its Community, pp. 16-20. Florida State

Museum, Gainesville.

------. 2001. Pituophis melanoleucus mugitus. Digging behavior. Herpetological

Review 32:109.

Freitag, S., C. Hobson, H. C. Briggs, and A. S. Van Jaarsveld. 1998. Testing for potential

survey bias: the effect of roads, urban areas and nature reserves on a southern

African mammal data set. Animal Conservation 1:119-127.

Fretwell, S. D. 1972. Populations in a Seasonal Environment. Princeton University press,

Princeton, New Jersey, USA.

Fretwell, S. D., and H. J. Lucus Jr. 1970. On territorial behaviour and other factors

influencing habitat distribution in birds. Acta Biotheoretica 19:16-36.

Frost, C.C. 1993. Four centuries of changing landscape patterns in the longleaf pine

ecosystem. Proceedings of the Tall Timbers Fire Ecology Conference 18:17-43.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 103

. 2006. History and future of the longleaf pine ecosystem. In S. Jose, E. Jokela,

and D. Miller (eds.), The longleaf pine ecosystem, ecology, silviculture, and

restoration, pp. 9-48. Springer, New York.

Frost, C. C., J. Walker, and R. K. Peet. 1986. Fire-dependent savannas and prairies of the

Southeast: original extent, preservation status and management problems. Pp. 348-

357 In Wilderness and natural areas in the eastern United States: a management

challenge. Kulhavy, D. L., and R. N. Conner (Eds.). Center for Allied Studies,

School of Forestry, Stephen F. Austin University, Nacogdoches, Texas.

Gerald, G. W., M. A. Bailey, and J. N. Holmes. 2006a. Movements and activity range

sizes of Northern Pinesnakes ( Pituophis melanoleucus melanoleucus) in Middle

Tennessee. Journal of Herpetology 40:503-510.

. 2006b. Habitat utilization of Pituophis melanoleucus melanoleucus (Northern

Pinesnakes) on Arnold Air Force Base in middle Tennessee. Southeastern

Naturalist 5:253-264.

Gibbons, J. W., D. E. Scott, T. J. Ryan, K. A. Buhlmann, T. D. Tuberville, B. S. Metts, J.

L. Greene, T. Mills, Y. Leiden, S. Poppy, and C. T. Winne. 2000. The global

decline of reptiles deja vu amphibians. Bioscience 50:656-666.

Gibbons, J. W., and R. D. Semlitsch. 1981. Terrestrial drift fences with pitfall traps: an

effective technique for quantitative sampling of animal populations. Brimleyana

7:1-16.

Glitzenstein, J. S., W. J. Platt, and D. R. Streng. 1995. Effects of fire regime and habitat

on tree dynamics in North Florida longleaf pine savannahs. Ecological

Monographs 4:441-476.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Goertz, J. W. 1964. The influence of habitat quality upon density of cotton rat

populations. Ecological Monographs 34:359-381.

Green, D. M. 2003. The ecology of extinction: population fluctuation and decline in

amphibians. Biological Conservation 111:331-343.

Gregory, P. T., J. M. Mccartney, and K. W. Larsen. 1987. Spatial patterns and

movements. In R. Seigel, J. Collins and S. Novak (eds.). Snakes: Ecology and

Evolutionary Biology, pp. 366-396. McGraw-Hill, New York.

Guisan, A. and W. Thuiller. 2005. Predicting species distribution: offering more than

simple habitat models. Ecology Letters 8:993-1009.

Guyer, C., and M. A. Bailey. 1993 Amphibians and reptiles of longleaf pine

communities. In S. Hermann (ed.), The longleaf pine ecosystem: ecology,

restoration and management, pp. 130-158. Tall Timbers Research Station,

Tallahassee, Florida.

Hardy, D. L., and H. W. Greene. 1999. Surgery on rattlesnakes in the field for

implantation of transmitters. Sonoran Herpetologist 12:26-28.

Heard, G. W., D. Black, and P. Robertson. 2004. Habitat use by the inland carpet python

(Morelia spilota metcalfev. Pythonidae): Seasonal relationships with habitat

structure and prey distribution in a rural landscape. Austral Ecology 29:446-460.

Hels, T., and E. Buchwald. 2001. The effect of road kills on amphibian populations.

Biological Conservation 99:331-340.

Hewitt, G. M. 1999. Post-glacial re-colonization of European biota. Biological Journal

of the Linnean Society 68:87-112.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Hey, J., R. S. Waples, M. L. Arnold, R. K. Butlin, and R. G. Harrison. 2003.

Understanding and confronting species uncertainty in biology and conservation.

Trends in Ecology and Evolution 18:597.

Himes, J. G. 2000. Burrowing ecology of the rare and elusive Louisiana Pinesnake,

Pituophis ruthveni (Serpentes: Colubridae). Amphibia-Reptilia 22:91-101.

Himes, J. G. and L. M. Hardy. 2002. Movement patterns and habitat selection by native

and repatriated Louisiana Pinesnakes (Pituophis ruthveni): implications for

conservation. Herpetological Natural History 9:103-116.

Himes, J. G., L. M. Hardy, D. C. Rudolph, and S. J. Burgdorf. 2002. Growth rates and

mortality of the Louisiana Pinesnake. Journal of Herpetology 36:683-687.

Homer, C., C. Huang, L. Yang, B. Wylie, and M. Coan. 2004. Development of a 2001

national land cover database for the United States. Photogrammetric Engineering

and Remote Sensing 70:829-840.

Hopkins, C. D. 1977. Methods f or generating land suitability maps: a comparative

evaluation. Journal of the American Institute of Planners 43:386-400.

Huey, R. B., C. R. Peterson, S. J. Arnold, and W. P. Porter. 1989. Hot rocks and not-so-

hot rocks: retreat site selection by garter snakes and its thermal consequences.

Ecology 70:931-944.

Jaarola, M., and H. Tegelstrom. 1995. Colonization history of north European field voles

(Microtus agrstis) revealed by mitochondrial DNA. Molecular Ecology 4:299-

310.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 106

Jennings, S. B., N. D. Brown, and D. Sheil. 1999. Assessing forest canopies and

understory illumination: canopy closure, canopy cover and other measures.

Forestry 72:59-74.

JMP Version 5.1. SAS Institute Inc., Cary, North Carolina, USA.

Johnson, D. 1980. The comparison of usege and availability measurements for evaluating

resource preference. Ecology 61:65-71.

Jones, C. G., J. H. Lawton, and M. Shachak. 1994. Organisms as ecosystem engineers.

Oikos 69:373-386.

Katnik, D. D., and R. B. Wielgus. 2005. Landscape proportions versus monte carlo

simulated home ranges for estimating habitat availability. Journal of Wildlife

Management 69:20-32.

Kery, M. 2002. Inferring the absence of a species- a case study of snakes. Journal of

Wildlife Management 66:330-338.

Kincaid, B. W. and G. N. Cameron. 1982. Effects of species removal on resource

utilization in a Texas community. Journal of Mammology 63:229-235.

Knick, S. T. and D. L. Dyer. 1997. Distribution of black-tailed jackrabbit habitat

determined by GIS in southwestern Idaho. Journal of Wildlife Management 6:75-

85.

Komarek, E. V., 1965. Effects of fire on temperate forests and related ecosystems:

Southeastern United States. Pp. 251-277 In Fire and Ecosystems. Kozlowski, T.

T. and E. E. Ahlgren (Eds.). Academic Press, New York, New York, USA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 107

Kumazawa, Y., H. Ota, M. Nishida, and T. Ozawa. 1996. Gene rearrangements in snake

mitochondrial genomes: Highly concerted evolution of control-region-like

sequences duplicated and inserted into a tRNA gene cluster. Molecular Biology

and Evolution 13:1242-1254.

Landers, J. L., D. H. van Lear, and W. D. Boyer. 1995. The longleaf pine forests of the

southeast: requiem or renaissance? Journal of Forestry 93:39-44.

Legge, J. T., R. Roush, R. Desalle, A. P. Vogler, and B. May. 1996. Criteria for

establishing evolutionarily significant units in Cryan’s Buckmoth. Conservation

Biology 10:85-98.

Lindenmayer, D. B., J. Fischer, A. Felton, R. Montague-Drake, A. D. Manning, D.

Simberloff, K. Youngentob, D. Saunders, D. Wilson, A. M. Fleton, C. Blackmore,

A. Lowe, S. Bond, N. Munro, and C. P. Elliott. 2007. The complementarity of

single-species and ecosystem-oriented research in conservation research. Oikos

116:1220-1226.

Lipps, G. 2005. A framework for predicting the occurrence of rare amphibians: A case

study with the green salamander. M.S. Thesis. Bowling Green State University,

Bowling Green, Ohio.

Litt, A. R., L. Provencher, G. W. Tanner, and R. Franz. 2001. Herpetofaunal responses to

restoration treatments of longleaf pine sandhills in Florida. Restoration Ecology

9:462-474.

Madsen, T., and R. Shine. 1996. Seasonal migration of predators and prey: Pythons and

rats in tropical Australia. Ecology 77:149-156.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 108

Malczewski, J. 1996. A GIS-based approach to multiple criteria group decision­

making. International Journal of Geographic Information Systems 8:955-971.

Manly, B.F.J. 1991. Randomization and Monte Carlo Methods in Biology. Chapman and

Hall, New York, New York, USA.

Marsh, D. M. and P. C. Trenham. 2001. Metapopulation dynamics and amphibian

conservation. Conservation Biology 15:40-49.

Mayden, R. L. 1997. A hierarchy of species concepts: the denouement in the saga of the

species problem. In M. F. Claridge (ed.), Species: the units of biodiversity, pp.

381-424, Chapman and Hall, London.

McCoy, E. D., and H. R. Mushinsky. 1992. Studying a species in decline: Changes in

populations of the gopher tortoise on federal lands in Florida. Florida Scientist

55:116-124.

McCullough, D. R., editor. 1996. Metapopoulations and Wildlife Conservation. Island

Press, Washington, D. C., U.S.A.

McDearman, W. 2005. Gopher Tortoise (Gopherus polyphemus) soil classification for

the federally listed range. U.S. Fish and Wildlife Service, Jackson, Mississippi.

Means, D. B. 2006. Vertebrate faunal diversity of longleaf pine ecosystems. In S. Jose, E.

Jokela, and D. Miller (eds.), The longleaf pine ecosystem, ecology silviculture,

and restoration, pp. 157-213. Springer, New York.

Means, D. B., C. K. Dodd, S. A. Johnson, and J. G. Palis. 2004. Amphibians and fire in

longleaf pine ecosystems: Response to Schurbon and Fauth. Conservation

Biology 4:1149-1153.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Means, D. B., J. G. Palis, and M. Baggert. 1996. Effects of slash pine silviculture on a

Florida population of flatwoods salamander. Conservation Biology 10:426-437.

Meshaka, W. E., and J. N. Layne. 2002. Herpetofauna of a long-unbumed sandhill habitat

in South-Central Florida. Florida Scientist 65:35-50

Miller, J. H. and K. V. Miller. 1999. Forest Plants of the Southeast and Their Wildlife

Uses. Southern Weed Science Society, Champaign, Illinois.

Miller, R. G., Jr. 1981. Simultaneous Statistical Inference. McGraw Hill, N.Y.

Mushinsky, H. R., and E. D. McCoy. 1994. Comparison of gopher tortoise populations

on islands and on the mainland in Florida. Pps. 37-49 In Biology of North

American tortoises. Bury, R. B. and D. J. Germano (Eds.). U.S. National Biological

Survey, Fish and Wildlife Research 13, Washington D.C. USA.

Noel, J. M., W. J. Platt, and E. B. Moser. 1998. Structural characteristics of old- and

second-growth stands of longleaf pine (Pinus palustris) in the Gulf Coastal region

of the U.S.A. Conservation Biology 12:533-548.

Noss, R. F., M. A. O’Connell, and D. D. Murphy. 1997. The Science of Conservation

Planning. Island Press: Washington D.C., U. S. A.

O’Brien, S. J. and E. Mayr. 1991. Bureaucratic mischief- recognizing endangered

species and subspecies. Science 251:1187-1188.

Olsen, M., and J. Lee. 2007. Pituophis melanoleucus lodingi. Tree Climbing.

Herpetological Review 38:210.

Outcalt, K. W. 1997. Status of the longleaf pine forests of the West Gulf Coastal

Plain. Texas Journal of Science 49:5-12.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 110

Palumbi, S. R. 1996. Nucleic acids II: the polymerase chain reaction. In D. M. Hillis, C.

Moritz, and B. K. Mable (eds.), Molecular Systematics, pp. 205-320, Sinauer,

Sunderland, Massachusetts.

Pearson, D. E. and L. F. Ruggiero. 2003. Transect versus grid trapping arrangements for

sampling small-mammal communities. Wildlife Society Bulletin 31:454-459.

Pechmann, J. H. K., D. E. Scott, R. D. Semlitsch, J. P. Caldwell, L. J. Vitt, and J. W.

Gibbons. 1991. Declining amphibian populations: The problem of separating

human impacts from natural fluctuations. Science 23:892-895.

Posada, D., and K. A. Crandall. 1998. Modeltest: testing the model of DNA substitution.

Bioinformatics 14:817-818.

Reddy, S., and L. M. Davalos. 2003. Geographic sampling bias and its Implications for

conservation priorities in Africa. Journal of Biogeography 30:1719-1727.

Reichling, S. B. 1995. The taxonomic status of the Louisiana Pinesnake, Pituophis

melanoleucus ruthveni (Serpentes: Colubridae). Journal of Herpetology

29: 186-198.

Reinert, H. K., and D. Cundall. 1982. An improved surgical implantation method for

radio-tracking snakes. Copeia 3:702-705.

Richter, S. C. and R. A. Seigel. 2002. Annual variation in the population ecology of the

endangered gopher frog, Rana sevosa Goin and Netting. Copeia 4:962-972.

Rodda, G. 1993. Where’s Waldo (and the snakes)? Herpetological Review 24:44-45.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 111

Rodriguez-Robles, J. A., and J. M. De Jesus-Escobar. 1999. Molecular systematics of

New World lampropeltinine snakes (Colubridae): Implications for biogeography

and evolution of food habits. Biological Journal of the Linnean Society 68:355-

385.

______. 2000. Molecular systematics of new world gopher, bull, and pinesnakes

{Pituophis. Colubridae), a transcontinental species complex. Molecular

Phylogenetics and Evolution 14:35-50.

Rogers, A. R., and H. Harpending. 1992. Population growth makes waves in the

distribution of pairwise genetic differences. Molecular Biology and Evolution

9:552-569.

Ronquist, F., and J. P. Huelsenbeck. 2003. MrBayes 3: Bayesian phylogenetic inference

under mixed models. Bioinformatics 12:1572-1574.

Roux, D. J., K. H. Rogers, H. C. Biggs, P. J. Ashton and A. Sergeant. 2006. Bridging the

science-management divide: moving from unidirectional knowledge transfer to

knowledge interfacing and sharing. Ecology and Society 11(1): 4.

Rouse, Jeremy. 2006. Spatial ecology of Sistrurus catenatus catenatus and Heterodon

platirhinos in a rock-barren landscape. M.S. thesis. University of Guelph, Canada.

Row, J. R., and G. Blouin-Demers. 2006. Kernels are not accurate estimators of home-

range size for herpetofauna. Copeia: 797-802.

______. 2006b. Thermal quality influences effectiveness of thermoregulation, habitat

use, and behavior in milk snakes. Oecologia 148:1-11.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Rudolph, D. C., S. J. Burgdorf, R. N. Conner, C. S. Collins, D. Saenz, R. R. Schaefer, T.

Trees, C. M. Duran, M. Ealy, and J. G. Himes. 2002. Prey handling and diet of

Louisiana Pinesnakes {Pituophis ruthveni) and black Pinesnakes {Pituophis

melanoleucus lodingi), with comparisons to other selected colubrid snakes.

Herpetological Natural History 9:57-62.

Rudolph, D. C., Burgdorf, S. J., Schaefer, R. R, Conner, R. N., and Zappalorti, R. T.

1998. Snake mortality associated with late season radio-transmitter implantation.

Herpetological Review 29:155-156.

Russell, K. R., D. C. Guynn, and H. G. Hanlin. 2002. Importance of small isolated

wetlands for herpetofaunal diversity in managed, young growth forests in the

Coastal Plain of South Carolina. Forest Ecology and Management 163:43-59.

Sala, O. E., F. S. Chapin, J. J. Armesto, E. Berlow, J. Bloomfield, J. Dirzo, E. Huber-

Sanwald, L. F. Huenneke, R. B. Jackson, A. Kinzig, R. Leemans, D. M. Lodge,

H. A. Mooney, M. Oesterheld, N. L. Poff, M. T. Sykes, B. H. Walker, M. Walker,

and D. H. Wall. 2000. Global biodiversity scenarios for the year 2100. Science

287:1770-1774.

Schemske, D. W., B. C. Husband, H. M. Ruckelshaus, C. Goodwillie, M. Parker, and J. G

G. Bishop. 1994. Evaluating approaches to the conservation of rare and

endangered plants. Ecology 75:584-606.

Schneider, S., D. Roessli, and L. Excoffier. 2000. ARLEQUIN. A software for

population genetics data analysis, version 2.0. Genetics and Biometry

Laboratory, Department of Anthropology, University of Geneva, Geneva.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 113

Schurbon, J. M. and J. E. Fauth. 2003. Effects of prescribed burning on

amphibian diversity in a Southeastern U. S. national forest. Conservation Biology

17:1338-1349.

Seone, J., J. Bustamante, and R. Diaz-Delgado. 2005. Effect of Expert Opinion on the

Predictive Ability of Environmental Models of Distribution. Conservation

Biology 19:512-522.

Sheldrick, B.H. and C. Wang. 1993. Particle size distribution. Pp. 499-557 In Soil

Sampling and Methods of Analysis. Carter, M. R. (Ed.). Lewis Publishers, Ann

Arbor, Michigan, USA.

Shine, R., and M. Fitzgerald. 1996. Large snakes in a mosaic rural landscape: the ecology

of carpet pythons Morelia spilota (serpents: pythonidae) in coastal eastern

Australia. Biological Conservation 76:113-122.

Smith, D. and C. Qualls. 2004. Herpetofaunal survey of longleaf pine forests in South

Mississippi to determine the status and distribution of the mimic glass lizard and

black Pinesnake. Report to the Mississippi Department of Wildlife, Fisheries, and

Parks. Jackson, Mississippi, USA. 46 pp.

Smith, L. L., and C. K. Dodd, Jr. 2003. Wildlife mortality on U.S. Highway 441 across

Paynes Prairie, Alachua County, Florida. Florida Scientist 66:128-140.

Smith, L. L., D. A. Steen, J. M. Stober, M. C. Freeman, S. W. Golladay, L. M. Conner

and J.Cochrane. 2006. The vertebrate fauna of Ichauway, Baker County, Georgia.

Southeastern Naturalist 4:599-620.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 114

Standora, M. 2002. Landscape level GIS modeling of Eastern Massasauga Rattlesnake

(,Sistrurus catenatus catenatus) habitat in Michigan. M.S. Thesis. Indiana-Purdue

University, Ft. Wayne, Indiana.

Steen, D. A., M. J. Aresco, S. G. Beike, B. W. Compton, E. P. Condon, C. K. Dodd, H.

Forrester, J. W. Gibbons, J. L. Greene, G. Johnson, T. A. Langden, M. J. Oldham,

D. N. Oxier, R. A. Saumere, F. W. Schueler, J. M. Sleeman, L. L. Smith, J. K.

Tucker, and J. P. Gibbs. 2006. Relative vulnerability of female turtles to road

mortality. Animal Conservation 9:269-273.

Swofford, D. L. 2002. PAUP* 4.0: Phylogenetic Analysis Using Parsimony (*and other

methods). Version 4.b.l0. Sinauer Associates, Sunderland, Massachusetts.

Taberlet, P., L. Fumagalli, A. Wust— Saucy, and J. Cosson. 1998. Comparative

phylogeography and postglacial colonization routes in Europe. Molecular

Ecology 7:453-464.

Tear, T. H., J. M. Scott, P. H. Hayward, and B. Griffith. 1993. Status and prospects for

success of the Endangered Species Act: a look at recovery plans. Science

262:976-977.

______. 1995. Recovery plans and the endangered species act: are criticisms supported

by data? Conservation Biology 9:182-195.

Ter Braak, C. J. F. 1986. Cononical correspondence analysis: a new eigenvector

technique for multivariate direct gradient analysis. Ecology 67:1167-1179.

Tomlin, C. D. 1990. Geographic Information Sytems and Cartographic Modeling.

Prentice Hall College Division, Englewood Cliffs, New Jersey.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Tuberville, T. D., J. R. Bodie, J. B. Jensen, L. LaClaire, and J. W. Gibbons. 2000.

Apparent decline of the southern hog-nosed snake, (Heterodon simus). Journal of

the Elisha Mitchell Scienteific Society 116:19-40.

Tuberville, T.D., J.D. Wilson, M.E. Dorcas, and J.W. Gibbons. 2005. Herpetofaunal

species richness of southeastern national parks. Southeastern Naturalist 4:537-

569.

Varner, J. M. and J. S. Kush. 2004. Remnant old-growth longleaf pine (Pinus palustris

Mill.) savannas and forests of the southeastern USA: Status and threats. Natural

Areas Journal 24:141-149.

Vignieri, S. N., E. M. Hallerman, B. J. Bergstrom, D. J. Hafner, A. P. Martin, P. Devers,

P. Grobler, and N. Hitt. 2006. Mistaken view of taxonomic validity undermines

conservation of an evolutionarily distinct mouse: a response to Ramey et al.

(2005). Animal Conservation 9:237-243.

Vitousek, P. M., H. A. Mooney, J. Lubchenco, and J. M. Melillo. 1997. Human

domination of Earth’s ecosystems. Science 277:494-499.

Waldron, J. L., S. H. Bennett, S. M. Welch, M. E. Dorcas, J. D. Lanham, and W.

Kalinowsky. 2006. Habitat specificity and home range size as attributes of species

vulnerability to extinction: a case study using sympatric rattlesnakes. Animal

conservation 9:414-420.

Waldrop, T. A., D. L. White, and S. M. Jones. 1992. Fire regimes for pine-grassland

communities in the southeastern United States. Forest Ecology and Management

47:195-210.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Ware, S., C. Frost, and P. Doerr. 1993. Southern mixed hardwood forest: The former

longleaf pine forest. In W. Martin, S. Boyce, and A. Echtemacht (eds.),

Biodiversity of the Southeastern United States, pp. 447-493. John Wiley & Sons,

New York.

Wilson, D. S., H. R. Mushinsky, and R. A. Fischer. 1997. Species profile: Gopher

tortoise (Gopherus polyphemus) on military installations in the southeastern United

States. U.S. Army Corps of Engineers Technical Report (SERDP-97-10). 36 pp.

Woodroffe, R., and J. R. Ginsberg. 1998. Edge effects and the extinction of populations

inside protected areas. Science 280:2126-2128.

Wright, A. H., and Wright, A. A. 1957. Handbook of snakes of the United States and

Canada, Vol. II, Comstock Publishing Associates, Ithaca, New York.

Yale, A. L., M. E. Dorcas and J. W. Gibbons. 1999. Herpetofaunal diversity in coastal

Plain communities of South Carolina. The Journal of the Elisha Mitchell

Scientific Society 115:270-280.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.