Louisiana State University LSU Digital Commons

LSU Historical Dissertations and Theses Graduate School

2000 Host and Parasite Interactions Among Three Symbionts: Pocket , Chewing Lice, and Endosymbiotic Bacteria. David Lee Reed Louisiana State University and Agricultural & Mechanical College

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_disstheses

Recommended Citation Reed, David Lee, "Host and Parasite Interactions Among Three Symbionts: Pocket Gophers, Chewing Lice, and Endosymbiotic Bacteria." (2000). LSU Historical Dissertations and Theses. 7384. https://digitalcommons.lsu.edu/gradschool_disstheses/7384

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Historical Dissertations and Theses by an authorized administrator of LSU Digital Commons. For more information, please contact [email protected]. INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy submitted. Thus, som e thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy subm itted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer arid continuing from left to right in equal sections with small overlaps.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6” x 9n black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to ordesr.

Bell & Howell Information and Learning 300 North Zeeb Road, Ann Arbor, Ml 48106-1346 USA 800-521-0600

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. HOST AND PARASITE INTERACTIONS AMONG THREE SYMBIONTS: POCKET GOPHERS, CHEWING LICE, AND ENDOSYMBIOTIC BACTERIA

A Dissertation

Submitted to the Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College in partial fulfillment of the requirement for the degree of Doctor of Philosophy

in

The Department of Biological Sciences

by David L. Reed B.S., University of North Carolina, Wilmington, 1991 M.S., Louisiana State University, 1994 December 2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. UMI Number: 9998703

UMI___ ®

UMI Microform 9998703 Copyright 2001 by Bell & Howell Information and Learning Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code.

Bell & Howell Information and Learning Company 300 North Zeeb Road P.O. Box 1346 Ann Arbor, Ml 48106-1346

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

I would like to thank my advisor, Mark S. Hafner, for the countless hours

that he invested in my education. I benefited enormously from his friendship,

guidance, and knowledge. My committee members, Mark S. Hafner, Fred H.

Sheldon, J. Michael Fitzsimons, J. V. Remsen, and William Font, encouraged me

at every step along the way, and for their help, I am grateful. Fred Rainey and

Naomi Ward-Rainey were very helpful in the areas of systematic and

evolutionary microbiology. I appreciate their expertise and willingness to share

their knowledge. Shannon Allen, Mark S. Hafner, and Melanie B. Smith

contributed significantly to the work presented in chapters one and two. I am

indebted to Dwanda Lewis, Myles Digby, Justin Lyman, and Rebecca Miller who

cloned and sequenced bacterial DNA for chapters three and four. I simply could

not have accomplished my research goals if not for these collaborators.

I would like to thank my parents for giving me the early education

necessary to excel in life. They taught me the importance of a well-rounded

education and I am forever grateful to them. I would like thank my wife, Jamie,

for tolerating a Ph.D. candidate who was more often at his office than at home.

Lastly, I'd like to thank one of the most important groups in my graduate training.

From my earliest mentors, Jim Demastes and Theresa Spradling, to my most

recent friends and colleagues, I thank the graduate students at the LSU Museum of

Natural Science. There is no more important source of knowledge and

inspiration in the Museum of Natural Science than its graduate students. A

special thank you to each of you. Lastly, I would like to thank Dr. Charles M.

Fugler for being a great mentor, friend, and colleague. He is deeply appreciated

and equally missed.

ii

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

ACKNOWLEDGEMENTS...... ii

LIST OF TABLES...... v

LIST OF FIGURES...... vi

ABSTRACT...... viii

INTRODUCTION...... 1

CHAPTER 1. MAMMALIAN HAIR DIAMETER AS A POSSIBLE MECHANISM FOR HOST SPECIALIZATION IN CHEWING LICE...... 4 Introduction...... 4 Materials and Methods...... 5 Results...... 10 D iscussion ...... 16

CHAPTER 2. SPATIAL PARTITIONING OF HOST HABITAT BY CHEWING LICE OF THE GENERA GEOMYDOECUS AND THOMOMYDOECUS (PHTHIRAPTERA: TRICHODECTIDAE)...... 23 Introduction...... 23 Materials and Methods...... 24 R esults...... 27 D iscussion...... 31

CHAPTER 3. BACTERIAL DIVERSITY IN CHEWING LOUSE HOSTS...... 36 Introduction...... 36 Materials and Methods...... 46 Results...... 52 D iscussion...... 58

CHAPTER 4. BACTERIAL DIVERSITY IN CHEWING LOUSE EGGS...... 67 Introduction...... 67 Materials and Methods...... 69 Results and Discussion...... 71

CHAPTER 5. STUDIES OF COPHYLOGENY: CHEWING LICE AND ENDOSYMBIOTTC BACTERIA...... 76 Introduction...... 76 Materials and Methods...... 78 Results...... 81 D iscussion...... 97

SUMMARY AND CONCLUSIONS...... 105

LITERATURE CITED...... 107

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. APPENDIX: SPECIMENS EXAMINED...... 117

VITA...... 119

iv

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

Page

1.1 Model II regression analyses (major axis method) of the relationship between hair diameter (Diam) and body mass (Mass) ...... 11

2.1 Mean guard hair diameter (pm) for each of the ten body regions in the three pocket specimens examined...... 29

2.2 Total surface area (cm2) of each pelage region (Fig. 2.1) and number of observed and expected lice per region...... 30

3.1 Bacteria associated with chewing lice identified to the level of genus by means of BLAST searches of GenBank ...... 53

3.2 Endosymbiotic bacteria identified from nine taxa of chewing lice by means of environmental extraction, PCR, sequencing, and searching GenBank databases via BLAST nucleotide recognition search protocol...... 56

4.1 Types of bacteria found in DNA extracts of chewing louse eggs. Bacteria were identified by BLAST searches of the Genbank DNA sequence database ...... 72

5.1 Uncorrected pairwise distances (P-distances) for 27 taxa of the gamma subclass of the Proteobacteria ...... 85

5.2 Uncorrected P-distances for 42 taxa of Staphylococcus...... 91

v

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

Page

1.1 Electron micrograph (left) of a chewing louse (Geomydoecus aurei) attached to the hair shaft of a pocket gopher (Thomomys bottae)...... 6

1.2 Major axis regressions performed on ln(mass) and In(diameter) for 18 of ...... 12

1.3 Major axis regression through the origin of independent contrasts of ln(mass) and In(diameter) for members of the family Geomyidae ...... 13

1.4 Major axis regression through the origin of independent contrasts of ln(mass) and In(diameter) for pocket gophers in the genus Thomomys...... 14

1.5 Regression (major axis method) performed on In(mass) and In(diameter) for eight individuals of Thomomys bottae...... 15

1.6 Linear regression (general linear model) performed on chewing louse groove width and pocket gopher hair diameter...... 17

2.1 External surface of a pocket gopher showing the distribution of Geomydoecus and Thomomydoecus lice...... 26

2.2 Comparative density of lice in the 10 regions of pocket gopher pelage...... 32

3.1 Regression of the number of types of bacteria found as a function of the total number of clones sequenced for each of the seven gopher-Iouse host pairs listed in Table 3.2...... 60

3.2 Rarefaction curve showing the number of clones identified from Geomydoecus expansus compared to the total number of clones surveyed ...... 62

3.3 Rarefaction curve showing the number of clones identified from Geomydoecus geomydis compared to the total number of clones surveyed ...... 63

5.1 Maximum likelihood phylogeny for gamma Proteobacteria ...... 84

vi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5.2 Fifty percent majority rule consensus tree of 1000 bootstrap replicates using the parsimony optimality criterion for species of gamma Proteobacteria...... 88

5.3 Maximum likelihood phylogeny for Staphylococcus species...... 90

5.4 Fifty percent majority rule consensus tree of 1000 bootstrap replicates using the parsimony optimality criterion Staphylococcus species...... 96

vii

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

This study investigates the patterns of cophylogeny documented

between pocket gophers (Rodentia: Geomyidae) and their ectoparasitic

chewing lice (Phthiraptera: Trichodectidae). The first two chapters

investigate the mechanisms that reinforce cospeciation in this system.

Chapter one examines whether mammalian hair diameter is a resource

tracked by chewing lice, and thus reinforces the patterns of cospeciation. I

found that hair diameter of pocket gophers is correlated with gopher body

size and the body size of the ectoparasitic chewing lice. Hair diameter also is

correlated with size of the rostral groove, which chewing lice use to grasp

hair shafts of pocket gophers. In chapter two I examine whether hair

diameter is the resource by which coexisting chewing louse species partition

available habitat. Although the pattern of spatial partitioning was clearly

evident, hair diameter does not seem to be the mechanism by which lice

partition their habitat. Alternative resources could include temperature,

humidity, or the distribution of sebaceous glands throughout the gopher

pelage. Chapters three, four, and five investigate the bacteria associated with

the pocket gopher-chewing louse system. I used the culture-independent

method of amplifying DNA sequences to determine the identification of

bacteria extracted from samples of chewing lice. This method yielded

approximately 35 distinct lineages of bacteria associated with this system. I

also amplified DNA sequences of bacteria from the eggs of chewing lice to

v iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. determine whether any bacteria were transferred vertically in insects

through the eggs. Several species of bacteria were repeatedly seen in these

amplifications, which could be explained by the incorporation of bacteria in

the insect eggs. Further study would be required to provide conclusive

evidence of direct vertical transmission of bacteria in chewing louse eggs.

The final chapter of this dissertation looks in detail at two groups of bacteria

associated with chewing lice (gamma-Proteobacteria and Staphylococcus

species). Complete sequencing of thel6S rRNA gene demonstrated that

multiple species of both groups are associated with the pocket gopher-

chewing louse system. Studies of cophylogeny between the species of

Staphylococcus and their hosts were inconclusive and require more

intensive taxon sampling.

ix

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

So, naturalists observe, a flea

Has smaller fleas that on him prey;

And these have smaller still to bite 'em;

And so proceed ad infinitum.

Jonathan Swift (1667-1745) in Poetry, a Rhapsody.

Since 1984, Mark Hafner and his colleagues have studied cophylogeny

(also termed cospeciation) in a group of (pocket gophers) and their

insect ectoparasites (chewing lice; for a review of this work see Hafner,

Demastes, Spradling, and Reed in press). One portion of my research

explores the possible mechanisms that may have reinforced if not caused the

pattern of cospeciation documented in this host-parasite system. In chapter

one, I examine the relationship between mammalian hair diameter, body

mass, and chewing louse rostral groove diameter. I show a significant,

positive allometric relationship between mammalian hair diameter and

body mass. I also show a significant positive relationship between pocket

gopher hair diameter and the rostral groove dimensions of chewing lice.

Lice use this rostral groove to grasp hairs of their host. Coupled with

previous evidence of a strong allometric relationship between rostral groove

width and louse body size, these findings suggest that hair diameter of the

host is an important determinant of body size in chewing lice. When

1

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. viewed in light of the extreme host specificity of chewing lice (Reed and

Hafner 1997), one can postulate that the fit between gopher hair diameter

and louse rostral groove diameter may reinforce patterns of cospeciation.

Certain taxa of chewing lice of the genera Geomydoecus an d

Thomomydoecus coexist on pocket gophers of the genus Thomomys. In

chapter two, I investigate the spatial distribution of coexisting louse species

on their hosts and explore possible mechanisms of resource partitioning by

chewing lice. Chewing lice appear to partition available host resources

spatially, with Geomydoecus occurring primarily on the lateral and dorsal

regions of the host, and Thomomydoecus occurring primarily on the lateral

and ventral regions. Although spatial partitioning of the host habitat is

evident, it does not appear to be explained by hair diameter. Ecological and

behavioral interactions between these two genera of chewing lice may have

important effects on gopher-louse cospeciation.

My research extended the gopher-louse study of cophylogeny to a

third taxonomic group, the endosymbiotic bacteria of chewing lice. The

asociality and patchy distribution of pocket gophers, compounded by the

extreme isolation of chewing louse populations on individual hosts, suggest

that the bacterial endosymbionts of chewing lice may show patterns of

geographic variation that mirror those of their hosts. Many biologists

assume, either implicitly or explicitly, that microbial species are ubiquitous

(or nearly so) and that chance plays a major role in determining microbial

community structure at any specific locality (e.g. Ward et al. 1998). Recent

2

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 16S rRNA-based molecular surveys of microbial communities seem to

support the assumption of bacterial ubiquity, in that organisms with

similar—sometimes identical—16S rRNA sequences have been recovered

from geographically distant localities (Staley 1999).

My research has examined this assumption in the context of a well-

defined host-parasite system by testing the alternative hypothesis that

bacterial lineages show phylogenetic structuring that parallels that of their

hosts (cophylogeny). Chapter three is an examination of the diversity of

bacteria found within chewing lice that parasitize pocket gophers. Many of

these bacteria likely are short-term transients in the gopher-louse system,

perhaps acquired by the lice through diet. In contrast, other bacteria may be

obligate endosymbionts that perform functions necessary for louse survival.

Many endosymbiotic bacteria are transferred from female insects to their

progeny via the insect ovum. Chapter four describes the search for bacteria

in the eggs of chewing lice. Transovarial inoculation of insects is a complex

mechanism that likely evolved slowly. Therefore, bacterial species found in

the eggs of chewing lice likely are long-term associates of the lice and may

show patterns of cophylogeny with their louse hosts. The fifth, and final

chapter, is a study of cophylogeny between chewing lice and a selected group

of endosymbiotic bacteria of the genera Staphylococcus and Acinetobacter.

The results of this study are discussed in the context of those from the

previous four chapters in a closing section titled "Summary and

Conclusions".

3

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

MAMMALIAN HAIR DIAMETER AS A POSSIBLE MECHANISM FOR HOST SPECIALIZATION IN CHEWING LICE

INTRODUCTION

Body size of host and parasite often are correlated positively. Harvey

and Keymer (1991) and Morand et al. (2000) demonstrated this trend for

chewing lice (Phthiraptera: Trichodectidae) and their pocket gopher hosts

(Rodentia: Geomyidae). Specifically, Harvey and Keymer (1991) used the

comparative method to demonstrate that increased body size in pocket

gophers is associated invariably with increased size of their ectoparasites.

They suggested that lice grow larger on larger hosts because those hosts

presumably live longer and allow their lice more time to grow. Although

intriguing, this explanation seems unlikely, given that generation time of

chewing lice (about 40 days; Rust 1974) is almost an order of magnitude less

than generation time of even the shortest-lived species of pocket gopher

(about one year; Nowak 1999).

Other potential explanations exist for the body-size correlation

documented by Harvey and Keymer (1991). For example, ability of the host

to detect and destroy ectoparasites may scale with host body size and thereby

place an upper limit on body size of the parasite. Evolutionary changes in

the body mass of the host also might alter habitat of chewing lice in terms of

temperature, humidity, hair length and diameter, and other habitat

4

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. parameters. Because lice are extremely host specific (Price and Emerson

1972; Reed and Hafner 1997) and inextricably tied to their host for survival

(Kellogg 1913; Marshall 1981), it seems likely that lice would show finely

tuned adaptations for life on their host. Characteristics of the hair,

particularly hair diameter, should be important components of the

environment of the louse (Fig. 1.1) because of the pivotal role that hair plays

in louse feeding, locomotion, ovipositing, and survival (Murray 1957a;

1957b).

Mammal pelage consists of two basic types of hairs: guard hairs,

which are relatively long and thick, and wool hairs (or underfur), which are

shorter, thinner hairs (Mayer 1952). My observations of chewing lice reveal

that lice spend most of their time moving among guard hairs and are

seldom seen attached to wool hairs. Accordingly, I focused this

investigation on mammalian guard hairs. I examined the relationship

between the diameter of guard hair and body mass at several taxonomic

levels in mammals (interordinal, intrafamilial, intrageneric, and

intraspecific) to determine the generality of the hair size-body size

relationship in mammals. I also investigated whether hair diameter of the

host was correlated with host and parasite body size in geomyid rodents.

MATERIALS AND METHODS

A single adult individual from each of 18 species of non-geomyid

mammals (representing 17 families in 9 orders; Appendix I) was

5

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Fig. 1.1. Electron micrograph (left) of a chewing louse {Geomydoecus aurei) attached to the hair shaft of a pocket gopher (Thomomys bottae). Magnified view (right) of rostral groove and hair shaft.

6

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. examined to assess variation in hair diameter across different orders of

mammals. Taxa were selected randomly from the Louisiana State

University Museum of Natural Science Collection of Mammals, except that

species with conspicuous spines or quills were avoided and range of body

size was maximized as much as possible (Appendix I). Because of difficulty

in sampling so many widely divergent taxa, some specimens were not

collected from the wild (Appendix I) and season of collection was not

standardized.

Guard hairs (n = 20 per individual) were removed from the nape region of

museum study skins. The nape region was selected to standardize the

sampling procedure and reduce the likelihood of damage to haiirs from

grooming. Hairswere mounted on microscope slides with Permount® and

secured with cover slips. Mathiak (1938) determined that the greatest

diameter of most mammalian guard hairs was found roughly one-half the

distance from the root to the tip. Accordingly, I measured hair diameter

about midway between the root and tip by using a light microscope fitted

with an ocular micrometer scale. Mass of each in this study was

taken directly from the specimen tag or was estimated based on inform ation

provided by Nowak (1999).

Fourteen species in the family Geomyidae (Appendix I) were

examined to assess variation in hair diameter within a single family of

mammals. Guard hairs (n = 200 per specimen) were sampled from

throughout the gopher pelage and prepared for light microscopy. A single

7

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. individual from each of seven species in the genus Thom om ys (A ppendix I)

was examined to assess variation in hair diameter within a single genus of

pocket gophers. Guard hairs (n = 200 per specimen) were sampled from

throughout the pelage and prepared for microscopy. Likewise, eight

individuals from a single species, T hom om ys bottae, were examined to

assess variation in hair diameter within a single species of pocket gopher.

Guard hairs (n = 200 per specimen) were sampled from throughout the

gopher pelage and prepared for microscopy.

Width of the rostral groove was measured for adult lice collected

from the same individual gophers from which hair diameter was measured

and with about equal representation of male and female lice. Louse samples

were: Geomydoecus scleritus (n = 11) fro m pinetus; Geomydoecus

panamensis (n = 11) from Orthogeomys cavator; Geomydoecus setzeri (n = 8)

from O. underwoodi; Geomydoecus aurei (n = 10) from Thomomys bottae;

Thomomydoecus minor (n = 6) from T hom om ys bottae; and Geomydoecus

oregonus (n = 11) from Thomomys bulbivorus. Lice were cleared for light

microscopy by soaking in the following series of solutions (10-20

min/solution): 50% EtOH, 60% EtOH, 70% EtOH, 10% KOH, 80% EtOH, 90%

EtOH, 100% EtOH, and xylene. Lice were mounted on microscope slides,

secured with a coverslip, and allowed to dry for 24 h. Rostral groove width

was measured with a light microscope fitted with an ocular micrometer.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Because comparisons across species boundaries potentially are

confounded by phylogenetic relationships, independent contrasts

(Felsenstein 1985; Harvey and Pagel 1991) were used to remove phylogenetic

effects from these data. The computer program CAIC (Comparative

Analysis of Independent Contrasts; Purvis and Rambaut 1995) uses a

phylogenetic hypothesis to generate independent contrasts of data, which are

then analyzed statistically. Because most mammalian ordinal relationships

are unclear, comparisons among orders of mammals were not transformed

using CAIC. Such comparisons generally are considered independent

because the phylogenetic distance between terminal taxa was large. Hair

diameter and body mass measurements for pocket gophers were

transformed into independent contrasts by using a composite geomyid

phylogeny based on phylogenetic studies by Hafner et al. (1994), Smith (1998),

and Spradling (1997). No phylogenetic hypotheses were available for taxa

below the level of species; therefore, comparisons within Thomomys bottae

were not transformed into independent contrasts. Model II regression

analyses (major axis method) were performed with the SYSTAT statistical

analysis software package (SYSTAT, Inc. 1992). Model II regression is

appropriate when two variables lack a clear dependent-independent

relationship and both are measured with error (LaBarbera 1989; Martin and

Barbour 1989; Silva 1998). Regressions of independent contrasts were

constrained through the origin, as required by CAIC to retain n - 2 degrees of

freedom (Garland et al. 1992; Purvis and Rambaut 1995).

9

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

The regression analysis of hair diameter and body mass for 18 species

representing 9 orders of mammals (Fig. 1.2) revealed a positive relationship

(P < 0.05; Table 1.1). The allometric coefficient of this relationship (a = 0.13 ±

0.03) was low, which indicates that body mass increased more rapidly than

hair diameter among mammals examined, which ranged in body size from

a 3.5-g bat ( Pipistrellns) to a 600-kg bear ( Ursus).

Regression analysis of independent contrasts of hair diameter and

body mass for 14 species of pocket gophers (Fig. 1.3) revealed a similar

allometric trend (a = 0.25 ± 0.05; P < 0.05; Table 1.1). Thus, when analyzed at

the family level (and controlling for phylogenetic relationships within the

family), larger species of pocket gophers tend to have thicker guard hairs.

Examination of the relationship between hair diameter and body

mass within a single genus of pocket gophers (Thom om ys; Fig. 1.4) show ed

the same trend evident in the analyses at higher taxonomic levels. In

individuals representing seven species ofThomomys , hair diameter showed

a significant allometric relationship with body mass (a = 0.20 ± 0.07; P < 0.05;

Table 1.1).

Regression analysis of hair diameter and body mass for eight

individuals of Thomomys bottae (Fig. 1.5) also showed a significant, positive

relationship (a =0.33 ± 0.04; P < 0.05; Table 1.1). Thus, larger individuals of

10

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. <0.05 Cl P 95% 0.24-0.43 <0.05 0.02-0.37 <0.05 0.13-0.36 <0.05 0.06-0.19 0.03 SE 0.13 0.25 0.05 ln(Mass) 0.33 0.04 * Regression model Slope ln(Diam) = B * ln(Mass) 0.20 0.07 ln(Diam) = A + B * ln(Mass) (8) ln(Diam) = A + B (7) Taxon (») Table 1.1. Model II regression analyses (major axis method) of the relationship between hair diameter (Diam) and Thomomys bottae Thomomys Geomyidae (14) ln(Diam) = B * ln(Mass) Non-geom yid Mammalia (18) listed in Appendix I). body mass (Mass) at four taxonomic levels in mammals (depicted graphically in Figs. 1.2 through 1.5; taxa examined are

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

=4.s 4

S-, jU 3 "S 6 .5 ■5 t - i

1 2 4 6 8 1 0 1 2 1 4

Body mass (in In g)

Fig. 1.2. Major axis regressions performed on ln(mass) and ln(diameter) for 18 species of mammals. Regression models are in Table 1.1, and taxa examined are listed in Appendix I.

12

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 .3

re ■ J G u - i — .. ° g 0.2 cni C302 r-C U i — ■ c c o cr- ° fll u 0.1 g a -o £ fljc »p^ is O.T5

0.1 0 0.1 0.2 0 .3 0 .4 0 .5

Independent contrast of body mass (in In g)

Fig. 1.3. Major axis regression through the origin of independent contrasts of ln(mass) and ln(diameter) for members of the family Geomyidae. Regression models are in Table 1.1, and taxa examined are listed in Appendix I.

13

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. • • 0.08 -

04 -

0

TJ -0.04 -

-0.08 0 0.1 0.2 0.3 0.4 0.5

Independent contrasts of body mass (in In g)

Fig. 1.4. Major axis regression through the origin of independent contrasts of ln(mass) and In(diameter) for pocket gophers in the genus Thomomys. Regression models are in Table 1.1, and taxa examined are listed in A ppendix I.

14

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 4

3.5

3

2.5 3 4 5 6 Body mass (in In g)

Fig. 1.5. Regression (major axis method) performed on ln(mass) and In(diameter) for eight individuals of Thomomys bottae. Individuals designated by squares (collected in native desert-scrub habitat) and triangles (collected in alfalfa fields) are those analyzed by Patton and Brylski (1987) and Smith and Patton (1988). These individuals are discussed further in the text. Regression models are in Table 1.1, and taxa examined are listed in Appendix I.

15

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. T. bottae tend to have thicker guard hairs than smaller individuals of the

same species.

The regression analysis of louse groove width and gopher hair

diameter (Fig. 1.6) revealed a positive relationship between those variables

(P < 0.05). The regression coefficient (a = 1.09 ± 0.09) suggested a near­

isometric relationship between groove width and hair diameter, which

indicates that the two structures varied proportionately. Importantly, the Y-

intercept of the regression (Fig. 1.6) w as close to zero ((3 = -3.9 /xm), w h ic h

suggests that width of the rostral groove of a chewing louse was very similar

in actual dimensions to maximum width of the guard hairs of its host.

DISCUSSION

Hair Diameter and Body Size

My analyses document a consistent negative allometric relationship

between hair diameter and body mass in mammals, regardless of the

taxonomic level examined. The low allometric coefficient of this

relationship (ranging from a = 0.13 to 0.33; Table 1.1) indicates that larger

mammals tend to have guard hairs that are larger in absolute diameter but

proportionately smaller than guard hairs of smaller mammalian species.

This consistent relationship between hair diameter and body size in

mammals is reminiscent of the negative allometric relationship observed

for many other mammalian features that scale with body size (e.g., brain

size, longevity, metabolism) and suggests that in most species of mammals

hair diameter may be constrained within certain boundaries by simple

16

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

70 • v*-O 'e' £ 03U , 3 — e 03 E 3 C8 O I—I

20 20 40 503060 70 80 90 Hair Diameter of Pocket Gopher (^m)

Fig. 1.6. Linear regression (general linear model) performed on chewing louse groove width and pocket gopher hair diameter for six species of lice collected from five species of pocket gophers Geomys ( pinetus, Orthogeomys cavator, O. underxvoodi, Thomomys bottae, and T. bulbivorus). Regression models are shown in Table 1.1, and taxa examined are listed in Appendix I.

17

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. growth laws. This is not to suggest, however, that hair lacks important

functional significance in most mammalian species or that adaptation has

not played an important role in evolution of specialized hairs, such as

spines or quills (specifically excluded from this analysis). Although hair

diam eter, per se, may be rigidly constrained in many (or most) species of

mammals, other aspects of mammalian pelage, including detailed

microstructure of the hair, hair length, shape, color, and density (number of

hairs per follicle and follicle density) are less likely to scale with body size

and, thus, may have evolutionary flexibility in a wide variety of adaptive

contexts.

The analysis of hair diameter and body-mass relationships within a

single species of pocket gopher (T. bottae; Fig. 1.5) em phasizes the tig h t

linkage between these two variables in this species. This relationship

suggests that a change in mean body mass within a lineage of pocket gophers

over time (e.g., in response to climate change or an increase or decrease in

food resources) will be accompanied by a corresponding change in hair

diameter. To test this prediction, I included in my analysis six individuals of

T. bottae that also were examined by Patton and Brylski (1987) and Smith

and Patton (1988). Those studies compared body-size relationships between

gophers collected from native desert-scrub habitat between 1937 and 1971

(indicated by squares in Fig. 1.5) to those collected from the same geographic

region during 1984 and 1985, many years after the native vegetation had

been converted to irrigated alfalfa fields (indicated by triangles in Fig. 1.5).

18

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Smith and Patton (1988) showed that the pocket gophers collected in the

alfalfa fields (presumably direct descendants of the populations sampled

decades earlier in native habitat) were significantly larger than their

ancestors in overall body mass and other body dimensions. This increase in

body mass is evident in my analysis (Fig. 1.5), and it is clear that the increase

in overall body size in these pocket gophers was accompanied by an increase

in hair diameter.

Hair Diameter and Parasite Size

Larger pocket gophers host larger chewing lice (Harvey and Keymer

1991; M orand et al. 2000), and the above analyses corroborate the prelim inary

evidence presented by Morand et al. (2000) showing that larger pocket

gophers also have thicker guard hairs. To establish a meaningful

connection between hair diameter and louse body size, it is important to

focus on some aspect of the body of the louse that interacts directly with the

hair of the gopher. For this, I chose the rostral groove of the chewing louse,

which is used to grasp the hair of the gopher (Fig. 1.1). Given that the rostral

groove is a rather rigid structure, I predicted a close fit between rostral

groove width in chewing lice and hair diameter in pocket gophers under the

assumption that a louse with a very narrow rostral groove would be unable

to grasp a thick hair, whereas a louse with a very wide groove may slip from

a narrow hair. Murray's (1957a, 1957b) studies of chewing lice ( Damalinia )

on sheep (Ovis) showed that lice kept in laboratory colonies in which thin

glass fibers were used as artificial hairs were extremely sensitive to width of

19

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. the glass fibers. If fibers were too thick, lice were unable to grasp them

between the gonopod and abdomen and could not lay eggs. Similarly,

differences in hair shape (and perhaps diameter) between humans of

African descent and those of European descent have been used to explain

reduced susceptibility of Africans to head lice (Pediculits humanus) from

Europe and reduced susceptibility of Europeans to head lice of the same

species from Africa (United States Centers for Disease Control 1984).

Presumably, lice in the two regions have evolved differential abilities to

grasp flat hairs typical of humans of African descent versus round hairs

typical of humans of European descent.

My results corroborate preliminary findings of Morand et al. (2000),

which show a close fit between hair diameter in pocket gophers and rostral

groove width of chewing lice when analyzed at interspecific and intergeneric

levels (Fig. 1.6). This finding, coupled with Morand et al.'s (2000) discovery

of a relationship between body size and groove width in chewing lice,

suggests that interspecific variation in body size of lice may be determined

largely by interspecific variation in diameter of gopher hair. If small species

of lice (with narrow rostral grooves) are unable to grasp thick hairs of large

pocket gophers and if large lice tend to avoid hosts with thin hairs, this

could help explain the high level of host-specificity observed in chewing lice

at zones of contact between gopher species of very different body sizes (M. S.

Hafner, pers. comm.). This potentially obligate relationship between groove

size and hair diameter also may explain why certain species of lice are

20

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. unable to survive on certain species of hosts in laboratory transfer

experiments. For example, experiments by Reed and Hafner (1997) have

shown that chewing lice transferred between species of pocket gophers of

similar body size often are able to establish successful breeding colonies on

foreign (non-native) hosts. Lice that occur naturally on a large species of

host occasionally are able to survive and reproduce on a smaller species of

host, but the reverse does not seem to be true. This suggests that the rostral

groove of lice from the smaller hosts may be too narrow to grasp thick hairs

of the larger host species. A similar study of body-size relationships between

bird lice and their hosts has shown that feather size is a crucial factor in

determining success of lice experimentally introduced onto new host taxa

(D. Clayton, pers. comm.).

Preliminary analyses of individual, sexual, and ontogenetic variation

in hair diameter in pocket gophers (chapter two) suggest that variation at

these levels may be within the normal range of tolerance for louse rostral

grooves. This pattern would explain why a young pocket gopher does not

"outgrow" its parasites as the gopher increases in body size (and, thus, hair

diameter) ontogenetically. The pattern also would explain why male and

female pocket gophers, which often show marked sexual dimorphism in

size (Patton and Brylski 1987), nevertheless host the same species of chewing

louse.

My analyses demonstrate a negative allometric relationship between

hair diameter and body mass among mammals when analyzed at multiple

21

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. taxonomic levels. This relationship is evident in the family Geomyidae,

even at the intraspecific level. These data also reveal a significant positive

relationship between hair diameter of pocket gophers and rostral groove

width of their chewing lice. In fact, I show a nearly exact fit between hair

diameter and groove width for host-parasite pairs. Given that lice

eventually die if removed from their host, it seems likely that groove width

would be under strong selective pressure to conform to host hair diameter.

If groove width in lice scales with overall body size, then selection for

optimal groove width could indirectly constrain body size in lice. However,

the causal mechanism driving the positive relationship between host and

parasite body size (as originally documented by Harvey and Keymer 1991)

remains untested. Further studies involving direct observation of chewing

lice transferred to non-native hosts may elucidate the causal mechanism

underlying the empirical observation that larger species of pocket gophers

tend to host larger species of chewing lice.

22

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

SPATIAL PARTITIONING OF HOST HABITAT BY CHEWING LICE OF THE GENERA GEOMYDOECUS AND THOMOMYDOECUS (PHTHIRAPTERA: TRICHODECTIDAE)

INTRODUCTION

Ectoparasitic chewing lice, Geomydoecus and Thomomydoecus spp.

(Phthiraptera: Trichodectidae), live their entire lives exclusively on pocket

gophers of the family Geomyidae (Marshall 1981; Hellenthal and

Price 1984). Chewing lice are wingless, obligate parasites that can survive

only a short time when removed from their host (Kellogg 1913; Marshall

1981). One species of louse often is confined to a single species of host

(Emerson and Price 1981), which suggests a long-term, perhaps obligate,

association between each host-parasite pair. In many instances, this long­

term association has resulted in parallel cladogenesis between the pocket

gopher and chewing louse lineages. This pattern, termed "cophylogeny," is

well documented for certain lineages of pocket gophers and their associated

chewing lice (Hafner and Nadler 1988; Demastes and Hafner 1993; Hafner et

al. 1994).

Although most individual pocket gophers host populations of a

single species of louse, three species within Thomomys host representatives

of two genera of chewing lice (Geomydoecus and Thomomydoecus;

Hellenthal and Price 1984), with species of both genera usually coexisting o n

an individual host. The principle of competitive exclusion (Gause 1934),

23

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. suggests that stable coexistence of two species of lice on an individual host

relies on partitioning some aspect of the resources provided by the host

(Durden 1987).

During my initial studies of chewing louse distribution on pocket

gophers, I noticed that guard hair diameter seems to vary predictably with

body region, and perhaps provides a mechanism for resource partitioning by

chewing lice. Lice attach to gopher hairs by means of a head groove located

on the rostrum (Fig. 1.1). Secure attachment is necessary for survival of the

parasite because of the louse's absolute dependence on the host. In chapter

one of this dissertation I show a significant positive relationship between

hair diameter in several genera of pocket gophers and rostral groove width

of their chewing lice, and I also demonstrate that the head groove of a louse

is of the appropriate size to grip tightly onto the hair shaft of its natural host.

In the present study, I investigate hair diameter as a potential

mechanism for resource partitioning that may result in stable coexistence

between two species of chewing lice on an individual pocket gopher.

Specifically, I test the hypothesis that louse species of Geomydoecus and

Thomomydoecus are able to coexist by partitioning the host's resources

spatially based on hair diameter.

MATERIALS AND METHODS

Three specimens of Thomomys bottae connectens w ere trapped on 17

March 1997 in Albuquerque, Bernadillo County, New Mexico. Two males

and one female were collected with traps designed by Baker and Williams

24

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. (1972). Pocket gophers were collected on a single day from a single locality to

limit variation in louse population density caused by weather patterns,

reproductive condition of the host, or other seasonal or geographic factors.

Each gopher was killed by placing it in an air-tight container saturated

with chloroform. This process also immediately immobilized and killed the

resident louse population of each gopher. The gopher was then incised

medially along the abdomen, and the entire skin was removed and pinned

to a piece of cardboard, with the fur side down. Unnecessary movement of

the skin was avoided to reduce accidental displacement of the lice. The skin

was frozen on a block of dry ice, then pressed between two pieces of

cardboard, wrapped tightly in aluminum foil and frozen in an ultracold

freezer (-75° C).

W hile frozen, the gopher skin was cut into 10 regions (Fig. 2.1):

anterior ventral, cheek, dorsal head, lateral nape, lateral, nape, posterior

dorsal, posterior ventral, rump, and ventral head. Samples from the right

and left sides of the body were pooled for each region, and each region was

placed individually in a plastic bag to avoid loss of lice and contamination by

lice from other regions. Each section of the gopher pelt was brushed

vigorously, and lice were collected in a 1.5-ml cryotube. Adult lice were then

identified, using a dissection microscope, as either Geomydoecus aurei or

Thomomydoecus minor. Only adults were used in this analysis because

visual identification of juvenile lice is problematic. Each pelage region was

measured (in cm2) so that the number of lice per region could be

25

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. lice. Thomomydoecus Head Ventral and Ventral Ventral Anterior Posterior 'Literal Geomydoecus Nape Rump Head Dorsal Dorsal Posterior Check Nape Lateral Lateral Ventral Ventral Anterior Head Posterior Ventral Head Ventral Ventral Ventral Anterior Posterior Nape Lateral Lateral Check Rump Head Nape Dorsal Dorsal Posterior Ihcck Nape Lateral Lateral Geomydoecus Thomomydoecus Ventral Ventral Head Anterior Posterior Ventral Figure 2.1. External surface of a pocket gopher showing the distribution of Darkly shaded regions contained morecontained lice than fewer expected lice than in allexpected three pocketin at leastgopher 2 of specimensthe pocket examinedgopher specimens(Table 2.1). examined. Lightly shaded regions contained more lice than expected in 2 of the 3 pocket gopher specimens examined. Unshaded regions

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. standardized (lice/cm 2). The total surface area (the 10 regions combined) and

overall louse density (lice/cm2) were calculated for each of the three gopher

specimens. If one assumes a null model of even louse distribution, the

expected number of Geomydoecus and Thomomydoecus for each region

based on the size of the region and the mean density for that particular

gopher was determined. Chi-square analyses were used to test whether the

observed numbers of lice were significantly different from expected numbers

based on an even distribution.

Ten guard hairs were taken from each of the 10 regions from all three

gophers (chewing lice normally grasp guard hairs rather than underfur; pers.

obser.). The hairs were mounted on microscope slides, and the mid-point

diameter of each hair was measured (in ym) with a light microscope fitted

with an ocular micrometer. Analysis of variance (ANOVA) and a Duncan

post-ANOVA test (SAS Institute 1994) were used to detect significant

differences in mean hair diameter among the ten regions. Specimens are

deposited in the New Mexico Museum of Natural History (NMMNH 2378,

2379, and 2380).

RESULTS

Mean hair diameter (pooled data for the three hosts) varied from

33.61 ± 1.75 ym in the rump region to 45.22 ± 1.53 ym in the lateral-nape

region (Table 2.1). The Duncan post-ANOVA test identified five groups of

regions within which mean hair diameter was not significantly different

27

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. (Table 2.1). In general, pocket gopher hair was smaller in diameter on the

dorsal surfaces and larger in diameter on lateral and ventral surfaces.

The overall mean density of Geomydoecus aurei (pooled data for all

three hosts; Table 2.2) was 0.34 lice/cm2 (range 0.28-0.50 lice/cm2), and

overall mean density for Thomomydoecus minor was 0.94 lice/cm2 (range

0.28-1.38 lice/cm2). Chi-square analysis of louse distribution revealed that

none of the populations of either species was distributed evenly over the

gopher pelage (all chi-square values exceeded the critical value of 27.88,

P<0.001, df = 9; Table 2.2).

Individuals of Geomydoecus aurei were found in all 10 regions of the

gopher pelage, although some regions contained very few individuals (see

pooled data, Table 2.2). Likewise, at least one individual of

Thomomydoecus minor occurred in all regions except the dorsal-head

region. Geomydoecus aurei occurred in greater abundance than expected on

dorsal and lateral surfaces of the hosts (Fig. 2.1a), whereas Thomomydoecus

minor was found in greater abundance than expected on lateral and ventral

surfaces (Fig. 2.1b). For Geomydoecus, the regions of high abundance

(shaded regions in Fig. 2.1a) comprised only 34% of the total surface area of

the gopher yet contained 78% of all Geomydoecus individuals (Table 2.2). In

contrast, the regions of high abundance for Thomomydoecus (shaded

regions in Fig. 2.1b), which also comprised 34% of the total surface area of

the gopher, contained only 55% of all Thomomydoecus individuals.

28

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table 2.1. Mean guard hair diameter (pm) for each of the ten body regions in the three pocket gopher specimens examined. The Duncan post-ANOVA test reveals five groups (designated A-E) within which mean hair diameter was not significantly different. For each region, louse taxa found in greater abundance than expected (Fig. 2.1) are indicated.

Mean hair Duncan Greater abundance Region diameter (pm) grouping than expected

Lateral nape 45.22 ± 1.53 A Geomydoecus aurei

Cheek 44.90 ± 1.76 A Geomydoecus aurei

Posterior ventral 42.66 ± 1.28 A B Thomomydoecus minor

Anterior ventral 41.40 ± 1.60 AB Thomomydoecus minor

Lateral 39.53 ± 1.83 B both species

Ventral head 38.72 ± 1.51 B C neither species

Dorsal head 38.31 ± 1.54 B C D neither species

Nape 35.23 ± 1.53 C D E Geomydoecus aurei

Posterior dorsal 34.32 ± 1.67 D E Geomydoecus aurei

Rump 33.61 ± 1.75 E neither species

29

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table 2.2. Total surface area (cm2) of each pelage region (Fig. 2.1) and number of observed and expected lice per region for the two coexisting species of chewing lice. Data are pooled for the three hosts examined.

Thomomydoecus minor Geomydoecus aurei

Region Area (cm2) Observed Expected Observed Expected

Anterior ventral 58.89 80 59 11 20

Cheek 40.38 2 37 12 14

Dorsal head 35.63 0 35 6 14

Lateral nape 33.01 21 26 29 11

Lateral 56.01 128 50 26 18

Nape 30.01 16 29 14 10

Posterior dorsal 35.25 28 35 62 13

Posterior ventral 67.00 71 62 12 21

Rum p 165.75 153 151 4 54

Ventral head 27.88 17 32 9 10

Total 549.78 516 516 185 185

30

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Therefore, differences are large in both density and distribution of the two louse

taxa (Fig. 2.2).

DISCUSSION

Geomydoecus aurei and Thomomydoecus minor are not evenly distributed

throughout the pelage of their host (Thomomys bottae) and show a tendency to

subdivide the available habitat dorsoventrally (Figs. 2.1, 2.2). Future studies will

determine whether this pattern changes geographically, seasonally, or with the age

or reproductive condition of the host. When one considers that lice showed the

same distributional pattern on the male and female hosts examined in this study,

sex of the host does not appear to influence louse distribution, at least for non-

reproductive hosts (as in this study).

Given that both genera of lice were found throughout the gopher pelage

(with the single exception of the absence of Thomomydoecus in the dorsal-head

region; Table 2.2), it is clear that the taxa are interactive (as defined by Brooks 1980)

and are able to transit, if not forage and reproduce in, regions of pelage with hairs

of different diameters. In fact, regions of high abundance for Geomydoecus aurei

encompass almost the entire range of hair diameters (from 34.32 ym to 45.22 ym,

Table 2.1). In contrast, regions of high abundance for Thomomydoecus minor

include only regions with hairs of intermediate diameter (from 39.53 ym to 42.66

ym, Table 2.1).

Despite the general dorsoventral trend in hair diameter (Table 2.1) and a

similar dorsoventral trend in louse distribution (Table 2.2; Fig. 2.1), the broad

overlap in hair diameters used by the two species (Table 2.1)

31

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. without prohibited reproduction Further owner. copyright the of permission with Reproduced pelage. Asterisks indicate regions that contained more lice than expected in all 3 all in expected than 2.1). Fig. lice 2.1, more (Table ined contained exam that specimens regions gopher indicate pocket Asterisks pelage. Louse density (lice/cm 2) 0 2 1.5 Figure 2.2. Comparative density of lice in the 10 regions of pocket gopher gopher pocket of regions 10 the in lice of density Comparative 2.2. Figure . . 5 5 - - * 03 •-t 0 03 > n O ^ p •n ' o 3 •-* o < 03 Li ' o 3 0 ^ p O rD C/3 o 3 - ^ p *0 < ■-n Li Regions of pocket gopher pocket of Regions -a f* 3 CO ’ o CO O o i S l-t •t .-r- *73 CL 32 D T ro S3 3 S3 ■-I re S3 n 03 . O ft> 03 O 3* n P^ 3 < | I [ | re S3 z hmmdeu minor Thomomydoecus emdeu aurei Geomydoecus 3* n re T 7 re o e r S3 CO . a S3 a —1 I I

suggests that hair diameter, alone, is insufficient to explain habitat

partitioning in this louse community. It is possible, however, that

Thomomydoecus lice are less efficient than their competitors at grasping

hairs of extremely large or small diameter, yet are superior competitors in

regions of intermediate hair diameter. It is also possible that the two louse

species differ in their ability to evade grooming pressure from the host,

which may vary dorsoventrally. Waage (1979) suggested that areas of

overlap in the distribution of co-occurring ectoparasites may receive greater

grooming pressure from the host because of higher overall density of

parasites. Waage contended that removal of parasites from these areas of

overlap (by grooming) would reinforce spatial partitioning of the

ectoparasites by removing the ability of parasites to cross corridors between

areas of exclusive habitation. However, my observations of captive pocket

gophers suggest that they groom only infrequently, and I think it is more

likely that the habitat partitioning observed in this study results from

differential responses on the part of the two louse species to other

microhabitat features such as temperature or humidity gradients or location

and density of sebaceous glands of the host (Murray 1957a; 1957b). It seems

reasonable to postulate that the dorsal and ventral surfaces of a pocket

gopher constitute very different microhabitats, and this hypothesis will be

examined in future studies. In addition, future studies of Thomomys bottae

populations that host only one of these two species will reveal the extent to

33

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. which competition may be influencing the distribution of these species

when they coexist.

The fact that hair diameter may have little of no influence on louse

distribution at the level of the individual host does not automatically falsify

the hypothesis that hair diameter may be an important causal factor

influencing louse distribution at higher phylogenetic levels, e.g., among

different species and genera of hosts (Reed 1994; Page and Hafner 1996). For

example, I have shown (chapter one) dramatic differences in hair diameter

among different genera of pocket gophers and have documented a close

relationship between hair diameter in the hosts and rostral groove

dimensions of their chewing lice. Artificial transfer studies by Reed and

Hafner (1997) suggest that lice that normally parasitize species of pocket

gophers with narrow hairs may be unable to grasp the wider hairs of larger

species of pocket gophers. Finally, studies by Murray (1957a; 1957b) show

that hair diameter in sheep may influence ovipositing in their chewing lice.

Together, these studies suggest that hair diameter may be a coarse-grained

determinant of chewing louse distribution, wherein lice are unable to

transfer between hosts with large differences in hair diameter (e.g., Reed and

Hafner 1997), but are tolerant of lower levels of variation, such as those

observed at the individual and intraspecific host levels. It follows that some

other, as yet unknown, environmental parameter, perhaps temperature or

humidity, may be the fine-grained determinant of louse distribution at the

individual and intraspecific host levels. If so, differential responses to these

34

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. factors by different species of lice may enable stable coexistence of multiple

species on a single host individual.

35

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

BACTERIAL DIVERSITY IN CHEWING LOUSE HOSTS

INTRODUCTION

The Hosts: Pocket Gophers of the Rodent Family Geomyidae

Pocket gophers (Rodentia: Geomyidae) are fossorial rodents whose

geographical range extends from southern Canada through northwestern

Colombia. The family consists of six genera and approximately 400

recognized species and subspecies. Evolutionary relationships among

genera of pocket gophers are well characterized (Hafner 1982; Hafner and

Nadler 1988; 1990; Honeycutt and Williams 1982). These rodents are

extremely asocial and live in isolation except during brief mating

encounters. Each individual gopher occupies an extensive burrow system

that is plugged with soil at all entrances, and populations of gophers are

small and isolated from other conspecific populations. Different species of

pocket gophers are rarely found in the same region, and then only along

narrow zones of parapatry. These natural history characteristics prevent

widespread transfer of parasites among individuals in a population, among

conspecific populations, or among species.

The Parasites: Chewing Lice of the Phthirapteran Family Trichodectidae

Chewing lice of the genera Geomydoecus and Thomomydoecus

(Phthiraptera: Trichodectidae) are restricted to pocket gopher hosts.

36

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Chewing lice are small, hemimetabolous insects no greater than 1 mm in

length. They feed on skin detritus and the glandular secretions of pocket

gophers (Marshall 1981). Trichodectids are obligate ectoparasites that spend

their entire life cycle on the host. The 120 recognized species and subspecies

of these two genera are partitioned into four subgenera and 26 species

complexes (Hellenthal and Price 1991; 1994). The morphological

of trichodectid lice has been studied extensively (e.g., Hellenthal and Price

1984; Price et al. 1985). The most recent and comprehensive systematic

treatment of trichodectid lice is that of Page et al. (1995), in which 58

morphological characters from adults and first instars were analyzed

cladistically. The louse phylogeny generated by Page and his colleagues

supports the traditional species complexes proposed by R. Hellenthal and R.

Price over the past two decades. This phylogeny also is consistent with

molecular-based louse phylogenies proposed by Hafner and Nadler (1988)

and Hafner et al. (1994). Previous studies have shown that these lice exhibit

a high degree of host specificity (Patton et al. 1984; Timm 1983). Reed and

Hafner (1997) showed that chewing lice can survive on non-native hosts,

but their ability to colonize new hosts diminishes as the phylogenetic

distance increases between the natural host and the surrogate host.

Transmission of chewing lice is thought to occur only through direct host-

to-host contact (Tim m 1983). H ow ever, D em astes et al. (1998) w ere able to

show that transmission was not strictly maternal. The combination of low

parasite vagility and obligate contact-transmission of lice limits

37

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. opportunities for colonization of new host species. The aforementioned

life-history characteristics of pocket gopher hosts further retard the

colonization of new hosts by chewing lice. The absence of widespread

transfer of lice among gopher species has, in part, led to the pattern of

cophylogeny as documented by Hafner et al. (1994), and results to date

suggest that cospeciation is common in this host-parasite assemblage (e.g.,

Demastes and Hafner 1993; Hafner and Nadler 1988; Hafner et al. 1994; Page

1990b). More vagile species of lice, such as the phthirapteran lice of avian

hosts, show more evidence of host-switching and less cospeciation (e.g.,

Johnson and Clayton in press) than do trichodectid chewing lice associated

with pocket gophers.

The Endosymbionts: Bacteria

The term symbiosis originates from the Greek sym (together) and bios

(life). Smith and Douglas (1987) refined this definition and restricted it to

include only "long-term associations in close proximity." The definition

today encompasses intracellular and extracellular symbionts, and

mutualistic as well as parasitic relationships between symbionts. Douglas

(1989) reviewed the symbioses of insects and bacteria and determined that

approximately 10% of known insect species contain non-parasitic

microorganisms. More recently, Werren et al. (1995) estimated that 17% of

insect species harbor the bacterium Wolbachia. Modem molecular

techniques are demonstrating that Douglas' (1989) figures were vastly

underestimated. It is becoming clear that most insects host multiple species

38

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. of microorganisms that are required for specialized functions in the host,

such as production of vitamins, synthesis of essential amino acids, digestion

of complex foods, and development of offspring during the instar stage.

Prior to my research, endosymbiotic bacteria were not known from

trichodectid lice of pocket gophers. Ries (1931) used traditional microscopy

to search for bacterial endosymbionts in the chewing lice associated with

pocket gophers. He was unsuccessful in finding bacteria, although this is

possibly an artifact of the microscopy techniques available at the time of

study. Eberle and McLean (1983) documented bacterial symbionts in

mycetomes of the human body louse (Ped.icu.lus). These bacteria migrate

from specialized cells located near the louse midgut to lateral oviducts in

adult lice, where they are incorporated into the eggs and passed

transovarially to offspring. Saxena and Agarwal (1985) showed that some

bird lice harbor endosymbiotic bacteria that may aid in the digestion of

keratin-based feathers. My preliminary analyses using PCR, cloning, and

sequencing revealed many bacterial taxa associated with trichodectid

chewing lice. Although the biological interaction between these symbionts

and their host is not yet known, it is likely that one or more of them provide

the louse with tissue-degrading enzymes (lice feed on host skin detritus) or

essential nutrients (Dadd 1985). None of the louse-associated endosymbiotic

bacteria has been cultured outside of its host.

39

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Endosymbiotic Relationships: Insects and Bacteria

Numerous relationships have been described between invertebrates

and endosymbiotic bacteria (Douglas 1989; Ishikawa 1989; Moran and Telang

1998; Nardon and Grenier 1989; Saffo 1992; for review see Baumann and

Moran 1997). One of the best studied endosymbiont systems is that of the

gut-inhabiting protozoa and bacteria of termites (Ohkuma and Kudo 1996;

for review see Brune and Friedrich 2000). Termites (Reticulitermes:

Rhinotermitidae) have an enlarged hindgut (or paunch) that harbors a

complex assemblage of microorganisms (Berchtold et al. 1994; Ohkuma and

Kudo 1996). This intestinal microflora aids in lignocellulose digestion,

methanogenesis, nitrogen fixation, and the maintenance of a low redox

potential to prevent entry of foreign bacteria (Ohkuma and Kudo 1996).

Wasps of the genus Trichogramma host the cytoplasmically inherited

bacterium Wolbachia that usually causes complete parthenogenesis in the

host (Schilthuizen and Stouthamer 1997). The intracellular bacteria found

in cockroaches (multiple families: Blattaria) are a monophyletic subgroup of

the Flavobacter-Bacteroides group that shows evidence of cophylogeny with

its host (Bandi et al. 1994). Aksoy et al. (1995) found mycetome-associated

endosymbionts in tsetse files (Glossina: Glossinidae) that formed a

monophyletic lineage within the y-3 subdivision of the Proteobacteria.

Suffice it to say that a wide diversity of bacterial lineages have evolved

symbiotic relationships with invertebrates, particularly insects.

40

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. The most extensive study of insect-bacterium symbiosis is that of pea

aphids (Homoptera: Aphididae) and their endosymbiotic bacteria of the

genus Buchnera. Unterman et al. (1989) sequenced the 16S rRNA genes for

the primary (p) and secondary (s) endosymbionts of aphids and determined

that the p-symbionts were in the gamma subgroup of the Proteobacteria.

They determined that the association between the symbionts was ancient,

perhaps as old as 420 million years (Myr). Munson et al. (1991) determined

that the s-symbionts formed a monophyletic clade in the Enterobacteriaceae.

The genus Buchnera was described for the p-symbionts of aphids by M unson

et al. (1991). Moran et al. (1993) compared the phylogenetic histories of the

endosymbionts and their hosts and determined that there was complete

congruence between the phylogenies because of a long history of

cospeciation. In addition, Moran et al. (1993) refined the age of the

association to 160-280 Myr based on fossil evidence of the host. This study

also provided the first convincing evidence of rates of molecular evolution

for bacterial lineages (1-2% sequence divergence per Myr). Previously,

Ochman and Wilson (1987) proposed a rate of 0.7-0.8% per Myr, but they had

no fossil evidence on which to base their calibration.

M unson et al. (1993) dem onstrated th at the genom e of Buchnera

closely resembles that of its free-living relatives (e.g., E. coli) in size and

composition and does not show the reduced genome characteristics of other

well-known endosymbiotic bacteria (typically non-essential genes are lost in

endosymbiotic bacteria). Subsequent research by Moran and her colleagues

41

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. has focused primarily on rates of evolution in Buchnera. For example,

Moran et al. (1995) found that the nucleotide substitution rate was 36 times

greater in Buchnera than in its aphid hosts and that Buchnera evolved twice

as fast as other closely related bacteria. Brynnel et al. (1998) investigated rates

of evolution in the tu f gene of Buchnera to refine the calibration of

nucleotide evolution in bacterial lineages. Again, Brynnel and colleagues

found that evolutionary rates in Buchnera were faster than those of closely

related bacteria. More recently, studies of rates of evolution have been

expanded to pairs of closely related sister species outside the aphid system.

For example, Ochman et al. (1999) demonstrated that experimental estimates

of rates of bacterial evolution (based on laboratory strains) were in conflict

with rates estimated from naturally occurring bacteria. Ochman et al. (1999)

examined many lineages of bacteria (simultaneously) and found clock-like

rates of change when the rates were calibrated with information from fossil

pea aphids.

Relatively few individual bacteria are passed transovarially from

female insects to their offspring. This recurring bottleneck increases the

likelihood of endosymbiotic bacteria retaining slightly deleterious

mutations. This likelihood is further increased by low population densities

attributed to the endosymbiotic lifestyle and the absence of genetic

recombination (Buchnera have been shown to have little or no

recombination or horizontal gene transfer; Moran and Baumann 1994).

"Muller's ratchet" refers to a phenomenon in which slightly deleterious

42

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. mutations gradually accumulate in small, asexually reproducing

populations. Moran (1996) described a classic example of Muller's ratchet in

Buchnera and used this phenomenon to explain why Buchnera species

evolve more rapidly than closely related free-living bacteria. The

accumulation of slightly deleterious mutations also was investigated in the

ribosomal RNA of Buchnera and other endosymbiotic bacteria by Lambert

and Moran (1998). They found that accumulations of these mutations

destabilized the secondary structure of ribosomal RNA in all endosymbiont

lineages relative to their free-living relatives. Further, Lambert and Moran

(1998) demonstrated that this instability has evolved separately in each

endosymbiont lineage and thus appears to be a predictable result of the

endosymbiotic lifestyle.

Baumann et al. (1995; 1997) studied the obligate nature of aphid-

endosymbiont associations. They showed that aphids treated with antibiotic

agents grow more slowly, show decreased adult weight, and fail to

reproduce. All require 10 essential amino acids (arginine, histidine,

isoleucine, leucine, lysine, methionine, threonine, tryptophan, valine, and

phenylalanine), which are usually acquired in the 's natural diet.

Douglas and Prosser (1992) demonstrated that tryptophan was lacking in the

pea aphid diet but was produced for them by the endosymbiotic Buchnera.

Many other endosymbiotic bacteria overproduce essential amino acids

which are lacking in the host's diet (Dadd 1985; Aksoy 1995). Buchnera also

shows other features characteristic of the endosymbiotic lifestyle. For

43

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. example, the Buchnera genome contains single copies of genes coding for

rRNA (typical of slow growing bacteria) and have high levels of GroEL (a

chaperon involved in protein folding) which is commonly found at high

levels in endosymbiotic bacteria (Baumann et al. 1997).

Clearly, Nancy Moran, Paul and Linda Baumann, and their

collaborators have provided the scientific community with a model system

for the study of insect-bacterium endosymbioses. Much of the work that lies

ahead in the gopher-louse-bacterium system will provide independent tests

of their findings and offer additional insight into the evolution of complex

endosymbiotic relationships.

Environmental Sampling: A Culture-Independent Method

The two traditional methods for detection and characterization of

endosymbiotic bacteria—light microscopy and culturing—offer little promise

in the gopher-louse system. A previous search using traditional light

microscopy failed to find endosymbiotic bacteria in the trichodectid lice of

pocket gophers (Ries 1931), and most bacteria known today cannot be

cultured in the laboratory. Even if successful, both of these methods are

likely to underestimate the bacterial diversity found within a host.

I used an alternative approach for detection and characterization of

endosymbiotic bacteria, an approach known as environmental sampling.

This approach uses the intrinsic power of the Polymerase Chain Reaction

(PCR) to amplify DNA sequences from bacteria found in an environmental

sample. By extracting and selectively amplifying the DNA found in a gram

44

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. of soil, a milliliter of seawater, or an entire chewing louse, one can generate

billions of copies of bacterial DNA. This technique has been used

extensively in studies of microbial diversity and evolution (Woese 1994; for

review see Pace 1997), and it permits culture-independent detection and

phylogenetic placement of unknown bacteria (Hugenholtz et al. 1998; W ard

et al. 1998). Because this method circumvents the need to culture bacteria in

the laboratory, the investigator is no longer constrained to the relatively few

bacterial taxa that are cultivable. Although culture-independent methods

have broadened our knowledge of bacterial distribution, diversity, and

evolution, the method has certain limitations. For example, several studies

have shown that chimeric 16S rRNA sequence artifacts can be formed when

PCR is used to amplify DNA sequences from mixed populations of bacteria

(Komatsoulis and Waterman 1997; Wang and Wang 1997). Suzuki and

Giovannoni (1996) showed that certain bacterial taxa were more likely than

others to amplify under certain PCR conditions, causing a bias in the kinds

and relative numbers of bacteria detected. Additionally, PCR biases can arise

merely as the result of differences in 16S rRNA gene copy number (Farrelly

et al. 1995; Rainey et al. 1996). T anner et al. (1998) surveyed n e g a tiv e D N A

extraction procedures and documented that ambient contamination was

easily amplified during PCR with the universal bacteria primers typically

used in the culture-independent approach. Tanner and colleagures

cautioned that this contamination is unavoidable and recommended use of

negative controls to identify potential contaminants. Despite each of these

45

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. legitimate concerns, the culture-independent sampling technique holds

great promise for documenting the large number of undescribed bacterial

taxa found in environmental samples (Pace 1997).

This study used the environmental sampling approach to survey

chewing lice for endosymbiotic bacteria. Chewing lice ( Geomydoecus) w ere

collected from eight species of pocket gophers in four genera ( ,

Geomys, Orthogeomys, an d Thomomys). Because DNA extractions from

these specimens contained DNA from both chewing lice and bacteria,

primers specific to bacterial 16S rRNA sequences were used to preferentially

amplify bacterial DNA. Cloning and cycle sequencing generated DNA

sequences for the putative bacterial endosymbionts. These were compared

to known bacterial DNA sequences to identify the unknown bacteria.

MATERIALS AND METHODS

Collection of Specimens

Chewing louse samples were collected from their pocket

gopher hosts by euthanizing the gopher with chloroform in an airtight

canister and then brushing the fur vigorously. All ectoparasites were

collected on aluminum foil beneath the pocket gopher host. The louse

samples were stored in liquid nitrogen, transferred to the laboratory, and

deposited in the LSU Museum of Natural Science Collection of Genetic

Resources. The following lice were collected: Geomydoecus oklahomensis

from Geomys bursarius halli (host number LSUMZ 31463); and G. b. major

(LSUMZ 31448); Geomydoecus scleritus from Geomys pinetus;

46

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Geomydoecus expansus from Cratogeomys castanops (LSUMZ 31455);

Geomydoecus panamensis from Orthogeomys cavator (LSUMZ 29253);

Geomydoecus setzeri from O. cherriei (LSUMZ 29539); Geomydoecus

costaricensis from O. heterodus (LSUMZ 29501); Geomydoecus chapini fro m

O. hispidus (LSUMZ 29232); Geomydoecus setzeri from O. underwoodi

(LSUMZ 29493); and Geomydoecus centralis from Thojjiomys bottae

(LSUMZ 29569).

Extraction Protocol

Adult chewing lice were washed twice in a solution of 400/d saline

EDTA buffer (containing 150mM NaCl, lOmM EDTA, pH=8.0), 10/d of 25%

SDS (sodium dodecyl sulfate), and 5/d of lOmg/ml lysozyme to remove

bacteria from the outer surfaces. The lice were then placed in 1.5ml

microcentrifuge tubes along with 400/d of saline EDTA buffer, and 5/tl

lysozyme (lOmg/ml) and were crushed with sterile micropestles. Two

negative controls were also used; one contained all the reagents used in the

DNA extraction process and the other contained the same reagents plus a

micropestle to insure that nothing was contaminated with bacteria. The

extraction solution was incubated at 37°C for 30 minutes. Five microliters of

proteinase-K (15mg/ml) and 10/d of SDS (25%) were added, and the solution

was incubated at 55°C for 30 minutes.

Genomic DNA was extracted by adding 400/d of phenol/chloroform,

vortexing, and spinning in a microcentrifuge for two minutes at 14,000 rpm.

The supernatant, which contained the extracted DNA, was carefully

47

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. removed, and the extraction procedure was repeated two more times. One

milliliter of 95% ethanol (EtOH) was added, and the solution was incubated

on ice for 24 hours to precipitate the DNA. Centrifugation at 14,000 rpm for

20 minutes concentrated the DNA into a pellet. The EtOH was decanted,

and the microcentrifuge tubes were dried in a vacuum centrifuge for 15

minutes. The resulting pellets were resuspended in 50/d of TE buffer.

Polymerase Chain Reaction

The polymerase chain reaction was used to amplify copies of bacterial

16S rRNA from genomic DNA. Amplifications were performed using

primers 27-f (5'-GAG TTT GAT CCT GGC TCA G-3') and 1525-r (5'-AGA

AAG GAG GTG ATC CAG 00-3') which were designed for their ability to

anneal to many types of bacteria. Genomic DNA (2 /d), 3/d of each primer

(10/tM), 3/d of deoxynucleoside-triphosphate (dNTP) mixture (dATP, dGTP,

dCTP, and dTTP, each 1/iM), 3/d of MgCl2 (25mM), and 1 unit of Taq D N A

polymerase were combined in a 50/d PCR reaction. Negative PCR controls,

which contained no DNA template, were used to test for contamination of

the PCR reagents. Thirty-five thermal cycles were performed, each with a 1

minute denaturation period of 94°C, a 1 minute annealing period of 56°C,

followed by a 1 minute extension period of 72°C. After 35 cycles, a single

extension time of 10 minutes at 72°C was used to facilitate polymerase

activity and extend PCR products. Amplification products were visualized

with ethidium bromide after electrophoresis in a 1% agarose gel. Samples

were electrophoresed adjacent to a DNA ladder to determine the fragment

48

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. size of the amplified DNA. Reactions containing fragments of the

appropriate size (ca. 1,530 base pairs "bp") were cleaned with Qiaquick™ spin

column PCR purification kits (Qiagen®, Valencia, CA) as prescribed by the

manufacturer.

Cloning

Cleaned PCR fragments were ligated into pCR®4-TOPO® plasmid

vectors (Invitrogen® TOPO TA Cloning® Kit for Sequencing, Carlsbad, CA).

Ligations were performed at room temperature with topoisomerase as

prescribed by the manufacturer. Ligated plasmids containing PCR fragment

inserts were transformed into TOPIO One Shot® competent £.coli cells as

prescribed by the manufacturer. They were incubated at 37°C for 30 m inutes

and plated in 50/zl, 100/d, and 250/d aliquots onto ampicillin-resistant

(50/ig/ml) Luria broth (LB) agar plates. The plates were incubated overnight

at 37°C. Bacterial colonies were picked the next day based on either

blue/white screening or presence/absence screening for clones that

contained a plasmid insert. Positive clones were collected and transferred to

5ml culture tubes containing LB broth and 50/ig/ml ampicillin. The

cultures were shaken overnight at 37°C.

Two methods were used to isolate cloned plasmid DNA from the

genomic DNA of £. coli. The first method used S.N.A.P. mini-prep kits

(Invitrogen®, Carlsbad, CA) to separate plasmid DNA from £.coli genom ic

49

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. DNA by centrifugation. The second method employed an additional PCR

step using primers located in the plasmid (M13f: 5'-GTA AAA CGA CGG

CCA G-3' and M13r: 5'-CAG GAA ACA GCT ATG AC-3'). These primers

flanked the DNA template insertion site and matched the plasmid sequence

exactly. Both procedures produced high quality PCR fragments.

Cycle Sequencing and BLAST Searches

Cycle sequencing was performed on cloned plasmid vectors after

mini-preparations or after reamplification with plasmid primers. Both

procedures generated highly accurate and unambiguous DNA sequences.

Cycle sequencing was performed by using the ABI PRISM™ Big Dye™ kit (PE

Applied Biosystems, Foster City, CA) as prescribed by the manufacturer, with

the following exceptions. The "ready reaction mix" provided in the kit was

diluted with 2.5X sequencing buffer (200mM Tris, 5mM MgCl2, pH=9.0). The

reactions contained 2/zl "ready reaction mix" (not the prescribed 8/d), 2.8/d of

sequencing buffer (2.5X), 3.2/d of primer (.5/iM), and 2/d of DNA template.

The thermal cycling protocol consisted of a denaturation phase of 96° for 10

sec, followed by an annealing phase of 48-50° for 5 sec, followed by an

extension phase of 60° for 4 min. The cycling protocol was repeated a total of

25 times. This system uses dye-labeled terminators that fluoresce upon laser

contact during electrophoresis on an ABI 377-XL automated sequencer.

Two plasmid primers (M13f and M13r) and four internal 16S rRNA

primers (536f: 5'-CAG CMG CCG CGG TAA TWC-3', 1114f: 5'-GCA ACG

50

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. AGC GCA ACC C-3', 519r: 5'-GWA TTA CCG CGG CKG CTG-3', and 960r: 5'-

GCT TGT GCG GGY CCC CG-3') were used. Cycle sequencing products were

cleaned with the ethanol/sodium acetate procedure outlined in the ABI Big

Dye manual. Automated DNA sequencing was performed with an ABI™

377-XL DNA Sequencer and provided sequences approximately 700 bp in

length.

Sequencing reactions using primers M13f and M13r generated

approximately 1,000 bp of the 16S rRNA gene (base pairs 1-500 and 1,050-

1,550). These fragments were compared to the National Center for

Biotechnology Information database (Genbank) by using BLAST searches.

Complete 16S rRNA sequences (-1,550 bp) were required for phylogenetic

study of several bacterial lineages. Therefore, some clones were sequenced

in both forward and reverse directions by using the 6 primers listed above.

The computer program Sequencher® 3.1 (Gene Codes Corporation, Ann

Arbor, MI) was used to proofread and join contiguous fragments of DNA

sequence into a single consensus sequence for each cloned sample.

Completely sequenced clones with similar identifications could then be

aligned with Sequencher 3.1. Divergent sequences were aligned by using

ClustalX (Thompson et al. 1997).

BLAST searches were used to query GenBank (NCBI National Center

for Biotechnology Information, http://www.ncbi.nlm.nih.gov) for DNA

sequences that showed high sequence similarity to the unknown bacterial

DNA sequences amplified from chewing louse extractions. The results were

51

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. returned via e-mail or directly over the World Wide Web. The unknown

and known sequences (called high-scoring segment pairs, or HSPs) are

aligned and compared statistically at NCBI. The distribution of a randomly

chosen set of HSPs follows a Poisson distribution. The BLAST algorithm

uses this distribution to determine the probability associated with the

similarity between HSPs. The BLAST program reports both probability

values (P-values) and scaled E-values, the latter being easier to interpret.

The E-value decreases as sequence similarity increases. For example, E-

values of 5 and 10 have corresponding P-values of 0.993 and 0.99995.

However, when E < 0.01, P-values and E-value are nearly identical. The

BLAST program ranks and lists all HSPs by E-values. The top BLAST result

(if significant at E < 0.01) was used as temporary identification for the

unknown bacterium until further study could reveal its accurate taxonomic

placement. The BLAST program also provides pairwise alignments of HSPs,

values of percent sequence similarity, and information regarding gaps in the

alignment of the two sequences.

RESULTS

BLAST searches of 339 16S rRNA sequences (-500 bp each) identified

35 lineages of bacteria in eight divergent groups of the domain Eubacteria

(Table 3.1). Many of the searches found no significantly similar DNA

sequences in GenBank and thus, are not listed in Table 3.1 (n=112/339).

These could represent PCR artifacts, chimeric sequences of multiple

organisms (Komatsoulis and Waterman 1997; Robison-Cox et al. 1995;

52

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table 3.1. Bacteria associated with chewing lice identified to the level of genus by means of BLAST searches of GenBank.

Types of Bacteria Genus Habitat

Thermus/Deinococcus group • T herm us g ro u p Meiothermus

Proteobacteria • Alpha Proteobacteria Rhizobiaceae Ensifer Methylobacterium P l a n t s Air conditioning units Rhodospirillaceae Phaeospirillum Rhodospirillum D e a d . S ea Hot springs W aste water Soil and salt marsh Sphingomonas group Sphingomonas Plant roots S o il River sediments • Beta Proteobacteria Burkholderia group Burkholderia Soil OEurope & Japan) Cystic Fibrosis patients Rotting bark Plant roots Blood cultures Rhizobacterium S o il Plant roots Comamonadaceae Acidovorax S o il Activated sludge Leptothrix Activated sludge Alcaligenaceae Alcaligenes W hirlpool baths Bone marrow Cerebrospinal fluid S o il Activated sludge P l a n t s Rhodocyclus Azoarcus Ultramicrobacterium group Ultramicrobacterium

Gamma Proteobacteria Psuedomonas group Pseudomonas Plant roots Oil degredation S o il M u sh ro o m s Activated sludge Groundwater sources (table cont.)

53

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Types of Bacteria Genus Habitat

• Gamma Proteobacteria (cont.) Psuedomonas group (cont) Acinetobactev Cerebrospinal fluid Blood, urine, pus Nose, throat, vagina M o u se sk in S o il Activated sludge S e a w a t e r Legionellaceae Legionella Hospital patients Xanthomonas group Xanthomonas Hospital respirator Human urine Various plants • Delta Proteobacteria Desulfuromonas group Geobacter S o il F irm ic u te s • Actinobacteria Actinomycetales Brevibacterium Oil degredation Cheese production Gordonia S o il Aquatic sediments Biofilters (waste treatment) Janibacter — Kitasatospora — Micrococcus S o il Oil degredation S k in Nocardioides P la n ts S o il S e a w a t e r Contaminated ground water Propionibacterium S k in Swiss cheese

• Bacillus/Clostridium group Bacillus/Staph, group Facklamia Blood, urine, etc. Powdered tobacco (snuff) Staphylococcus S k in Hymenoptera in amber Clostridiaceae Peptostreptococcus H u m a n g u t m icrobe Streptococcaceae Streptococcaceae Throat, vagina Vertebrate lungs A v ia n c ro p Verrucomicrobia • Verrucomicrobiales Verrucomicrobiaceae Verrucomicrobium —

Planctomycetales Planctomyces S p o n g e

Undescribed Eubacteria Undescribed taxa Various habitats

54

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Suzuki and Giovannoni 1996; Wang and Wang 1997), sequences for species

of bacteria not yet archived in GenBank, or fragments that failed to sequence

correctly. Some searches found significantly similar sequences in GenBank

that were themselves labeled "unknown" (e.g., "unknown Eubacterium"

and "unknown Proteobacterium"). These "unknown" sequences represent

species that have yet to be described. The louse-associated sequences often

showed high sequence similarity with the "unknown" sequences in

G enB ank.

Table 3.2 shows pocket gopher and chewing louse hosts, their general

collection locality, and the number and type of bacteria found in DNA

extracts from chewing lice. Many of the 23 lineages of bacteria that were

found in only one or two louse species (e.g., Acidovorax, Brevibacterium,

and Ensifer) likely represent transient bacteria acquired through the louse

diet. Twelve bacterial lineages (Table 3.2) were found in three or more

species of chewing lice and warrant further investigation. However the

"unknown Proteobacterium" and "unknown Eubacterium" categories

(Table 3.2) likely contain several distantly related species of bacteria and

cannot be considered single "lineages." Six lineages (e.g., Alcaligenes,

Azoarcus, Burkholderia, Methylobacterium, Propionibacterium, and

Staphylococcus) were found in four or more DNA extracts from chewing

lice. Staphylococcus-like clones had the highest prevalence and were found

in seven of eight louse species (Table 3.2). Many Staphylococcus spp. are

skin-associated bacteria (Table 3.1) so it is not surprising to find them

55

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3 1 1 1 1 2 1 of Hosts Number 1 1 1 1 1 1 1 1 1 1 2 1 6 4 1 7 4 14 3 of 20 4 22 6 (table cont.) Clones Number

1 1 1 minor bottae centralis Thom om ys New Mexico Geomydoecus Thomomydoecus 2 heterodus Costa Rica costariccnsis Geomydoecus Orthogcomys 1 7 3 setzeri Costa Rica undcrwoodi Geomydoecus Orthogcomys

1 3 1 5 9 1 2 3 cavator Costa Rica panam ensis Geomydoecus Orthogcomys l Florida Geomys pinetus scleritus Geomydoecus 1 1 2 3 9 Geomys Missouri bursarius geom ydis Geomydoecus 1 1 1 1 1 1 4 2 5 8 expansus castanops New Mexico Geomydoecus Cratogeomys

Table 3.2. Table 3.2. Endosymbiotic bacteria identified from nine taxa of chewing lice by means of environmental extraction, Kilasalosporia Legionella Leplothrix Lnicmioslic Gardania Geobacter Brcvibacterium Burkholderia Ensifer Janibacter Facklamia Azoarcus Locality: Acidovorax Aciin'tobacler Alcaligenes Host: Chewing Louse Gopher Host: PCR, PCR, sequencing, and searching GenBank databases via nucleotide BLAST recognition search protocol. "Unk." = unknown.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 1 1 1 1 1 1 4 4 5 7 1 3 3 2 3 5 1 2 2 1 2 2 5 1 1 1 1 1 7 5 1 2 9 2 2 1 1 2 2 24 16 43 13 227 17 24 29 329 21 45 13 48 46 1 2 1 1 8 9 1 1 l 1 1 10 19 16 35 1 1 9 3 4 1 51 17 1 1 3 1 1 2 2 l 2 3 4 5 1 1 16 18 77 78 l 48

Bacterial Types Identifiable Verrucomicrobium Examined Clones Xantlwimmas Total Clones llllramicrolHictcrium Unk. ProteobacteriumUnk. Eubacterium 4 Staphylococcus Strqitococcus Rlmlospirillum Sphignomonas Pseudomonas Rliizobactcrium Planctomyces Propionibaclerium Phacospirillum Peptostreptococcus Nocardioides O c h ro M u m Mciallienmis Micrococcus Melhylobacterium

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. associated with insects that feed on skin detritus. Representatives of three bacterial

genera were found in five {Burkholderia) or six {Methylobacterium, and

Propionibacterium) species of chewing lice. Some species ofBurkholderia and

Methylobacterium are known plant associates (Table 3.1), so the clones found in the

chewing louse samples could be transient bacteria ingested by the lice. Bacteria

associated with plants (especially roots) are likely to come in contact with lice that live

in the fur of subterranean, herbivorous rodents. Species in the genera Acinetobacter,

Legionella, Pseudomonas, and Xanthomonas are closely related members of the gamma

subgroup of the Proteobacteria. Many of the endosymbiotic bacteria of insects are

found in this subgroup of Proteobacteria.

One species of chewing louse {Geomydoecus expansus) hosted eight of the most

common types of bacteria (Table 3.2). The absence of one of the common bacterial

species {Azoarcus) in G . expansus may represent incomplete taxon sampling.

Geomydoecus centralis and Thomomydoecus minor (collected from Thomomys bottae)

hosted a combined total of only six types of bacteria. The predominance of only a few

species could be caused by PCR bias, local extinction of other bacteria, or incomplete

taxon sampling. In total, 227 of 329 clones were tentatively identified as one of 35

types of bacteria.

DISCUSSION

Sampling Strategy

Ribosomal DNA sequences from 35 types of bacteria were amplified and

sequenced from DNA extracted from chewing louse samples. Certain types of bacteria

58

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. could be found in most of the chewing louse hosts, whereas others were found only in a

single host (Table 3.2). Bacteria detected in only a few hosts could, in fact, be

associated with all eight louse species but may have been missed in my study due to

incomplete taxon sampling. This is a likely possibility, considering that approach was

intentionally wide in scope and relatively shallow in depth. An exhaustive survey,

including the sequencing of more clones from each DNA extraction, might help to

reduce taxon sampling bias. To test this, I plotted the number of different types of

bacteria found in my study against the total number of clones sequenced, which varied

from one chewing louse extraction to another (Fig. 3.1). Although Figure 3.1 shows

that more taxa are detected when more clones are sequenced, the low slope of the line

(a = 0.1844) indicates that the discovery of additional taxa will require examination of

very large numbers of additional clones. If I had sampled all of the chewing louse

extracts exhaustively, then the number of bacterial taxa detected would have had no

relationship with the number of clones sequenced. This pattern clearly demonstrates

that the sampling effort in this study did not detect all of the bacterial lineages in

chewing louse samples. I used rarefaction curves to obtain rough estimates of the total

number of clones required to adequately sample bacterial taxa in two chewing louse

hosts. The rarefaction curves tally the number of new taxa of bacteria detected (Y-axis)

as one increases the number of clones examined (X-axis). For example, if a sampling

59

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 50

'O au > 40 ■ © 4>

a >> 30 . H

’E (j « 20 ■

©

I I ■ i ■ ■ - t 10 20 30 40 50 60 70 80 90

Total Number of Clones Sequenced

Figure 3.1. Regression of the number of types of bacteria found as a function of the total number of clones sequenced for each of the seven gopher-louse host pairs listed in Table 3.2.

60

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. effort of 77 (Fig. 3.2) or 78 (Fig. 3.3) clones was sufficient to detect all (or nearly all)

bacterial lineages found in association with a chewing louse host, one would expect to

see rarefaction curves in Figs. 3.2 and 3.3 gradually become asymptotic to some value

on the Y-axis. However, Figs. 3.2 and 3.3 instead show that new bacteria are being

detected even as the last few clones are sequenced. It would appear that many

additional clones will need to be examined before concluding that I have an adequate

representation of the bacterial community associated with chewing lice.

Polymerase Chain Reaction Bias

It is important to note that the presence of certain bacteria in an experimental

extraction may affect the ability to amplify the DNA of other bacteria. It is not possible

to assess bacterial species abundance by merely counting the number of clones of a

particular bacterium because it has been shown that bacterial DNA sequences occurring

in low abundance in an extract can amplify more rapidly than more abundant DNA

sequences if the PCR conditions are better suited to the rarer sequence (Suzuki and

Giovannoni 1996). Information on presence or absence of bacteria amplified from

mixed-template PCR may be biased in the same way. Although use of primers from

conserved regions of the 16S rRNA permits studies of bacterial diversity, conserved

primers also introduce biases. I suggest that in future studies, conserved primers should

be used initially to detect and identify novel types of bacteria. However, after the initial

61

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

Total Number of Clones Sequenced

Figure 3.2. Rarefaction curve showing the number of clones identified from Geomydoecus expansus compared to the total number of clones surveyed.

62

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. without prohibited reproduction Further owner. copyright the of permission with Reproduced G eomydoecus geomydis eomydoecus G

Types of Bacteria Identified iue33 Rrfcincreso n h u e fcoe dniidfo from identified of clones ber num the ing show curve Rarefaction 3.3. Figure 10 20 30 • ■ 10 Total Number of Clones Sequenced Clones of Number Total compared to the total num ber o f clones surveyed. clones f o ber num total the to compared 20 30 40 63 50 070 60 090 80 survey, primers specific to target bacterial taxa should be used, especially if the

researcher intends to search for those bacteria in multiple hosts. This procedure may

help to avoid some of the problems associated with mixed-template PCR.

The Ubiquity of Bacteria

Of the 35 types of bacteria found in this study, eight are known to occur in

association with plants and 12 have been recovered from soil samples (Table 3.1).

Many species of bacteria are thought to be ubiquitous species worldwide. Six genera

detected in chewing louse extracts (Acinetobacter, Acidovorax, Alcaligenes,

Comamonas, Leptothrix, and Pseudomonas) also were found in a study of activated

sludge from a waste water treatment plant. Other bacteria amplified from chewing lice

have been detected in seawater, deep sea sediments (including the Marianas trench), hot

springs, and desert soils. Considering this remarkable diversity of habitats, it is

reasonable to assume that certain species of bacteria detected in this study may be

geographically widespread, if not ubiquitous. The fact that many of these

geographically and ecologically distinct organisms have nearly identical 16S rDNA

sequences is even more remarkable. Perhaps the best evidence of ubiquity in bacteria is

reviewed by Staley and Gosink (1999). They are interested in psychrophilic, free-living

bacteria that inhabit polar sea ice. These bacteria live on opposite poles of the earth and

cannot live in the intervening warm tropical waters. Despite the lack of an apparent

means of genetic exchange, their results suggest that there has been recent gene flow

64

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. between the two polar ice cap populations of bacteria. In short, we know that some

species of bacteria show evidence of gene flow over incredible distances, whereas

others (e.g., the endosymbiotic Buchnera associated with pea aphids) have virtually no

means of genetic exchange over comparatively short distances.

The Endosymbionts of Trichodectid Chewing Lice

I have documented that there are at least 35 lineages of bacteria associated with

the pocket gopher and chewing louse system. However, many of these are likely soil

and plant microbes that are transients ingested by chewing lice. Some may be skin

bacteria associated with the louse's mammalian host, whereas others may be a part of

the natural gut flora of the chewing louse. Whether any of these bacteria are truly

endosymbiotic—as are those found in mycetomes of other insects—remains untested. To

document that bacteria are truly endosymbiotic, one must: 1) find the endosymbiont in

every host; 2) fail to find the endosymbiont in the environment as a free-living

bacterium; 3) use in situ hybridization to probe the host for DNA sequences unique to

the putative endosymbiont to document that it is living within the host; 4) remove the

endosymbiont and document negative effects on the host; and 5) re-infect these

bacteria-free hosts with the endosymbiont and show reversal of the negative effects;.

These tests will document that the symbiont resides in the host and is an integral part of

the hosts existence.

65

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. It is beyond the scope of this dissertation to determine whether the 35

lineages of bacteria found in chewing lice are, indeed, true endosymbionts.

However, an evaluation of bacterial sequence similarity and phylogeny may

be used to infer endosymbiosis. For example, if a lineage of bacteria shows

phylogenetic structuring that matches that of its host, the most probable

cause is cospeciation. Cospeciation, in turn, suggests a long history of

association between the bacteria and its host. In the present study, it is

possible that a lineage of bacteria found in chewing lice is actually

cospeciating with the pocket gopher and is only a transient in the gut of the

louse. It is conceivable that one of the skin-associated bacteria of pocket

gophers has cospeciated with its host. If so, this bacterial lineage likely

would show evidence of cophylogeny with the chewing louse lineage, given

that chewing lice are known to have cospeciated with their pocket gopher

hosts (Hafher et al. 1994) as well. In an effort to determine whether any of

the above lineages show evidence of cospeciation, I sequenced the entire 16S

rRNA gene for clones identified as Acinetobacter, Legionella, Pseudomonas,

Staphylococcus, an d Xanthomonas (chapter 5). These species were selected

for the initial analysis because Staphylococcus was detected in a wide

diversity of hosts (Table 3.2), and the other lineages are part of the gamma

subgroup of Proteobacteria that contains many endosymbiotic bacteria.

Several other lineages detected in chewing lice also warrant further study

(e.g., Burkholderia an d Propionibacterium), but were not investigated here.

66

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 4

BACTERIAL DIVERSITY IN CHEWING LOUSE EGGS

INTRODUCTION

Most insects host internal bacteria, some of which are transient

visitors, whereas others are long-term endosymbionts. There are numerous

well-described insect-bacteria associations in which the endosymbiotic

bacteria are required for the survival of their host (Douglas 1989). Pea

aphids, for example, require an endosymbiotic bacterium (Buchner a) to

reproduce, and they fail to produce offspring if they are treated with

antibiotics that kill their endosymbionts (Douglas 1989; Ishikawa 1989).

Termites must have a large assemblage of microorganisms to aid in the

digestion of cellulose (Douglas, 1989). It is often crucial for insects to have

the required complement of endosymbionts immediately upon emerging

from the egg (e.g., when the bacteria provide essential nutrients). Thus,

bacteria are transmitted transovarially in pea aphids (Douglas 1989; Ishikawa

1989), tsetse flies (Aksoy et al. 1997), hydrothermal vent invertebrates (Cary

et al. 1993), cockroaches (Sacchi et al. 1988), leafhoppers (Purcell et al. 1986),

human body lice (Eberle and McLean 1983), Phthirapteran bird lice (Saxena

and Agarwal 1985), and many other invertebrates. In some insects (e.g.

cockroaches, ants, and certain beetles) the endosymbionts migrate to the host

ovarioles where they coat the outer surface of developing eggs. Some

67

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. hatching insects eat their own egg case as a source of energy and thus, are

inoculated with necessary microorganisms (Douglas 1989). In other insects,

the endosymbionts migrate to reproductive structures and actually become

incorporated inside the developing egg, which insures transmission to the

offspring (Douglas 1989).

In human body lice ( Pediculus), mycetome-associated bacteria leave

their specialized cells via a single opening and migrate along the digestive

tract to the oviduct Eberle and McLean (1982; 1983). The endosymbionts

migrate directly to the oviducts, even when the oviducts have been

experimentally moved to unnatural locations within the insect. In weevils,

one species of endosymbiont is permanently associated with the host's

oocytes (reviewed by Douglas 1989). These symbionts are present in mature

oocytes, but leave to infect developing oocytes in adult weevils. Although

these endosymbionts are passed through the eggs to offspring of both sexes,

they eventually are extirpated in the males. There are also mycetome-

associated endosymbionts in weevils that have a more typical mode of

vertical transmission in which the symbionts migrate from their specialized

cells to the ovaries to be incorporated into the eggs. Thus, weevils are host

to at least two distinct lineages of endosymbionts that have evolved

different mechanisms of vertical transmission.

The complex mechanisms involved in transovarial transmission of

endosymbionts doubtlessly evolved over countless generations. As a result

of this intimate, long-term relationship between these highly specialized

68

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. host-parasite associates, many of them also show evidence of cophylogenv. I

am interested to know if the eggs of chewing lice contain bacteria that are

transmitted transovarially. If so, these bacteria are almost certainly long­

term endosymbionts, and are therefore good candidates for the study of

cophylogeny. Although Saxena and Agarwal (1985) documented the

presence of mycetocyte-associated bacteria in the ovaries of some

Phthirapteran bird lice, and Saxena et al. (1985) discussed the history of

bacteria associated with the ovaries of many species of lice (including

mammal lice in the family Trichodectidae), no study has determined

whether mycetocyte-associated bacteria are present in the trichodectid lice

that live on pocket gophers. In this study, I used PCR, cloning, and DNA

sequencing to search for 16S rRNA sequences from bacteria associated with

louse eggs.

MATERIALS AND METHODS

Twenty chewing louse eggs from Geomydoecus eioingi were collected

from the fur of their host, Geomys bursarius, and deposited into each of two

1.5ml microcentrifuge tubes. Each tube of 20 lice was then washed four

times in a solution containing 400/rl saline EDTA buffer (containing 150mM

NaCl, lOmM EDTA, pH=8.0), 5/d lysozyme (lOmg/ml), and 10/d SDS (25%).

This treatment dislodged and removed bacteria from the outer surface of the

egg. The solution used to wash the egg was removed after each treatment

and transferred to a separate sterile microcentrifuge tube; the louse eggs

remained in the same 1.5ml microcentrifuge tube throughout the entire

69

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. process. Extreme care was taken not to rupture the eggs or accidentally

remove them during the wash cycles. After the washes, the eggs were

crushed with a micropestle. Two negative controls were used during DNA

extraction. The first contained all of the reagents used in the DNA

extraction process, and the second contained the same reagents plus a

micropestle (identical to the one used to crush the eggs in the experimental

samples) to insure that none of the extraction reagents or tools was

contaminated with bacteria. Each of the four wash solutions, the controls,

and the eggs were then digested with proteinase K as described in the

previous chapter.

Genomic DNA was then extracted by adding 400/d of

phenol/chloroform, vortexing, and spinning in a micro centrifuge for two

minutes at 14,000 rpm. The supernatant, which contains the extracted DNA,

was carefully removed and the extraction procedure was repeated two more

times. One milliliter of 95% EtOH was added, and the solution was

incubated on ice overnight to precipitate genomic DNA. Centrifugation at

14,000 rpm for 20 minutes concentrated the DNA into pellets. The EtOH

was decanted, and the microcentrifuge tubes were dried in a vacuum

centrifuge for 15 minutes. The resulting pellets were resuspended in 30/d of

TE buffer. The PCR protocol and primers (27-f and 1525-r) described in the

previous chapter were used to amplify bacterial 16S rRNA sequences from

the wash solutions and the louse eggs. Cloning and DNA sequencing also

followed the procedure outlined in the previous chapter. BLAST searches

70

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. in GenBank provided tentative identifications for unknown bacterial

sam ples.

RESULTS AND DISCUSSION

The PCR reactions produced bright, distinct DNA bands for both the washes

and the eggs, whereas the negative controls only produced faint smears of

DNA that could not be cloned. I judged that this slight contamination was

insufficient to cause spurious PCR amplifications in the experimental

extracts. Seventy-four clones were sequenced from environmental samples

of chewing louse eggs. Most of the bacterial clones from the eggs (50 of 74)

were bacteria in the genus Acinetobacter, and Staphylococcus-like bacteria

were found in 16 of 74 clones (Table 4.1). A third lineage in the genus

Pseudomonas was found in 8 of 74 clones. Sixty-five clones were sequenced

from the wash solutions used to remove bacteria from the outside of the

eggs. Acinetobacter accounted for 23 of these 65 clones, Staphylococcus 22 of

the 65 clones, an d Afipia, Pseudomonas, and Dyadobacterium were found in

a few clones each (Table 4.1).

From these data it appears that several types of bacteria (notably

Acinetobacter, Afipia, Dyadobacterium, Pseudomonas, and Staphylococcus)

are present on the outer surface of chewing louse eggs. If I assume that my

washing technique effectively removed the external bacteria from the eggs,

then Acinetobacter, Staphylococcus, and Pseudomonas likely occur inside

chewing louse eggs. It is also possible, however, that the washing protocol

71

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table 4.1. Types of bacteria found in DNA extracts of chewing louse eggs. Bacteria were identified, by BLAST searches of the Genbank DNA sequence database.

Bacteria Found Egg1 Egg 2 Wash 1 Wash 3

Acinetobacter 18 32 16 7

Afipia 0 0 10 0

Dyadobacterium 0 0 4 1

Pseudomonas 5 3 4 1

Staphylococcus 15 1 0 22

Total 38 36 34 31

72

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. did not remove all bacteria from the outer surface of the egg. If so, then

Acinetobacteria, Staphylococcus, and Pseudomonas must adhere strongly to

the outer surface of louse eggs, as these taxa were the last removed by

washing. Results of the preceding analysis cannot be used to test these two

hypotheses, so I designed another protocol to search for bacterial DNA inside

louse eggs.

Twenty louse eggs were washed as before, but this time in a series of

10 washes. The first, fifth, and tenth washes were amplified by using PCR, as

were the eggs themselves and the appropriate controls. All three washes

produced PCR products of the target length (ca. 1,530 bp), just as in the

previous analysis. However, the extract containing only chewing louse eggs

repeatedly failed to produce PCR bands (despite the use of numerous PCR

protocols), which suggested the absence of bacterial DNA within the egg. To

determine that the DNA extraction protocol was effective, I amplified the

egg extracts using primers that anneal to the cytochrome b gene of chewing

lice (obviously chewing louse eggs should contain chewing louse DNA).

Certainly there is more host (chewing louse) DNA in a louse egg than

endosymbiont DNA. These PCR reactions produced single, bright DNA

bands that, when sequenced, were found to be louse cytochrome b sequences.

Although this protocol does not preclude a methodological problem with

bacterial extractions from chewing louse eggs (discussed below), it is

consistent with the hypothesis that the chewing louse eggs contain no

internal bacterial DNA.

73

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. A constant concern in any attempt to amplify endosymbiont DNA is

the far greater concentration of host DNA compared to that of the

endosymbiont. The resulting swamping effect can be overcome during PCR

by use of highly specific DNA primers that preferentially anneal to bacterial

DNA. However, when PCR reactions fail, one cannot rule out low DNA

copy number as the cause of the failure. I tried to quantify the amount of

DNA in each of the extractions in this study, including the washes that

amplified successfully, but all samples contained DNA in such low

quantities that they failed to register when examined with a fluorometer. It

is conceivable that the failure of PCR reactions containing egg extracts is the

result of either no DNA template (i.e., sterile eggs) or the result of low copy

number of bacterial DNA. It is not possible to distinguish between these two

possibilities at this time. We know from other studies that the number of

bacteria transferred through the egg is always low (Douglas, 1989), so the

hypothesis of low copy number seems likely. However, we also know that

in the absence of a long-term endosymbiosis with transovarially inherited

bacteria, the eggs of most animals are, in fact, sterile at the time of

deposition.

Endosymbionts are sometimes transmitted vertically among insect

hosts by adhering to the outer surface of the insect egg (Douglas 1989). This

mechanism for vertical transmission of endosymbionts may be

evolutionarily intermediate between (primitive) contact transmission and

(derived) transovarial transmission. My results are consistent with the

74

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. hypothesis that bacteria are found on the outer surface of chewing louse

eggs, but not within. However, conclusive evidence is lacking because my

tests cannot distinguish between bacteria that adhered to the egg's surface

while the egg was inside the louse versus bacteria that adhered to the egg's

surface after the egg was laid. It is interesting to note that each of the taxa of

bacteria found in association with chewing louse eggs also was found in

abundant numbers in the louse extracts (see previous chapter). Further

analyses using microscopy and in situ hybridization may reveal whether the

eggs contain, or are coated with, truly endosymbiotic bacteria. In situ

hybridization also could be used to identify bacteria associated with the

ovarioles of chewing lice to further understand vertical transmission of

these bacteria.

75

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

STUDIES OF COPHYLOGENY: CHEWING LICE AND ENDOSYMBIOTIC BACTERIA

INTRODUCTION

Host-parasite systems are intrinsically interesting to evolutionary

biologists because they often represent long and intimate associations

between two or more groups of organisms that are distantly related and

quite dissimilar biologically. This long history of association often leads to

reciprocal adaptations in the hosts and their parasites (classical coevolution

or coadaptation) as well as contemporaneous cladogenic events in the two

lineages (cospeciation or cophylogeny). Comparative phylogeneticists are

particularly interested in the phenomenon of cophylogeny because

cospeciation events identify temporal links between the host and parasite

phylogenies, and thus provide an internal time calibration for comparative

studies of rates of evolution in the two groups.

Hafner and Nadler (1990) proposed a protocol for investigation of

cophylogeny that was designed to remedy many of the problems that

hampered earlier studies. They argued that the minimal requirements for a

valid test of cophylogeny were: 1) independent host and parasite

phylogenies; 2) well-corroborated phylogenies based on tree-building

algorithms with explicit assumptions; and 3) rigorous statistical tests of

similarity between the host and parasite phylogenies. In addition to these

76

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. basic requirements, they emphasized that a study of cophylogeny would be

enhanced significantly by comparative investigation of molecular (as

opposed to morphological) differentiation in the hosts and their parasites.

This framework for investigation of cophylogeny has been used in

numerous published studies, and the number of molecular-based studies of

cophylogeny has increased rapidly (Page in press).

Previous studies of host-endosymbiont cophylogeny, particularly

those involving aphids and their bacterial endosymbionts (Moran et al. 1993;

1995; Munson et al. 1991; Unterman et al. 1989), have shown remarkable

levels of concordance between host and symbiont phylogenies. Using host

divergence dates estimated from fossil evidence, Moran et al. (1993) were

able to infer bacterial divergence dates in the absence of a fossil record. This

inference would not have been valid without firm evidence of cophylogeny.

Similarly, Moran et al. (1995) used cophylogeny as a framework to compare

relative rates of molecular change in aphids and their endosymbiotic

bacteria, which circumvented the need for estimates of absolute rates of

change. This latter study (Moran et al. 1995) was similar in design to the

Hafner et al. (1994) study of rates of molecular change in gophers and lice,

and the results of both studies showed that rate of evolution in the parasite

was faster than that of the host.

I investigated the phylogenetic relationships of several lineages of

endosymbiotic bacteria associated with the gopher-louse system. To do this,

I sequenced the 16S rRNA gene (1,531 bp) for clones extracted from chewing

77

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. louse samples. The bacteria studied were closely related to the following

genera: Acinetobacter, Facklamia, Legionella, Pseudomonas, Staphylococcus,

and Xanthomonas. The gamma subclass of the Proteobacteria (which

contains Acinetobacter, Facklamia, Legionella, Pseudomonas, an d

Xanthomonas) has been implicated in numerous endosymbiotic

relationships with insects. Members of the genus Staphylococcus are n o t

known to be endosymbiotic, but were prevalent in most chewing louse and

louse egg extracts, and thus warrant further investigation. DNA sequences

for these taxa were used to generate phylogenetic hypotheses that could be

compared to well established host phylogenies. This is the first step in

determining whether any of these lineages have shared a long history of

association with chewing lice or pocket gophers.

MATERIALS AND METHODS

The specimens collected and methods of DNA extraction, PCR,

cloning, and sequencing, are provided in chapter 3. Contiguous DNA

fragments were assembled with Sequencher 3.1 and consensus sequences

were generated for each clone. Clones of 16S rRNA sequence were aligned

with the computer program ClustalX (Thompson et al. 1997) to determine

the optimal alignment. ClustalX is a global alignment program that aligns

similar sequences first, based on a guide tree, before aligning more divergent

ones. DNA sequences for outgroup taxa were downloaded from the

National Center for Biotechnology Information (GenBank). Sequences

identified as Staphylococcus were analyzed separately from those identified

78

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. as gamma-Proteobacteria. Complete sequences are given in Appendices II

and HI.

Phylogenetic Inference

I used the best-fit maximum-likelihood (ML) method described by

Cunningham et al. (1998) to choose the most appropriate ML model of

nucleotide evolution. This method incrementally increases the number of

parameters in the ML model until the addition of new parameters no longer

increases significantly the fit between the model and the data. I used the

computer program ModelTest (Posada and Crandall 1998) as a guide to select

the best-fit ML model. This program calculates ML scores with 64 nested ML

models and determines the best-fit model based on hierarchical likelihood-

ratio tests. The selected model and parameter estimates were then entered

in PAUP* (Phylogenetic Analysis Using Parsimony, Swofford 2000). A

heuristic search was performed, with random sequence addition and TBR

branch swapping. The resulting tree topology was then used to re-estimate

the ML parameters. Another heuristic search was performed (with 10

random sequence additions) with the new parameter estimates. If the

resulting tree was different from the previous tree based on -In likelihood

score then, the process was repeated until the resulting tree score did not

change in successive iterations. This successive approximation generates

the best estimates for the ML parameters and helps to insure that tree

searching does not stop at a local optimum. I used multiple outgroups in

my analyses selected from well established phylogenies of 16S rRNA

79

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. sequences (Rainey et al. 1994; Lambert et al. 1998). For comparative

purposes, I used the maximum-parsimony (MP) method, as implemented in

PAUP*. Bootstrap analyses (1,000 replicates) were generated to determine

the relative level of support for individual nodes.

Comparing Host and Parasite Phylogenies

Bacterial phylogenies generated in the above analyses were compared

statistically to well-established louse phylogenies generated in previous

analyses (Hafner et al. 1994; Hafner and Page 1995; Page et al. 1995; Page and

Hafner 1996). Component analysis (e.g., Page 1990a; 1994) treats the parasite

phylogeny as a lineage, rather than as a set of codes, and maps the parasite

tree onto the host tree. The major limitation of this method, as

implemented in COMPONENT (Page 1993a), is its poor ability to deal with

host-switching events. Recognizing this, Page extended the algorithm for

reconciling trees to incorporate host switching, and he implemented the

method in the computer program TreeMap (http:// taxonomy.

zoology.gla.ac.uk). A major advantage of component analysis

(COMPONENT and TreeMap) is that it allows statistical tests of similarity

between host and parasite phylogenies that take into consideration both

cladistic topology and branch lengths. For these reasons, component

analysis was used to determine the amount of similarity between host and

parasite phylogenies.

Although significant topological concordance of host and parasite (or

host and symbiont) trees is unlikely to result from chance, it is possible that

80

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. recent host switching may produce spurious congruence, especially if closely

related parasites preferentially colonize hosts that are closely related.

Similarly, incongruence between host and parasite phylogenies may result

from differential survival of multiple parasite lineages (lineage sorting),

rather than host switching (Page 1993b). I mapped the bacterial phylogeny

onto the phylogeny of their chewing louse hosts with the computer program

TreeMap, which reconciled the host and parasite trees by maximizing the

number of cospeciation events and minimizing the number of sorting

events (i.e., host-switches, duplications, and extinctions). Identification of

probable host-switching events may reveal whether colonization of new

hosts by bacteria is simply opportunistic (nearest neighbor), or whether these

symbionts are tracking a particular resource in the host environment.

RESULTS

Maximum Likelihood Models

The computer program ModelTest (Posada and Crandall 1998) picked

the Tamura-Nei (Tamura and Nei 1993) model of nucleotide evolution for

both the Staphylococcus and gamma-Proteobacteria data sets. The Tamura-

Nei m odel allow s for tw o rates of transitions (A<->G an d C«-»T), one rate for

tranversions, and it allows unequal base frequencies. ModelTest determined

that the addition of both an invariant sites parameter and a variable sites

parameter (according to a gamma distribution) significantly increased the fit

of the model. An invariant sites parameter assumes that some nucleotide

positions do not vary across taxa in the study. The variable sites parameter

81

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. assumes that some nucleotide sites are more variable than others, and this

pattern can be estimated by a gamma distribution. The shape of the gamma

distribution is described by a shape parameter, a. W hen alpha is small (e.g.,

a = 0.05) most sites are evolving slowly but a few sites are changing rapidly.

As alpha increases (e.g., a = 10) the distribution becomes centered around a

value of 1.0 and all sites are evolving at nearly the same rate. Lower alpha

values are most common (a = 0.03 to 0.07) in DNA sequence data.

ModelTest does not perform likelihood ratio tests between all

pairwise comparisons of the 64 nested models. Rather it uses a flow diagram

that begins with the most specific model (Jukes-Cantor) and tests the

addition of new parameters. For instance, if the addition of unequal base

frequencies to a Jukes-Cantor model is found to produce a significant

difference in negative log likelihood (-InL) score, the new model (in this case

an F-81 model; Felsenstein 1981) is chosen over the null model (Jukes-

Cantor). Sometimes the structure of the tests used in ModelTest can result

in over-parameterized estimates of the best-fit ML model. To reduce the risk

of choosing a model that is too parameter-rich, I tested the model chosen by

ModelTest (Tamura-Nei plus invariant sites plus gamma, or TrN+I+G)

against many permutations of simpler models. The removal of any

parameter associated with the TrN+I+G model produced significantly worse

-InL scores based on likelihood ratio tests (p < 0.05), thus showing support

for the best-fit model selected by ModelTest.

82

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Gamma-Proteobacteria Analysis

The parameter estimates provided by ModelTest were used in the first

iteration of ML analyses in PAUP*. A heuristic search provided a single ML

tree estimate upon which new ML parameters were estimated. These varied

slightly from those generated previously by ModelTest and were used to

perform another heuristic search (10 replicates). In this analysis, the

resulting tree topologies varied among the ten replicates and the resulting

parameter estimates varied again from the previous estimates. Because it is

advantageous to re-estimate the parameters on increasingly better trees until

the estimates stabilize, the parameters were estimated again and another

heuristic search was performed (10 replicates). The resulting parameter

estimates did not change during this iteration and the 10 heuristic search

replicates converged on the same tree topology. The best ML tree (Fig. 5.1)

showed that the clones extracted from chewing louse samples were nested

within the gamma-Proteobacteria sequences downloaded from GenBank.

The uncorrected sequence divergence ranged from 0.0 to 16.8 percent (Table

5.1). Parsimony analysis (equally weighted) produced an almost identical

topology (redundant taxa were primed from this analysis). Bootstrap

support for clades was either very low (shown as polytomies) or moderate

(Figure 5.2). Figure 5.2 shows the 50% majority rule consensus tree of all

bootstrap replicates (n = 1,000) and polytomies represent clades found in

fewer than 50% of bootstrap replicates.

83

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. —10 Changes Acinetobacter sp.l X 8 1 6 5 9 Acinetobacter johnsonii X 8 1 6 6 3 f ex. Geomydoecns expansus 8 8 -6

r Acinetobacter calcoaceticus X 8 1 6 6 1

I Acinetobacter calcoaceticus X 8 1 6 5 7

— Acinetobacter Iwoffii X 8 1 6 6 5

— ex. Geomydoecus geomydis 9 1 -3 1

Acinetobacter haemolyticus X 8 1 6 6 2

ex. Geomydoecus geomydis 9 1 -1 1

- ex. Geomydoecus geomydis 349 -1 9

• ex. Geomydoecus geomydis 3 4 9 -1 5 ex. Geomydoecus expansus 8 8 -1 7

ex. Geomydoecus geomydis 9 1 -1 8

ex. Geomydoecus geomydis 9 1 -2 8

Acinetobacter baumannii X 8 1 6 6 7

HAcinetobacter baumannii X 8 1 6 6 0

r Acinetobacter junii X 8 1 6 6 4

*■ Acinetobacter junii X 8 1 6 5 8

* 'inetobacter radioresistens X 8 1 6 6 6

Acinetobacter calcoaceticus X 8 1 6 6 8

------ex. Geomydoecus expansus 8 8 -5 -2

ex. Geomydoecus geomydis B -22

------ex. Geomydoecus expansus 88 -5 -1

ex. Geomydoecus geomydis B-26

— ex. Geomydoecus geomydis 9 1 -1 9

■ ex. Geomydoecus geomydis 349-11

Pseudomonas aeruginosa A F 2 3 7 6 7 8

Figure 5.1. Maximum likelihood phylogeny for gamma Proteobacteria using the Tamura-Nei + I + G model (see text for details of m odel).

84

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0.0432 0.1584 0.0426 0.0404 0.1850 0.1106 0.1147 0.0439 0.0371 (table cont.) 0.0282 0.12350.1563 0.1222 0.1127 0.1174 0.0302 0.0452 7 8 0.0303 0.0468 0.1562 0.0274 0.0274 0.0247 0.0247 0.1764 0.1764 0.0302 0.0282 0.0282 0.0433 0.1503 0.1503 0.1497 0.0254 0.0254 0.0200 0.0282 0.0137 0.0220 0.0193 0.0303 0.0193 0.03630.03910.0199 0.0282 0.0488 0.0248 0.0014 0.0445 0.0350 0.0445 0.0350 0.0432 0.0363 0.0282 0.0323 0.0062 0.0275 0.0275 0.0467 0.0227 0.0193 0.0275 0.0275 0.0426 0.1167 0.1281 0.1235 0.1235 0.1145 0.0363 0.0391 0.0199 0.0323 0.0227 0.0254 0.0254 0.1525 0.1525 • 0.1508 0.03770.1249 0.0377 0.1249 0.0227 0.1293 0.1235 0.0363 0.0234 0.0234 0.1167 0.0364 0.0351 0.0234 0.0234 0.0316 0.0000 0.0371 0.1494 0.1115 0.1154 0.1154 0.1184 0.1127 0.11900.0414 0.1201 0.0324 0.1201 0.0324 0.1232 0.1174 1 2 3 4 5 6 0.0357 0.0364 0.0275 0.0303 0.03030.1536 0.0337 0.1470 0.0220 0.1543 0.0220 0.1543 0.0186 0.1540 0.0268 0.0268 0.0419 0.03640.1360 0.0302 0.1265 0.0501 0.0323 0.1237 0.0296 0.0386 0.0323 0.1519 0.0296 0.0344 0.1299 X81666 0.0426 X81662 0.0323 0.0371 0.0281 0.0281 X81657 0.0412 X81668 X81661 0.0412 0.0364 X81667 0.0357 0.0316 X81660 AF237678 0.1320 0.1135 349-15 91-31 0.0379 91-28 0.0317 0.0345 91-11 0.0323 0.0379 0.0289 0.0289 0.0151 88-5-1 0.1785 0.1765 0.1775 0.1775 0.1775 X81663 0.0289 0.0358 0.0351 0.0351 X81665 X81658 0.0391 0.0289 0.0226 0.0226 0.0268 0.0336 0.0336 X81664 X81659 Table 5.1. Uncorrected pairwise distances (P-distances) for 27 taxa of the gamma subclass of the Proteobacteria. Geomydoecus geomydis-349-19 Geomydoecus geomydis-91-\8 Geomydoecus expansus-88-5-2 Geomydoecus geomydis-91-19 Acinetobacter calcoaceticus Acinetobacter johnsonii Acinetobactersp.l Acinetobacter radioresistens Acinetobacter baumannii Acinetobacter baumannii Geomydoecus geomydis-B-26 Geomydoecus geomydis- Geomydoecus expansus-88-6 Geomydoecus expansus- Geomydoecus expansus-88-17 Geomydoecus geomydis- Geomydoecus geomydis- Geomydoecus geomydis-349-11 Pseudomonas aeruginosa Geomydoecus geomydis-B-22 Acinetobacter junii Acinetobacter calcoaceticus Acinetobacter junii Acinetobacter haemolyticus Geomydoecus geomydis- Acinetobacter Iwoffii Acinetobacter calcoaceticus Bacterial Taxa 11 Numbers following taxon names refer to National Center for Biotechnology1 Information (NCBI) accession numbers 14 15 12 13 10 or clones numbers. 2 3 19 18 16 17 8 5 4 21 23 24 7 9 6 20 22 25 26 27

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. o CM CM CO 0 0 CO CO o s c m i n CO p N NO TP ON ON r H o Cm o 1-H r-H r H o 1-H o o o i-h o 1-H 1-H o o u o © O © © d © d d O) 3

CM CO 00 VO in o vO Cm o Cm in in CO Tp NO CM O in i n t P t p Tp m tP CO CO in Tp t p r-H iH i-h i-h i-h 1“H i-h rH P-H i-h O d d © O d d d o d

CM ao in o m vO CM CO Cm NO Cm 0 0 Cm ON 00 Cm O n in CM sO o in t p tP t p oo TP CO t p in tP in 1-H i- h r-H ^ h r-H ^ h p-h i- h i- h © d o d d d O d d d d

CM 0 0 O s t P CO in o 0 0 Os oo in CM GO CO CO TP ^H CMCM O n t p p H oo Cm Tp Tp o o Cm o O y—* o 1“H i - h ^ H o © o i- h o t—« r—• o o 1-H O* © o ’ d o ’ o ’ o ’ o ’ o ’ o ’ d d

NO CM CM CO Cm ON 00 o 0 0 ON in VO CM 0 0 in CO in CO CO CM ON t p CO 0 0 oo o Tp Tp o r-H o Cm o o f-H f“H o p H o 1-H ^ H o o o r-H o f-H f-H o o 1-H d © d d © o ’ d o d d d d o ’

CM Cm 0 0 1-H ^ H Cm m in CM Cm fH CM Tp ON o o p H CO CM CM 1“H tp O 0 0 CO NO 1-H 1“H CM CM in Tp CM CM CM Cm CM CO f-H CM CM TP f-H 1“H 1-H f-H 1-H f-H i- h 1-H r-H o o f-H 1-H o o ' o ’ o ’ d o ’ o o ’ d o ’ o ’ d o ’ d o ’

Tp vO ON vO ON vO 0 0 vO 1-H Os Tp CO NOCm T p 0 0 O n 0 0 CO ON r-H VO ON o 00 in CM Os Tp TP CM CM CM m Tp CO CM CM 00 CM 1“H CM CM CM CM r-H o o f-H 1-H o o O 1“H © PH f-H © O p H © o o © o* o ’ o ' © o ' o ’ o ’ o © o ’ o ’

o CO in CM Cm CM in O n O n CM 00 in NO 1-H o oo m Tp ON 0 0 in CO 1-H o 00 NO 00 CO ON in CO CM CO CM CM CM in in CO CO CM CM CM 1-H 1-H CO CO CM o 1-H O O 1-H 1-H o O O 1-H O 1-H 1-H o o d d O o ' o ’ o ' d O O d O d O d o ’ O

sO CO Cm NO NO Cm 0 0 NO O s o c o ON i n vO VO t p o o 1-H c o CM Cm Cm CO ON ON OO ON O n vO Cm 1-H Cm 1-H i n CO CM CM O O in Tp o o o Cm o 1-H CM o o CM O o r-H O O 1-H 1-H o o o t h p 1-H 1-H O p © © O d © d © © © d o d d o d d d

in 1-H o r-H in r-H Cm o in CM 0 0 GO GO CO TP 1-H NO NO ON TP CO in Cm NO in TP ON OO NO NO NO TP o CO ON CO CM O CO CM CM CM in in CM CM CM N CM CM CO CM CM OO p 1-H O O 1“H 1-H o O o f H o 1-H 1-H o O 1-H o ’ o ’ d d o o d d d d d d d d d d o d

onNn^in«Noo»OriNn«inifits INn«lD'eNe09iHrlpHplrlrtpCrlrlnWNPIN«NNN

86

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

ON vO »-H N CN

CN lft ON N O CNi-* OO

NO CO o m «—• CN i-h CO N i-h i-h O O 0*0

CN ON O CO 00 in i-H i - h co CN i h 1/1 N O 1-H T~« o o o o o

CN c o T}1 OO o ON N* CN t>* oo CN o t*H 1-H O ^ H CN I-H 1-H o o I-H o ’ o* 0 * 0 * o

CO CO CO on 0 0 CO CO ON oo NO co oo t v nO NO N N VO 1-H I-H 1-H 1-H I-H H o o o o o o

ON CO CN t v 00 oo o CO CO r-H m c o o n oo t v o |H 1-H 1-H o f-H f-H o 1-H 1-H o o ^H o o d o 0 * 0 ©

CN y—, in O n i-h CN NO NO ON in CN NO in On CN FHNO ^H1-H FHi-h O CN o FHo r*H FH OO r-H o o o* © o ’ 0 * 0 o ’

OfHC40^tAvONC09\OiNCNO«U)vOts • NO^'lftvO^oOGvr'HpHiNHrtfN^iNr^NW NNNM W N

87

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Acinetobacter sp.l X 8 1 6 5 9

82 Acinetobacter johnsonii X 8 1 6 6 3

ex. Geomydoecus expansus 8 8 -6

Acinetobacter calcoaceticus X 8 1 6 5 7 64 Acinetobacter Iwoffii X 8 1 6 6 5

57 Acinetobacter haemolyticus X 8 1 6 6 2 ex. Geomydoecus geomydis 3 4 9 -1 9

52 ex. Geomydoecus geomydis 3 4 9 -1 5

ex. Geomydoecus geomydis 9 1 -2 8 67 ex. Geomydoecus expansus 8 8 -1 7

ex. Geomydoecus geomydis 9 1 -3 1

ex. Geomydoecus geomydis 9 1 -1 1 100 51 Acinetobacter baumannii X 8 1 6 6 0

Acinetobacter junii X 8 1 6 5 8

100 Acinetobacter radioresistens X 8 1 6 6 6

Acinetobacter calcoaceticus X 8 1 6 6 8

ex. Geomydoecus expansus 8 8 -5 -2 65 65 90 ex. Geomydoecus geomydis B -2 2

72 ex. Geomydoecus expansus 8 8 -5 -1

ex. Geomydoecus geomydis B -2 6

Pseudomonas aeruginosa A F 2 3 7 6 7 8

ex. Geomydoecus geomydis 3 4 9 -1 1

ex. Geomydoecus geomydis 9 1 -1 9

Figure 5.2. Fifty percent majority rule consensus tree of 1000 bootstrap replicates using the parsimony optimality criterion for species of gamma Proteobacteria. Numbers at nodes represent bootstrap support.

88

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

The parameter estimates provided by ModelTest were used in the first

iteration of ML analyses in PAUP*. A heuristic search provided a single ML

tree estimate upon which new ML model parameters were estimated. These

varied only slightly from those generated previously by ModelTest. They

were used to perform another heuristic search (10 replicates) which

provided the same topology in each of the ten replicates. The resulting

parameter estimates did not deviate from the previous estimates and

therefore the process of iteration was terminated. The best ML tree (Fig. 5.3)

showed that the clones extracted from chewing louse samples were nested

both within and outside the Staphylococcus sequences downloaded from

GenBank. The uncorrected sequence divergence ranged from 0.0 to 12.4

percent (Table 5.2). Parsimony analysis generated a similar phylogenetic tree

that was well supported. Figure 5.4 is the 50% majority rule bootstrap tree

that was generated in the parsimony analysis, and again, bootstrap support

was either very low (values less than 50% are shown as polytomies) or

moderate based on 1,000 replicates (Figure 5.4). Some of the clones in this

analysis were closely related to known taxa whereas others seem to be

phylogenetically distinct groups. Several clones isolated from chewing lice

found on the pocket gopher Thomomys bottae are basal to the known

Staphylococcus species shown in Figures 5.3 and 5.4. When 16S rRNA

sequences from these clones were queried against the NCBI database using

BLAST searches, the most similar taxa were those used in this analysis. In

89

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. — 5 Changes ex. Thomomydoecus minor -21 ex. Thomomydoecus m inor -19 ex. Thomomydoecus minor -2 7 ex. Thomomydoecus minor -2 5 ex. Thomomydoecus minor -1 7 _ ex. Thomomydoecus minor -3 0 - e x . Geomydoecus setzeri-7 - ex. Geomydoecus geomydis-6 - e x . Thomomydoecus minor -2 0 e x . Geomydoecus centralis-11 . ex. Geomydoecus centralis-1 ex. Thomomys bottae*-2 e x . Thomomydoecus minor -8 e x . Thomomys bottae*-7 ex. Thomomys bottaeM e x . Geomydoecus expansis-21 ex. Geomydoecus geomydis-17 ex. Thomomys bottae*-5 I e x . T hom om ys bottae*-10 ex. Geomydoecus expansis-7 Staphylococcus succinus A F 0 0 4 2 2 0 e x . Geomydoecus expansis-9 ex. Geomydoecus expansis-1 r e x . Geomydoecus expattsis-10 - ex. Geomydoecus expansis-5 - e x . Geomydoecus expansis-8 Staphylococcus gallinarum D 8 3 3 6 6 — Staphylococcus saprophyticus L 2 0 2 5 0 e x . Geomydoecus panamensis-9 Staphylococcus cohnii A B 0 0 9 9 3 6 Staphylococcus pasteuri A B 0 0 9 9 4 4 CStaphylococcus wameri A J2 7 6 8 1 0 ex. Geomydoecus expansis-21 ex. Geomydoecus setzeri-3 £ex. Geomydoecus expansis-25 — Staphylococcus aureus X 6 8 4 1 7 1- Staphylococcus sciuri A F 0 4 1 3 5 8 Staphylococcus pulvereri A B 0 0 9 9 4 2 1*1 Staphylococcus vitulinus A B 0 0 9 9 4 6 •— Staphylococcus lentus D 8 3 3 7 0 Bacillus subtilis A J2 7 6 3 5 1 • Macrococcus bovicus Y 1 5 7 1 4

Figure 5.3. Maximum likelihood phylogeny for Staphylococcus species using the Tamura-Nei + I + G model (see text for details of model). Clones marked with an asterisk could be from either of two hosts (Thomomydoecus minor o r Geomydoecus centralis).

90

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7 0.0071 0.0045 0.0039 0.0311 0.0227 0.0229 0.1208 0.0234 0.0266 0,0045 0.0233 0.0343 0.0350 0.0188 0.0227 0.0361 0.0214 Thomomys Thomomys bottae. (table cont.) 0.0110 0.0097 0.0071 0.02750.0273 0.0263 0.0248 0.0259 0.0246 0,0220 0.0259 0.0233 0.0259 0.02530.0315 0.0251 0.0228 0.0232 0.1135 0.1126 0.0162 0.0136 0.0084 0.0045 5 6 0.0297 0.03500.0337 0.0104 0.0343 0.0091 0.0295 0.0248 0.0095 0.0318 0.0360 0.1129 0.1237 0.0317 0.02460.0080 0.0220 0.0324 0.0087 0.0312 0.0308 0.0331 0.0084 0.0304 0.0291 0.0239 (listed by louse taxon). Numbers following 0.0363 0.0356 0.0376 0.0363 0.0291 0.0337 0.0323 0.0310 0.0094 0.0304 0.0291 0.0304 0.0292 0.1005 0.0992 0.0344 0.0311 0.0298 0.0155 0.0130 Staphylococcus 0.0019 0.0026 0.0350 0.1136 0.1143 0.0357 0.0102 0.0108 0.0343 0.0350 0.0337 0.0337 0.0324 0.0337 0.0337 0.0343 0.0330 0.03300.0298 0.0318 0,0998 0.0304 0.0052 0.0032 0.0045 0.0323 0.0303 0.0310 0.0376 0.03560.0369 0.0337 0.0343 0.0330 0.0343 0.0323 0.0330 0,0317 0.0328 0.0308 0.0315 0.0302 0.0356 0.0337 0.0343 0.0330 0.0330 0.03130.0092 0.0293 0.0071 0.0300 0.0078 0.0287 0.0064 0.0298 0.03560.0356 0.0336 0.0343 0.0330 0.0356 0.03360.0389 0.0343 0.0369 0.0330 0.0339 0.0319 0.0325 0.0350 0.0330 0.03570.03300.0356 0.03370.0336 0.0310 0.0344 0.0317 0.0317 0.0331 0.03040.0362 0.0317 0.0342 0.0252 0.03220.0107 0.03220.0114 0.0302 0.0087 0.0329 0.0094 0.0308 0.0315 0.0100 0.0215 0.0195 0.0330 0.0310 0.0317 0.0304 0.0213 0.1018 0.0330 0.0311 0.0317 0.0324 0.03170.0343 0.0297 0.0323 0.0304 0.0330 0.0350 0.0330 0.0337 0.0324 0.0253 0.0356 0.0336 0.0343 0.0330 0.0259 0.0234 1 2 3 4 0.1111 0.1143 0.0304 0.0039 0.0643 0.0616 0.0350 0.0331 0.0337 0.0324 0.0078 0.01360.0597 0.0122 0.0473 0.0085 0.0609 0.0339 0.0590 0.0582 0.0629 0.0583 0.0349 0.0329 0.0100 0.0107 0.0948 0.0623 0.0603 0.0595 0.0609 9 6 0.0630 0.0363 0.0344 0.0350 17 0.0608 21 0.0595 10 24 0.0622 9 0.0642 5 0.0589 25 4 0.0595 8 0.0570 24 25 17 0.0316 20 0.0623 19 27 30 0.0329 0.0310 0.0575 8 0.0635 14 1 0.0623 7 3 0.0622 2 5 7 0.0596 4 0.0590 15

Table 5.2. Uncorrected P-distances for 42 taxa of Bacterial Taxa Thomomydoecus minor Thomomydoecus minor Thomomydoecus minor Thomomydoecus minor Geomydoecus setzeri Geomydoecus geomydis Thomomydoecus minor Thomomydoecus minor Geomydoecus centralis Geomydoecus expansus Bacillus subtilis Thomomydoecus minor Thomomydoecus minor Thomomys bottae* Geomydoecus expansus Geomydoecus geomydis Geomydoecus expansus Staphylococcus gallm rum Geomydoecus centralis Staphylococcus Jentus Geomydoecus expansus Geomydoecus setzeri Staphylococcus pasteuri Staphylococcus warneri Geomydoecus panamensis Geomydoecus expansus Staphylococcus cohnii S. saprophyticus Staphylococcus sciuri Geomydoecus expansus Thomomys bottae* Thomomys bottae* Geomydoecus expansus Geomydoecus expansus 7 Staphylococcus succinus Thomomys bottae Staphylococcus pulvereri Staphylococcus vitulinus Thomomys bottae* Geomydoecus expansus Staphylococcus aureus Macrococcus bovicus 5 7 10 15 taxon names refer to clones numbers. Asterisks1 denote clones that came from4 lice of the gopher 6 3 8 9 11 12 13 16 17 2 21 20 36 14 37 18 19 25 31 38 39 40 42 22 23 24 26 27 34 30 32 29 35 41 25 33

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. C v O O O v CO rH oo00 CN| cooo rH CM NO Tf tn CM Cm1-H 00 NO o CO CM V O C S fH CM OS 00 NO NO tM 00 On tM ON rH rH oo cm o f"H rH 00 Ov CmOn o w o o fH o rH o o o o CM CM o CM co o CO o CO |-H rH CM o l-H u p o o o o o o o o o o o o o o O o o o o o o o o i-h rH o d d o d d d d d d d d d o ’ d d d d d d d d d d d d 3 (0

NO t-« O CMCMON O n CO NO CO NO CM ON in oooo ON CO NO rH CM NO o tS*Cm ON NO Cm r r CM 1-H Cm CMin in NO Cm Cm oo CM NO CMCm CO Cm rH Cm NO o CM ON CM O CM CM COCM CM CM CM CM CM CM CM o i-h CM 1-H CO CM CO CM CO CM fH CO o rH o OOO o O o o o o o O O o o o o o O o O O O p p»-h i -h d o o o * o d o o* o* d o d d O d d d o d o ' d o’d o o ’ o ’ d

oo00ON ON CmTt

ON o O n CO Cm 0 0 CO rH rH tn o CO NO CO in CM Cm oo H NO CO CO 0 0 00 NO l-H CM ON |-H Cm ON CM O ON in l-H o 0 0 Cm ON rH o oo CM O CM o T f CO CO1-H 00 rH CO CM CM CM o CM ?-» O o o CM rH o o o o CM CM 1-H CM CO o CO o co rH CO o rH o o p p o o o p o o p o o o o o o O O p o O o o p o pl-H rH o’ d d o’ o’ o’ o’d d o' d o ’ d o’ d o o ’ o ’ d o d o ' o' o* o’ d o’ d d

CO 1—* NO 0 0 NO O ^ in CO oooo CM 0 0 o NO CO CM CM T* 0 0 NO tn in m CO CM ON NO r H o NO Cm NO O n GO H}

^HH*CMrHrHrn CM 00Cm Onin oooorH00 oo Onoooo NOoo CM COco Cmo in CO OOOn in On00ON ^ 00CMONCONOin vo0000ONONtn TJ1o co oo 00 in CMCM o Cm rHrHo CMrHCM rH rHCMrHo o o rH rHCMCMrHCMCOrHCOrHCOCMCMCOr H rn o O o O o p o o O o p po p o o o o o o O p O o o p O O p l-H r H d d d d d o d d d d d d d d d d d d d d d d d d d d d d d d d

in m Cm ON CM CM CM Cm o 0 0 ON o T}«oo m oo CO NO OO CM NO CM CO CO Cm o c o ON CO ON O rH O n o O n no O o l-H rH o oo ON o o ON NO NO O rH o CM o c o CO CM CO O n CM o rH CM O CM CM CM n f S CM CM o rH rH l-H CM CM CM CM l-H CM CO CM CO CM c o CM CM CO O rH o o O o OOOOO OO p OO o o O o o O p OO O p O o o O p 1-H l-H d d d d d d o d d d d d d d d d o* o ’ d d d d d d d d d d d d d d

tn oo Cm CM CM o CM rH r ji o Cm CM Cm o Cm •**» r* Cm ON QO in CO l-H O n in o •M* o NO CO Hji Cm rH CM m CM CM Cm 00 r i NO CM tn rH l-H O n CM CM ON in OO ON CM CM CM CM rH o ON O O CM CM o CM CM CM r n CM CM CM o rH n CM CM CM CM CM CM l-H 1-H CO CM CO CM CO CM CM CO rH rH O O O O o O O OOO o O o O o O O OO o o o p pO o o o o O p 1-H l-H d o ’ d o ’ d d d o d d d d d d d d o ’ d d d d d d d d d o ’ o ’ d d d o* d

OO NO 00 OOO CMin Tt< m n o CO 1-H CMCOCm oo r n O n CMin O n in NO 00 CMo tji Cm CO r H in c m in o CMc o o 1-H O CM CMin CMCOCM o o 1-H CMr n Cm OO CMl-H CO r H Ht* in c o r H O o o o CMCMo COCMCO 1-H CMCMCMo l-H r H CMCMCMCMCO CMr H CMCOCMCOCMCOCMCMCO r H CM o o o o O o o o poooOooOooOO O O o p o O o O o o O p r H r H o d d d d o d d o d d d d d d d d d d d d d d d d d d d d d d d d d

O r H CM CO to NO Cm CO O n o rH CM CO in NO Cm 00 ON o rH CMCO in NO 00 On o fH CM HPin'finvoM nffirti-iON r H rH rH rH r * rH r n r n ^ rH CM CM CM CM CM CMCM CM CM CM COCOCOCOCO co COCOCO CO

92

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. c CO T p r -i ON NO TP r - »1-H T p oot v CO 0 0 CO vO 0 0 CM LO tM ooO n in TT O n o o 0 0 o o ON i- h cm Cm tM O n O CM o o oCM CM o CM CO o CO o CM i- h rH CMo i- h U o o o o o o o o o o © o o O rH i—» J£ o o o o o o © o o o © o oo ’ o o © 3 n3

ONCMONin CM oo in ONCMCOin CMo o ONo TP NO NO CMin 1-H CMNOONNOo NOo Cm 00 On in T p r H i- h r H ♦H CMCM r H CMi-H CO r H CO r n r H CMo r H © o © O o o p o p O o O p O o O r H i - h o o ’ © o o ’ o ’ o ’ © o ’ o o o ’ o ’ o* o OO o

ON in CO in r H O n o T p 00 CM in O n o CM tP CM Cm VO NO r n co in VO C m OO in CM CONO o N O C m r H oo 00O n NO T p CM o l-H r r r H CMCM r H 1-H CO r H CO r H CO r n r CM o r H o o o Cr O o o O o o o p O o o OO r H t- h o*o’ © © O © o*o ’ oo ’ o ’ o o ’ o© ’ o’ O O ©

TpCm in 00 in r H tn ONtn CO co inO n o o o Cm vO O n o CM o On Cm 00 O' o 00 m in 00 CMONCMON TpCO O CMONCO r H O r H r n r r CMCMCMr H 1—H CO 1-H CO r H CO CMCMCOo r n CM OO OO o o o o O o OO O O p o O o 1-H © © o* © © © o ’ o ’ o o ’ O* © o o’o’o’o’©o ©

r H r H tn CO vO o inCMCm ooCMNO00 oo00 ooVOCMCO o O 00 CMON r n o o 0O CMO CMr p o H p COCO 1-H CO COCO CM r H r n o o o CMCM CMCO o coo co1-H 1-H CO o 1-H O o O o o p o o O p o p o op o o pr H rH o o O o o o o o ’ o*o ’ O o o o o ’ o o ’o o’ oo ’

00 TP HpCOCO in CMoo00 CO r H CMVOCMCMCMco1-H o o Hp in CO 1-H o CMHP tn tnr n Cm TpVOCMin 00 CMCm in CMTP NO1-H O o CM CMCMr H o o oCO CM1-H CMCO o CO r H CO 1-H CMCO o CM CM o p Oo o o o o O Oo O o p o o O op o1—1I— o’ o ’ o o’o o o o o ’ o’ oo ’ o* o o o ’ oo ’ o* o oo

ONCm tp m oo ONTp CMoo CO r H CM1-H oo Tp00 in coCm TpON T-H CMO n O n CmvO oo ON ON C m m o CMCMO n CMr H r H CMO n O n 00 ON r H o t-H r H r n o o oCMCM CMCO O coo CO 1-H CMo 1-H O O p p oo p O o O op O o O o op p oo 1-H o o o oo o o 0 o O o o o o o o o o o o o o o

CM vO CO O n O n COo o Cm COvO TP o Hpr n Cm 00 oo oo in ONCM00CO O n COo tP CMCMO n O n O n o TP O n o tM o ONo CO O n o tn VOm CO O i- h r H r H r H r H o O c? r H CMr H o COo COo CM1-H r H CMo r n OO O o o o o O p OO o o p o p o p o o o O 1*H 1-H o o ’ o ’ o ’ o o o o* o o o o* o* o o o o ’ o ’ o* o ’ o o o

O CM CMONO n CO VOCOvO CMO n 1-H m 0O coONCOVO r H CMvO O TP Cm ON ^P Cm O n Cm Cm m Tpin vO Cm Cm ONCm TpvO CMCm COCm r H tM vOO CMON CM CM CMCMCM CMCMCMCMCMCMO r H CMr H COCMCOCMCOCMl-H COO OOO o OO o oO p o o o O O O p O o O O p 1-H r n o o o o’o ’ o o o’ o O o ’ o o oO ’ o ’ o o ’ o ’ o o’o o*o

OHNO^I/1\ON«OO n O CM CO m vO C>H0 0 O n o rH CMCO in VO tv COOn o rH CM nfMO^lfivONOOO'pHriiHP^HfHHfHrljH^HN CMeM CMCMCMCMCMCMCMCOCO CO COCO CO COCOCOCOTP TP TP

93

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. O n o o cm ON CM ON Cm c NC t p O n ON ^ p NO in CO o CO CM o N ON o co © © o © o © o u o o o o ’ © © o .2 X>cc

, ■CM ON in CM Tt* 00 CO «—«N o CM N Tp CO o CO 1—< CO o co o o p o o o o o ' o o ' o o © O

Cm CM 1—* o CM in CM CO CO o NO t t ON o c TT CM co CO o CO o CO CM o rM O n co o o o o o o o © O o o o ' o o o ' o ’ o* o

Tp o 1—■* Cm Tp ON Tp 00 CM CM ON O n in NO o CM CM CM CM CM CM CM o CM 00 O CO o o © o O O o o o *—• o ' o ' o o o o ' o ' o ’ o ’ o ’

TpT- O 00 oo oo O n CO CM TpNO Cm o o o co ON O NO NO o in CO <—* CO o CM FCM oo ON O o p o o o o O o O o ' o 'o ’ o ’ o ' o o o o ’ o ' ©

t—»00 00 o in Tp o o in CO «—> o VO 0O CM ON in oo CO vO in vO o CM CM CM CM CM CM CM T“^ CM o CO O o o o p o o o O o *—* *—• o © o ’ o ' © © o ' © o o o ©

in in on CM00 oo CMCm f—1COCM 00 in Cm CO CMTT Cm 1“^ OOvO COON ON f—HCM1—4CO CMCO CMCOCM COo 1—* CM o O © o o O O p O O o «—* o o o o o ©o'o' o o o o o ’

in Cm CM 0 0 CM vO 0 0 0 0 0 0 0 0 vO CM CO o 0 0 CM o CM Tp o TT00 CO r—< 0 0 CO 00 oo CM CM 1-^ CM CO O CO o CO f—* 1-^ CO o CM o O p O o p p o o o o o o ' o ’ o o ’ o* o o o ' o ' o ' © o © o

VO ON o ,,in o CM , CM CM ON in Cm TP O CM CM o CM co o TP 00 CO Cm o CM Cm cm O CM CM CM co o CO o CO T-*!—• co O r-^ CM pp oooo oooo OO o r-< *—« o ' o o o o ’ o o o ' © o o ' o o o ’ o

vO CO CM Cm TP ''p 00 VO in ' p in in CO Cm ON vO O ^ Cm CM ON 1-^ CM O CO Cm CM o Cm O n vO o O O CM CM o CM CO O CO O CO CM o CM OOOO p O O O p o o o o O o o o O o O o o o © © o o ' © o ’ o

CM a o ON o CM CO Tp in Cm 00 On o r>HCM CM CM CM CO CO CO CO CO CO CO CO CO CO *p Tp ^P

94

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

00 CO

CM CO Cs* CO CO p ©

t n i n O T— ©CO p 1—

o © 0.0183 0.0284 0.0826 0.1021 0.0285 0.0071 0.0743 0.0193 0.0977 0.0277 0.1057 0.1193 0.0263 0.0964 0.1182 0.0489 0.0775 0.0689

r^rlCO^invOtNOOOM-^r^t-ti^^r^rHTHr^rNCMCMCMCMCNieMrJeMCMC'leOCOCOeoeOCOCO^^^^'^^

95

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - ex. Thomomydoecus minor -24 - ex. Thomomydoecus minor -19 - ex. Thomomydoecus minor -27 - ex. Thomomydoecus minor -25 - ex. Thomomydoecus minor -17 - ex. Thomomydoecus minor -30 - ex. Geomydoecus setzeri-7 66 - ex. Geomydoecus geontydis-6 100 - ex. Thomomydoecus minor -20 90 - ex. Geomydoecus centralis-14, - ex. Geomydoecus centralis-1 90 - ex. Thomomys bottae*-2 - ex. Thomomydoecus minor -8 67 - ex. Thomomys bottae*-7 - ex. Thomomys bottae*-4 - ex. Geomydoecus expansis-21 76 - ex. Geomydoecus geomydis-17 64 - ex. Thomomys bottae*-5 63 81 98 - ex. Thomomys bottae*-10 ■ ex. Geomydoecus expansis-7 • Staphylococcus succinus AF004220 54 ■ ex. Geomydoecus expansis-9 ■ ex. Geomydoecus expansis-4 ■ ex. Geomydoecus expansis-10 93 ■ ex. Geomydoecus expansis-5 80 ■ ex. Geomydoecus expansis-8 99 ■ Staphylococcus gallinarum D83366 92 • ex. Geomydoecus panamensis-9 ■ Staphylococcus cohnii AB009936 ■ Staphylococcus saprophyticus L20250 60 • ex. Geomydoecus expansis-24 100 ■ ex. Geomydoecus setzeri-3 76 ■ ex. Geomydoecus expansis-25 • Staphylococcus aureus X68417 85 • Staphylococcus pasteuri AB009944 Staphylococcus wameri AJ276810 99 Staphylococcus pulvereri AB009942 61 Staphylococcus vitulinus AB009946 Staphylococcus lentus D83370 Staphylococcus sciuri AF041358 Bacillus subtilis AJ276351 Macrococcus bovicus Y15714

Figure 5.4. Fifty percent majority rule consensus tree of 1000 bootstrap replicates using the parsimony optimality criterion for Staphylococcus species. Clones marked with an asterisk could be from either of two hosts (Thomomydoecus minor or Geomydoecus centralis).

96

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. other words, these clones have 16S rRNA sequences that represent lineages

distinct from those surveyed elsewhere and deposited in the NCBI database

(GenBank).

DISCUSSION

Culture Independent Methods

The effectiveness of 16S rRNA genes for detecting phylogenetic signal

among Eubacteria has been debated widely. Since culture-independent

sampling methods came into practice, (Hugenholtz and Pace 1996,

Hugenholtz et al. 1998, and Pace 1997), the number of ribosomal DNA

sequences in nucleotide databases has grown rapidly. However, some

researchers have cast doubt on the effectiveness of the 16S rRNA gene in

phylogenetic studies of closely related organisms. David Ward and his

colleagues (Ward et al. 1998) showed that the 16S rRNA gene was

insufficient to resolve the phylogenetic relationships of distinct ecotypes of

bacteria associated with microbial mats communities. Furthermore, the

ecologically and genetically distinct populations studied by Ward and his

colleagues (differentiated based on internal transcribed spacer DNA

sequences) had identical 16S rRNA sequences over their entire length

(>1,500 bp). However, studies such as those by Paul Baumann and his

colleagues (e.g., Baumann et al. 1997) found sufficient variation in the 16S

rRNA gene to resolve the phylogenetic relationships of apparently closely

related species. The discrepancy may be in the designation of bacterial

"species."

97

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Because there are so many 16S rRNA sequences for Eubacteria in

genetic databases, researchers using culture-independent analyses are

constrained to use ribosomal sequences to identify unknown samples. As

microbiologists add other genes to these databases, there will be decreased

dependence on ribosomal sequences. Regardless of their usefulness in

phylogenetics, ribosomal DNA sequences provide an effective "name tag"

for unidentified bacteria. In my work, the 16S rRNA gene showed that the

pocket gopher-chewing louse assemblage harbored many types of bacteria

(see chapter 3). More specific analyses (e.g. the analysis of gamma-

Proteobacteria and Staphylococcus) showed that DNA samples extracted

from chewing louse samples can also contain numerous closely related taxa.

Bacteria sampled from lice cluster within and among well known species of

bacteria (Figs. 5.1 - 5.4). However, many of these species were described based

on the amount of 16S rRNA sequence similarity between unknown samples

and previously sequenced species. David Ward (1998) has discussed the

flaws in this phenetic classification procedure and has encouraged

microbiologists to use a more natural species concept. Therefore, the

delineation of "species" for the taxa of bacteria associated with chewing lice

(e.g., Acinetobacter or Staphylococcus) is problematic due to lack of

information on the natural history of these organisms.

Analyses of the gamma-Proteobacteria

Bacterial 16S rRNA sequences that were tentatively identified as

gamma-Proteobacteria (using BLAST searches of GenBank) were completely

98

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. sequenced and analyzed cladistically. Known gamma-Proteobacteria were

analyzed along with the clones sampled from chewing lice. Some clones

grouped within the dade of Acinetobacter species downloaded from

GenBank (Figure 5.1). Three clones (88-5-1, 88-5-2, and B-22) were closely

related to the Acinetobacter spedes, but were found to be sister to them

rather than nested within Acinetobacter. W hen the com plete 16S rR N A

sequences for these clones were queried against GenBank, the most closely

related taxa were the Acinetobacter species used in this analysis rather than a

group nested between Acinetobacter and the outgroup taxa. These clones

could represent either a new lineage of Acinetobacter or a distinct

undescribed lineage. Three additional clones appear outside the

Acinetobacter dade (B-26, 91-19, and 349-11; Fig. 5.1). BLAST searches of

GenBank determined that the dosest taxa to these three outliers are

members of the genus Pseudomonas. Previously, these clones had been

assigned to the genera Legionella, Pseudomonas, and Xanthomonas based

on partial 16S rRNA sequences. Figure 5.1 was rooted with the only

putative outgroup taxon known at the time {Pseudomonas aeruginosa).

Figure 5.2 (analyzed after Figure 5.1) used the three outliers and P.

aeruginosa as outgroup taxa. It is evident that the relationship among the

four individuals has changed in Fig. 5.2 and warrants further investigation.

To understand the phylogenetic relationship and taxonomic placement of

these taxa, better sampling and a new series of outgroup taxa will be

required.

99

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Bootstrap analysis (Figure 5.2) shows some ambiguity in the

relationships of the known species of gamma-Proteobacteria. It appears that

the low bootstrap support for some clades is a product of the nucleotide

substitution patterns inherent in the ribosomal DNA sequences of these

taxa. There are long regions of invariable sites punctuated by regions of high

variability (Appendix II). Where nucleotide changes occur between taxa, the

changes contain homoplasious information and provide limited

phylogenetic signal. As a result, some clades remain unresolved. A genetic

marker that evolves at a rate faster than the 16S rRNA gene may help to

clarify these relationships.

Analyses of Staphylococcus

Clones that were similar to Staphylococcus based on preliminary

BLAST searches of GenBank were completely sequenced and analyzed by

maximum likelihood and parsimony analyses. Eleven species of

Staphylococcus were downloaded from GenBank along with two outgroup

taxa, Bacillus subtilis and Macrococcus bovicus. Many of the clones extracted

from chewing lice clustered within the known Staphylococcus taxa (Fig. 5.3).

Whereas other clones (Thomomydoecus minor 24, 19, 27, 25, and 17) appear

as basal lineages relative to the known Staphylococcus species (Fig. 5.3). Still

other clones (those from Thomomydoecus minor, Geomydoecus setzeri, G.

geomydis, G. centralis, and those from the gopher taxon Thomomys bottae)

form a unique clade nested within the known Staphylococcus species (Fig.

5.3). The most basal lineages of clones (Fig. 5.3) could be unique to the

100

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Thomomys bottae assemblage or an artifact of incomplete taxon sampling.

The complete 16S rRNA sequence for every clone in this analysis was

queried against GenBank and the closest matches were those found in the

main clade of Staphylococcus species. These clones appear to form two basal

lineages (all from Thomomydoecus minor) and one derived lineage (those

from Thomomydoecus minor, Geomydoecus setzeri, G. geomydis, G.

centralis, and those from the gopher taxon Thomomys bottae) related to

Staphylococcus. For the same reasons stated previously, it would be

premature to formally name these taxa as distinct species based solely on

their 16S rRNA sequences. However, both of the basal lineages and the

derived lineage appear to be quite distinct from the taxa of Staphylococcus

for which 16S rRNA sequences are known (Fig. 5.4).

Bootstrap analysis, based on 1,000 replicates using the parsimony

optimality criterion, shows weak to moderate support for most of the clades

in the analysis of Staphylococcus and Staphylococcus-like clones . It seems

that the 16S rRNA gene is able to resolve interspecific relationships in this

clade, but it fails to resolve relationships among more closely related taxa. A

genetic marker that evolves at a rate faster than the 16S rRNA gene would

be beneficial in future analyses of these bacteria.

Cophylogeny

Of the two phylogenetic analyses presented in this chapter, only the

one involving species related to Staphylococcus (Fig. 5.3) could be examined

for patterns of cophylogeny. The clones sampled from chewing lice in the

101

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. analysis of the gamma-Proteobacteria (Fig. 5.1) came from only two hosts

and studies of cophylogeny must contain at least three hosts as there is only

one phylogenetic relationship possible for two taxa. Patterns of cophylogeny

often are cryptic to the investigator, therefore, statistical tests of similarity

between host and parasite phylogenies are the only objective way to discern

these patterns. In the case of the Staphylococcus species, there are five

chewing louse hosts {Thomomydoecus minor, Geomydoecus centralis, G.

expansus, G. p anam ensis, and G. setzeri) and 29 bacterial clones sampled

from the five hosts. Preliminary analyses of chewing louse and

Staphylococcus tree topologies (using TreeMap) suggest that there are only

three cospeciation events and many host switching events that encompass

the entire tree topology depicted in Figures 5.3 and 5.4. However, patterns of

cophylogeny may exist in localized areas of the phylogeny. For example,

certain subclades within the Staphylococcus assemblage may show

cophylogeny with their insect hosts even though the overall assemblage

does not. The difference, of course, is the scale at which one compares host-

parasite pairs. Despite little evidence of cospeciation between

Staphylococcus and their chewing louse hosts in this analysis, there could be

evidence of cospeciation between one or two clades of Staphylococcus and

their hosts. Currently, investigation at lower phylogenetic levels is

hampered by incomplete taxon sampling. Future studies should sample

specific Staphylococcus clades intensively by designing PCR primers that

102

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. preferentially anneal to only a few Staphylococcus species, which make it easier to

amplify the target sequences.

Chewing Louse and Bacteria Associations

Because this study was the first to document the association of bacteria with

chewing lice of the family Trichodectidae, it is not surprising that this analysis generated

more questions than it answered. I have now documented that multiple taxa of both

Acinetobacter and Staphylococcus are associated with the pocket gopher-chewing louse

system. Whether these taxa represent multiple species of bacteria remains to be tested.

At present, I cannot say whether either lineage is truly endosymbiotic. To test the

hypothesis of endosymbiosis, future research should use microscopy techniques in

combination with in situ hybridization, which causes target species of bacteria to

fluoresce. This would allow the researcher to target certain species of bacteria to

determine if they are present within the chewing louse host. It is possible that the

bacteria surveyed in this study are only transient associates of chewing lice. Despite

efforts to sterilize the surface of chewing lice prior to DNA extraction, it is conceivable

that the bacteria described herein are actually associated with the pocket gopher or the

underground burrow system in which the assemblage lives. Undoubtedly, we are only

scratching the surface of a complex interaction between a mammal, an insect, and

numerous microorganisms. This particular host-parasite system (gophers, lice, and

Eubacteria) is exceptionally promising because of the high level of population isolation

103

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. documented for the mammal and insect lineages. Future work must

include base-level characterization of the microbial community associated

with pocket gophers, chewing lice, and their subterranean burrow systems.

In addition, knowledge of the basic natural history of the bacteria associated

with the gopher-louse system is important so that future studies of

cospeciation can focus on lineages of bacteria that have biologically

meaningful interactions with the (gopher-louse) system, rather than

transient bacteria or forms introduced by contamination.

104

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. SUMMARY AND CONCLUSIONS

This dissertation examined the possible mechanisms of cospeciation

in a host-parasite system and surveyed new associates in a model system for

the study of cophylogeny. Pocket gophers and their ectoparasitic chewing

lice have shared a long history of association. Because of low host and

parasite vagility, lineages of chewing lice have been stranded on lineages of

pocket gophers for millennia. Through time, these lineages have speciated

in parallel. This pattern of cophylogeny has likely been reinforced by many

mechanisms, including the evolution of an intricate fit between the louse

rostral groove and the hair shaft of its natural host (chapter 1). Whereas the

fit between louse rostral grooves and gopher hair shafts may reinforce

patterns of cophylogeny at interspecific levels, it does not seem to be a m eans

of habitat partitioning by competing species of chewing lice that live on the

same pocket gopher host (chapter 2).

Certain species of chewing lice can coexist with other louse species,

but this situation is the exception, rather than the rule in the gopher-louse

system. For example, two genera of lice (Thomomydoecus an d

Geomydoecus) occasionally co-occur on pocket gophers of the genus

T h o m o m ys. In chapter 2, I showed that those coexisting louse species were

spatially segregated over the pelage of the gopher, but the mechanism by

which they partition their habitat does not appear to be differential ability to

grasp hairs of different diameter.

105

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. The last three chapters of this dissertation involve the search for new

symbionts in the gopher-louse association, namely bacterial taxa associated

with the chewing lice. In chapter 3, I used culture-independent methods

pioneered by Carl Woese and Norm Pace to identify 35 bacterial lineages that

were found in association with pocket gophers, chewing lice, and their

subterranean burrow system. In chapter 4, I demonstrated that there was

bacterial DNA associated with the eggs of chewing lice. Vertical

transmission of endosymbionts through eggs is a common means of

dispersal for endosymbiotic bacteria of insects. In the final chapter, I

determined the phylogenetic relationships of two lineages of bacteria

associated with this system. I sequenced the complete 16S rRINA gene for all

clones related to the genera Staphylococcus an d Acinetobacter. I used

parsimony and maximum likelihood analyses to generate phylogenetic

hypotheses for the bacteria that were compared to those of their hosts

(chewing lice). Unfortunately, little can be said about w hether bacteria show

evidence of cophylogeny with their hosts until taxon sampling is more

complete. It also will be necessary to determine whether thtese bacteria are

true endosymbionts or merely transients in this system. Future research in

this area will focus on determining which of the 35 bacterial lineages are

truly endosymbiotic, further sampling of Staphylococcus a n d Acinetobacter,

and identifying additional lineages of bacteria that almost certainly exist, but

have yet to be discovered.

106

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

Aksoy, S., A. A. Pourhosseini, and A. Chow. 1995. M ycetome endosym bionts of tsetse flies constitute a distinct lineage related to Enterobacteriaceae. Insect Molecular Biology 4:15-22.

Aksoy, S. 1995. Molecular analysis of the endosymbionts of tsetse flies: 16S rDNA locus and over-expression of a chaperonin.Insect Molecular Biology 4:23-29.

Aksoy, S., X. Chen, and V. Hypsa. 1997. Phylogeny and potential transmission routes of midgut-associated endosymbionts of tsetse (Diptera: Glossinidae). Insect Molecular Biology 6:183-190.

Baker, R. J., and S. L. Williams. 1972. A live trap for pocket gophers. Journal o f Wildlife Management 36:1320-1322.

Bandi, C., G. Damiani, L. Magrassi, A. Grigolo, R. Fani, and L. Sacchi. 1994. Flavobacteria as intracellular symbionts in cockroaches. Procedings of the Royal Society of London B 257:43-48.

Baumann, P., L. Baumann, C. Lai, and D. Rouhbakhsh, N. A. Moran, and M. A. Clark. 1995. Genetics, physiology, and evolutionary relationships of the genus Buchnera: Intracellular symbionts of aphids. Annual Review of Microbiology 49:55-94.

Baumann, P. and N. A. Moran. 1997. Non-cultivable microorganisms from symbiotic associations of insects and other hosts. Antonie von Leeuwenhoek 72:39-48.

Baumann, P., N. A. Moran, and L. Baumann. 1997. The evolution and genetics of aphid endosymbionts. Bioscience 47:12-20.

Berchtold, M., W. Ludwig, and H. Kunig. 1994. 16S rDNA sequence and phylogenetic position of an uncultivated spirochete from the hindgut of the termite Mastotermes darwiniensis Froggatt. FEMS Microbiology Letters 123:269-274.

Brooks, D. R. 1980. Allopatric speciation and non-interactive parasite community structure. Systematic Zoology 29:192-203.

Brune, A., and M. Friedrich. 2000. Microecology of the termite gut: structure and function on a microscale. Current Opinion in Microbiology 3:263-269.

Brynnel, E. U., C. G. Kurland, N. A. Moran, and S. G. E. Anderson. 1998. Evolutionary rates for tu f genes in endosymbionts of aphids. M olecular Biology and Evolution 15:574-582.

107

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Cary, S. C., W. Warren, E. Anderson, and S. J. Giovannoni. 1993. Identification and localization of bacterial endosymbionts in hydrothermal vent taxa with symbiont-specific polymerase chain reaction amplification and in situ hybridization techniques. Molecular Marine Biology and Biotechnology 2:51-62.

Cunningham, C. W., H. Zhu, and D. M. Hillis. 1998. Best-fit maximum likelihood models for phylogenetic inference: empirical tests with known phylogenies. Evolution, 52:978-987.

Dadd, R. H. 1985. N utrition: organisms, Pp.315-319, in, Comprehensive insect physiology, biochemistry, and pharmacology, vol. 4. (G. A. Kerkut and L. I. Gilbert, eds.). Pergamon Press, Inc., Oxford, N. Y.

Demastes, J. W. and M. S. Hafner 1993. Cospeciation of pocket gophers and their chewing lice. Journal of Mammalogy 74:521-530.

Demastes, J. W., M. S. Hafner, D. J. Hafner, and T. A. Spradling. 1998. Pocket gophers and chewing lice: a test of the maternal transmission hypothesis. Molecular Ecology 1065-1069. 7:

Douglas, A. E. 1989. Mycetocyte symbiosis in insects. Biological Review Cambridge Philosophical Society 64:409-434.

Douglas, A. E., and W. A. Prosser. 1992. Synthesis of the essential amino acid tryptophan in the pea aphid Acyrthosiphon( pisum) symbiosis. Journal o f Insect Physiology 38:565-568.

Durden, L. A. 1987. Interspecific associations between epifaunistic arthropods on the eastern chipmunk, Tamias striatus, and an overview of arthropod coexistence on mammals in general. Journal of the Tennessee Academy of Science 62:44-51.

Eberle, M. W., and D. L. McLean 1982. Initiation and orientation of the symbiote migration in the human body louse Pediculus humanus L. Journal of Insect Physiology 28:417-422.

Eberle, M. W., and D. L. McLean 1983. Observation of symbiote migration in human body lice with scanning and transmission electron microscopy. Canadian Journal of Microbiology 29:755-762.

Emerson, K. C., and R. D. Price. 1981. A host-parasite list of the Mallophaga on m am m als. Miscellaneous Publications of the Entomological Society of America 12:1-72.

Farrelly, V., F. A. Rainey, and E. Stackebrandt. 1995. Effect of genomic size and rrn gene copy number on PCR amplification of 16S rRNA genes from a mixture of bacterial species. Applied and Environmental Microbiology 61:2798-2801.

108

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Felsenstein, J. 1981. Evolutionary trees from DNA sequences: a maximum likelihood approach. Journal of Molecular Evolution 17:368-376.

Felsenstein, J. 1985. Phylogenies and the comparative method. Am erican Naturalist 125:1-15.

Garland, T. Jr., P. H. Harvey, and A. R. Ives. 1992. Procedures for the analysis of comparative data using phylogenetically independent contrasts. Systematic Biology 41:18-32.

Gause, G. F. 1934. The struggle for existence. The Williams and Wilkins Company, Baltimore, Maryland, 163 pp.

Hafner, M. S. 1982. A biochemical investigation of geomyoid systematics (Mammalia: Rodentia). Z. Zool. Syst. Evolutions Forsch. 20:118-130.

Hafner, M. S., and R. D. M. Page. 1995. M olecular phylogenies and host-parasite cospeciation. Philosophical Transactions of the Royal Society of London 349:77-83.

Hafner, M. S., J. W. Demastes, T. A. Spradling, and D. L. Reed. In press. Cophylogeny between pocket gophers and chewing lice. In: (R.D.M. Page, ed.) Cophylogeny, Univ. Chicago Press.

Hafner, M. S., and S. A. Nadler. 1988. Phylogenetic trees support the coevolution of parasites and their hosts. Nature 332:258-259.

Hafner, M. S., and S. A. Nadler. 1990. Cospeciation in host-parasite assemblages: comparative analysis of rates of evolution and timing of cospeciation events. Systematic Zoology 39:192-204.

Hafner, M. S., P. D. Sudman, F. X. Villablanca, T. A. Spradling, J. W. Demastes, and S. A. Nadler. 1994. Disparate rates of molecular evolution in cospeciating hosts and parasites. Science 265:1087-1090.

Harvey, P. H., and A. E. Keymer 1991. Comparing life histories using phylogenies. Philosophical Transactions of the Royal Society of London Series 332:31- B 39.

Harvey, P. H., and M. D. Pagel. 1991. The comparative method in evolutionary biology. Oxford University Press, Oxford, United Kingdom. 239 pp.

Hellenthal, R. A., and R. D. Price 1984. Distributional associations among Geomydoecus and Thomomydoecus lice (Mallophaga: Trichodectidae) and pocket gopher hosts of the Thomomys bottae group (Rodentia: Geomyidae). Journal of Medical Entomology 21:432-446.

109

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Hellenthal, R. A., and R. D. Price. 1991. Biosystematics of the chewing lice of pocket gophers. Annual Review of Entomology 36:185-203.

Hellenthal, R. A., and R. D. Price. 1994. Two new subgenera of chewing lice (Phthiraptera: Trichodectidae) from pocket gophers (Rodentia: Geomyidae), with a key to all included taxa. Journal of Medical Entomology 31:450-466.

Honeycutt, R. L., and S. L. Williams. 1982. Genic differentiation in pocket gophers of the genus , with comments on intergeneric relationships in the subfamily Geomyinae. Journal of Mammalogy 63:208-217.

Hugenholtz, P., B. M. Goebel, and N. R. Pace. 1998. Impact of culture-independent studies on the emerging phylogenetic view of bacterial diversity. Journal o f Bacteriology 180:4765-4774.

Hugenholtz, P., and N. R. Pace. 1996. Identifying microbial diversity in the natural environment: a molecular phylogenetic approach. Trends in Biotechnology 14:190-197.

Hugenholtz, P., C. Pitulle, K. L. Hershberger, and N. R. Pace. 1998. Novel Division level bacterial diversity in a Yellowstone hot spring. Journal of Bacteriology 180:366-376.

Ishikawa, H. 1989. Biochemical and molecular aspects of endosymbiosis in insects. International Review of Cytology 116:1-45.

Johnson, K. P., and D. H. Clayton. In press. Cophylogeny at the community level: Comparing phylogenies of wing and body lice to that of columbiform hosts. In: (R.D.M. Page, ed.) Cophylogeny, Univ. Chicago Press.

Kellogg, V. L. 1913. Distribution and species-forming of ectoparasites. The American Naturalist 47:129-158.

Komatsoulis, G. A., and M. S. Waterman 1997. A new computational method for detection of Chimeric 16S rRNA artifacts generated by PCR Amplification from mixed bacterial populations. Applied and Environmental Microbiology 63:2338-2346.

LaBarbera, M. 1989. Analyzing body size as a factor in ecology and evolution. Annual Review of Ecology and Systematics 20:97-117.

Lambert, J. D., and N. A. Moran. 1998. Deleterious mutations destabilize ribosomal RNA in endosymbiotic bacteria. Proceedings of the National Academy of Sciences 95:4458-4462.

Lambert, L. H., T. Cox, K, Mitchell, R. A. Rossello-Mora, C. Del Cueto, D. E. Dodge, P. Orkland, and R. J. Cano. 1998. Staphylococcus succinus sp. nov., isolated from Dominican amber. International Journal of Systematic Bacteriology 48:511-518.

110

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Marshall, A. G. 1981. The Ecology of Ectoparasitic Insects. Academic Press, London. 459 pp.

Martin, R. D., and A. D. Barbour. 1989. Aspects of line-fitting bivariate allometric analyses. Folia Primatologica 53:65-81.

Mathiak, H. A. 1938. A key to the hairs of the mammals of southern Michigan. The Journal of Wildlife Management 2:251-268.

Mayer, W. V. 1952. The hair of California mammals with keys to the dorsal guard hairs of California mammals. The American Midland Naturalist 48:480-512.

Moran, N. A. 1996. Accelerated evolution and Muller's ratchet in endosymbiotic bacteria. Proceedings of the National Academy of Sciences 93: 2873-2878.

Moran, N. A., M. A. Munson, P. Baumann, and H. Ishikawa. 1993. A molecular clock in endosymbiotic bacteria is calibrated using the insect hosts. Proceedings of the Royal Society of London Series 253:167-171. B

Moran, N. A., C. D. von Dohlen, and P. Baumann 1995. Faster evolutionary rates in endosymbiotic bacteria than in cospeciating insect hosts. Journal o f Molecular Evolution 41:727-731.

Moran, N. A., and P. Baumann. 1994. Phylogenetics of cytoplasmically inherited microorganisms of arthropods. Trends in Ecology and Evolution 9:15-20.

Moran, N. A., and A. Telang. 1998. Bacteriocyte-associated symbiosis of insects. Bioscience 48:295-304.

Morand, S., M. S. Hafner, R. D. M. Page, and D. L. Reed. 2000. Comparative body size relationships in pocket gophers and their chewing lice. Biological Journal of the Linnean Society 70:239-249.

Munson, M. A., P. Baumann, M. A. Clark, L. Baumann, N. A. Moran, D. J. Voegtlin, and B. C. Campbell. 1991. Evidence for the establishment of aphid-Eeubacterium endosymbiosis in an ancestor of four aphid families. Journal of Bacteriology 173:6321-6324.

Munson, M. A., L. Baumann, and P. Baumann. 1993. Buchnera aphidicola (a prokaryotic endosymbiont of aphids) contains a putative 16S rRNA operon unlinked to the 23S rRNA-encoding gene: sequence determination, and promoter and terminator analysis. Gene 137:171-178.

Murray, M. D. 1957a. The distribution of the eggs of M am m alian Lice on their hosts I. Description of the oviposition behaviour. Australian Journal of Zoology 5:13-18.

I l l

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Murray, M. D. 1957b. The distribution of the eggs of Mammalian Lice on their hosts. H. Analysis of the oviposition behaviour of Damalma ovis (L.). Australian Journal of Zoology 5:19-29.

Nardon, P., and A. M. Grenier. 1989. CRC Handbook on Endocytobiosis in Coleoptera: Biological, biochemical, and genetic aspects. Pp 175-216 in: Insect endocytobiosis: Morphology, physiology, genetics, and evolution. (W. Schwemmler and G. Gassner, eds.) 272pp.

Nowak, R. M. 1999. Walker's mammals of the world. The Johns Hopkins University Press, Baltimore, Maryland 1629 pp.

Ochman, H., S. Elwyn, and N. A. Moran. 1999. Calibrating bacterial evolution. Proceedings of the National Academy of Sciences 96:12638-12643.

Ochman, H., and A. C. Wilson. 1987. Evolution in bacteria: evidence for a universal substitution rate in cellular genomes. Journal of M olecular Evolution 26:74-86.

Ohkuma, M., and T. Kudo. 1996. Phylogenetic Diversity of the Intestinal Bacterial Community in the Termite Reticulitermes speratus. Applied and environmental microbiology 62:461-468.

Pace, N. R. 1997. A molecular view of microbial diversity and the biosphere. Science 276:734-740.

Page, R. D. M. 1990a. Component analysis: A valiant failure? Cladistics 6:119-136.

Page, R. D. M. 1990b. Temporal congruence and cladistic analysis of biogeography and cospeciation. Systematic Zoology 39:205-226.

Page, R. D. M. 1993a . COMPONENT 2.0. Tree comparison software for use with Microsoft® Windows™. The Natural History Museum, London, UK.

Page, R. D. M. 1993b. Genes, organisms, and areas: the problem of multiple lineages. Systematic Biology 42:77-84.

Page, R. D. M. 1994. Maps between trees and cladistic analysis of historical associations among genes, organisms, and areas. Systematic Biology 43:58- 77.

Page, R. D. M. in press. Cospeciation. University of Chicago Press.

Page, R. D. M., and M. S. Hafner. 1996. Molecular phylogenies and host-parasite cospeciation: gophers and lice as a model system. Pp. 255-270, in: New Uses for New Phylogenies (P. H. Harvey, et al., eds.). Oxford Univ. Press, 349 pp.

112

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Page, R. D. M., R. D. P. Price, and R. A. H ellenthal. 1995. Phylogeny of Geomydoecus and Thomomydoecus pocket gopher lice (Phthiraptera: Trichodectidae) inferred from cladistic analysis of adult and first-instar morphology. Systematic Entomology 20:129-143.

Patton, J. L., and P. V. Brylski. 1987. Pocket gophers in alfalfa fields: causes and consequences of habitat-related body size variation. American Naturalist 130:493-506.

Patton, J. L., M. F. Smith, R. D. Price, and R. A. Hellenthal. 1984. Genetics of hybridization between the pocket gophers Thomomys bottae and Thomomys tozonsendii in northeastern California. Great Basin Naturalist 44:431-440.

Posada, D., and K. A. Crandall (1998). Modeltest: testing the model of DNA substitution. Bioinformatics 14:817-818.

Price, R. D. and K. C. Emerson 1972. A new subgenus and three new species of Geomydoecus (Mallophaga: Trichodectidae) from Thom om ys (Rodentia: Geomyidae). Journal of Medical Entomology 9:463-467.

Price, R. D., R. A. Hellenthal, and M. S. Hafner. 1985. The Geomydoecus (Mallophaga: Trichodectidae) from the Central American pocket gophers of the subgenus Macrogeomys (Rodentia: Geomyidae). Proceedings of the Entomological Society of Washington 87:432-443.

Purcell, A. H., T. Steiner, F. Megraud, and J. Bove. 1986. In vitro isolation of a transovarially transmitted bacterium from the leafhopper Euscelidius variegatus (Hemiptera: Cicadellidae). Journal of Invertebrate Pathology 48:66-73.

Purvis, A., and A. Rambaut. 1995. Comparative analysis by independent contrasts (CAIC): an Apple Macintosh application for analyzing comparative data. Computer Applications for Biosciences 11:247-251.

Rainey, F. A., E. Lang, and E. Stackebrandt. 1994. The phylogenetic structure of the genus Acinetobacter. FEMS Microbiology Letters 124:349-354.

Rainey, F. A., N. L. Ward-Rainey, P. H. Janssen, H. Hippe, and E. Stackebrandt. 1996. Clostridium paradoxum DSM 7308T contains multiple 16S rRNA genes with heterogeneous intervening sequences. Microbiology 142:2087- 2095.

Reed, D. L. 1994. Possible mechanism for cospeciation in pocket gophers and lice, and evidence of resource partitioning among coexisting genera of lice. M. S. Thesis. Louisiana State University, Baton Rouge, Louisiana, 74 pp.

113

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reed, D. L., and M. S. Hafner. 1997. H ost specificity of chewing lice on pocket gophers: a potential mechanism for cospeciation. Journal of Mammalogy 78:655-660.

Ries, E. 1931. Die symbiose der Lause und Federlinge. Zeitschrift fiir Morph. Okol. Tiere 20:233-367.

Robison-Cox, J. F., M. M. Bateson, and D. M. Ward. 1995. Evaluation of nearest- neighbor methods for detection of chimeric small-subunit rRNA sequences. Applied and Environmental Microbiology 61:1240-145.

Rust, R. W. 1974. The population dynam ics and host utilization of Geomydoecus oregonus, a parasite of Thomomys bottae. Oecologia (Berlin) 15:287-304.

Sacchi, L., A. Grigolo, M. Mazzini, E. Bigliardi, B. Baccetti, and U. Laudani. 1988. Symbionts in the oocytes of Blattella germanica L. (Dictyoptera: Blattellidae): Their mode of transmission. International Journal of Insect Morphology and Embryology 17:437-446.

Saffo, M. B. 1992. Invertebrates in endosymbiotic associations. Am erican Zoologist 32:57-565.

SAS Institute Inc. 1994. SAS/STAT user's guide, release 6.03 edition. SAS Institute, Cary, North Carolina, 1,028 p.

Saxena, A. K., and G. P. Agarwal. 1985. Mycetocytes in Aegypoecus perspicuus (Phthiraptera). Current Science 54:763-764.

Saxena, A. K., G. P. Agarwal, S. C handra, and O. P. Singh. 1985. Pathogenic involvement of Mallophaga. Zeitshcrift fiir Angewandte Entomologie 99:294-301.

Schilthuizen, M. and R. Stouthamer. 1997. Horizontal transmission of parthenogenesis-inducing microbes in Trichogramma wasps. Proceedings o f the Royal Society of London 264:361-365. B

Silva, M. 1998. Allometric scaling of body length: elastic or geometric similarity in mammalian design. Journal of Mammalogy 79:20-32.

Smith, D. C. and A. E. Douglas. 1987. The biology of symbiosis. Edward Arnold, London. 302 pp.

Smith, M. F. 1998. Phylogenetic relationships and geographic structure in pocket gophers in the genus Thomomys. Molecular Phylogenetics and Evolution 9:1-14.

Smith, M. F., and J. L. Patton 1988. Subspecies of pocket gophers: Causal bases for geographic differentiation in Thomomys bottae. Systematic Zoology 37:163- 178.

114

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Spradling, T. A. 1997. Relative rates of molecular evolution in rodents and their symbionts. Ph.D. dissertation, Louisiana State University, Baton Rouge. 174 pp.

Staley, J. T. 1999. Bacterial Diversity: a time for place. American Society for Microbiology News 65:681-687..

Staley J. T., and J. J. Gosink. 1999. Poles apart: Biodiversity and biogeography of sea ice bacteria. Annual Review of Microbiology 53:189-215.

Suzuki, M. T., and S. J. Giovannoni. 1996. Bias caused by template annealing in the amplification of mixtures of 16S rRNA genes by PCR. Applied and Environmental Microbiology 62:625-630.

Swofford, D. L. 2000. PAUP*. Phylogenetic Analysis Using Parsim ony (‘and O ther Methods). Version 4. Sinauer Associates, Sunderland, Massachusetts.

SYSTAT, Inc. 1992. Statistics, version 5.2 edition. SYSTAT, Inc. Evanston, Illinois.

Tamura, K., and M. Nei. 1993. Estimation of the num ber of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution 10:512-526.

Tanner, M. A., B. M. Goebel, M. A. Dojka, and N. R. Pace. 1998. Applied and Environmental Microbiology 64:3110-3113.

Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanm ougin, and D. G. Higgins. 1997. The ClustalX windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Research 24:4876-4882.

Timm, R. M. 1983. Farenholz's Rule and resource tracking: A study of host- parasite coevolution. Pp. 225-266, in: Coevolution (M. Nitecki, ed.), Univ. Chicago Press.

United States Centers for Disease Control. 1984. Controlling head lice. 2nd ed. United States Department of Health and Human Service, Centers for Disease Control, (CDC) 84-8397:1-15.

Unterman, B. M., P. Baumann, and D. L. McLean. 1989. Pea aphid symbiont relationships established by analysis of 16S rRNAs. Journal of Bacteriology 171:2970-2974.

Waage, J. K. 1979. The evolution of insect/vertebrate associations. Biological Journal of the Linnean Society 12:187-224.

115

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Wang, G. C.-Y., and Y. Wang. 1997. Frequency of formation of chimeric molecules as a consequence of PCR coamplification of 16S rRNA genes from mixed bacterial genomes. Applied and Environmental Microbiology 63:4645-4650.

Ward, D. M. 1998. A natural species concept for prokaryotes. Current Opinions in Microbiology 1:271-277.

Ward, D. M., M. J. Ferris, S. C. Nold, and M. M. Bateson. 1998. Natural view of microbial biodiversity within hot spring cyanobacterial mat communities. Microbiology and Molecular Biology Reviews 62:1-18.

Werren, J. H., D. W indsor, and L. R. Guo. 1995. D istribution of Wolbachia am ong neotropical arthropods. Proceedings of the Royal Society of London. Series B 262:197-204.

Woese, C. R. 1994. There must be a prokaryote somewhere: microbiology's search for itself. Microbiological Reviews 58:1-9.

116

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. APPENDIX: SPECIMENS EXAMINED

Specimens examined: Specimens are deposited in the Museum of Natural Science, Louisiana State University (LSUMZ) and the Museum of Vertebrate Zoology, University of California, Berkeley (MVZ).

Interordinal comparisons

O rder Family Species Specimen number Sex Mass (g)

Insectivora Soricidae Cryptotis parva LSUMZ 23789 U 4.0 T alpidae Scalopus aquations LSUM Z 6981 M 113.4 C hiroptera Pteropodidae Pteropus LSUMZ 17726 M 900b Vespertilionidae Pipistrellus hesperus LSUMZ 22041 M 3.5 Primates Cercopithecidae Colobus guereza LSUMZ 26271' F 8,000b Loris id ae Nycticebus LSUMZ 28897* M 650b Lagomorpha Leporidae Lepus alleni LSUMZ 13465 M 3,500b Ochotonidae Ochotona princeps LSUMZ 35909 F 129.9 Rodentia Heteromyidae Dipodomys ordii LSUMZ 25321 M 72 M uridae Baiomys taylori LSUMZ 4629 F 8.4 Sciuridae Sciurus niger LSUMZ 28476 F 1,000 Artiodactyla Bovidae Taurolragus oryx LSUM Z 36155* U 600,000b Bovidae Madoqua LSUM Z 36156* M 5,000b Hyracoidea Procaviidae Procavia capensis LSUM Z 34649* M 4,000b Perissodactyla T apiridae Tapirus bairdii LSUMZ 6977 M 250,000b Equidae Equus LSUMZ 36157* U 400,000b C arnivora M ustelidae Mustela frenata LSUMZ 28014 M 207 U rsidae Ursus arctos LSUMZ 36158* U 600,000b

(appendix cont.)

117

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Intrafamilial comparisons Rodentia Geomyidae Geomys breviceps LSUMZ 31451 M 180 Geomyidae G. pinetus LSUMZ 23655 M 360 Geomyidae G. tropicalis LSUMZ 34345 M 235 Geomyidae Orthogeomys cavator LSUMZ 29253 M 875 Geomyidae O. cherriei LSUMZ 29539 M 455 Geomyidae O. heterodus LSUMZ 29265 M 620 Geomyidae O. underwoodi LSUMZ 28368 M 260 Geomyidae Thomomys bottae LSUMZ 35992 M 210 Geomyidae T. bulbivorus LSUMZ 31313 M 410 Geomyidae T. mazama LSUMZ 31398 F 87 Geomyidae T. monticola LSUMZ 31411 M 94 Geomyidae T. talpoides LSUMZ 34387 F 91 Geomyidae T. townsendii LSUMZ 31264 M 300 Geomyidae T. umbrinus LSUMZ 34362 M 115

Intrageneric comparisons Rodentia Geomyidae Thomomys bottae LSUMZ 35992 M 210 Geomyidae T. bulbivorus LSUMZ 31313 M 410 Geomyidae T. mazama LSUMZ 31398 F 87 Geomyidae T. monticola LSUMZ 31411 M 94 Geomyidae T. talpoides LSUMZ 34387 F 91 Geomyidae T. townsendii LSUMZ 31264 M 300 Geomyidae T. umbrinus LSUMZ 34362 M 115

Intraspecific comparisons Rodentia Geomyidae Thomomys bottae LSUMZ 35992 M 210 Geomyidae T. bottae (juvenile) LSUMZ 20936 M 37.5 Geomyidae T. b. perpes MVZ 166402 M 236 Geomyidae T. b. perpes MVZ 140497 M 96 Geomyidae T. b. perpallidus MVZ 170521 M 189 Geomyidae T. b. perpallidus M VZ 170565 M 200 Geomyidae T. b. perpallidus M VZ 80595 M 74.2 Geomyidae T. b. perpallidus MVZ 80606 M 89.0 Body mass estim ated from Nowak (1999). ’ Specimens not collected from the wild.

118

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. VITA David Lee Reed was bom in Charlotte, North Carolina, on 5 December, 1968. He attended school in Charlotte until 1987 when he graduated from East Mecklenburg High School. He attend the University of North Carolina, Wilmington where he received his bachelor of science degree in biological sciences in 1991. He then attended Louisiana State University and received his master of science degree in zoology in December, 1994. David then worked for two years as a biologist at The Florida Aquarium in Tampa, Florida. He is currently finishing his doctoral research in zoology at Louisiana State University and anticipates graduation in December, 2000. He has accepted a postdoctoral research position in the Department of Biology, at the University of U tah.

119

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. DOCTORAL EXAMINATION AND DISSERTATION REPORT

Candidates David L. Reed

Major Fields Zoology

Title of Dissertations Host and Parasite Interactions Among Three Symbionts: Pocket Gophers, Chewing Lice, and Endosymbiotic Bacteria

Approveds

Major Professor

Dean of tneMiraduate School

EXAMINING COMMITTEE:

7$ _ - CQ - ______

Date of nation:

September 29, 2000

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