Louisiana State University LSU Digital Commons

LSU Master's Theses Graduate School

2008 Comparisons of bacterial community within the abdomens of Formosan subterranean , fresh- and alcohol-stored, from their native (China) and introduced (U.S.) range Huei-Yang Ho Louisiana State University and Agricultural and Mechanical College

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

Recommended Citation Ho, Huei-Yang, "Comparisons of bacterial community within the abdomens of Formosan subterranean termites, fresh- and alcohol- stored, from their native (China) and introduced (U.S.) range" (2008). LSU Master's Theses. 3982. https://digitalcommons.lsu.edu/gradschool_theses/3982

This Thesis 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 Master's Theses by an authorized graduate school editor of LSU Digital Commons. For more information, please contact [email protected].

COMPARISONS OF BACTERIAL COMMUNITY WITHIN THE ABDOMENS OF FORMOSAN SUBTERRANEAN TERMITES, FRESH- AND ALCOHOL-STORED, FROM THEIR NATIVE (CHINA) AND INTRODUCED (U.S.) RANGE

A Thesis

Submitted to the Graduate Faculty of the Louisiana State University and Agricultural and Mechanical College in partial fulfillment of the requirements for the degree of Master of Science

In

The Department of Biological Sciences

by Huei-Yang Ho B.S., Louisiana State University, 2005 December 2008

ACKNOWLEDGEMENTS

This research was supported through funding from the Board of Regents.

I would like to express my heartfelt gratitude to my major professor, Dr. Claudia

Husseneder for her unwavering patience and understanding and for being readily available to give advice, share her experience and provide guidance and encouragement in the course of my career here. Thank you for the numerous selfless sacrifices. I would also like to extend a great thank you to Dr. Lane Foil for exceeding the expectations required of a committee member, in putting aside a great amount of time to provide guidance through the finer details of improving my thesis. The encouragement and suggestions from Dr. Meredith Blackwell had been very helpful. To her I owe a special thank you for being tremendously accommodating in her schedule and her prompt response to any needs that arose. Her eagerness to lend assistance at every opportunity is greatly appreciated. For setting aside time from his tight schedule to critique on my presentation and providing expert advice on my research, thank you Dr. Gary

King. Once again, I would like to thank all my committee members for their invaluable ideas and suggestions which challenged me to keep seeking for answers. It had indeed been a humbling journey of discovering the ever increasing room for improvement.

Thank you to Ryan Callegan and Dr. Albert Lee for helping to clarify on certain technical aspects of my project. Ryan Callegan had been very helpful at fielding questions pertaining to my research and providing fresh perspectives on graduate studies. I would also like to thank Dr.

Dee Colby, Dr. Gabriel Aluko, Dr. Rachael Collier, Dawn Simms, Jennifer Delatte, Matt Kelly,

Katie and Casey for various forms of assistance provided, directly or indirectly, and for facilitating a conducive environment for completing my project.

ii

Not to be forgotten are the contribution of statistical analyses advice from Michael

McKenna and Drs Charles Monlezun and Bin Li. Thank you for the countless number of hours spent figuring out ways to help interpret the data and for the patience in explaining the statistics used. I would also like to thank Cindy Henk and Ying Xiao for their assistance in the microscopy work.

Last but definitely not least, I would like to express my appreciation to my brother, Huei-

Jin Ho for helping me to secure journal articles that were unavailable through LSU. Words do not do justice to the support my parents, Kah-Keong Ho and Gooi-Chee Thow, have provided, without whom I would not have come this far in my studies.

iii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ...... ii

LIST OF TABLES ...... vi

LIST OF FIGURES ...... vii

ABSTRACT ...... viii

CHAPTER 1. INTRODUCTION ...... 1

CHAPTER 2. LITERATURE REVIEW ...... 5 2.1. Classification and Distribution of FST ...... 5 2.2. Economic Importance of the FST and Control Methods Used for Subterranean Termites .... 6 2.3. Termite Eusociality and Dependency of Lower Termites on Their Gut Endosymbionts ...... 9 2.4. The Gut Environment of Subterranean Termites and Their Endosymbionts ...... 10 2.5. Previous Community Studies of Gut in FST ...... 12 2.6. Practicality of Using Ethanol to Preserve FST Samples from Remote Locations ...... 14

CHAPTER 3. MATERIALS AND METHODS ...... 16 3.1. Termite Collection and Dissection...... 16 3.2. DNA Extraction ...... 16 3.3. 16S Ribosomal RNA Gene Amplification ...... 17 3.4. Cloning ...... 18 3.5. Amplified Ribosomal DNA Restriction Analysis (ARDRA) ...... 18 3.6. DNA Sequencing ...... 20 3.7. 16S Ribosomal RNA Gene Sequence Analysis ...... 20 3.8. Classification of Ribotypes and Initial Phylogenetic Analysis ...... 21 3.9. Chimeric/anomalous Sequence Analysis ...... 21 3.10. Bacteria Ribotype Richness, Diversity and Similarity Analysis ...... 23 3.11. Statistical Analysis for Bacteria Compositions in the Fresh vs Alcohol-stored Louisiana FST Samples ...... 25 3.12. Statistical Analysis for Bacteria Compositions in the Louisiana vs China FST Samples .. 26

CHAPTER 4. RESULTS ...... 27 4.1. Applicability of Using ARDRA for Studying Bacteria Communities in FST Samples ...... 27 4.2. Identification of Artifact Sequences in the Bacteria Community of the FST ...... 29 4.3. Identities of Bacteria in the FST ...... 31 4.4. Comparison Between the Fresh and Alcohol-stored Louisiana FST Samples ...... 42 4.4.1. Effects of Alcohol Storage on DNA Degradation of the Bacteria in the FST Colonies ...... 42 4.4.2. Alcohol Storage Effects on the Biodiversity of Ribotypes ...... 43 4.4.3. Alcohol Storage Effects on the Proportions of Bacteria Classes in the Louisiana FST Colonies ...... 46 4.5. Comparison Between the Alcohol-stored Louisiana and China FST Samples...... 49

iv

4.5.1. Differences in the Biodiversity of Ribotypes Between the Alcohol-stored Louisiana and China FST Samples ...... 49 4.5.2. Differences in the Bacteria Class Proportions Between the Alcohol-stored Louisiana and China FST Samples ...... 50

CHAPTER 5. DISCUSSION ...... 52 5.1. ARDRA Analysis...... 52 5.2. Analysis of Artifact Sequences in the Bacteria Community of the FST Samples ...... 53 5.3. Identities of Bacteria in the FST ...... 55 5.4. Alcohol Storage Effects on the Bacteria Communities in the Louisiana FST Samples ...... 59 5.5. Comparison of Bacteria Community in the Alcohol-stored FST Samples from the Native and Introduced Ranges ...... 62

CHAPTER 6. CONCLUSION...... 66

REFERENCES ...... 68

APPENDIX: LIST OF FULL-LENGTH 16S RIBOSOMAL RNA GENE SEQUENCES (RIBOTYPES), ACCORDING TO DOMINANCE, FOUND IN THE NINE FST CLONE LIBRARIES...... 82

VITA ...... 91

v

LIST OF TABLES

4.1. RFLP-types of ninety-two 16S rRNA gene clones from Hunan FST colony...... 28

4.2. List of confirmed chimerics and anomalous sequences after comparing the results from different settings within Bellerophon, the modified Bellerophon on Greengenes website, Chimera Check and Mallard...... 32

4.3. List of bacteria ribotypes with sequence similarities of 97% or higher to known bacteria sequences from the DDBJ/EMBL/GenBank databases, according to their abundance...... 35

4.4. DNA concentration of DNA extracts from 50 FST workers and 16S PCR products for the 9 FST colonies...... 43

4.5. List of the number of clones sequenced and the actual number of ribotypes, singletons, doubletons and unique ribotypes found in the nine FST colonies and one from Japan (Shinzato et al., 2005)...... 44

4.6. The estimated percentage of the total number of ribotypes captured and ribotype diversity based on different indices from the nine FST colonies and one from Japan (Shinzato et al., 2005)...... 45

4.7. The three similarity indices (Chao-Jaccard Abundance, Bray-Curtis and Morisita-Horn) used to determine the number of shared ribotypes among the nine FST colonies and one from Japan (Shinzato et al., 2005)...... 47

vi

LIST OF FIGURES

4.1. RFLP-types of some 16S rRNA gene clones used in this study...... 28

4.2. Phylogenetic tree showing the placement of the anomalous or chimeric sequences among all the other ribotypes found in the nine FST colonies...... 34

4.3. Phylogenetic trees showing the 16S rRNA gene sequences from the FSTs is this study (Cf) and their highest match from the DDBJ/EMBL/Genbank databases… ...... 38

4.4. a) Extracted DNA from the nine FST colonies. b) PCR products of 16S rRNA gene sequences from the same nine FST colonies...... 42

4.5. Species accumulation curves generated based on the Mao Tau means of all nine FST colonies and one from Japan (Shinzato et al., 2005)...... 44

4.6. Proportions of bacteria classes from the nine FST colonies and one from Japan (Shinzato et al., 2005)...... 48

4.7. Proportions of (high G+C gram-positive), low G+C gram-positive and gram- negative bacteria in the nine FST colonies and one from Japan (Shinzato et al., 2005)...... 49

vii

ABSTRACT

The Formosan subterranean termite (FST), a pest species native to China and introduced to the U.S., is obligatorily dependent on its gut microbiota. Using high-throughput 16S rRNA gene sequencing, the effects of long-term alcohol storage and geographic location on the bacteria composition of the FST colonies were investigated. Initial studies found the use of amplified ribosomal DNA restriction analysis (ARDRA) to be unpractical due to its higher costs compared to the direct sequencing of 16S rRNA gene sequences. Using nine FST colonies consisting of fresh and alcohol-stored Lousiana FST colonies and alcohol-stored China FST colonies, 237 bacteria ribotypes were identified from 1876 clones based on a <97% sequence similarity criterion. Twenty-four of the ribotypes were artifact sequences and were excluded from subsequent analyses. Most of the remaining ribotypes were novel (70.89% of total ribotypes).

Termite-specific bacteria dominated the bacteria composition in the FSTs (66.45% of total clones). Only 3.34% of the total clones were similar to environmental bacteria. Thirteen bacteria phyla were represented: (42.91% of total clones), (30.49%),

Spirochaetes (11.30%), Actinobacteria (5.70%), (2.24%), Tenericutes (1.55%), candidate division Termite Group 1 (1.01%), candidate division TM7 (0.64%),

(0.59%), (0.48%), candidate division Synergistes (0.21%), candidate division

ZB3 (0.05%) and (0.05%). The Bacteroidetes ribotype previously identified to be dominant in FST from Japan, was also among the dominant phyla in all the FST colonies of this study (38.71% of total clones). Differential DNA degradation occurred in the alcohol-stored

FST samples, leading to higher proportions of the gram-positive bacteria such as Actinobacteria,

Bacilli and Clostridia and lower proportions of the gram-negative bacteria such as Bacteroidetes and compared to the fresh FST samples. Long-term alcohol storage of the FST led

viii

to the discovery of less abundant ribotypes when the predominant ribotype was reduced.

Geographic region did not show detectable influence on the FST bacteria composition. This was likely due to the multiple introduction of FST from China into the U.S. Future studies using T-

RFLP to sample the bacteria community from FST colonies randomized across each geographical region would be useful in confirming the observations from this study.

ix

CHAPTER 1. INTRODUCTION

The Formosan subterranean termite (FST) originated most likely from China (Kistner,

1985) and has spread to various regions throughout the world along shipping trade routes. Its introduction into the continental United States from China was probably mediated through Japan and Hawaii (Su and Tamashiro, 1987). The destructive activities of the FST have huge economic implications worldwide. For example, annual losses in the U.S. alone cost more than

USD 1.5 billion due to control cost of the FST (Su and Scheffrahn, 1998).

The subterranean termite gut is known to contain a wide assemblage of microbes consisting of protists, fungi, archaea and bacteria (König et al., 2006). The workers of FST, which are responsible for feeding the colony, do not produce sufficient enzymes to digest the woody materials on which they feed solely upon (Nakashima et al., 2002). Therefore, termite survival is dependent upon the activity of microorganisms within the gut to aid in the digestion of the food. The protists within the FST gut digest the wood diet of the termites and are necessary for the survival of the FST (Ohkuma et al., 2000). The other microbes supplement the nutrient-poor diet of termites with nitrogen through nitrogen fixation (Ohkuma et al., 1999) and recycling (Potrikus and Breznak, 1981). Microbial symbionts of termites also provide energy through acetogenic reduction of CO2 (Breznak and Brune, 1994), provide resistance against pathogens (Dillon and Dillon, 2004) and may be involved in influencing nestmate recognition

(Matsuura, 2001).

The protists in the FST have already been well characterized but the bacteria community remains largely unknown. Therefore, defining the bacteria community is important in understanding the termite-endosymbiont relationship as well as for developing new strategies for the control of the FST. Current research efforts are shifting the emphasis towards the

1

development of more environmental-friendly termite control practices such as biological control methods. Termite control could be achieved through the disruption of the gut microbial composition, such as elimination of key bacteria species which are found in the FSTs regardless of the geographic origin of the termite colony. Identification of gut bacteria common to termites across different regions can also serve as a basis for paratransgenesis, where a suitable bacteria candidate is engineered to carry foreign genes into the termite host system (Husseneder and

Grace, 2005). The bacteria vector could be used to express genes detrimental to termites’ survival.

Thus far, only FST colonies from the introduced range (Japan and Louisiana) were used for comprehensive bacteria community studies and only a few colonies were sampled

(Husseneder et al., 2005; Shinzato et al., 2005). The purpose of this research was to provide a comprehensive study on the variation in the bacteria composition among FST colonies within and between the introduced (U.S.) and native ranges (China).

The first objective of this study was to evaluate the applicability of amplified rDNA restriction analysis (ARDRA) in studying the bacteria community in the FST. ARDRA has been widely used as a cheap and quick method to determine bacteria phylogeny and to identify bacteria species with known sequences (Heyndrickx et al., 1996). In bacteria community studies, ARDRA is used to screen out clones with similar 16S rRNA gene sequences before performing DNA sequencing. The applicability of ARDRA in bacteria community studies is dependent on its ability to: 1) distinguish between most bacteria ribotypes in the clone library and 2) provide cost advantage over the use of high-throughput sequencing without prior screening of the clone library to eliminate similar sequences. To determine if ARDRA fulfilled

2

both criteria in this study, a clone library constructed using the 16S rRNA gene sequences isolated from one of the FST samples was sequenced and compared with the ARDRA results.

The second objective of this study was to screen for artifact sequences (chimeric and anomalous sequences) in the 16S rRNA gene clone libraries from the different FST samples to ensure that none of the ribotype sequences were artifacts. Anomalous sequences are artifact sequences that are non-chimeric and are unexpectedly divergent from a 16S rRNA gene sequence. Chimeric sequences are artificial PCR amplicons that result from the combination of

DNA fragments from different DNA templates (Wang and Wang, 1996; Wang and Wang, 1997).

Chimeric sequences are formed when a partially-synthesized DNA strand anneals to a different

DNA template during the PCR elongation cycle, consequently DNA extension resumes using the new DNA template. The accumulation of chimeric sequences in public databases (Ashelford et al., 2005; Ashelford et al., 2006; Hugenholtz and Huber, 2003), which resulted from the increasing use of culture-independent techniques in bacteria community studies, has led to the development of various methods for detecting chimeric sequences. In this study, nearest- neighbor methods using programs such as Bellerophon, Chimera Check and Mallard, in conjunction with signature analysis and BLAST analysis were used to screen out artifact sequences.

Since the effects of long-term alcohol storage on the bacteria community within termites has not been investigated, our third objective was to compare the bacteria composition in FST colonies stored long-term in alcohol with bacteria composition of fresh FST colonies. Bacteria ribotype richness and diversity, bacteria class proportions, as well as the bacteria ribotypes shared among the colonies were analyzed statistically to determine if there were any differences between the alcohol-stored and the fresh FST colonies.

3

For the fourth objective, the bacteria composition in the alcohol-stored samples from the native range of the FST, China, and the introduced range of the FST, Louisiana was compared.

Our hypothesis was that the large-scale physical separation of the FST colonies would allow the bacteria communities to develop separately, resulting in differences in the bacteria compositions between the FST colonies of the different geographical ranges.

4

CHAPTER 2. LITERATURE REVIEW

2.1. Termite Classification and Distribution of FST

Termites (Isoptera) are a group of eusocial insects. The name Isoptera means “equal- winged” and refers to the two pairs of wings of the alates (reproductive termites) which are equal in size and similar in appearance (Kumari et al., 2006). Recently, Isoptera was proposed to be grouped under the Blattodea (cockroaches) (Inward et al., 2007), which are closely related to the

Mantodea (mantids) and together, these groups form the superorder Dictyoptera, the most basal group of winged insects (Deitz et al., 2003). The subdivision of termites into seven families:

Mastotermitidae, Hodotermitidae, Termopsidae, Kalotermitidae, Serritermitidae,

Rhinotermitidae and Termitidae (Eggleton, 2001) has been widely used but recent molecular studies show that a revision in the classification system may be needed (Inward et al., 2007).

Of more than 2700 termite species worldwide (König et al., 2006), at least 200 species are known to damage buildings (Su and Scheffrahn, 1998). Subterranean termites

(Rhinotermitidae) represent 80% of the economically significant species and of those, the largest group is Coptotermes (Su and Scheffrahn, 1998). The Formosan subterranean termite (FST),

Coptotermes formosanus Shiraki, is one of two economically important subterranean termites to have been introduced in more than five regions worldwide (Su and Scheffrahn, 1998). The FST is found between the latitudes 35oN and S, spanning regions in China, Taiwan, Japan, Guam,

Midway, Hawaii, and the continental United States (Su and Tamashiro, 1987).

From their likely site of origin in China (Kistner, 1985), the FST was believed to have been transported to Japan about 300 years ago, to Hawaii almost 100 years ago, and finally to the continental U.S.A approximately 50 years ago (Su and Tamashiro, 1987). In the continental

U.S.A., the FST was first collected in Charleston, South Carolina, in 1957 (Chambers et al.,

5

1988). The earliest discovery of the FST in Louisiana was in 1966 (La Fage, 1987).

Infestations, however, were already well established in Lake Charles and New Orleans (La Fage,

1987), indicating much earlier introduction.

2.2. Economic Importance of the FST and Control Methods Used for Subterranean Termites

The FST causes structural damage to buildings and also affects agriculture, horticulture and forestry (Su and Tamashiro, 1987). Most recent estimates of annual losses due to FST were

USD 1.5 billion in the U.S.A. (Su and Scheffrahn, 1998), USD 0.8 billion in Japan (Tsunoda,

2003) and 0.8 billion RMB (~USD 150 million) in China (Zhong and Liu, 2003). Subterranean termites pose a unique problem for control efforts because their cryptic foraging within underground tunnels make detection of their presence and treatment difficult (Su and Scheffrahn,

1998).

The earliest termite control methods used insecticides that served only as foraging barriers between the sources of infestations and the structures to be protected (Randall and

Doody, 1934). Environmental and public health concerns were raised over the formerly used cyclodiene termiticides, which were cheaper and more persistent compared to the termiticides used today (Perrott et al., 2004). Metabolic inhibitors such as borates (Grace, 1990), dechlorane

(Paton and Miller, 1980), hydramethylnon (Su et al., 1982), A-9248 (Su and Scheffrahn, 1988) and sulphuramid (Su and Scheffrahn, 1988; Su and Scheffrahn, 1991; Su et al., 1982) were incorporated into baits and caused reduction in termite populations but could not completely eliminate the termite infestations. In the 1970s, biological agents against termites using nematode (Fujii, 1975) and fungi (Lai et al., 1982) were developed but were found to be ineffective. Two classes of insect growth regulators, juvenoids (juvenile hormone analogs or

JHAs) and chitin synthesis inhibitors (CSIs), have been evaluated for FST control. Field trials 6

using JHAs such as methoprene and hydroprene were not successful (Hrdy et al., 1979) while fenoxycarb treatments only led to a decrease in termite foraging activity (Jones, 1989). The CSI hexaflumuron has been shown to cause significant ecdysis inhibition in a wide range of subterranean termite species, including Coptotermes (Su and Scheffrahn, 1993). Other control methods include wood treatments and physical barriers employing gravels or stainless steel mesh

(Su and Scheffrahn, 1998). Many of the above termite control methods are still being used today.

Two of the main methods currently used in termite control are liquid termiticides and baits. Non-repellant liquid termiticides such as imidacloprid, fipronil and chlorfenapyr are applied around the perimeter of a structure to form a barrier (Potter, 1997). A disadvantage of this method is its inability to affect the termite colonies located some distance from the treated area. Secondly, the termites exposed to the toxicant may die before returning to the nest to spread it to other nestmates through contact or grooming. Thirdly, if the termiticide is not applied correctly, termites can to bypass the barrier and infest the structures. The alternative control method uses a monitoring-baiting program in which baits containing hydramethylnon, hexaflumuron, noviflumuron or sulfluramid are set up at stations in which termites are detected

(Su and Scheffrahn, 1998). The advantage of this method is that it actively targets the termite colony. Furthermore, the chemicals used in baits are slow-acting so that the termites can live long enough to transmit the toxicant to their nestmates (Su and Scheffrahn, 1998). However, this method also has its limitations. The efficacy of this system is dependent on the number of termites feeding on the bait and transmitting the toxicant to other individuals within the colony.

The termites may have limited contact with the bait due to high availability of natural food

7

resources. Furthermore, the toxicants are diluted as they are transmitted from individual to individual.

Aside from using termite control methods that have direct toxicity to the termites, targeting the microbial community within their guts could also lead to the demise of the termite hosts. A possible termite control method is through the use of pathogens to infect the termites, leading to their decreased fitness level or death (Connick et al., 2001; Osbrink et al., 2001).

Another termite control method is the disruption of the bacteria composition in the termite gut by eliminating key bacteria species or groups found in the FSTs, leading to the decrease in fitness of the FST or their death. The eradication of protozoa in the FST and another subterranean termite species, Reticulitermes flavipes, has been shown to lead to the starvation of the termites after which death ensues (Husseneder and Collier, In press; Mauldin and Rich, 1980).

A novel termite control method is the use of a paratransgenesis system (Husseneder and

Grace, 2005). Paratransgenesis is the use of an organism such as a bacteria vector to shuttle foreign genes into the body of a host organism (Husseneder and Collier, In press). It can serve as a biological control method that can self-perpetuate and propagate throughout the termite population. A termite gut bacterium, Enterobacter cloacae, had been successfully transformed with a recombinant plasmid (pEGFP) which was used to deliver and spread foreign genes in FST

(Husseneder and Grace, 2005). This provided the proof of concept for paratransgenesis. The bacteria was introduced into a few FST workers through feeding and propagated to other termites through the bacteria-fed termites. The bacteria population was then shown to persist in the termite guts. The efficiency of the bacteria propagation even when using a low amount of inoculation demonstrates the viability of this method in the use of termite control, as soon as the bacterium has been genetically engineered to express genes detrimental to termites. However,

8

one of the first prerequisites for using this method is that a bacteria common to all FST colonies, yet is specific only to the FST, has to be identified.

2.3. Termite Eusociality and Dependency of Lower Termites on Their Gut Endosymbionts

Termites are the only major non-hymenopteran insects that exhibit eusociality (Eggleton,

2001). Eusociality is defined by members of a society that express reproductive division of labor, cooperative brood care and overlap of adult generations (Thorne, 1997). As polymorphic insects, termites have different functional castes within each colony. Workers make up the bulk of the colony and are responsible for food-foraging, feeding other colony members, nest- building, brood-caring and grooming (Kumari et al., 2006). The other functional castes (ie. soldiers, reproductives) and early developmental stage termites are incapable of feeding directly on woody materials and thus, are solely dependent on the workers to feed them.

The diet of subterranean termites consists of woody materials. Composed of lignocellulosic compounds derived from plants cell walls, these materials are refractory to digestion because of their highly complex matrices (Li et al., 2006). Termites of the family

Termitidae have lost dependency on endosymbiotic protozoa found in the guts of the other termite families (Kumari et al., 2006) through the development of alternative cellulose degradation methods such as the production of higher levels of endogenous cellulase and the use of symbiotic fungi. Subterranean termites (Rhinotermitidae) such as the FST produce cellulase enzymes endogenously in their salivary glands, foregut and midgut (Nakashima et al., 2002); however, endogenous cellulases are insufficient to support the termite’s metabolic needs. The termite’s survival relies on the existence of the microbes, in large part the protozoa, within its gut which produce most of the cellulases required for the wood digestion. To meet the nutritional

9

demand of the termites, the protozoa must act in synergy with the endosymbiotic archaea, bacteria and yeasts within the termite gut (Breznak and Brune, 1994).

The gut microbes of termites may be acquired through several routes. In the incipience of a termite colony, the newly-hatched larvae acquire their inoculation of microbial endosymbionts through feeding by the reproductive pair (Kumari et al., 2006). The pool of essential endosymbionts is propagated and maintained among nestmates via trophallaxis as the colony expands. Individual termites undergo several molts throughout their lifespan, and all the gut contents including the endosymbionts, are lost when the internal gut lining is shed

(Brugerolle and Radek, 2006). To replenish its supply of microbial load, the termite needs to be fed through proctodeal trophallaxis, a form of coprophagy where fluid is transferred from the anus of the donor to the recipient (Nalepa et al., 2001). Other possible sources of gut microbes include the termite’s diet and their grooming behavior. They can obtain their gut microbes through regular body contact with their subterranean nest, which is constructed using their own feces among other materials (Kumari et al., 2006), and also the surrounding environment where they forage for food.

2.4. The Gut Environment of Subterranean Termites and Their Endosymbionts

Within the digestive system of subterranean termites, microbes are concentrated in the hindgut (Bignell, 2000) but are also found in the midgut. The pH in the different regions of the digestive system in subterranean termites ranges from 5.5 to 7.5 (Bignell and Anderson, 1980).

Despite the small hindgut volume in FST of about 0.08 – 0.16µl (Breznak and Pankratz, 1977), the elaborate hindgut structures create several microenvironments. The hindgut of FST is subdivided into five segments: the proctodeal segment (P1), enteric valve (P2), paunch (P3), colon (P4) and rectum (P5) (König et al., 2006). The paunch houses three flagellate species

10

found exclusively in FST (Yamin, 1979): Pseudotrichonympha grassii (Hypermastigida),

Spirotrichonympha leidyi (Trichomonadida) and Holomastigotoides hartmanni (Oxymonadida)

(Ohkuma et al., 2000). In addition, the paunch serves as a microbial fermentation chamber.

Compared to a cow’s rumen, the termite paunch is about 108 times smaller, resulting in a 500 times larger oxygen influx per unit volume (Brune, 1998). Consequently, there is a microoxic periphery in the paunch at the epithelial surface with a steep decrease in oxygen gradient to complete anoxia towards the center of the paunch (Brune et al., 1995). To maintain an anaerobic environment critical for the survival of the obligate anaerobes and flagellates, the constant supply of oxygen through the epithelium has to be consumed by facultative aerobes.

The bacteria community that inhabits the gut of lower termites is composed of great bacteria species diversity. The density of in the termite gut is estimated to be as much as 106-107 cells per µl (Schultz and Breznak, 1978). Besides attaching to the gut epithelium, the bacteria endosymbionts find different niches by attachment to the protozoa surface, residing within the protozoa cytoplasm and being freely suspended in the contents of the termite gut (Breznak and Pankratz, 1977). Even though the protozoa are mainly responsible for digesting the lignocellulosic compounds in wood-feeding termites, cellulolytic bacteria may contribute significantly to the breakdown of wood. In addition, methanogens and acetogens use the end products released by the protozoa, H2 and CO2, to produce methane and acetate, respectively; thereby reducing the waste products from the gut environment that would be detrimental to the termites if accumulated in high volumes (Breznak, 2000). To adapt to the scarcity of N in their diet, the termites carry gut bacteria that are capable of conserving N supply.

Nitrogen-fixing bacteria convert N2 to NH3, a nitrogen source that can be assimilated by the termites (Slaytor, 1969). Uricolytic bacteria in the termite gut are able to extract the N atoms

11

from the uric acid excreted by the termites (Potrikus and Breznak, 1981). The normal flora in the termite guts has also been suggested to protect the termite hosts against invasion by environmental pathogens (Dillon and Dillon, 2004). Environmental bacteria picked up by the termites have to compete against the resident gut bacteria which are adapted to the exotic niche within the termite gut environment. Other bacteria roles include vitamin supplementation

(Martens et al., 2002) and balancing the gut pH levels (Bignell and Anderson, 1980). Bacteria composition can also influence nestmate recognition (Matsuura, 2001).

The social structure of termites was proposed to have arisen from the termite colony’s dependence on the gut endosymbionts of the workers (Thorne, 1997). In the symbiont transfer hypothesis, eusociality of termites was thought to have resulted from the termite’s need to acquire endosymbionts through trophallaxis from their nestmates upon hatching and after every molt. Individual termites would be incapable of surviving outside of a social structure in which the termites obtained their endosymbionts through living together in groups with overlapping generations. Although the symbiont transfer hypothesis applies mainly to the relationship between the lower termites and their gut protists, bacteria which make up a large portion of the termite gut flora, are crucial to the termite gut microecosystem.

2.5. Previous Community Studies of Gut Bacteria in FST

Breznak and Pankratz (1977) performed one of the earliest studies on the bacteria in the guts of FST from Louisiana. Using light and electron microscopy, they identified 7 distinctively different bacteria morphotypes and the location of different bacteria in the termite gut. In subsequent culture-dependent studies, aromatics-degrading bacteria (Harazono et al., 2003) and sulfate-reducing bacteria (Kuhnigk et al., 1996) were isolated from the guts of FST.

12

A more comprehensive study on the bacteria community in the gut of the FST was done by Shinzato et al. (2005). They constructed a 16S rRNA gene library using a FST colony from

Japan. The rRNA gene, a housekeeping gene in prokaryotes, has well-conserved regions in the

DNA sequence which make good targets for binding of universal primers. Concomitantly, certain regions in the 16S rRNA gene accumulate variation as the bacteria diverge genetically.

Both the conserved and variable regions in the rRNA gene allow the gene to be used for bacteria species identification. Shinzato et al. (2005) screened 261 clones using amplified rDNA restriction analysis (ARDRA), also known as restriction fragment length polymorphism (RFLP), and found 49 unique RFLP-types. The RFLP-types were defined by 16S rRNA genes that produced different RFLP banding patterns. The most dominant RFLP-type was a Bacteroides bacterium constituting 70% of the total clones. Most of the RFLP-types were from the phyla

Clostridiales and Spirochaetes. The RFLP-types were categorized within 10 bacteria phyla and many of them were clustered with gut bacteria identified in other termite species. Though more comprehensive than previous culture-dependent studies, Shinzato et. al’s (2005) culture- independent study of the bacteria community in the gut of FST was performed for only one termite colony.

To investigate the variations between termite colonies, Husseneder et al. (2005) used two

FST colonies from Louisiana for 16S rRNA gene sequencing. As in Shinzato et al.’s (2005) study, the same Bacteroides bacterium was dominant in both the Louisiana termite colonies

(>75%). Of the 105 individual 16S rRNA genes sequenced, 12 ribotypes, defined as a 16S rRNA gene sequence with <97% similarity to other known 16S rRNA gene sequence, belonging to 4 bacteria phyla were identified. In addition, bacteria were cultured from one of these colonies and another three colonies from Hawaii. Most of the bacteria isolated through culture

13

belonged to the orders Lactobacillales, and Enterobacteriaceae. Only one of the cultured bacteria matched a ribotype previously found in FST, reflecting a huge discrepancy between bacteria identified using culture-dependent and -independent techniques. The same bacteria has since been characterized and designated as Pilibacter termitis (Higashiguchi et al.,

2006). Since the previous studies only studied the bacteria community from a few FST colonies and they were only from the introduced range (Japan and U.S.), our study would use at least three colonies from each geographic region and also include FST colonies from the native range

(China).

2.6. Practicality of Using Ethanol to Preserve FST Samples from Remote Locations

Using 95% ethanol is considered to be the most practical choice of preservation for DNA from field work samples that is left unextracted from the cells (Flournoy et al., 1996; Harry et al.,

2000; Kilpatrick, 2002; King and Porter, 2004; Mandrioli et al., 2006; Post et al., 1993; Reiss et al., 1995). A quick ethanol penetration of the tissue rapidly removes water and oxygen and preserves the tissues by preventing autolytic processes from damaging the DNA (King and

Porter, 2004). Studies show that the DNA preservation of different insects varied widely when stored in 95% ethanol. In one study, amplification of low copy number genes in scorpions and spiders yielded no products after storage in 95% ethanol for 6 weeks (Vink et al., 2005). Reiss et al. (1994) found that beetles stored in 95% ethanol at room temperature preserved the DNA well for about 6 weeks and started showing a decrease in DNA yield after 73 days. However, in a study where blackfly was stored at 4oC in 95% ethanol, Post et al. (1993) could recover DNA even after 371 days whereby 29% of the DNA had minimal DNA shearing.

While 95% ethanol has been shown to be capable of preserving the DNA of insects, the

DNA preservation of the insect endosymbionts is not well-known. Thus far, a study on the

14

bacteria community in the guts of FST stored in 95% ethanol at room temperature for one week did not lead to changes in the bacteria biodiversity (Husseneder et al., 2005). A study in another termite species, Microcerotermes sp., when stored in 95% ethanol at room temperature for 30 days also showed no significant change in the bacteria biodiversity, even though the yield of

DNA extracts from the termite gut decreased by 50%. In our study, we wanted to determine if differential DNA degradation occurred in the bacteria community of the FST samples.

15

CHAPTER 3. MATERIALS AND METHODS

3.1. Termite Collection and Dissection

Three FST colonies from Louisiana, United States were freshly collected from Lake

Charles (sample 1) and Chalmette Battlefield (sample 2) and City Park (sample 3), both from

New Orleans and in the summers of 2005, 2004 and 2003, respectively. The guts of fifty FST workers were dissected on the same day from each FST colony under sterile conditions.

Additional workers from the FST colony from City Park were subsequently preserved in 95% ethanol (Sigma-Aldrich Corp., St. Louis, Mo.) for 5 years at room temperature until DNA extraction was performed in 2008.

Three FST colony samples from China, collected from Hunan province (sample 7),

Sanshui (sample 8) and Zhongshan (both from Guangdong province) (sample 9), and another two FST colony samples from New Orleans, Louisiana, U.S., collected from Armstrong Park and

French Quarter, were immediately preserved in 95% ethanol at room temperature during fall of

2001 and shipped to Louisiana State University. The abdomens of fifty FST workers were dissected from the Hunan and Zhongshan FST colonies during the fall of 2005 and Sanshui FST colony during the fall of 2006. During the spring of 2008, fifty worker abdomens were dissected from the Armstrong Park (sample 5) and French Quarter (sample 6) FST colonies preserved in

95% ethanol at room temperature. Another fifty worker abdomens were dissected from the FST colony from City Park which was earlier preserved in 2003 (sample 4).

3.2. DNA Extraction

Upon dissection, all the samples were suspended in 200μl of TE buffer (10mM Tris-HCl,

1mM EDTA, pH 8.0). Using a sterilized pestle, the guts or abdomens from each of the nine dissections were crushed and homogenized into the buffer. The mixture was centrifuged at

16

5000xg for 10 min to harvest the microbial cells and the supernatant was decanted before resuspending the pellet in lysis buffer (20mM Tris-HCl (pH 8.0), 2mM sodium EDTA, 1.2%

Triton X-100, 20mg/ml lysozyme). The mixture was incubated overnight at 37oC to ensure complete lysing of the bacteria cells. The remaining DNA extraction procedure was performed using the Qiagen DNeasy Tissue Kit (Qiagen, Valencia, CA) according to the manufacturer’s protocol. DNA concentration was quantified using the Thermo Scientific NanoDrop 1000 spectrophotometer (NanoDrop Technologies, Wilmington, DE) and the quality of the DNA was determined by using a 1% agarose gel electrophoresis (Figure 4.4).

3.3. 16S Ribosomal RNA Gene Amplification

The 16S rRNA genes were amplified using bacteria-specific primer pair 27F (5’-

AGAGTTTGATCCTGGCTCAG-3’) (Suzuki and Giovannoni, 1996) and 1492R (5’-

GGTTACCTTGTTACGACTT-3’) (Suzuki and Giovannoni, 1996). A PCR Optimizer kit

(Invitrogen, San Diego, CA), which included PCR buffers of varying pH and Mg2+ concentrations, was used to optimize the initial 16S rRNA gene PCR reactions. The optimized

PCR conditions, which were used in subsequent reactions, consisted of 1X PCR buffer (60mM

Tris-HCl, 15mM (NH4)2SO4, pH 8.5 and 2.5mM MgCl2), 0.25mM dNTP, 1U AmpliTaq DNA polymerase (Roche, Indianapolis, IN), 0.25μg of each primer with 1-10ng of DNA template from the fresh FST samples and 30-160ng from the alcohol FST samples. Higher amounts of DNA template were used in the alcohol-stored FST samples to account for the additional genomic

DNA introduced by the FST fat body tissue from the abdomen. The dNTPs were added only after heating the PCR reaction tubes for 2-3 min at 80oC. The 80oC incubation step was included to remove secondary structures in the DNA templates as addition of excess dNTPs may hinder the denaturation of the secondary structures (Innis et al., 1990). Reaction mixtures were

17

incubated in a PTC-200 thermal cycler (MJ Research, Reno, NV) using the following program:

94oC for 2 min; 25 cycles of denaturation (94oC for 1 min), annealing (55oC for 2 min) and extension (72oC for 3 min); and a final 7 min extension at 72oC. The use of a longer elongation cycle during PCR has been attributed to a decrease in formation of chimeric sequences

(Meyerhans et al., 1990; Wang and Wang, 1996). Presence of PCR products was confirmed using 1% agarose gel electrophoresis and the Thermo Scientific NanoDrop 1000 spectrophotometer (NanoDrop Technologies, Wilmington, DE). PCR products were purified using the UltraClean PCR Clean-Up kit (MoBio, Solana Beach, CA).

3.4. Cloning

The 16S PCR products were cloned using the TOPO TA Cloning kit (Invitrogen, San Diego,

CA). A dilution series of the transformation assay was plated on Luria-Bertani (LB) agar plates containing 50μg/ml kanamycin with 40μl of 40mg/ml X-Gal spread on the agar surface. The dilution series ensured that a plate with countable colonies was obtained while the X-Gal was used for blue-white screening to select for clones with 16S PCR product inserts.

3.5. Amplified Ribosomal DNA Restriction Analysis (ARDRA)

Up to 101 bacterial 16S rRNA gene sequences isolated from previous studies of

Coptotermes formosanus (Harazono et al., 2003; Husseneder et al., 2005; Noda et al., 2005;

Shinzato et al., 2005) were downloaded from Genbank and used to perform in silico digests with

Vector NTI Advance V10.1.1 software (Invitrogen). The simulated restriction digests were carried out using different combinations of the restriction enzymes AluI, TaqI, HhaI, HaeIII and

RsaI. The fragments were analyzed using BioNumerics V3.0 software (Applied Maths, St

Martens Latem, Belgium) to select for the most suitable restriction enzymes. Band sizes within the range of 70-1400bp were analyzed. Based on the in-silico restriction digest analysis, double

18

digests using the RsaI and HaeIII enzyme combination were determined to be effective in distinguishing most of the 16S rRNA gene sequences and so, were used in the actual restriction digests.

Ninty-two clones from the Hunan termite colony were randomly selected from the 16S rRNA gene clone library using sterile toothpicks and spread over patch areas in grid-marked LB agar plates. The grid-marked LB agar plates were incubated overnight at 37oC. This served to increase the cell growth for each clone to obtain sufficient amounts for plasmid extraction and

PCR amplification. Colonies from each patch area were transferred into 0.6ml microfuge tubes

o containing 100μl H2O. Plasmids from the clones were extracted by boiling the cells at 98 C for

10 minutes and centrifuging the tubes at 13, 200 rpm for 1 min.

The 16S rRNA gene inserts in the clone plasmids were isolated by PCR amplification of the sequences flanking the inserts using primer set M13F (5’-GTAAAACGACGGCCAG-3’) and

M13R (5’-CAGGAAACAGCTATGAC-3’). The remaining procedures were the same as those described in the 16S PCR protocol except that a different PCR buffer was used for the PCR reactions (60mM Tris-HCl, 15mM ammonium sulfate, 2.0mM MgCl2, pH 9.0). Double digests using the two restriction enzymes, RsaI and HaeIII, were performed for 2 hours at 37oC in 20μl reaction volumes containing 15μl of the M13 PCR products solution, 2μl of the commercially supplied buffer, 3μl of Millipore water and 0.25μl of each restriction enzyme. The restriction fragments were separated on a 3% Genepure Sieve GQA agarose (ISC BioExpress, Kaysville,

UT) in 1X Tris-Borate-EDTA (TBE) buffer for 3 hours at 200V and chilled with ice. Gels were stained with ethidium bromide, visualized using a UV transilluminator and digitalized with

Kodak 1D LE V3.5 software (Kodak, New Haven, CT). Gel images were subsequently analyzed using BioNumerics V3.0 software to compare the restriction fragment length polymorphism

19

(RFLP) patterns. The number of different RFLP patterns (richness) and the frequency of each pattern (evenness) were ascertained. All ninety-two clones were subsequently sequenced to compare the bacteria ribotypes (97% sequence similarity in the 16S rRNA gene sequences) to the

RFLP-types.

3.6. DNA Sequencing

Clones from Lake Charles, Chalmette Battlefield and City Park (fresh FST from

Louisiana); Armstrong Park, French Quarter, City Park (alcohol-preserved FST from Louisiana); and Hunan, Sanshui and Zhongshan (alcohol-preserved from China) were inoculated into 96- wells microtiter plates containing 200μl LB kanamycin broth in each well and incubated overnight at 37oC while shaking at 250rpm. Of the culture, 100μl was mixed with 100μl of LB kanamycin broth with 30% glycerol before sealing the plates and sending them out to the

Interdisciplinary Center for Biotechnology Research (ICBR) at the University of Florida for high-throughput forward- and reverse-directional single-pass DNA sequencing. Sequence lengths of 1200-1400bps were obtained for the 16S rRNA genes.

3.7. 16S Ribosomal RNA Gene Sequence Analysis

The forward and reverse 16S rRNA gene fragments were contiged using the

ContigExpress module within the Vector NTI Advance V10.1.1 software (Invitrogen, San Diego,

CA) before the full-length sequences were compared to the GenBank/EMBL/DDBJ databases

(www.ncbi.nlm.nih.gov) using Basic Local Alignment Search Tool (BLAST) to find matches to known bacteria. An inventory of the sequence identities was compiled. Using the AlignX module within the Vector NTI Advance V10.1.1 software (Invitrogen, San Diego, CA), sequences that matched similar identities were compared using the ClustalW algorithm to make pairwise and multiple alignments using the default settings. The sequences were classified into

20

different ribotypes using the <97% sequence similarity criterion (Stackebrandt and Goebel,

1994).

3.8. Classification of Ribotypes and Initial Phylogenetic Analysis

Using the default settings on the naïve Bayesian rRNA classifier (Wang et al., 2007) on the RDP website (Cole et al., 2007), the sequences were classified into the known major bacteria lineages based on a confidence level of 80%. The ribotypes were named according to the prefix

Cf (Coptotermes formosanus) followed by a first numeral indicating the phyla classification of the ribotype on the RDP website and a second numeral assigned to each ribotype (Cf1-xx-10-xx:

Actinobacteria, Bacteroidetes, Cyanobacteria, Spirochaetes, Planctomycetes, Proteobacteria,

Firmicutes, candidate division TM7, Verrucomicrobia and unclassified ribotypes, respectively).

Classification of the sequences was further confirmed using BLAST and reclassified based on the latest publications. Sequence divergence among the ribotypes was calculated using the

Kimura 2-parameter model and the phylogenetic tree was constructed using the neighbor-joining method as implemented in MEGA 3.1. Highly divergent sequences, in which the distance matrices could not be calculated, were excluded from the phylogenetic tree analysis. Bootstrap replicates of 1000 were performed for each phylogenetic tree to assess the reproducibility of the branching patterns.

3.9. Chimeric/anomalous Sequence Analysis

To screen for chimeric sequences, the ribotype sequences were analyzed using nearest- neighbor methods as implemented in the programs such as Bellerophon, Chimera Check and

Mallard and confirmed based on signature analysis and BLAST results. The Bellerophon program (Huber et al., 2004) was executed through the Greengenes website (DeSantis et al.,

2006b), where the sequences were pre-aligned in the NAST format (DeSantis et al., 2006a), and

21

through the Bellerophon server website under the four different correction algorithms of Kimura,

Jukes-Cantor, Huber-Hugenholtz and “none”. The Bellerophon program implemented through the Greengenes website had been modified to include weighting the likelihood of a sequence being chimeric according to the similarity of the parent sequences. Putative chimerics are determined by a divergence ratio of >1.1 and fragment-to-parent levels of similarity of >90%

(DeSantis et al., 2006b). In the program implemented through the Bellerophon server, the

Kimura and Jukes-Cantor distance corrections account for the saturation of nucleotide substitutions in the sequences, making long distance relationships more prominent while Huber-

Hugenholtz correction strongly differentiates between highly similar sequences (Huber et al.,

2004). Sequences were determined to be putative chimerics if the preference score was >1

(Huber et al., 2004). Independent analysis of the ribotype sequences were also performed using the Chimera Check program of the Ribosomal Database Project (RDP) website (Cole et al.,

2003). The Chimera Check analysis was stringent, whereby ribotype sequences were considered to be putative chimeras based on either one of two criteria. Firstly, the Maximum Improvement

Score (MIS) had to be above 55 (Robisoncox et al., 1995). Secondly, both fragments left and right to the sequence breakpoint had to have a higher similarity score to two different sequences in the database than to the similarity score of the whole sequence to a sequence in the database.

Lastly, chimeric and anomalous sequences were detected using the Mallard program (Ashelford et al., 2006), and individually inspected with the Pintail program (Ashelford et al., 2005) according to the recommended protocols, using 99.9% as the cutoff line for identifying unusually high deviation from expectation (DE) values. Mallard and Pintail programs detect anomalous sequences based on the same algorithms. Mallard is used to analyze multiple clone sequences in a library at a time while Pintail is necessary for manual inspection of the individual sequences.

22

Chimeric sequences are a category of anomalous sequences which have fragments of the sequence match different 16S rRNA gene sequences. Other anomalous sequences include those that are highly divergent from their closest sequence match and do not appear to be a 16S rRNA gene. The ribotype sequences that were detected as putative chimeric sequences in more than one of the different settings under Bellerophon were determined to be false positives if they were not detected by the Chimera Check or Mallard programs. Putative chimerics identified from the different programs were confirmed to be chimeric sequences if they were detected by at least two programs, were misclassified in the phylogenetic tree when compared to the classification based on the naïve Bayesian rRNA classifier on the RDP website (based on signature analysis), were outliers in the initial phylogenetic tree and if the individual fragments of the putative chimeric sequences matched with divergent sequences when compared to public databases using BLAST.

Ribotype sequences determined to be chimeric, along with the anomalous sequences, were excluded from the subsequent phylogenetic trees which were constructed using the methods described above, according to the phyla grouping of the ribotypes. Artifact sequences were also excluded from the ribotype richness, diversity and similarity analyses.

3.10. Bacteria Ribotype Richness, Diversity and Similarity Analysis

Rarefaction analyses based on the different algorithms and indices were performed for bacteria communities from each of the FST colony, averaged over 100 randomization runs using the EstimateS V7.5 software (Colwell, 2005). Mao Tau (Mao et al., 2005) is an algorithm used to measure the ribotype richness in the FST colonies. To estimate the percentage of total ribotypes captured in each colony, the non-parametric estimators of ribotype richness, Chao1 (Chao et al.,

2001), ICE (Colwell and Coddington, 1994), ACE (Chao et al., 2000) and Jack1 (Heltshe and

Forrester, 1983) were used. The Chao1 and Jack1 estimators place more weight on the

23

singletons and doubletons; ACE considers the abundance of ribotypes in a sample while ICE is based on the presence-absence of the ribotypes.

Bacteria ribotype diversity was calculated using the Simpson’s index of diversity

(Simpson, 1949) and the Shannon index (Shannon and Weaver, 1949). The Simpson’s reciprocal index (1/D) in EstimateS was converted to the Simpson’s index of diversity (1-D) for use as a measure of the ribotype diversity, where a greater number represents a higher diversity within a scale of 0 - 1. It shows the likelihood that two ribotypes randomly drawn from the same colony are different. The Shannon index has a scale between 0 - 4.5, where a higher value represents a higher diversity. The Shannon index weighs both the ribotype richness and abundance while the Simpson’s index of diversity is biased towards the dominant ribotypes.

The similarity of bacteria composition among the FST colonies was determined using the Chao-

Jaccard Abundance index (Chao et al., 2005), Bray-Curtis index (Bray and Curtis, 1957) and the

Morisita-Horn index (Horn, 1966). Similarity among the colonies was defined by the number of shared bacteria ribotypes, in terms of ribotype richness and abundance, between two colonies when performing a pairwise comparison. The Chao-Jaccard Abundance index is a modification of the classic Jaccard index and is less sensitive to the sample size because it accounts for the unseen shared ribotypes. The Bray-Curtis index is dominated by the most frequent ribotypes while the Morisita-Horn index considers both the ribotype richness and abundance. The values for the Chao-Jaccard Abundance, Bray-Curtis and Morisita-Horn indices range from 0 to 1, where 0 means no similarity exists between the two samples while 1 shows perfect similarity.

Comparison of the different richness, diversity and similarity indices across the FST colonies were made and the results for each of the FST colony were interpolated according to the colony with the lowest number of clone samples.

24

3.11. Statistical Analysis for Bacteria Compositions in the Fresh vs Alcohol-stored Louisiana FST Samples

(1) The DNA concentration from the FST samples and concentration of 16S rRNA PCR products from each of the six Louisiana FST colonies were compared between the fresh and alcohol-stored FST samples using a two-tailed Mann-Whitney U test implemented through the

SPSS V16 software (SPSS Inc., Chicago, IL). The DNA concentrations were determined to be significantly different between the fresh and alcohol-stored samples if the p-value was ≤ 0.05.

(2) The bacteria ribotype richness and diversity were compared between the fresh and alcohol- stored FST samples using the 95% confidence intervals of the rarefaction analyses using different indices to determine if they overlapped. (3) The two-tailed Mann-Whitney U test was used to determine if the similarities among the bacteria communities of the Louisiana FST samples of the same storage condition were higher than similarities of the bacteria communities of the Louisiana FST samples from different storage conditions. A p-value of ≤0.05 would indicate a significant difference between the FST colonies with the same storage conditions and those of different storage conditions. (4) To further investigate the effects of long-term alcohol storage on the bacteria communities, the proportion of bacteria classes between the three fresh

FST colonies and the three alcohol-stored FST colonies from Louisiana were tested using the

Kruskal-Wallis test to see if they were different. Class proportions for the FST colonies were different if the p-value was ≤0.05. The Mann-Whitney U test was used to compare the individual proportions of bacteria classes in the fresh FST colonies with the alcohol-stored FST colonies. The individual class proportions were different if the p-value was ≤0.05.

25

3.12. Statistical Analysis for Bacteria Compositions in the Louisiana vs China FST Samples

Steps 2-4 from the comparison between the fresh and alcohol-stored Louisiana FST samples were used, but the fresh and alcohol-stored Louisiana FST samples were replaced with the alcohol-stored Louisiana and China FST samples.

26

CHAPTER 4. RESULTS

4.1. Applicability of Using ARDRA for Studying Bacteria Communities in FST Samples

The first objective was to determine if using ARDRA to screen the 16S rRNA gene clone libraries would save costs and time by reducing sequencing redundancy. As one of the criteria,

ARDRA needs to be effective in distinguishing the banding patterns of the majority of the ribotypes, defined by having a less than 97% sequence similarity in their 16S rRNA gene sequences. Based on in-silico digest using some 16S rRNA gene sequences previously discovered from the guts of termites, none of the tested restriction enzymes distinguished all the sequences when used alone. Double-digest reaction using HaeIII and RsaI differentiated all the

16S rRNA gene sequences tested and therefore was used in the ARDRA. Of the 92 clones from the Hunan alcohol-stored FST colony that were analyzed using ARDRA, 3 were omitted due to poor sequence quality and another 9 were excluded due to poor quality of the PCR products.

The remaining 80 clones consisted of 26 ribotypes as determined by DNA sequencing but showed only 25 distinguishable RFLP-types. RFLP patterns of the clones consisted of 3 to 10 bands ranging from 70-1400bp (Figure 4.1). Only two of the ribotypes, Cf4-20 and Cf4-62, had

RFLP-types that could not be distinguished and were grouped under RFLP-type H21 (Table 4.1).

These two ribotypes, consisting of 12 clones, had 96.9% sequence similarity and clustered together within the Clostridia group (Figure 4.3). ARDRA was effective in distinguishing 92% of the ribotypes.

As a second criterion for ARDRA’s applicability, it had to be cheaper than high- throughput sequencing. The material and labor costs for performing ARDRA added up to approximately $5.00/sample while high-throughput sequencing costs around $6.80/sample, assuming a labor cost of $12/hour. To be cost effective, the clone library needs to consist mostly

27

of clones with the same ribotypes so that duplicate clones can be excluded from sequencing.

Unique ribotypes must not comprise more than 25% of a 16S rRNA gene clone library for

ARDRA to be cost efficient since the the price of ARDRA is three fourths the price of performing sequencing. Considering the number of unique ribotypes comprised one third of the

80 clones that were evaluated and that ARDRA was more labor-intensive, pre-screening the clone libraries using ARDRA would not be a practical option. ARDRA was not used in analyzing the remaining 16S rRNA gene libraries.

1,114bp

67bp

M M M

Figure 4.1. RFLP-types of some 16S rRNA gene clones used in this study (M: DNA molecular weight marker VIII (Roche Applied Science, Indianapolis, IN)).

Table 4.1. RFLP-types of ninety-two 16S rRNA gene clones from Hunan FST colony.

Ribotype # of clones RFLP-type Cf4-05 2 H01 Cf4-13 1 H02 Cf2-03 5 H03 Cf4-29 1 H04 Cf4-69 1 H05

(Table continued) 28

Cf2-21 1 H06 Cf2-02 1 H07 Cf4-03 4 H08 Cf4-66 2 H09 Cf4-06 2 H10 Cf4-34 1 H11 Cf7-43 1 H12 Cf1-12 1 H13 Cf1-11 2 H14 Cf2-30 6 H15 Cf4-55 1 H16 Cf8-01 1 H17 Cf2-31 1 H18 Cf7-18 1 H19 Cf4-32 27 H20 Cf4-20 4 H21a Cf4-62 8 H21a Cf4-57 1 H22 Cf4-67 1 H23 Cf4-30 3 H24 Cf8-04 1 H25 Poor PCR products 9 Poor sequences 3 a. Bold labels indicate RFLP-types that could not be differentiated.

4.2. Identification of Artifact Sequences in the Bacteria Community of the FST

Of the 2358 clones that were analyzed from the all nine termite colonies, 482 clones were omitted due to poor quality sequences or the inability to align the forward and reverse fragments of the 16S rRNA gene sequence. Of the remaining 1876 clones, 237 bacteria ribotypes were identified.

To ensure that none of the bacteria ribotypes used in further analyses were artifact sequences, chimeric and anomalous sequences were detected using the Bellerophon program,

Chimera Check and Mallard. The different programs yielded 96 putative chimeras altogether

(Table 4.2). Many of the putative chimeric sequences detected using the four different correction

29

algorithm settings within Bellerophon were different. When the Jukes-Cantor and Kimura corrections were implemented, no putative chimerics were detected (not shown). When no correction was applied, Bellerophon yielded 18 putative chimeras. The use of the Huber-

Hugenholtz correction setting increased the number of putative chimerics to 23 sequences.

Chimeric sequence analysis using the modified Bellerophon through the Greengenes website resulted in the identification of 21 putative chimerics. Chimera Check identified up to 47 putative chimerics while Mallard found 22 anomalous sequences.

Most of the putative chimeric sequences were found in only one of the three programs used, indicating that a high number of false positives were detected. When the putative chimeric and anomalous sequences were compared to the classification of the ribotypes according to their phyla, Mallard was the only program that detected putative anomalous sequences which matched the outliers in the phylogenetic tree (Fig. 2.1). The ribotype Cf4-43 was detected as a putative chimeric by both Bellerophon and Chimera Check. Upon manual inspection using Pintail, was found to be a two-fragment chimeric sequence with a breakpoint at around 750bp. Even though

Cf10-18, a 1809bp sequence, was not detected as a putative chimeric in any of the three programs, it was designated as an anomalous sequence because it was highly divergent from the other ribotype sequences and when inspected manually using Pintail, was detected as a questionable sequence. Upon further inspection, the sequence was found to have been erroneously aligned.

Altogether, 18 ribotypes consisting of 39 clones were determined to be anomalous sequences and 6 ribotypes consisting of 13 clones were chimeric sequences. The chimeric sequences comprised 0.57% of the total clones in the fresh Louisiana FST samples, 0.87% in the alcohol-stored Louisiana FST samples and 0.56% in the alcohol-stored China FST samples. Of

30

the putative chimeric sequences detected by the Bellerophon program with the different settings, the modified Bellerophon and Chimera Check, 73 of the sequences were determined to be false positives. Mallard was the most suitable program for detecting the artifact sequences in this study since it found 91.6% of the artifact sequences and none of the sequences detected were false positives. All the chimeric and anomalous sequences were excluded from subsequent analyses.

4.3. Identities of Bacteria in the FST

With the exclusion of the chimeric and anomalous sequences, the remaining 213 bacteria ribotypes comprising 1824 clones were classified into 13 bacteria phyla: Bacteroidetes (42.91% of the total clones), Firmicutes (30.49%), Spirochaetes (11.30%), Actinobacteria (5.70%),

Proteobacteria (2.24%), Tenericutes (1.55%), candidate division Termite Group 1 (1.01%), candidate division TM7 (0.64%), Verrucomicrobia (0.59%), Planctomycetes (0.48%), candidate division Synergistes (0.21%), candidate division ZB3 (0.05%) and Cyanobacteria (0.05%)

(Figure 4.3). Within each phylogenetic tree, the bacteria ribotypes were not clustered according to the storage conditions of the FST samples (fresh vs. alcohol) or the geographic region of the

FST samples (Louisiana vs. China).

The most abundant ribotype, Cf2-30, which consisted 38.71% of the clones analyzed, was classified within the phylum Bacteroidetes. The most diverse phylum was the Firmicutes, with 72 bacteria ribotypes belonging to and Clostridia. Ribotype Cf4-32 from the

Clostridia group was the second most highly represented ribotype at 147 clones (8.21% of total clones). The third most abundant ribotype, Cf4-07, was also within the Clostridia class and found only in the alcohol-stored FST samples from Louisiana and China and made up 6.85% of the total clones. The rest of the ribotypes occurred in much smaller proportions, comprising less

31

Table 4.2. List of confirmed chimerics and anomalous sequences after comparing the results from different settings within Bellerophon, the modified Bellerophon on Greengenes website, Chimera Check and Mallard. a Ribotype Bellerophon CC Mallard Final status Breakpoint Description Nonea HHa GGa Cf10-01 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-02 Xb X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-03 X X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-05 X X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-12 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-13 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-14 X X X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-15 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-16 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-17 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-18 Anomalousc 5' and 3' end fragments aligned in reversed directions

Cf10-19 X Anomalous Unusually divergent from a 16S rRNA gene sequence (1809bp)

Cf10-20 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf10-23 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf2-24 X X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf2-38 X X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf2-05 X Anomalous 5' and 3' end fragments aligned in reversed directions

Cf4-72 X X X X Anomalous 550 Poor sequence quality up to 466bp on the 5' end

Cf10-04 X X X Chimeric 925 Two fragment chimera, with 5' end Spirochaetes and 3' end Clostridia in origin

Cf2-42 X X X X Chimeric 925 Two fragment chimera, with 5' end Spirochaetes and 3' end Bacteroidetes in origin

Cf7-74 X X X X X Chimeric 975 Two fragment chimera, with 5' end Spirochaetes and 3' end Bacteroidetes in origin Cf1-20 X X X X X Chimeric 875 Two fragment chimera, with 5' end Actinobacteria and 3' end Spirochaetes in origin Cf4-84 X X X X Chimeric 850 Two fragment chimera, with 5' end Clostridia and 3' end Actinobacteria in origin

Cf4-43 X X Chimeric 750 Two fragment chimera, with both fragments Clostridia in origin

Cf4-46 X X Non-anomalous

Cf4-58 X X Non-anomalous (Table continued) 32

Cf4-70 X X Non-anomalous

Cf4-83 X X Non-anomalous

Cf6-03 X X Non-anomalous

Cf6-04 X X Non-anomalous

Cf6-16 X X Non-anomalous

Cf7-63 X X Non-anomalous a Putative chimerics found in only one of the programs or different settings within Bellerophon were not shown. Number of sequences not shown: None - 3; Huber-Hugenholtz (HH) - 10; Greengenes (GG) - 14; Chimera Check (CC) - 38 b "X" indicates the sequences that were detected by the programs. c Anomalous sequence was not detected by any of the programs.

33

Figure 4.2. Phylogenetic tree showing the placement of the anomalous or chimeric sequences (ribotype labels) among all the other ribotypes found in the nine FST colonies. The phyla represented by the letters are candidate division ZB3a, Cyanobacteriab and Planctomycetesc. The phylogenetic tree was constructed using the neighbor-joining method with 1,000 replicates of bootstrapping. Nodes with bootstrap support of ≥70% (hollow circles) and ≥95% (solid circles) are marked. Cf10-12, Cf10-13, Cf10-14, Cf10-15, Cf10-16, Cf10-17, Cf10-19, Cf10-20 and Cf2-38 were excluded because the Kimura distance could not be calculated with those sequences. The scale bar represents a 20% sequence divergence.

34

than 1.92% of the total clones. The three most abundant ribotypes, Cf2-30, Cf4-32 and Cf4-07, were previously found in the FST. None of the ribotypes found in this study identified closely with any potential FST pathogens.

Of the 213 ribotypes, 151 were novel (70.89% of total ribotypes). Even though the novel ribotypes made up the bulk of the 16S rRNA gene libraries, they made up only 30.21% of the total clones. There were 85 singletons and 37 doubletons. The single- and doubletons made up

57.28% of the total ribotypes but were represented in only 8.72% of the total number of clones.

Sixty-two of the 213 ribotypes identified had sequence similarities of 97% or higher to known bacteria sequences on the DDBJ/EMBL/GenBank databases (Table 4.3). Of those, 33 ribotypes

(63.49% of total clones) were previously identified from the guts of C. formosanus, 11 ribotypes

(2.96% of total clones) were from other termite guts and the remaining 18 ribotypes (3.34% of total clones) were endo- or ectosymbionts of other organisms or from the environment. Most of the individuals from the bacteria community were specific to the FST and other termites.

Table 4.3. List of bacteria ribotypes with sequence similarities of 97% or higher to known bacteria sequences from the DDBJ/EMBL/GenBank databases, according to their abundance. Top Match Accession # of Ribotypea Number Clones Description/Isolation source

1 Cf2-30 AB062769.1 706 gut of Coptotermes formosanus 1 Cf4-32 AB062845.1 147 gut of Coptotermes formosanus 19 Cf4-07 AY533171.1 125 gut of Coptotermes formosanus 1 Cf7-72 AB062829.1 35 gut of Coptotermes formosanus 1 Cf1-12 AB062844.1 22 gut of Coptotermes formosanus 22 Cf2-32 AY571962.1 12 gut of Coptotermes formosanus 1 Cf4-30 AB062842.1 10 gut of Coptotermes formosanus 1 Cf1-22 AB062818.1 9 gut of Coptotermes formosanus 1 Cf4-98 AB062774.1 8 gut of Coptotermes formosanus 1 Cf5-01 AB062813.1 8 gut of Coptotermes formosanus 1 Cf6-10 AB062831.1 8 gut of Coptotermes formosanus 1 Cf4-29 AB062841.1 6 gut of Coptotermes formosanus 1 Cf6-19 AB062810.1 6 gut of Coptotermes formosanus (Table continued) 35

18 Cf7-40 AF068347.1 6 gut of Coptotermes formosanus 1 Cf7-65 AB062768.1 6 gut of Coptotermes formosanus 1 Cf2-13 AB062826.1 5 gut of Coptotermes formosanus 1 Cf4-94 AB062805.1 5 gut of Coptotermes formosanus 1 Cf7-31 AB062840.1 5 gut of Coptotermes formosanus 1 Cf4-23 AB062775.1 3 gut of Coptotermes formosanus 1 Cf4-97 AB062811.1 3 gut of Coptotermes formosanus 1 Cf7-19 AB062808.1 3 gut of Coptotermes formosanus 1 Cf7-24 AB062823.1 3 gut of Coptotermes formosanus 1 Cf7-36 AB062846.1 3 gut of Coptotermes formosanus 1 Cf2-16 AB062847.1 2 gut of Coptotermes formosanus 1 Cf4-28 AB062836.1 2 gut of Coptotermes formosanus 1 Cf7-16 AB062808.1 2 gut of Coptotermes formosanus 18 Cf7-39 AF068347.1 2 gut of Coptotermes formosanus 20 Cf2-01 AB055736.1 1 gut of Coptotermes formosanus 1 Cf2-14 AB062835.1 1 gut of Coptotermes formosanus 1 Cf4-24 AB062773.1 1 gut of Coptotermes formosanus 1 Cf4-25 AB062827.1 1 gut of Coptotermes formosanus 1 Cf7-30 AB062839.1 1 gut of Coptotermes formosanus 1 Cf7-32 AB062843.1 1 gut of Coptotermes formosanus 2 Cf4-41 AB088952.1 29 gut of Reticulitermes speratus 2 Cf4-52 AB088985.1 9 gut of Reticulitermes speratus 21 Cf4-64 AB198466.1 3 gut of Reticulitermes speratus 2 Cf9-02 AB089122.1 3 gut of Reticulitermes speratus 2 Cf4-37 AB100475.1 2 gut of Reticulitermes speratus 2 Cf4-44 AB088967.1 1 gut of Reticulitermes speratus 2 Cf4-56 AB100465.1 1 gut of Reticulitermes speratus 3 Cf9-01 AY587231.1 2 gut of Reticulitermes flavipes 4 Cf6-05 DQ453130.1 2 gut of the termite Odontotermes formosanus Shiraki 5 Cf10-22 AB298053.1 1 endosymbiont of Snyderella sp. in Cryptotermes secundus 6 Cf4-50 AB231042.1 1 gut of Hodotermopsis sjoestedti 7 Cf4-70 AB244437.1 35 larvae of Myrmeleon bore 8 Cf1-19 EU137440.1 5 black-tailed prairie dog 9 Cf6-08 AY959000.1 2 microbes on the human vaginal epithelium 10 Cf6-01 AM410705.1 2 causes septicemia 11 Cf4-69 AF467426.1 1 ectoparasitic chewing lice of pocket gophers 22 Cf6-04 EU047556.1 1 ability to solubilize insoluble mineral phosphate 12 Cf6-18 AJ292676.1 1 polychlorinated biphenyl-polluted soil 22 Cf6-16 EF592491.1 1 cellulose-producing bacterium 13 Cf3-01 EU017026.1 1 chloroplast of Musa acuminata 14 Cf6-13 AM084010.1 1 denitrifying bacteria from a municipal wastewater treatment plant 15 Cf2-36 EF025213.1 1 intestine of turkey

36 (Table continued)

22 Cf6-14 CP000539.1 1 nitroaromatic compound-degrader 16 Cf4-01 DQ105968.1 1 P-decomposer in the sediment of Guanting reservoir in Beijing 22 Cf2-35 AB298723.2 1 rice straw residue in a methanogenic reactor of cattle farm waste 17 Cf6-02 AY528708.1 1 Mimosa pudica 22 Cf4-02 EU057605.1 1 soil 22 Cf2-37 DQ337018.1 4 subsurface water of the Kalahari Shield, South Africa Cf6-17 EU137383.1 1 black-tailed prairie dog8

1. 2. 3. 4. Shinzato1. Shinzato et al., et al., 2005 2005;; 2. Hongoh Hongoh et al., 2003;et al., 3. (Stevenson 2003; et Stevensonal., 2004); 4. (Chou et etal., al., 2004;2007); 5. (Ikeda Chou-Ohtsubo et al.,et al., 2007; 2007); 6. (Nakajima5. et al., 2006); 7. (Nishiwaki et al., 2007)6. ; 8. (Jones et al., 2008); 9. (Hyman7. et al., 2005); 10. (Vaneechoutte et al., 8.2008) ; 11. (Reed and Hafner, Ikeda 2002)-Ohtsubo; 12. (Nogales et et al., 2001) 2007;; 13. (Jansen Nakajima et al., 2007) ;et 14. al.,(Heylen 2006; et al., 2006) Nishiwaki; 15. (Scupham, et 2007) al.,; 16.2007; (Li et al., JonesIn press) ;et 17. al., (Barrett 9. 10. 11. (Table continued) and2008; Parker, 2005)Hyman; 18. (Lilburn et al., et 2005; al., 1999) ; 19. Vaneechoutte (Higashiguchi et al., et 2006) al.,; 20.2008; (Ohkuma etReed al., 2002) and; 21. Hafner, (Nakajima 2002;et al., 2005) 12. Nogales et al., 2001; 13. Jansen et al., 2007; 14. Heylen et al., 2006; 15. Scupham, 2007; 16. (Li et al., In press; 17. Barrett and Parker, 2005; 18. Lilburn et al., 1999; 19. (Higashiguchi et al., 2006; 20. Ohkuma et al., 2002; 21. Nakajima et al., 2005; 22. Unpublished. a. Bacteria phyla represented: Cf1-xx – Actinobacteria, Cf2-xx – Bacteroidetes, Cf3-xx – Cyanobacteria, Cf4-xx – Spirochaetes, Cf6-xx – Proteobacteria, Cf7-xx – Firmicutes, Cf9-xx – Verrucomicrobia, Cf10-xx – Unclassfied ribotype

37

A. Bacteroidetes Cf2-15 Cf2-26 Cf2-10 50 clone BCf7-17 (AB062835) 97 clone RsTu1-13 (AB192208) 70 Cf2-06 61 Cf2-07 82 clone RsaM33 (AY571447) 99 clone RsaHf397 (AY571449) 98 62 clone RsaHf278 (AY571450) Cf2-14 100 clone BCf7-02 (AB062769) Cf2-04 52 61 79 Cf2-30 100 clone BCf1-03 (AB062769) Cf2-41 100 clone Rs-P65 (AB088938) Cf2-17 100 clone Rs-096 (AB100459) Cf2-02 97 98 Cf2-39 sp. AM15 (EU252503) 100 100 D. mossii strain CCUG 43457 (AJ319867) 100 Cf2-33 Bacteroidetes 72 Cf2-01 60 100 C. III of the termite bacteroides (AB055736) 100 clone RsaHw509 (AY571441) 81 Cf2-32 100 Bacterium S1 (AY571962) Cf2-13 100 clone BCf5-08 (AB062826) 98 clone Rs-K10 (AB088924) 100 clone Rs-B45 (AB088942) 96 Cf2-19 99 93 Cf2-23 100 Cf2-20 99 100 Cf2-21 Cf2-40 100 clone SJTU F 09 65 (EF398967) Cf2-18 100 clone Rs-106 (AB100460) 99 clone RsW01-084 (AB198508) 100 Cf2-28 100 Cf2-29 Cf2-16 100 clone BCf9-17 (AB062847) 75 Cf2-25 100 clone Cc3-055 (AB299528) Cf2-22 Cf2-09 56 100 clone RsaHw538 (AY571428) 91 Cf2-31 100 clone NkW01-014 (AB231051) Cf2-35 Flavobacteria 100 Flavobacteriaceae bacterium WH032 (AB298723) 65 Cf2-36 100 clone B5 B4 (EF025213)

0.02

Figure 3.3.4.3. Phylogenetic Phylogenetic trees trees showing showing the the 16S 16S rRNA rRNA gene gene sequences sequences from from the Formosan the FSTs in this studysubterranean (Cf) and termites their highest and their match highest from match the DDB on theJ/EMBL/Genbank DDBJ/EMBL/Genbank databases. databases. A. Bacteroidetes, A. B.Bacteroidetes, Firmicutes, B. Firmicutes, C. Spirochaetes, C. Spirochaetes, D. Actinobacteria, D. Actinobacteria, E. Proteobacteria,E. Proteobacteria, F. F. TM7, TM7, G. G. Verrucomicrobia and and Planctomycetes. Planctomycetes. The The phylogenetic phylogenetic trees treeswere wereconstructed constructed using the using the neighbor joining joining method method with with 1000 1000 bootstrap bootstrap replicates. replicates. Only Only bootstrap bootstrap values values of ≥50 of are ≥50 are indicated on thethe branchbranch nodes.nodes. The The scale scale bars bars represent represent 2% 2% difference difference in innucleotide nucleotide sequence. sequence. Ribotype sequences sequences from from other other studies studies were were marked marked with withcircles circles (filled (filled– termite – -specifictermite- specific bacteria, hollow – envienvironmentalronmental bacteria).

(Figure continued)

38

B. Firmicutes & Synergistes

Cf4-38 100 clone RsTz-92 (AB192177) Cf4-86 100 clone MgMjD-101 (AB234471) Cf4-50 100 clone HsW01-048 (AB231042) 81 Cf4-34 Cf4-35 100 clone Rs-P37 (AB088967) 53 52 Cf4-44 55 clone Rs-L03 (AB100484) clone UASB brew B8 (AF332709) Cf4-42 100 clone Rs-N94 (AB089040) Cf4-21 85 Cf4-16 100 clone p-878-a5 (AF371774) Cf4-41 75 100 clone Rs-N86 (AB088952) Cf4-40 100 Cf4-39 55 clone Rs-M04 (AB088979) Cf4- 02 100 Clostridium sp. CJT-3 (EU057605) 82 clone 5-1-N (AB197849) 96 Cf4-53 72 Cf4-11 99 clone 013C-H6 (DQ905456) Cf4-37 97 Cf4-85 52 100 clone Rs-L36 (AB100475) 99 clone HsW01-047 (AB231041) Cf4-88 100 53 Cf4-46 95 clone RsW02-034 (AB198524) 72 Cf4-66 95 clone AF17 (DQ394667) 100 Cf4-67 100 clone BRC147 (EF436433) Cf4-63 100 clone RsW01-073 (AB198499) 87 Cf4-12 85 clone MH87 (AY916298) 100 Cf4-77 99 clone aab21g12 (DQ815751) Cf4-48 95 94 Cf4-87 100 clone HsW01-031 (AB231039) Cf4-23 Clostridia 100 100 clone BCf10-04 (AB062775) 75 Cf4-96 89 clone Rs-J36 (AB089019) Cf4-91 99 90 Cf4-52 100 clone Rs-K66 (AB088985) Cf4-20 100 Cf4-62 98 clone Rs-Q64 (AB088970) 65 clone RsaHf311 (AY571396) 74 clone RsW02-061 (AB198536) Cf4-57 100 clone Rs-E61 (AB088987) 100 Cf4-25 100 clone BCf5-15 (AB062827) 78 Cf4-45 Cf4-30 71 98 100 clone BCf9-05 (AB062842) 96 Cf4-32 100 clone BCf9-13 (AB062845) Cf4-03 100 clone RsW02-061 (AB198536) Cf4-59 100 clone Rs-H32 (AB088972) 94 Cf4-14 Cf4-75 clone B5 N17 (EF025286) 57 Cf4-74 57 clone BOf1-23 (AB288883) 98 Cf4-98 100 100 clone BCf1-20 (AB062774) Cf4-24 100 clone BCf1-16 (AB062773) Cf4-58 79 clone Rs-F81 (AB088989) 71 Cf4-61 55 clone Rs-Q18 (AB088991) 100 Cf4-60 Cf4-94 59 100 clone BCf10-10 (AB062805) Cf4-29 64 100 clone BCf8-25 (AB062841) 100 Cf4 -83 Cf4-54 100 Cf4-55 98 89 clone Rs-E65 (AB126235) 93 Cf4-56 100 clone Rs-L24 (AB100465) Cf4-17 100 Cf4-22 100 Clostridium citroniae RMA 16102 (DQ279737) 93 clone PCD-27 (EF608543) Cf4-28 78 100 clone BCf7-04 (AB062836) 88 clone Rs-C88 (AB089066) 100 Cf4-33 Synergistes 100 Cf4-18 100 clone R38B21 (DQ009680) Cf4-04 100 Endosymbiont TC1 Trimyema compressum (AB118592) Cf4-13 100 Dehalobacter sp. E1 (AY766465) Cf4-10 99 Thermoactinomyces dichotomicus KCTC 3667 (AF138733) Cf4-01 100 Bacillus megaterium (DQ105968) Cf4-69 100 100 100 Staphylococcus sp. clone 87-8 (AF467426) Cf4-07 100 Pilibacter termitis strain TI-1 (AY533171) 98 74 Cf4- 08 99 Cf4-09 Bacilli Cf4-05 58 Cf4-79 99 69 Lactovum miscens anNAG3 (AJ439543) 100 clone C1N (DQ856510) Cf4-64 98 100 clone RsW01-031 (AB198466) Cf4 -81 Lactococcus sp. RT5 (DQ223877) Cf4-70 100 Lactococcus garvieae An1-1 (AB244437) 55 Cf4-97 100 clone BCf2-19 (AB062811)

0.02 (Figure continued) 39

C. Spirochaetes Cf7-06 75 clone rs443 (AJ419822) 66 Cf7-65 100 clone BCf1-01 (AB062768) 86 Cf7-08 67 Cf7-09 clone nc5 (AJ419819) 71 Treponema sp. ZAS-1 (AF093251) Cf7-01 clone HsDiSp319 (AB032009) 62 clone Rs-E18 (AB088893) clone mpsp2 (X89050) Treponema sp. SPIT5 (AM182455) Cf7-41 100 clone CFS6 (AF068345) 64 Cf7-51 Cf7-46 Cf7-72 100 clone BCf5-23 (AB062829) clone RsaHf236 (AY571482) 63 Cf7-12 53 95 Cf7-38 100 clone NkS4 (AB084953) Cf7-47 clone 290cost002-P3L-1157 (EF453804) clone 290cost002-P3L-1474 (EF454022) clone mp4 (AJ458946) clone Rs-A19 (AB088894) clone Rs-E21 (AB088903) Cf7-68 Cf7-42 98 Cf7-49 Cf7-24 100 clone BCf4-14 (AB419821) 99 Cf7-25 94 clone rs438 (AJ419821) 87 Cf7-63 62 clone RPK-52 (AB192256) Cf7-30 100 clone BCf7-24 (AB062839) Cf7-55 Cf7-56 Cf7-54 Cf7-29 100 Cf7-43 80 clone BCf6-24 (AB062834) Cf7-50 clone 290cost002-P3L-646 (EF454909) Cf7-39 100 clone CFS121 (AF068347) 99 Cf7-40 87 Cf7-60 63 Cf7-13 72 Cf7-58 Cf7-71 clone RsaHf365 (AY571483) Cf7-59 71 Cf7-73 clone 290cost002-P3L-1239 (EF453867) 100 clone 290cost002-P3L-517 (EF454847) Cf7-34 61 Cf7-35 83 Cf7-36 100 clone BCf9-14 (AB062846) Cf7-37 Cf7-17 93 Cf7-52 62 Cf7-20 Cf7-15 Cf7-18 Cf7-31 100 clone BCf8-03 (AB062840) Cf7-21 57 Cf7-27 Cf7-19 79 clone BCf5-01 (AB062825) 64 Cf7-22 Cf7-16 100 clone BCf11-05 (AB062808) 89 Cf7-32 100 clone BCf9-08 (AB062843) Cf7-53 99 Cf7-61 Cf7-03 Cf7-28 100 clone BCf6-15 (AB062832) Cf7-48 Cf7-64 100 clone RsTz-71 (AB192150)

0.02 (Figure continued)

40

D. Proteobacteria Cf6-13 100 Acidovorax sp. R-24667 (AM084010) 100 Cf6-14 100 Acidovorax sp. JS42 (CP000539) 85 Cf6-02 59 100 Burkholderia sp. MPUD4.5 (AY528708) Cf6-06 Betaproteobacteria 100 clone COB P3-26 (AY160866) 100 Cf6-07 100 clone DB-3 (DQ836748) 54 Cf6-10 100 clone BCf6-08 (AB062831) Cf6-18 100 clone WD2124 (AJ292676) Cf6-01 100 100 Acinetobacter junii ACI289 (AM410705) 68 Cf6-08 100 clone rRNA227 (AY959000) Cf6-04 99 Gammaproteobacteria 99 Enterobacter hormaechei strain TMPSB-T10 (EU047556) Cf6-05 100 99 Enterobacteriaceae bacterium Eant3-3 (DQ453130) 56 Cf6-16 70 Enterobacter sp. xw (EF592491) Cf6-11 100 clone MTG-93 (DQ307726) Cf6-03 Deltaproteobacteria 100 100 Desulfovibrio sp. ABHU2SB (AF056090) 100 Cf6-19 100 clone BCf11-19 (AB062810) Cf6-12 Epsilonproteobacteria 100 clone Rs-P71 (AB089112) 0.02 E. Actinobacteria Cf1-22 100 clone BCf3-22 (AB062818) 99 Cf1-10 100 Cf1-09 97 Cf1-11 73 Cellulosimicrobium cellulans SSCT73 (AB210965) Cf1-03 100 clone Rs-K09 (AB089078) 54 Cf1-19 52 100 clone Oh3137A10B (EU137440) Cf1-15 100 clone Rs-Q71 (AB089082) Cf1-08 100 clone fc3 (DQ303278) Cf1-02 100 clone Rs-J10 (AB089074) 100 Cf1-14 Cf1-12 100 clone BCf9-11 (AB062844) 100 74 Cf1-13 76 clone AP13U.307 (AM278923) Cf1-06 97 57 Cf1-17 54 Cf1-04 100 Cf1-05 84 Cf1-07 54 clone COB P3-21 (AY160874)

0.02 F. TM7 Cf8-01 91 Cf8-02 98 clone RsaM67 (AY571500) Cf8-03 100 clone RsW02-021 (AB198518) Cf8-04 100 clone Cc3-038 (AB299568)

0.02 G. Planctomycetes & Verrucomicrobia Cf9-05 100 clone vadinHA64 (U81738) 100 Cf9-04 97 Cf9-01 Verrucomicrobiae 100 Opitutaceae bacterium TAV1 (AY587231) Cf9-02 100 clone Rs-P07 (AB089122) Cf5-01 Planctomycetacia 100 clone BCf2-25 (AB062813)

0.02

41

4.4. Comparison Between the Fresh and Alcohol-stored Louisiana FST Samples

4.4.1. Effects of Alcohol Storage on DNA Degradation of the Bacteria in the FST Colonies

To investigate the effects of long-term storage of the FST samples on the integrity of the

bacterial DNA, the extracted DNA from the three fresh and three alcohol-stored Louisiana FST

colonies was compared (Lanes 1-6, Figure 4.4). In the fresh FST samples, the presence of sharp

bands measuring about 10,000bp in size showed good quality genomic DNA. A higher degree

of DNA smearing appeared on the gel in the alcohol-stored FST samples compared to the fresh

FST samples, indicating that the DNA was more fragmented. The higher quantity of DNA yield

in the alcohol-stored FST samples compared to the fresh FST samples (Table 4.4) was due to the

additional genomic DNA extracted from the termite fat body in addition to the termite gut

contents in the fresh FST samples. The amount of PCR products was higher in the fresh FST

samples compared to the alcohol-stored FST samples (p < 0.001, Mann-Whitney U, U < 0.001).

This confirmed that some degree of DNA degradation did occur in the alcohol samples.

10,000bp

1,500bp

1 2 3 M 4 5 6 7 8 9 1 2 3 M 4 5 6 7 8 9 LA Fresh LA EtOH China EtOH LA Fresh LA EtOH China EtOH

Figure 4.4. a) Extracted DNA from the nine FST colonies (M: DNA molecular weight marker). b) PCR products of 16S rRNA gene sequences from the same nine FST colonies. Although the PCR products were not visible on the agarose gel, presence of DNA was confirmed by the yield of measurable DNA concentrations using spectrophotometry (Table 4.4).

42

Table 4.4. DNA concentration of DNA extracts from 50 FST workers and 16S PCR products for the 9 FST colonies.

FST colony DNA concentration (ng/μl) DNA extracts 16S PCR products 1 6.52 17.69 Louisiana Fresh 2 3.82 11.15 3 1.12 8.57 4 71.79 2.09 Louisiana 5 35.08 4.67 EtOH 6 55.46 1.79 7 157.96 0.51 China 8 65.08 6.37 EtOH 9 55.69 2.95

4.4.2. Alcohol Storage Effects on the Biodiversity of Ribotypes

To compare the bacteria composition of the fresh and alcohol-stored Louisiana FST colonies, an analysis of the ribotype richness and diversity was performed. On average, 25.1% of the ribotypes found in the fresh Louisiana FST samples were found only in one colony compared to 21.5% in the alcohol-stored Louisiana FST samples (Table 4.5). The ribotype richness and diversity values were determined at a sample size of 123 clones for all the bacteria communities of the FST colonies according to the FST colony with the smallest sample size, which was the alcohol-stored FST colony from City Park (123 clones), to standardize comparisons across the FST colonies.

To compare the ribotype richness of the FST colonies, the Mao Tau rarefaction curves were used. Based on the Mao Tau curves, the number of ribotypes observed in the fresh FST colonies ranged from 25-37 ribotypes at 123 clones (Figure 4.5). In comparison, the alcohol- stored FST colonies from Louisiana had around 31-41 ribotypes. The 95% confidence intervals of the Mao Tau curves from the six Louisiana FST colonies overlap (not shown), which means that the species richness is not significantly different among the Louisiana FST colonies.

43

Table 4.5. List of the number of clones sequenced and the actual number of ribotypes, singletons, doubletons and unique ribotypes found in the nine FST colonies and one from Japan (Shinzato et al., 2005).

FST colony LA fresh LA EtOH China EtOH Japan 1 2 3 4 5 6 7 8 9 10 a Clones 350 248 276 123 161 177 172 238 131 250 Ribotypes 57 51 74 40 48 42 38 54 23 49 Singletons 41 35 49 33 30 17 19 34 12 34 Doubletons 7 8 9 3 8 8 9 7 5 7 Uniques 19 9 18 12 12 4 11 15 3 25 Uniques (%) 33.3 17.6 24.3 30.0 25.0 9.5 28.9 27.8 13.0 51.0 a Interpolation of the ribotype richness and diversity values for all the FST colonies was performed based on the number of clones sampled in the alcohol-stored City Park FST colony (colony 4).

60 3 50 5 1 2 LA Fresh 40 8 6 10 4 LA EtOH 30 7 China EtOH

20 Japan

9 Number ribotypesNumber of

10

0 123 0 100 200 300 400

Number of clones sampled Figure 4.5. Species accumulation curves generated based on the Mao Tau means of all nine FST colonies and one from Japan (Shinzato et al., 2005). Ribotype richness for all the FST colonies was compared at a sample size of 123 clones.

None of the rarefaction curves were starting to level off (Figure 4.5), which indicated that our sequencing effort needed to be increased substantially to capture the bulk of the ribotype richness in the FST colonies. The ribotype richness estimators, Chao1, ACE, ICE and Jack1, were used to estimate the percentage of total ribotypes captured from each of the FST colonies

44

(Table 4.6). The Chao1 method estimated that at least 32-48% of the total ribotypes in the fresh

Louisiana FST colonies were captured. Based on Jack1, the lower boundary of total ribotypes estimated to be captured in the fresh FST colonies ranged from 33-51%. The ACE and ICE indices both estimated that the lower boundary of the percentage of ribotypes captured from the fresh FST colonies ranged from 8-40%. The estimated lower bound of total ribotypes captured from the alcohol-stored Louisiana FST colonies was around the same range as the fresh

Louisiana FST colonies (Chao1 – 33-45%; Jack1 – 35-49%; ACE and ICE – 27-39%). Since the

95% confidence intervals for all the ribotype richness estimators in the fresh and alcohol-stored

Louisiana colonies overlap, the total number of ribotypes in the six FST colonies was determined to be not different.

Table 4.6. The estimated percentage of the total number of ribotypes captured and ribotype diversity based on different indices from the nine FST colonies and one from Japan (Shinzato et al., 2005).

LA fresh LA EtOH China EtOH Japan FST colony 1 2 3 4 5 6 7 8 9 10 Chao 1 33-100 32-95 48-100 41-79 45-93 33-67 30-66 38-90 18-42 30-92 ACE 8-100 18-91 40-100 38-71 39-84 27-55 26-55 24-87 12-36 13-86 ICE 8-100 18-91 40-100 38-71 39-84 27-55 26-55 24-87 12-36 13-85 Jack 1 33-48 35-49 51-68 45-59 49-64 35-48 32-45 41-55 19-28 32-46 Simpson (D) 0.51 0.44 0.19 0.06 0.05 0.12 0.11 0.11 0.24 0.49 1-D 0.49 0.56 0.81 0.94 0.95 0.88 0.89 0.89 0.76 0.51 Shannon 1.46 1.66 2.60 3.11 3.25 2.71 2.64 2.72 1.91 1.50

a The percentage of total ribotypes captured and diversity values for all the FST colonies was determined at 123 clones.

While the bacteria ribotype richness estimators provide information about the number of bacteria ribotypes, the Simpson index of diversity (1-D) and Shannon index give an indication of how evenly each ribotype is represented. The fresh Louisiana FST samples were less diverse compared to the alcohol-stored Louisiana FST samples (for both 1-D; Shannon: p = 0.05, Mann-

Whitney U, U < 0.001). The fresh Louisiana FST colonies were less diverse compared to the

45

alcohol-stored Louisiana FST colonies because they were dominated by the Bacteroides ribotype

Cf2-30 (61.84% of the fresh Louisiana FST clones) while the rest of the ribotypes were represented in much smaller numbers but in approximately equal abundances.

To determine the similarities among the bacteria communities of all the FST colonies, the

Chao-Jaccard Abundance, Bray-Curtis and Morisita-Horn indices were used. Similarities were higher among the fresh Louisiana FST colonies and among the alcohol-stored Louisiana FST colonies compared to the similarities between the fresh and alcohol-stored Louisiana FST colonies (Chao-Jaccard Abundance; Bray-Curtis; Morisita-Horn: p < 0.001, Mann-Whitney U, U

< 0.001; p = 0.001, Mann-Whitney U, U = 1.000; p = 0.025, Mann-Whitney U, U = 5.000)

(Table 4.7). This indicates that many of the ribotypes found in the fresh FST colonies were distinct from the alcohol-stored FST colonies. Both the Bray-Curtis and Morisita-Horn indices show that the fresh FST colonies were more similar to each other than the alcohol-stored FST colonies were to one another (p = 0.05, Mann-Whitney U, U <0.001; p = 0.05, Mann-Whitney U,

U < 0.001, respectively). Even though a higher number of ribotypes with similar abundances were shared between colonies of the same storage condition, the bacteria ribotypes did not form distinct clusters according to the storage conditions in the phylogenetic tree analyses (not shown), indicating that the ribotype sequences did not share higher similarities within each storage condition.

4.4.3. Alcohol Storage Effects on the Proportions of Bacteria Classes in the Louisiana FST Colonies

The effect of alcohol storage on the bacteria community of the FST samples was further investigated by observing the differences in proportions of the bacteria classes. The overall distribution of the bacteria classes from all nine FST colonies in this study and the FST colony from Shinzato et al.’s (2005) study is shown in Figure 4.6. 46

Table 4.7. The three similarity indices (Chao-Jaccard Abundance, Bray-Curtis and Morisita- Horn) used to determine the number of shared ribotypes among the nine FST colonies and one from Japan (Shinzato et al., 2005).

a. Lighter shading represents the pairwise comparison among the bacteria communities from colonies of the same storage condition or geographic region. b. Darker shading represents the pairwise comparison among the bacteria communities from colonies between different storage conditions or geographic regions.

47

100% Verrucomicrobiae 90% TM7 80% Planctomycetacia 70% Betaproteobacteria 60% Alphaproteobacteria 50%

40% Deltaproteobacteria

30% Actinobacteria

20% Bacilli

10% Clostridia

0% Spirochaetes 1 2 3 4 5 6 7 8 9 10 Bacteroidetes LA Fresh LA EtOH China EtOH Japan

Figure 4.6. Proportions of bacteria classes from the nine FST colonies and one from Japan (Shinzato et al., 2005). Bacteria classes with ≤10 clones were excluded.

The bacteria class proportions of each FST colony were different between the three fresh

FST samples and the three alcohol-stored FST samples (p < 0.001, Kruskal-Wallis test, χ2 =

46.35). Looking at the bacteria class proportions individually, the proportion of the

Bacteroidetes and Spirochaetes classes were more abundant in the fresh Lousiana FST colonies

(p =0.05, Mann-Whitney U, U < 0.001 for both classes) compared to the alcohol-stored

Louisiana FST colonies while the alcohol-stored FST colonies had a higher abundance of clones from the Actinobacteria, Bacilli and Clostridia classes (p =0.05, Mann-Whitney U, U < 0.001 for all three classes).

Since the major groups in the fresh FST samples, Bacteroidetes and Spirochaetes, were gram-negative bacteria and the major groups in the alcohol-stored FST samples, Actinobacteria,

Bacilli and Clostridia, were gram-positive bacteria, the remaining bacteria classes were classified according to their cell wall types for further investigation (Figure 4.7). The gram-negative bacteria proportions, which were predominant in the fresh FST colonies, were significantly

48

reduced in the FST samples stored long-term in alcohol (p =0.05, Mann-Whitney U, U < 0.001).

The gram-positive bacteria were able to withstand DNA degradation better than the gram- negative bacteria.

100%

90%

80% Gram-negative 70%

60%

50% Low G+C gram-positive 40%

30% Actinobacteria 20% (High G+C gram-positive) 10%

0% 1 2 3 4 5 6 7 8 9 10

LA Fresh LA EtOH China EtOH Japan

Figure 4.7. Proportions of Actinobacteria (high G+C gram-positive), low G+C gram-positive and gram-negative bacteria in the nine FST colonies and one from Japan (Shinzato et al., 2005).

4.5. Comparison Between the Alcohol-stored Louisiana and China FST Samples

4.5.1. Differences in the Biodiversity of Ribotypes Between the Alcohol-stored Louisiana and China FST Samples

On average, 21.5% of the ribotypes found in the Louisiana FST samples were unique from other FST samples compared to 23.3% uniques in the China FST samples (Table 4.5). The bacteria ribotype richness of the alcohol-stored Louisiana FST samples at sampling size of 123 clones as shown in the Mao Tau curves (Figure 4.5) ranged from 25-37 ribotypes.

Approximately 15-30 ribotypes were found in the China FST colony at a sample size of 123 clones. The ribotype richness between the Louisiana and China FST samples was not

49

significantly different as indicated by the overlapping 95% confidence intervals for the curves overlap (not shown).

The ribotype richness estimators approximated that at least a quarter of the total ribotypes in the three Louisiana FST samples were captured at 123 clones (Chao1; ACE; ICE; Jack1: 33-

45; 27-38; 27-38; 35-49). The lower boundaries for the total ribotypes captured in the China

FST samples were approximately in the same range (Chao1; ACE; ICE; Jack1: 18-38; 12-26; 12-

26; 19-41) (Table 4.6). The estimated proportion of total ribotypes captured overlapped for all the FST colonies which means that the total number of ribotypes in the Louisiana and China FST samples were not significantly different. The bacteria ribotype diversity was not significantly different between the Louisiana and China FST colonies (1-D; Shannon: p = 0.20, Mann

Whitney U, U = 1.00; p = 0.10, Mann Whitney U, U = 0.05) (Table 4.6).

The bacteria composition among colonies from the same geographic region and between geographic regions was not significantly different (Chao-Jaccard Abundance; Bray-Curtis;

Morisita-Horn: p = 0.194, Mann Whitney U, U = 9.50; p = 0.136, Mann Whitney U, U = 8.50; p

= 0.478, Mann Whitney U, U = 13.00) (Figure 4.5). No effects of geographic origin of the FST samples on the bacteria communities were detected. In further support of this, the bacteria ribotypes in the Louisiana and China FST samples did not form distinct clusters in the phylogenetic trees (not shown), indicating that the ribotype sequences were equally divergent across FST colonies of different geographic regions.

4.5.2. Differences in the Bacteria Class Proportions Between the Alcohol-stored Louisiana and China FST Samples

The bacteria class proportions were different between the three Louisiana FST samples and three China FST samples (p < 0.001, Kruskal-Wallis test, χ2 = 57.60). When the class proportions were analyzed individually, none of the bacteria class proportions were significantly 50

different between the Louisiana and China FST colonies (p ≥ 0.10, Mann-Whitney U, U ≥ 0.50 for all bacteria classes). The proportions of each bacteria class between the FST colonies of the two storage conditions showed a larger difference compared to the bacteria class proportions between FST colonies of the two geographic regions (Figure 4.6). Looking at the distribution of the bacteria according to their cell wall type in the Louisiana and China FST samples (Figure

4.7), no significant differences were found (p = 0.20, Mann-Whitney U, U = 1.00).

51

CHAPTER 5. DISCUSSION

5.1. ARDRA Analysis

Although ARDRA has been used in bacteria community studies in termites (Fisher, 2006;

Shinzato et al., 2005; Shinzato et al., 2007), its resolution power was not assessed with the direct sequencing of 16S rRNA gene clones. Using the same and only one restriction enzyme as in

Shinzato et al.’s (2005, 2007) studies, a few 16S rRNA gene sequences found in previous studies of FST could not be differentiated based on in-silico digests. It was likely that the use of only one restriction enzyme in the two studies resulted in an underrepresentation of the bacteria ribotypes in the termite guts. The resolution power of ARDRA, when using two restriction enzymes, was found to be almost comparable to the direct sequencing of 16S rRNA gene sequences in this study. The use of ARDRA gave a comparatively accurate representation of the bacteria ribotype as defined by the <97% sequence similarity criterion. In several other bacteria community studies, two to three restriction enzymes in separate digest reactions were also needed to distinguish all or most of the bacteria (De Baere et al., 2002; Lagace et al., 2004;

Stakenborg et al., 2005). Considering the high number of different bacteria ribotypes identified in this study, the use of ARDRA was found to be less suitable when compared to the direct use of high-throughput sequencing of the clones based on a cost-benefit analysis.

When sampling the 16S rRNA gene sequences in bacteria communities, we recommend running a pilot study using high throughput sequencing initially to determine the bacteria ribotype richness since bacteria richness varies from study to study. The use of ARDRA in bacteria community studies should only be considered if the number of novel bacteria ribotypes is lower than a fourth of the number of clones sampled, assuming that the costs of sequencing is about four times higher than the costs of ARDRA. In studies where the bacteria community is

52

expected to comprise a high number of bacteria that are closely related, it would be necessary to use at least two different restriction enzymes selected based on in-silico digests.

5.2. Analysis of Artifact Sequences in the Bacteria Community of the FST Samples

In 16S rRNA gene clone libraries, it is important to screen out artifact sequences. The increasing number of chimeric sequences accumulating in the public databases has led to the development of various programs to screen out chimeric sequences. The two most commonly used programs for detecting chimeric sequences in bacteria community studies, Chimera Check and Bellerophon, were compared with Mallard to determine the program that was most reliable in detecting anomalous sequences. Chimera Check, one of the most used programs in literature

(Robisoncox et al., 1995), was shown to be the least effective program for detecting anomalous and chimeric sequences in this study (Table 4.2), lacks reliability in detecting chimeric sequences among 16S rRNA gene sequences with higher than 84% sequence similarity

(Robisoncox et al., 1995). The four different settings within Bellerophon and the modified

Bellerophon detected different number of sequences as putative chimerics and some of the putative chimerics were different. The substitution models of Jukes-Cantor and Kimura within

Bellerophon failed to detect any chimeric sequences because they assume that the sequences were divergent. When the ribotype sequences were corrected for the saturation of nucleotide substitutions, slight indicators of chimerism in the nucleotide composition were removed, causing chimeric sequences to become undetectable. On the other hand, when the Huber-

Hugenholtz correction setting was applied, Bellerophon identified a high number of putative chimerics because that correction setting emphasizes the distance among similar sequences, increasing the number of false positives. Mallard, a recently developed program (Ashelford et al., 2006), was the only program of the three that was able to detect most of the chimeric and

53

anomalous sequences without any false positives. For 16S rRNA gene clone libraries with ribotype sequences that are highly similar, Mallard would be the most suitable program for eliminating artifact sequences, both chimeric and anomalous ones. However, the detection of artifact sequences should always be verified using at least one other program and another independent analysis method such as signature, secondary structure and "breaking and retreeing" analyses (Lilburn et al., 1999) because none of the methods are effective in detecting all the artifact sequences without any false positives when used as standalones (Ashelford et al., 2006).

In our study, artifact sequences made up 2.44% of the total clones while only 0.69% of the total number of clones was determined to be chimeric sequences. Among the screened libraries in Ashelford et. al’s (2006) study, two termite libraries (Hongoh et al., 2005; Hongoh et al., 2006) had less than 1.00% artifact sequences of the total clones, which was comparable to the number of artifact sequences found in our study. The number of chimeric sequences found in

Shinzato et al.’s (2005 & 2007) studies on two termites made up 4.2% and 3.4% of the total number of clones, respectively. Our lower number of chimeric sequences compared to Shinzato et al.’s (2005 & 2007) studies was likely due to the screening out of the false chimeric sequence positives.

In our study, the alcohol-stored FST samples had a higher degree of DNA degradation compared to the fresh FST samples. Presence of damaged DNA, caused by breaks, apurinic sites, and UV damage, has been attributed to the formation of chimeric molecules in the amplification of 16S rRNA gene sequences (Liesack et al., 1991; Paabo et al., 1990). Breaks in the DNA template could lead to a premature termination of PCR products, which anneal with homologous templates in subsequent PCR cycles to prime DNA synthesis and cause the formation of chimeric sequences (Paabo et al., 1990). However, we did not find a higher

54

occurrence of chimeric or anomalous sequences in our alcohol-stored FST samples compared to the fresh FST samples. Wang and Wang (1996) also did not find an increase in the frequency of chimeric sequence formation in damaged DNA.

Most of the anomalous sequences in this study were sequences that were wrongly aligned. Upon closer inspection, the sequences were found to share high similarities in the flanking regions. These sequences may have been caused by the insertion of primers or partially synthesized sequences into the clone plasmids instead of the complete 16S rRNA gene sequences. Subsequent amplification of the primers in the rolling circle amplification would have resulted in artifact sequences which were used as templates for the sequencing reactions.

5.3. Identities of Bacteria in the FST

The clone libraries from all nine colonies were combined for the analysis of the bacteria identities. The predominant ribotype in most of the FST samples, Cf2-30, was identical to BCf1-

03 (Shinzato et al., 2005) and CfPt1-2 (Noda et al., 2005). The Bacteroides bacteria is an intracellular endosymbiont of the protist Pseudotrichonympha grassii, a flagellate species found only in the guts of the FST (Noda et al., 2005). Besides this dominant Bacteroides bacteria, the class Bacteroidetes contains other ribotypes and is the most abundant group in this study.

Members of the Bacteroides are commonly found in the intestines of organisms and are known to be involved in the fermentation of carbohydrates and utilization of nitrogenous substances (Hentges, 1989). Bacteroides species have been shown to have many glucosidase activities such as β-1, 4 and -1, 6 xylosidase and glucosidase activities which are induced by the presence of hemicellulose (Reddy et al., 1984), so it is likely that the Bacteroides bacteria in the

FST guts play a role in the digestion of wood.

55

Ribotype Cf4-32 from the Clostridia class was the second most abundant ribotype and was highly similar to BCf9-13 (Shinzato et al., 2005). The Clostridia are known to be anaerobic, endospore-forming bacteria (Cummings and Macfarlane, 1997) and are found in habitats such as soil, intestines, necrotic wounds and corpses (Wells and Wilkins, 1996). Many clostridia degrade polysaccharides to produce acetone, alcohol, acetate, lactate, CO2, and hydrogen (Chen,

1995; Johnston and Goldfine, 1985; Mitchell, 1992; Rainey and Stackebrandt, 1993) and others can ferment nitrogenous or lipid compounds ((Elsden and Hilton, 1979). Acetogenic clostridial species have been isolated from a soil-feeding termite (Kane et al., 1991) but were also one of the major groups in termites based on culture-independent studies (Hongoh et al., 2006; Hongoh et al., 2003; Ohkuma and Kudo, 1996; Shinzato et al., 2005; Shinzato et al., 2007; Yang et al.,

2005).

Ribotype Cf4-07 was the third most abundant group. It identified closely with Pilibacter termitis, a lactic acid bacterium that was successfully cultured from a FST colony from Hawaii

(Higashiguchi et al., 2006) and BCf6-17, which was found in a FST colony from Japan

(Shinzato et al., 2005). are involved in roles such as sugar fermentation and pH regulation in the gut environment (Hutkins and Nannen, 1993). Lactic acid bacteria are typical and numerically significant carbohydrate-utilizing microorganisms in the guts of many wood- and soil-feeding lower and higher termites (Bauer et al., 2000; Tholen et al., 1997). The isolates proved to be aerotolerant and exhibit high potential rates of O2 reduction in the presence of fermentable substrates, which may explain why they are regularly encountered in the intestinal tracts of termites and other insects.

The Spirochaetes was the second most abundant class after the Bacteroidetes and also the most diverse class. Spirochaetes have been commonly found in the guts of different termite

56

species (Lilburn et al., 1999; Noda et al., 2003; Paster et al., 1996). Two spirochaetal strains successfully cultured from the gut of the termite Zootermopsis augusticollis were shown to be

CO2-reducing acetogens. Acetogens have an ability to produce acetate, an energy source for the termites (Leadbetter et al., 1999) under anaerobic conditions. Members of the Spirochaetes can also be nitrogen-fixers, as inferred from the isolation of nitrogenase iron-protein encoding gene

(nifH) from spirochaetes from termite guts and freshwater sediments (Lilburn et al., 2001).

Nitrogen-fixation contributes greatly to the nitrogen supply of the termite, since wood is a nitrogen-poor diet.

Actinobacteria, also known as Actinomycetes, are known to be distributed in both terrestrial and aquatic ecosystems and are commonly found in soil environments (Goodfellow and Williams, 1983). Actinobacteria can degrade plant, animal and microbial polymers in soil and litter and fix nitrogen in non-leguminous plants. Some Actinobacteria are important plant and animal pathogens (Goodfellow and Williams, 1983). Actinobacteria have been frequently found in other termite studies (Hongoh et al., 2006; Hongoh et al., 2003; Ohkuma and Kudo,

1996; Shinzato et al., 2005; Shinzato et al., 2007; Yang et al., 2005), and were dominant in the mound of the soil-feeding termite, Cubitermes niokoloensis (Fall et al., 2007).

Members of the Endomicrobia were found only within the Louisiana FST colonies.

Endomicrobia is a division within the candidate phylum Termite Group 1 (TG1) which contains a novel group of uncultured bacteria found exclusively in gut of lower termites and wood- feeding roaches (Ikeda-Ohtsubo et al., 2007; Stingl et al., 2004). These bacteria are specifically affiliated with the termite flagellates and have been found to be quite abundant in some termites

(Cook and Gold, 1998; Hongoh et al., 2003; Yang et al., 2005). Although their functions in the

57

termite gut are not known, Endomicrobia have been suggested to supply essential nitrogenous compounds to their host protists and the termites (Hongoh et al., 2008).

Few clones were found from the candidate division Synergistes, which have been detected in samples from different environmental sources (Vartoukian et al., 2007). The

Synergistes have often been misclassified within the phyla Firmicutes (Vartoukian et al., 2007), which is further supported by their clustering between the Clostridia and Bacilli. They are known to be anaerobic and have proteolytic activity (Vartoukian et al., 2007).

As in other bacteria community studies on termites (Brauman et al., 2001; Fall et al.,

2007; Hongoh et al., 2005; Hongoh et al., 2006; Hongoh et al., 2003; Minkley et al., 2006;

Ohkuma and Kudo, 1996; Shinzato et al., 2005; Shinzato et al., 2007), the bacteria ribotypes in this study were represented from a diverse range of phyla and functional groups. Each functional group is attributed with specific roles and forms an important component of the bacteria community and is crucial for the survival of the FST colony.

Most of the bacteria clones in this study were specific to the FST (Table 4.3) but a number of the ribotypes were previously found in the termite genus Reticulitermes. The FST and

Reticulitermes are closely related and both belong to the family Rhinotermitidae. Hongoh et al.

(2005) compared the bacteria communities in four species from the genus Reticulitermes

(Rhinotermitidae) to four species from the genus Microtermes (Termitidae) and found that the bacteria communities were specific to each termite genus. These observations suggest that there is an overlap of bacteria species in termite species that are closely related. Along with the support of other studies (Ohkuma et al., 2001), there is strong evidence that co-evolution occurred between the termites and their gut bacteria where the bacteria symbionts were acquired from a common ancestor and diverged independently over time.

58

None of the ribotypes were closely related to any potential termite pathogen. However, we did identify at least two bacteria ribotypes, both closely related to the dominant Bacteroides- related bacteria and Pilibacter termitis that were found n FST colonies from different regions.

These would be useful candidates for paratransgenesis, a proposed termite control method which uses a bacteria vector to carry genes that express toxins harmful to the termite host once the bacteria enters body of the termite host. In addition, some of the major bacteria classes identified, such as Bacteroidetes, Spirochaetes, Clostridia, Bacilli and Actinobacteria or even the two aforementioned bacteria, could be eliminated to disrupt the bacteria composition in the FST and thus, decrease the fitness or lead to the death of the termite colony.

5.4. Alcohol Storage Effects on the Bacteria Communities in the Louisiana FST Samples

In this study, the yield of amplification products of 16S rRNA genes was greatly reduced in the alcohol-stored FST samples compared to the fresh FST samples due to DNA degradation.

Similarly, Harry et al. (1998) found that DNA of bacteria in two types of soil stored up to a whole year in 95% ethanol at 4oC had DNA degradation of less than 5% and 50%. While DNA degradation was expected, the differential degradation of DNA has not been investigated before in gut endosymbionts of long-term stored insects.

Higher proportions of gram-positive bacteria (Clostridia, Bacilli and Actinobacteria) were found in the alcohol-stored FST samples compared to the fresh FST samples while the gram- negative bacteria (Bacteroidetes and Spirochaetes) proportions were lower. The proportion of gram-negative bacteria in our FST samples decreased by approximately three-fold when stored in alcohol for 3-5 years. These differences in bacteria composition between the fresh and alcohol-stored FST samples were not due to sampling variation in bacteria communities of the

FSTs from different locations because the differences were apparent even when the fresh City

59

Park FST sample and the same City Park FST sample stored for five years in alcohol were compared (Colonies 3 & 4 in Figure 4.6). The abundantly represented groups in the alcohol- stored FST samples were the same groups found in bacteria community studies of other long- term preserved specimens. In a study of ancient bacteria sampled in permafrost from different ages (0-8.1 MYA), DNA of Actinobacteria were found to be the most persistent, followed by the

DNA of Bacillalaceae and Clostridiaceae (Willerslev et al., 2004). The proportion of gram- positive bacteria increased approximately 50% in the permafrost samples 5-30 kyr of age compared to fresh permafrost samples and the whole bacteria community in the 300-600 kyr permafrost samples consisted of gram positive bacteria (Willerslev et al., 2004). The intestines of living humans were found to contain mainly of Clostridia, followed by the Bacteroides bacteria. In comparison, the intestines of freeze-dried human mummies were dominated by the presence of Clostridia bacteria (Rollo et al., 2007). Gram-negative bacteria, which made up

~40% of the bacteria community in the intestine of living humans, dropped to ~20% in the intestine of a 90-years-old mummy and were almost absent in a 3500-years-old Iceman mummy

(Rollo et al., 2007). This was comparable to our study, where the proportion of Bacteroidetes and other gram-negative bacteria decreased substantially in the alcohol-stored FST samples compared to the fresh FST samples. While the Bacteroidetes did not preserve well in long-term alcohol storage, their persistence in the alcohol samples could be attributed to their high numbers in the fresh FST samples.

It is not known why the gram-negative bacteria would be less resistant to DNA degradation compared to the gram-positive bacteria. The multiple layers of peptidoglycan in the cell wall of the gram-positive bacteria (Shockman and Barrett, 1983) may protect the DNA much

60

better from the environmental conditions than the thin peptidoglycan and lipopolysaccharide layers in the cell wall of the gram-negative bacteria (Salton and Kim, 1996).

Further evidence of differential degradation of bacteria DNA in the FST samples could be inferred from the higher bacteria ribotype diversity in the long-term alcohol-stored FST samples compared to the fresh samples. In addition, the majority of the bacteria ribotypes found after long-term alcohol storage were different from fresh samples. This finding was in contrast to previous findings. Husseneder et. al (2005) found no significant changes in the bacteria composition between the fresh FST samples and FST samples stored in 95% ethanol for a week.

In a study on the effects of some preservatives on the bacteria community in the gut of the termite Microcerotermes sp., storing the termites in absolute ethanol at room temperature for 30 days decreased the yield of extracted DNA from the termite guts by about 50% (Deevong et al.,

2006). However, amplification of the bacterial 16S rRNA genes did not lead to a difference in the yield of PCR products compared to those from fresh termite samples and there was also no significant difference in the bacteria composition. The two studies did not find any differential degradation of bacteria DNA because the termite samples used were stored only for the short- term. The bacteria communities in the alcohol-stored FST samples represent mainly the remaining bacteria after differential DNA degradation. The alcohol-stored samples revealed the less abundant bacteria ribotypes that would have been masked by the dominance of the

Bacteroides bacteria and do not reflect the remaining bacteria community of the FST in their natural environment.

Differences in the bacteria class proportions, bacteria cell wall type, bacteria diversity and the ribotypes observed in the bacteria composition between the fresh and alcohol-stored FST samples were not likely due to the presence of bacteria on the external surface of the alcohol-

61

stored FST worker abdomens or bacteria presence in the fat bodies of the FST workers. No bacteria were visible on the external surface of the alcohol-stored FST worker abdomens when visualized using scanning electron microscopy (SEM), which was a contrast to the high number of bacteria seen in the termite guts. Bacteriocytes, structures known to harbor bacteria within the fat body of certain insects, were lost in all termite lineages except for the species Mastotermes darwiniensis (Sacchi et al., 2000). Even though non-bacteriocyte intracellular bacteria were found in the fat body of eight termite species representing three termite families, they were only related to the Wolbachia, which belonged to the Alphaproteobacteria class (Bandi et al., 1997).

None of the bacteria ribotypes identified in this study belonged to the Alphaproteobacteria class.

Furthermore, the non-bacteriocyte bacteria found in the fat bodies of termites were scarce (Bandi et al., 1997). In comparison to the 106-107 cells per µl (Schultz and Breznak, 1978) found within the guts of termites, the bacteria on the body surface and in the fat body were negligible.

5.5. Comparison of Bacteria Community in the Alcohol-stored FST Samples from the Native and Introduced Ranges

Bacteria ribotype richness, diversity and colony similarity values from the pairwise comparisons were not significantly different between the alcohol-stored Louisiana and China

FST colonies and between the fresh Louisiana and Japan FST samples (Shinzato et al., 2005).

This observation was contrary to our expectation that the bacteria composition would be different in FST colonies separated by a huge geographical distance, especially between the introduced and native ranges. The similar ribotype richness and diversity across the FST colonies from the different geographic regions is an indication of the minimum number of bacteria species representing the major functional groups needed to maintain the gut ecology of the FST. The FST gut bacteria and their FST host have a co-dependent relationship (Brune and

Friedrich, 2000; Ohkuma et al., 2001). The termite gut provides the bacteria with a well 62

regulated environment where food is readily available. In exchange, the gut bacteria provide nutrition and protect their termite host against the invasion of pathogens (Dillon and Dillon,

2004).

The pH conditions that are maintained in the guts of the FST could also contribute to the similarity in ribotype richness and diversity in the FSTs of different regions. In a comparison of soil samples across North and South America to samples from other geographic regions, the bacteria diversity was found to be most affected by the soil pH at the continental scale, rather than to factors such as the mean annual temperature and latitudinal positions (Fierer and Jackson,

2006).

Even though the ribotype richness and diversity among the FST colonies were similar, most of the colonies within the same geographic regions and from different geographic regions had less than 70% similarity to one another when comparing bacteria ribotypes and abundances among colonies. The high variances among the FST samples was further reflected in the proportion of unique ribotypes among the FST samples, where approximately one quarter of the ribotypes found in each FST colony were unique (Table 4.5). In a study of the bacteria community in the gut of the lower termite Hodotermes mossambicus, differences were also found in the bacteria composition of the termites from different colonies of the same geographical region (Minkley et al., 2006).

The variances that existed among the bacteria communities of the sampled FST colonies could have been due to partial sampling. Since we might have captured less than 50% of the total bacteria community based on the lower boundaries of the ribotype richness estimators

(Table 4.6), there was a high probability that differences would be detected among the FST samples due to the high numbers of singletons and doubletons in each FST colony. However,

63

the variances among the FST samples remained high even when the single- and doubletons were removed, indicating that even the dominant bacteria groups were different. As such, it is not likely that partial sampling greatly influenced the differences in bacteria composition among the

FST colonies.

Variances in the bacteria community of the FST colonies can be attributed to two main biological factors. Firstly, at the incipience of each FST colony, the bacteria community of each colony is different from the other FST colonies because the first generation of termite workers receives its first inoculation of bacteria from the founding alate pair. Since the alates come from different environments and are capable of flying nearly one kilometer (Husseneder et al., 2006;

Messenger and Mullins, 2005) and much longer distances if aided by human transports (Su and

Tamashiro, 1987), it is likely that they carry different bacteria compositions and subsequently transfer a portion of their gut bacteria communities to their offspring. Contribution of bacteria communities from both the founding alates would give rise to a bacteria community that is different from those of the colony where the reproductives originated from.

Secondly, differences in the diet component of each FST colony would influence the composition of the bacteria community. Bacteria communities in the FST were shown to differ by 60% when comparing FST colonies fed with high molecular weight carbon sources and those fed with low molecular weight carbon sources (Tanaka et al., 2006). Within the gut of the wood- feeding higher termite Nasutitermes takasagoensis, Spirochaetes were predominant in the wood- and wood powder-fed termites, Bacteroidetes were predominant in the xylan-, cellobiose- and glucose-fed termites while Firmicutes was predominant in the xylose-fed termites (Miyata et al.,

2007). Although termites do not have such large diet variances in their native environment, it is likely that the different groups of bacteria are influenced by the variances in composition of the

64

wood diet. The bacteria composition is determined by the presence of bacteria that can utilize the materials in the diet of the FST. It is likely that the numbers of environmental bacteria are negligible in contributing to the variances in the bacteria communities of the FST. In this study, only two clones were highly similar to soil bacteria, and most of the bacteria ribotypes identified were specific to termites. Furthermore, studies using different species of soil-feeding termites found that the bacteria communities in the termite guts to be different from the mound of the termites (Fall et al., 2007; Schmitt-Wagner et al., 2003).

There was no detectable influence of the geographic regions of the FST colonies on the bacteria composition. A study on the bacteria communities in three soil-feeding termite species from the genus Cubitermes also did not find geographic region to be a good predictor of the bacteria composition (Schmitt-Wagner et al., 2003). An explanation for not seeing any geographic region effect is that the multiple termite introductions, likely originating from China and into the introduced range of the FST have made the bacteria compositions homogenous across the different regions. The introduction of the FST into the U.S. has been shown to occur a few times (Austin et al., 2006; Vargo et al., 2006; Wang and Grace, 2000). However, since the

FST colonies sampled in this study represented only a relatively limited area from each geographic region, future studies using T-RFLP to obtain the bacteria community fingerprints from more than three FST colonies randomized across each geographical location would be useful in confirming the observations from this study.

65

CHAPTER 6. CONCLUSION

The use of ARDRA with restriction enzymes selected based on in-silico analysis had a resolution power that was comparable to direct sequencing of the 16S rRNA gene sequences.

However, the applicability of the method also depended on the costs involved. Compared to automated high-throughput sequencing, ARDRA would only be cost effective if unique sequences comprised less than 25% of the bacteria richness, assuming that sequencing costs are four times higher than the costs of ARDRA per sample.

It was important to use different methods to screen out the artifact sequences because none of the standalone methods were able to detect all the artifact sequences. When studying a bacteria community containing many representatives with highly similar sequences, Mallard was found to be the most reliable program among the other commonly used nearest-neighbor methods such as ChimeraCheck and Bellerophon. Mallard was able to detect most of the artifact sequences while flagging the least number of false positives. Nearest-neighbor methods need to be confirmed using other methods such as signature, secondary structure and "breaking and retreeing" analyses.

There were no potential pathogens found in the FST samples. Two of the bacteria ribotypes identified were closely related to the dominant Bacteroidetes bacteria and Pilibacter termitis, which were found in most of the FST samples. These ribotypes would serve as suitable candidates for paratransgenesis, an alternative biological control method for the FSTs.

Long-term alcohol storage of FST samples (3-5 years) resulted in differential bacteria

DNA degradation. The proportions of Bacteroidetes and Spirochaetes, gram-negative bacteria groups, were lower in the fresh FST samples. Degradation of gram negative bacteria DNA led to a higher proportion of the gram-positive bacteria such as Clostridia, Bacilli and Actinobacteria in

66

alcohol FST samples compared to the fresh FST samples. Therefore, caution has to be exercised when studying bacteria community samples that have been stored over 30 days. Although gradual degradation of DNA is expected to occur, bacteria community studies need to consider the higher degree of DNA degradation that occur in gram-negative bacteria over time, especially when comparing studies that used different storage methods.

Comparison of bacteria composition in the alcohol-stored FST samples from the native

(China) and introduced range (Louisiana, U.S.) of the FST showed that different geographical regions had no detectable influence on the bacteria composition. This observation was likely due to the multiple introduction of the FST into the US, likely originating from China, facilitating the high variances among the FST colonies even within the same geographical region. Considering the FST colonies were sampled only from a relatively limited area from each geographic region, performing studies on FST colonies from multiple randomized locations across each geographic region would be needed to increase the statistical power to detect subtle differences in the FST colonies.

67

REFERENCES

Ashelford, K. E., Chuzhanova, N. A., Fry, J. C., Jones, A. J. and Weightman, A. J. (2005). At least 1 in 20 16S rRNA sequence records currently held in public repositories is estimated to contain substantial anomalies. Applied and Environmental Microbiology 71, 7724- 7736.

Ashelford, K. E., Chuzhanova, N. A., Fry, J. C., Jones, A. J. and Weightman, A. J. (2006). New screening software shows that most recent large 16S rRNA gene clone libraries contain chimeras. Applied and Environmental Microbiology 72, 5734-5741.

Austin, J. W., Szalanski, A. L., Scheffrahn, R. H., Messenger, M. T., Mckern, J. A. and Gold, R. E. (2006). Genetic evidence for two introductions of the Formosan Subterranean Termite, Coptotermes formosanus (Isoptera : Rhinotermitidae), to the United States. Florida Entomologist 89, 183-193.

Bandi, C., Sironi, M., Nalepa, C. A., Corona, S. and Sacchi, L. (1997). Phylogenetically distant intracellular symbionts in termites. Parassitologia 39, 71-75.

Barrett, C. F. and Parker, M. A. (2005). Prevalence of Burkholderia sp. nodule symbionts on four mimosoid legumes from Barro Colorado Island, Panama. Systematic and Applied Microbiology 28, 57-65.

Bauer, S., Tholen, A., Overmann, J. and Brune, A. (2000). Characterization of abundance and diversity of lactic acid bacteria in the hindgut of wood- and soil-feeding termites by molecular and culture-dependent techniques. Archives of Microbiology 173, 126-137.

Bignell, D. E. (2000). Introduction to Symbiosis. In Termites: Evolution, Sociality, Symbioses, Ecology, (eds T. Abe D. E. Bignell and M. Higashi), pp. 189-208. Netherlands: Kluwer Academic Publishers.

Bignell, D. E. and Anderson, J. M. (1980). Determination of Ph and Oxygen Status in the Guts of Lower and Higher Termites. Journal of Insect Physiology 26, 183-188.

Brauman, A., Dore, J., Eggleton, P., Bignell, D., Breznak, J. A. and Kane, M. D. (2001). Molecular phylogenetic profiling of prokaryotic communities in guts of termites with different feeding habits. Fems Microbiology Ecology 35, 27-36.

Bray, J. R. and Curtis, J. T. (1957). An Ordination of the Upland Forest Communities of Southern Wisconsin. Ecological Monographs 27, 326-349.

Breznak, J. A. (2000). Ecology of Prokaryotic Microbes in the Guts of Wood- and Litter-Feeding Termites. In Termites: Evolution, Sociality, Symbioses, Ecology, (eds T. Abe D. E. Bignell and M. Higashi), pp. 209-231. Netherlands. Kluwer Academic Publishers.

68

Breznak, J. A. and Brune, A. (1994). Role of Microorganisms in the Digestion of Lignocellulose by Termites. Annual Review of Entomology 39, 453-487.

Breznak, J. A. and Pankratz, H. S. (1977). In situ Morphology of Gut Microbiota of Wood-Eating Termites [Reticulitermes flavipes (Kollar) and Coptotermes formosanus Shiraki]. Applied and Environmental Microbiology 33, 406-426.

Brugerolle, G. and Radek, R. (2006). Symbiotic Protozoa of Termites. In Intestinal microorganisms of termites and other invertebrates (eds H. König and A. Varma), pp. 243-269. Berlin ; New York: Springer.

Brune, A. (1998). Termite guts: the world's smallest bioreactors. Trends in Biotechnology 16, 16-21.

Brune, A., Emerson, D. and Breznak, J. A. (1995). The Termite Gut Microflora as an Oxygen Sink - Microelectrode Determination of Oxygen and Ph Gradients in Guts of Lower and Higher Termites. Applied and Environmental Microbiology 61, 2681-2687.

Brune, A. and Friedrich, M. (2000). Microecology of the termite gut: structure and function on a microscale. Current Opinion in Microbiology 3, 263-269.

Chambers, D. M., Zungoli, P. A. and Hill, H. S. (1988). Distribution and Habitats of the Formosan Subterranean Termite (Isoptera, Rhinotermitidae) in South-Carolina. Journal of Economic Entomology 81, 1611-1619.

Chao, A., Chazdon, R. L., Colwell, R. K. and Shen, T. J. (2005). A new statistical approach for assessing similarity of species composition with incidence and abundance data. Ecology Letters 8, 148-159.

Chao, A., Hwang, W. H., Chen, Y. C. and Kuo, C. Y. (2000). Estimating the number of shared species in two communities. Statistica Sinica 10, 227-246.

Chao, A., Yip, P. S. F., Lee, S. M. and Chu, W. T. (2001). Population size estimation based on estimating functions for closed capture-recapture models. Journal of Statistical Planning and Inference 92, 213-232.

Chen, J. S. (1995). Alcohol-Dehydrogenase - Multiplicity and Relatedness in the Solvent-Producing Clostridia. Fems Microbiology Reviews 17, 263-273.

Chou, J. H., Chen, W. M., Arun, A. B. and Young, C. C. (2007). Trabulsiella odontotermitis sp. nov, isolated from the gut of the termite Odontotermes formosanus Shiraki. International Journal of Systematic and Evolutionary Microbiology 57, 696-700.

Cole, J. R., Chai, B., Farris, R. J., Wang, Q., Kulam-Syed-Mohideen, A. S., McGarrell, D. M., Bandela, A. M., Cardenas, E., Garrity, G. M. and Tiedje, J. M. (2007). The ribosomal database project (RDP-II): introducing myRDP space and quality controlled public data. Nucleic Acids Research 35, D169-D172.

69

Cole, J. R., Chai, B., Marsh, T. L., Farris, R. J., Wang, Q., Kulam, S. A., Chandra, S., McGarrell, D. M., Schmidt, T. M., Garrity, G. M. et al. (2003). The Ribosomal Database Project (RDP-II): previewing a new autoaligner that allows regular updates and the new prokaryotic . Nucleic Acids Research 31, 442-443.

Colwell, R. K. (2005). EstimateS: Statistical estimation of species richness and shared species from samples.

Colwell, R. K. and Coddington, J. A. (1994). Estimating Terrestrial Biodiversity through Extrapolation. Philosophical Transactions of the Royal Society of London Series B- Biological Sciences 345, 101-118.

Connick, W. J., Osbrink, W. L. A., Wright, M. S., Williams, K. S., Daigle, D. J., Boykin, D. L. and Lax, A. R. (2001). Increased mortality of Coptotermes formosanus (Isoptera : Rhinotermitidae) exposed to eicosanoid biosynthesis inhibitors and Serratia marcescens (Eubacteriales : Enterobacteriaceae). Environmental Entomology 30, 449-455.

Cook, T. J. and Gold, R. E. (1998). Organization of the symbiotic flagellate community in three castes of the eastern subterranean termite, Reticulitermes flavipes (Isoptera : Rhinotermitidae). Sociobiology 31, 25-39.

Cummings, J. H. and Macfarlane, G. T. (1997). Role of intestinal bacteria in nutrient metabolism. Clinical Nutrition 16, 3-11.

De Baere, T., de Mendonca, R., Claeys, G., Verschraegen, G., Mijs, W., Verhelst, R., Rottiers, S., Van Simaey, L., De Ganck, C. and Vaneechoutte, M. (2002). Evaluation of amplified rDNA restriction analysis (ARDRA) for the identification of cultured mycobacteria in a diagnostic laboratory. BMC Microbiol 2, 4.

Deevong, P., Hongoh, Y., Inoue, T., Trakulnaleamsai, S., Kudo, T., Noparatnaraporn, N. and Ohkuma, M. (2006). Effect of Temporal Sample Preservation on the Molecular Study of a COmplex Microbial Community in the Gut of the Termite Microcerotermes sp. Microbes and Environments 21, 78-85.

Deitz, L. L., Nalepa, C. and Klass, K.-D. (2003). Phylogeny of the dictyoptera re- examined (Insecta). Entomologische Abhandlungen (Dresden) 61, 69-91.

DeSantis, T. Z., Hugenholtz, P., Keller, K., Brodie, E. L., Larsen, N., Piceno, Y. M., Phan, R. and Andersen, G. L. (2006a). NAST: a multiple sequence alignment server for comparative analysis of 16S rRNA genes. Nucleic Acids Research 34, W394-W399.

DeSantis, T. Z., Hugenholtz, P., Larsen, N., Rojas, M., Brodie, E. L., Keller, K., Huber, T., Dalevi, D., Hu, P. and Andersen, G. L. (2006b). Greengenes, a chimera-checked 16S rRNA gene database and workbench compatible with ARB. Applied and Environmental Microbiology 72, 5069-5072.

70

Dillon, R. J. and Dillon, V. M. (2004). The gut bacteria of insects: Nonpathogenic interactions. Annual Review of Entomology 49, 71-92.

Eggleton, P. (2001). Termites and trees: a review of recent advances in termite phylogenetics. Insectes Sociaux 48, 187-193.

Elsden, S. R. and Hilton, M. G. (1979). Amino-Acid Utilization Patterns in Clostridial Taxonomy. Archives of Microbiology 123, 137-141.

Fall, S., Hamelin, J., Ndiaye, F., Assigbetse, K., Aragno, M., Chotte, J. L. and Brauman, A. (2007). Differences between bacterial communities in the gut of a soil-feeding termite (Cubitermes niokoloensis) and its mounds. Applied and Environmental Microbiology 73, 5199-5208.

Fierer, N. and Jackson, R. B. (2006). The diversity and biogeography of soil bacterial communities. Proceedings of the National Academy of Sciences of the United States of America 103, 626-631.

Fisher, M. L. (2006). Comparison of subterranean termite (Rhinotermitidae: Reticulitermes) gut bacterial diversity within and between colonies and to other termite species using molecular techniques (ARDRA and 16S rRNA gene sequencing.

Flournoy, L. E., Adams, R. P. and Pandy, R. N. (1996). Interim and archival preservation of plant specimens in alcohols for DNA studies. Biotechniques 20, 657-660.

Fujii, J. K. (1975). Effects of an entomogenous nematode, Neoaplectana carpocapsae Weiser, on the Formosan subterranean termite, Coptotermes formosanus Shiraki, with ecological and biological studies on C. formosanus. Honolulu: University of Hawaii.

Goodfellow, M. and Williams, S. T. (1983). Ecology of Actinomycetes. Annual Review of Microbiology 37, 189-216.

Grace, J. K. (1990). Oral Toxicity of Barium Metaborate to the Eastern Subterranean Termite (Isoptera, Rhinotermitidae). Journal of Entomological Science 25, 112-116.

Harazono, K., Yamashita, N., Shinzato, N., Watanabe, Y., Fukatsu, T. and Kurane, R. (2003). Isolation and characterization of aromatics-degrading microorganisms from the gut of the lower termite Coptotermes formosanus. Bioscience Biotechnology and Biochemistry 67, 889- 892.

Harry, M., Gambier, B. and Garnier-Sillam, E. (2000). Soil conservation for DNA preservation for bacterial molecular studies. European Journal of Soil Biology 36, 51-55.

Heltshe, J. F. and Forrester, N. E. (1983). Estimating Species Richness Using the Jackknife Procedure. Biometrics 39, 1-11.

71

Hentges, D. J. (1989). Anaerobes as normal flora. In Anaerobic infections in humans, (eds S. M. Finegold and W. L. George), pp. 37-53. San Diego: Academic Press.

Heylen, K., Vanparys, B., Wittebolle, L., Verstraete, W., Boon, N. and De Vos, P. (2006). Cultivation of denitrifying bacteria: Optimization of isolation conditions and diversity study. Applied and Environmental Microbiology 72, 2637-2643.

Heyndrickx, M., Vauterin, L., Vandamme, P., Kersters, K. and DeVos, P. (1996). Applicability of combined amplified ribosomal DNA restriction analysis (ARDRA) patterns in bacterial phylogeny and taxonomy. Journal of Microbiological Methods 26, 247-259.

Higashiguchi, D. T., Husseneder, C., Grace, J. K. and Berestecky, J. M. (2006). Pilibacter termitis gen. nov., sp nov., a lactic acid bacterium from the hindgut of the Formosan subterranean termite (Coptotermes formosanus). International Journal of Systematic and Evolutionary Microbiology 56, 15-20.

Hongoh, Y., Deevong, P., Inoue, T., Moriya, S., Trakulnaleamsai, S., Ohkuma, M., Vongkaluang, C., Noparatnaraporn, N. and Kudol, T. (2005). Intra- and interspecific comparisons of bacterial diversity and community structure support coevolution of gut microbiota and termite host. Applied and Environmental Microbiology 71, 6590-6599.

Hongoh, Y., Ekpornprasit, L., Inoue, T., Moriya, S., Trakulnaleamsai, S., Ohkuma, M., Noparatnaraporn, N. and Kudo, T. (2006). Intracolony variation of bacterial gut microbiota among castes and ages in the fungus-growing termite Macrotermes gilvus. Molecular Ecology 15, 505-516.

Hongoh, Y., Ohkuma, M. and Kudo, T. (2003). Molecular analysis of bacterial microbiota in the gut of the termite Reticulitermes speratus (Isoptera; Rhinotermitidae). Fems Microbiology Ecology 44, 231-242.

Hongoh, Y., Sharma, V. K., Prakash, T., Noda, S., Taylor, T. D., Kudo, T., Sakaki, Y., Toyoda, A., Hattori, M. and Ohkuma, M. (2008). Complete genome of the uncultured Termite Group 1 bacteria in a single host protist cell. Proceedings of the National Academy of Sciences of the United States of America 105, 5555-5560.

Horn, H. S. (1966). Measurement of Overlap in Comparative Ecological Studies. American Naturalist 100, 419-&.

Hrdy, I., Krecek, J. and Zuskova, Z. (1979). Juvenile hormone analogues: effects on the soldier caste differentiation in termites (Isoptera). Vestnik Ceskoslovenske Spolecnosti Zoologicke 43, 260-9.

Huber, T., Faulkner, G. and Hugenholtz, P. (2004). Bellerophon: a program to detect chimeric sequences in multiple sequence alignments. Bioinformatics 20, 2317-2319.

Hugenholtz, P. and Huber, T. (2003). Chimeric 16S rDNA sequences of diverse origin are accumulating in the public databases. Int J Syst Evol Microbiol 53, 289-293.

72

Husseneder, C. and Collier, R. E. (In press). Paratransgenesis in termites. In Insect Symbiosis, (eds K. Bourtzis and T. A. Miller). Boca Raton, Florida: CRC Press LLC.

Husseneder, C. and Grace, J. K. (2005). Genetically engineered termite gut bacteria ( Enterobactercloacae ) deliver and spread foreign genes in termite colonies. Applied Genetics and Molecular Biotechnology 68, 360-367.

Husseneder, C., Simms, D. M. and Ring, D. R. (2006). Genetic diversity and genotypic differentiation between the sexes in swarm aggregations decrease inbreeding in the Formosan subterranean termite. Insectes Sociaux 53, 212-219.

Husseneder, C., Wise, B. R. and Higashiguchi, D. T. (2005). Microbial diversity in the termite gut: A complimentary approach combining culture and culture-independent techniques.

Hutkins, R. W. and Nannen, N. L. (1993). Ph Homeostasis in Lactic-Acid Bacteria. Journal of Dairy Science 76, 2354-2365.

Hyman, R. W., Fukushima, M., Diamond, L., Kumm, J., Giudice, L. C. and Davis, R. W. (2005). Microbes on the human vaginal epithelium. Proceedings of the National Academy of Sciences of the United States of America 102, 7952-7957.

Ikeda-Ohtsubo, W., Desai, M., Stinglt, U. and Brune, A. (2007). Phylogenetic diversity of 'Endomicrobia' and their specific affiliation with termite gut flagellates. Microbiology-Sgm 153, 3458-3465.

Innis, M. A., Gelfand, D. H., Sninsky, J. J. and White, T. J. (1990). PCR Protocols: A Guide to Methods and Applications. San Diego: Academic Press, Inc.

Inward, D., Beccaloni, G. and Eggleton, P. (2007). Death of an order: a comprehensive molecular phylogenetic study confirms that termites are eusocial cockroaches. Biology Letters 3, 331-335.

Jansen, R. K., Cai, Z., Raubeson, L. A., Daniell, H., Depamphilis, C. W., Leebens- Mack, J., Muller, K. F., Guisinger-Bellian, M., Haberle, R. C., Hansen, A. K. et al. (2007). Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proceedings of the National Academy of Sciences of the United States of America 104, 19369-19374.

Johnston, N. C. and Goldfine, H. (1985). Phospholipid Aliphatic Chain Composition Modulates Lipid Class Composition, but Not Lipid Asymmetry in Clostridium butyricum. Biochimica Et Biophysica Acta 813, 10-18.

Jones, R. T., McCormick, K. F. and Martin, A. P. (2008). Bacterial communities of Bartonella-Positive fleas: Diversity and community assembly patterns. Applied and Environmental Microbiology 74, 1667-1670.

73

Jones, S. C. (1989). Field-Evaluation of Fenoxycarb as a Bait Toxicant for Subterranean Termite Control. Sociobiology 15, 33-41.

Kane, M. D., Brauman, A. and Breznak, J. A. (1991). Clostridium mayombei sp. nov, an H2/CO2 Acetogenic Bacterium from the Gut of the African Soil-Feeding Termite, Cubitermes speciosus. Archives of Microbiology 156, 99-104.

Kilpatrick, C. W. (2002). Noncryogenic preservation of mammalian tissues for DNA extraction: An assessment of storage methods. Biochemical Genetics 40, 53-62.

King, J. R. and Porter, S. D. (2004). Recommendations on the use of alcohols for preservation of ant specimens (Hymenoptera, Formicidae). Insectes Sociaux 51, 197-202.

Kistner, D. H. (1985). A New Genus and Species of Termitophilous Aleocharinae from Mainland China Associated with Coptotermes formosanus and Its Zoogeographic Significance (Coleoptera, Staphylinidae). Sociobiology 10, 93-104.

König, H., Frohlich, J. and Hertel, H. (2006). Diversity and Lignocellulolytic Activities of Cultured Microorganisms. In Intestinal microorganisms of termites and other invertebrates (eds H. König and A. Varma), pp. 271-301. Berlin ; New York: Springer.

Kuhnigk, T., Branke, J., Krekeler, D., Cypionka, H. and Konig, H. (1996). A feasible role of sulfate-reducing bacteria in the termite gut. Systematic and Applied Microbiology 19, 139-149.

Kumari, R., Sachdev, M., Prasad, R., Garg, A. P., Sharma, S., Giang, P. H. and Varma, A. (2006). Microbiology of Termite Hill (Mound) and Soil. In Intestinal microorganisms of termites and other invertebrates, (eds H. König and A. Varma), pp. 351-372. Berlin ; New York: Springer.

La Fage, J. P. (1987). Practical consideration of the Formosan subterranean termite in Louisiana: a 30-year-old problem

Lagace, L., Pitre, M., Jacques, M. and Roy, D. (2004). Identification of the bacterial community of maple sap by using amplified ribosomal DNA (rDNA) restriction analysis and rDNA Sequencing. Applied and Environmental Microbiology 70, 2052-2060.

Lai, P. Y., Tamashiro, M. and Fujii, J. K. (1982). Pathogenicity of 6 Strains of Entomogenous Fungi to Coptotermes formosanus (Isoptera, Rhinotermitidae). Journal of Invertebrate Pathology 39, 1-5.

Leadbetter, J. R., Schmidt, T. M., Graber, J. R. and Breznak, J. A. (1999). Acetogenesis from H2 plus CO2 by spirochetes from termite guts. Science 283, 686-689.

Li, C., Yuan, H. L. and Huang, H. Z. (In press). Vertical distribution characteristic of the P-decomposing bacteria in the sediment of Guanting Reservoir. College of Biological Science, China Agricultural University, 2 Yuanmingyuan Xi Lu, Haidian, Beijing 100094, China.

74

Li, L., Frohlich, J. and Konig, H. (2006). Cellulose Digestion in the Termite Gut. In Intestinal microorganisms of termites and other invertebrates (eds H. König and A. Varma), pp. 221-237. Berlin ; New York: Springer.

Liesack, W., Weyland, H. and Stackebrandt, E. (1991). Potential Risks of Gene Amplification by PCR as Determined by 16S rDNA Analysis of a Mixed-Culture of Strict Barophilic Bacteria. Microbial Ecology 21, 191-198.

Lilburn, T. C., Kim, K. S., Ostrom, N. E., Byzek, K. R., Leadbetter, J. R. and Breznak, J. A. (2001). Nitrogen fixation by symbiotic and free-living spirochetes. Science 292, 2495-2498.

Lilburn, T. G., Schmidt, T. M. and Breznak, J. A. (1999). Phylogenetic diversity of termite gut spirochaetes. Environmental Microbiology 1, 331-345.

Mandrioli, M., Borsatti, F. and Mola, L. (2006). Factors affecting DNA preservation from museum-collected Lepidopteran specimens. Entomologia Experimentalis Et Applicata 120, 239-244.

Mao, C. X., Colwell, R. K. and Chang, J. (2005). Estimating the species accumulation curve using mixtures. Biometrics 61, 433-441.

Martens, J. H., Barg, H., Warren, M. J. and Jahn, D. (2002). Microbial production of vitamin B-12. Applied Microbiology and Biotechnology 58, 275-285.

Matsuura, K. (2001). Nestmate recognition mediated by intestinal bacteria in a termite, Reticulitermes speratus. Oikos 92, 20-26.

Mauldin, J. K. and Rich, N. M. (1980). Effect of Chlortetracycline and Other Antibiotics on Protozoan Numbers in the Eastern Subterranean Termite Isoptera, Rhinotermitidae. Journal of Economic Entomology 73, 123-128.

Messenger, M. T. and Mullins, A. J. (2005). New flight distance recorded for Coptotermes formosanus (Isoptera : Rhinotermitidae). Florida Entomologist 88, 99-100.

Meyerhans, A., Vartanian, J. P. and Wainhobson, S. (1990). DNA Recombination during PCR. Nucleic Acids Research 18, 1687-1691.

Minkley, N., Fujita, A., Brune, A. and Kirchner, W. H. (2006). Nest specificity of the bacterial community in termite guts (Hodotermes mossambicus). Insectes Sociaux 53, 339-344.

Mitchell, W. J. (1992). Carbohydrate Assimilation by Saccharolytic Clostridia. Research in Microbiology 143, 245-250.

Miyata, R., Noda, N., Tamaki, H., Kinjyo, K., Aoyagi, H., Uchiyama, H. and Tanaka, H. (2007). Influence of feed components on symbiotic bacterial community structure in

75

the gut of the wood-feeding higher termite Nasutitermes takasagoensis. Bioscience Biotechnology and Biochemistry 71, 1244-1251.

Nakajima, H., Hongoh, Y., Noda, S., Yoshida, Y., Usami, R., Kudo, T. and Ohkuma, M. (2006). Phylogenetic and morphological diversity of Bacteroidales members associated with the gut wall of termites. Bioscience Biotechnology and Biochemistry 70, 211-218.

Nakajima, H., Hongoh, Y., Usami, R., Kudo, T. and Ohkuma, M. (2005). Spatial distribution of bacterial phylotypes in the gut of the termite Reticulitermes speratus and the bacterial community colonizing the gut epithelium. Fems Microbiology Ecology 54, 247-255.

Nakashima, K., Watanabe, H., Saitoh, H., Tokuda, G. and Azuma, J. I. (2002). Dual cellulose-digesting system of the wood-feeding termite, Coptotermes formosanus Shiraki. Insect Biochemistry and Molecular Biology 32, 777-784.

Nalepa, C. A., Bignell, D. E. and Bandi, C. (2001). Detritivory, coprophagy, and the evolution of digestive mutualisms in Dictyoptera. Insectes Sociaux 48, 194-201.

Nishiwaki, H., Ito, K., Shimomura, M., Nakashima, K. and Matsuda, K. (2007). Insecticidal bacteria isolated from predatory larvae of the antlion species Myrmeleon bore (Neuroptera : Myrmeleontidae). Journal of Invertebrate Pathology 96, 80-88.

Noda, S., Iida, T., Kitade, S., Nakajima, H., Kudo, T. and Ohkuma, M. (2005). Endosymbiotic Bacteroidales bacteria of the flagellated protist Pseudotrichonympha grassii in the gut of the termite Coptotermes formosanus. Applied and Environmental Microbiology 71, 8811-8817.

Noda, S., Ohkuma, M., Yamada, A., Hongoh, Y. and Kudo, T. (2003). Phylogenetic position and in situ identification of ectosymbiotic spirochetes on protists in the termite gut. Applied and Environmental Microbiology 69, 625-633.

Nogales, B., Moore, E. R. B., Llobet-Brossa, E., Rossello-Mora, R., Amann, R. and Timmis, K. N. (2001). Combined use of 16S ribosomal DNA and 16S rRNA to study the bacterial community of polychlorinated biphenyl-polluted soil. Applied and Environmental Microbiology 67, 1874-1884.

Ohkuma, M. and Kudo, T. (1996). Phylogenetic diversity of the intestinal bacterial community in the termite Reticulitermes speratus. Applied and Environmental Microbiology 62, 461-468.

Ohkuma, M., Noda, S., Hongoh, Y. and Kudo, T. (2001). Coevolution of symbiotic systems of termites and their gut microorganisms. RIKEN 41, 2.

Ohkuma, M., Noda, S., Hongoh, Y. and Kudo, T. (2002). Diverse bacteria related to the bacteroides subgroup of the CFB phylum within the gut symbiotic communities of various termites. Bioscience Biotechnology and Biochemistry 66, 78-84.

76

Ohkuma, M., Noda, S. and Kudo, T. (1999). Phylogenetic diversity of nitrogen fixation genes in the symbiotic microbial community in the gut of diverse termites. Applied and Environmental Microbiology 65, 4926-4934.

Ohkuma, M., Ohtoko, K., Iida, T., Tokura, M., Moriya, S., Usami, R., Horikoshi, K. and Kudo, T. (2000). Phylogenetic identification of hypermastigotes, Pseudotrichonympha, Spirotrichonympha, Holomastigotoides, and parabasalian symbionts in the hindgut of termites. Journal of Eukaryotic Microbiology 47, 249-259.

Osbrink, W. L. A., Williams, K. S., Connick, W. J., Wright, M. S. and Lax, A. R. (2001). Virulence of bacteria associated with the Formosan subterranean termite (Isoptera : Rhinotermitidae) in New Orleans, LA. Environmental Entomology 30, 443-448.

Paabo, S., Irwin, D. M. and Wilson, A. C. (1990). DNA Damage Promotes Jumping between Templates during Enzymatic Amplification. Journal of Biological Chemistry 265, 4718- 4721.

Paster, B. J., Dewhirst, F. E., Cooke, S. M., Fussing, V., Poulsen, L. K. and Breznak, J. A. (1996). Phylogeny of not-yet-cultured spirochetes from termite guts. Applied and Environmental Microbiology 62, 347-352.

Paton, R. and Miller, L. R. (1980). Control of Mastotermes darwiniensis Froggatt (Isoptera, Mastotermitidae) with Mirex Baits. Australian Forest Research 10, 249-258.

Perrott, R. C., Miller, D. M. and Mullins, D. E. (2004). Effects of competing food sources on subterranean termite, Reticulitermes spp., (Isoptera : Rhinotermitidae), consumption of hexaflumuron treated baits in laboratory assays. Sociobiology 44, 69-88.

Post, R. J., Flook, P. K. and Millest, A. L. (1993). Methods for the Preservation of Insects for DNA Studies. Biochemical Systematics and Ecology 21, 85-&.

Potrikus, C. J. and Breznak, J. A. (1981). Gut Bacteria Recycle Uric-Acid Nitrogen in Termites - a Strategy for Nutrient Conservation. Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences 78, 4601-4605.

Potter, M. F. (1997). Termites. In Mallis Handbook of Pest Control, (eds S. A. Hedges and D. Moreland), pp. 232-233: Mallis Handbook and Technical Training Company.

Rainey, F. A. and Stackebrandt, E. (1993). 16s rDNA Analysis Reveals Phylogenetic Diversity among the Polysaccharolytic Clostridia. Fems Microbiology Letters 113, 125-128.

Randall, M. and Doody, T. C. (1934). Poison dusts. I. Treatments with poisonous dusts. In Termites and termite control, (ed. C. A. Kofoid), pp. 463-476. Berkeley, CA: University of California Press.

77

Reddy, N. R., Palmer, J. K., Pierson, M. D. and Bothast, R. J. (1984). Intracellular Glycosidases of Human Colon Bacteroides ovatus B4-11. Applied and Environmental Microbiology 48, 890-892.

Reed, D. L. and Hafner, M. S. (2002). Phylogenetic analysis of bacterial communities associated with ectoparasitic chewing lice of pocket gophers: A culture-independent approach. Microbial Ecology 44, 78-93.

Reiss, R. A., Schwert, D. P. and Ashworth, A. C. (1995). Field Preservation of Coleoptera for Molecular-Genetic Analyses. Environmental Entomology 24, 716-719.

Robisoncox, J. F., Bateson, M. M. and Ward, D. M. (1995). Evaluation of nearest- neighbor methods for detection of chimeric small subunit ribosomal RNA sequences. Applied and Environmental Microbiology 61, 1240-1245.

Rollo, F., Luciani, S., Marota, I., Olivieri, C. and Ermini, L. (2007). Persistence and decay of the intestinal microbiota's DNA in glacier mummies from the Alps. Journal of Archaeological Science 34, 1294-1305.

Sacchi, L., Nalepa, C. A., Lenz, M., Bandi, C., Corona, S., Grigolo, A. and Bigliardi, E. (2000). Transovarial transmission of symbiotic bacteria in Mastotermes darwiniensis (Isoptera : mastotermitidae): Ultrastructural aspects and phylogenetic implications. Annals of the Entomological Society of America 93, 1308-1313.

Salton, M. J. R. and Kim, K. S. (1996). Structure. In Baron's Medical Microbiology, (ed. S. Baron). Galveston: The University of Texas Medical Branch.

Schmitt-Wagner, D., Friedrich, M. W., Wagner, B. and Brune, A. (2003). Axial dynamics, stability, and interspecies similarity of bacterial community structure in the highly compartmentalized gut of soil-feeding termites (Cubitermes spp.). Applied and Environmental Microbiology 69, 6018-6024.

Schultz, J. E. and Breznak, J. A. (1978). Heterotrophic Bacteria Present in Hindguts of Wood-Eating Termites [Reticulitermes flavipes (Kollar)]. Applied and Environmental Microbiology 35, 930-936.

Scupham, A. J. (2007). Examination of the microbial ecology of the avian intestine in vivo using bromodeoxyuridine. Environmental Microbiology 9, 1801-1809.

Shannon, C. and Weaver, W. (1949). The mathematical theory of communication. Urbana: University of Illinois Press.

Shinzato, N., Muramatsu, M., Matsui, T. and Watanabe, Y. (2005). Molecular phylogenetic diversity of the bacterial community in the gut of the termite Coptotermes formosanus. Bioscience Biotechnology and Biochemistry 69, 1145-1155.

78

Shinzato, N., Muramatsu, M., Matsui, T. and Watanabe, Y. (2007). Phylogenetic analysis of the gut bacterial microflora of the fungus-growing termite Odontotermes formosanus. Bioscience Biotechnology and Biochemistry 71, 906-915.

Shockman, G. D. and Barrett, J. F. (1983). Structure, Function, and Assembly of Cell- Walls of Gram-Positive Bacteria. Annual Review of Microbiology 37, 501-527.

Simpson, E. H. (1949). Measurement of Diversity. Nature 163, 688-688.

Slaytor, M. (1969). Energy Metabolism. In Biology of Termites, vol. 1 (eds K. Krishna and F. M. Weesner), pp. 307-331. New York

London: Academic Press.

Stackebrandt, E. and Goebel, B. M. (1994). A Place for DNA-DNA Reassociation and 16S ribosomal RNA Sequence Analysis in the Present Species Definition in Bacteriology. International Journal of Systematic Bacteriology 44, 846-849.

Stakenborg, T., Vicca, J., Butaye, P., Maes, D., De Baere, T., Verhelst, R., Peeters, J., de Kruif, A., Haesebrouck, F. and Vaneechoutte, M. (2005). Evaluation of amplified rDNA restriction analysis (ARDRA) for the identification of Mycoplasma species. Bmc Infectious Diseases 5, -.

Stevenson, B. S., Eichorst, S. A., Wertz, J. T., Schmidt, T. M. and Breznak, J. A. (2004). New strategies for cultivation and detection of previously uncultured microbes. Applied and Environmental Microbiology 70, 4748-4755.

Stingl, U., Radek, R., Yang, H. and Brune, A. (2004). The 'Endomicrobia': Cytoplasmic symbionts of termite gut protozoa form a separate phylum of prokaryotes. In Termite gut flagellates and their bacterial symbionts: Phylogenetic analysis and localization in situ, (ed. U. Stingl), pp. 49-63: Konstanz.

Su, N.-Y. and Scheffrahn, R. H. (1998). A review of subterranean termite control practices and prospects for integrated pest management programmes. Integrated Pest Management Reviews 3, 1-13.

Su, N.-Y. and Tamashiro, M. (1987). An Overview of the Formosan Subterranean Termite (Isoptera: Rhinotermitidae) in the World. In Biology and control of the Formosan subterranean termite, pp. 3-14.

Su, N. Y. and Scheffrahn, R. H. (1988). Toxicity and Feeding Deterrency of a Dihaloalkyl Arylsulfone Biocide, a-9248, against the Formosan Subterranean Termite (Isoptera, Rhinotermitidae). Journal of Economic Entomology 81, 850-854.

Su, N. Y. and Scheffrahn, R. H. (1991). Laboratory Evaluation of 2 Slow-Acting Toxicants against Formosan and Eastern Subterranean Termites (Isoptera, Rhinotermitidae). Journal of Economic Entomology 84, 170-175.

79

Su, N. Y. and Scheffrahn, R. H. (1993). Laboratory Evaluation of 2 Chitin Synthesis Inhibitors, Hexaflumuron and Diflubenzuron, as Bait Toxicants against Formosan and Eastern Subterranean Termites (Isoptera, Rhinotermitidae). Journal of Economic Entomology 86, 1453- 1457.

Su, N. Y., Tamashiro, M., Yates, J. R. and Haverty, M. I. (1982). Effect of Behavior on the Evaluation of Insecticides for Prevention of or Remedial Control of the Formosan Subterranean Termite (Isoptera, Rhinotermitidae). Journal of Economic Entomology 75, 188- 193.

Suzuki, M. T. and Giovannoni, S. J. (1996). Bias caused by template annealing in the amplification of mixtures of 16S rRNA genes by PCR. Appl Environ Microbiol 62, 625-30.

Tanaka, H., Aoyagi, H., Shina, S., Dodo, Y., Yoshimura, T., Nakamura, R. and Uchiyama, H. (2006). Influence of the diet components on the symbiotic microorganisms community in hindgut of Coptotermes formosanus Shiraki. Applied and Environmental Microbiology 71, 907-917.

Tholen, A., Schink, B. and Brune, A. (1997). The gut microflora of Reticulitermes flavipes, its relation to oxygen, and evidence for oxygen-dependent acetogenesis by the most abundant Enterococcus sp. Fems Microbiology Ecology 24, 137-149.

Thorne, B. L. (1997). Evolution of eusociality in termites. Annual Review of Ecology and Systematics 28, 27-54.

Tsunoda, K. (2003). Economic importance of Formosan termite and control practices in Japan. Sociobiology 41, 27-37.

Vaneechoutte, M., De Baere, T., Nemec, A., Musilek, M., van der Reijden, T. J. K. and Dijkshoorn, L. (2008). Reclassification of Acinetobacter grimontii Carr et al. 2003 as a later synonym of Acinetobacter junii Bouvet and Grimont 1986. International Journal of Systematic and Evolutionary Microbiology 58, 937-940.

Vargo, E. L., Husseneder, C., Woodson, D., Waldvogel, M. G. and Grace, J. K. (2006). Genetic analysis of colony and population structure of three introduced populations of the Formosan subterranean termite (Isoptera : Rhinotermitidae) in the Continental United States. Environmental Entomology 35, 151-166.

Vartoukian, S. R., Palmer, R. M. and Wade, W. G. (2007). The division "Synergistes". Anaerobe 13, 99-106.

Vink, C. J., Thomas, S. M., Paquin, P., Hayashi, C. Y. and Hedin, M. (2005). The effects of preservatives and temperatures on arachnid DNA. Invertebrate Systematics 19, 99-104.

Wang, G. C. Y. and Wang, Y. (1996). The frequency of chimeric molecules as a consequence of PCR co-amplification of 16S rRNA genes from different bacterial species. Microbiology-Uk 142, 1107-1114.

80

Wang, G. C. Y. and Wang, Y. (1997). Frequency of formation of chimeric molecules is a consequence of PCR coamplification of 16S rRNA genes from mixed bacterial genomes. Applied and Environmental Microbiology 63, 4645-4650.

Wang, J. S. and Grace, J. K. (2000). Genetic relationship of Coptotermes formosanus (Isoptera : Rhinotermitidae) populations from the United States and China. Sociobiology 36, 7- 19.

Wang, Q., Garrity, G. M., Tiedje, J. M. and Cole, J. R. (2007). Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Applied and Environmental Microbiology 73, 5261-5267.

Wells, C. L. and Wilkins, T. D. (1996). Clostridia: Sporeforming Anaerobic Bacilli. In Medical Microbiology, (ed. S. Baron), pp. 1273: University of Texas Medical Branch.

Willerslev, E., Hansen, A. J., Ronn, R., Brand, T. B., Barnes, I., Wiuf, C., Gilichinsky, D., Mitchell, D. and Cooper, A. (2004). Long-term persistence of bacterial DNA. Current Biology 14, R9-R10.

Yamin, M. A. (1979). Flagellates of the Orders Trichomonadida Kirby, Oxymonadida Grasse, and Hypermastigida Grassi and Foa Reported from Lower Termites (Isoptera Families Mastotermitidae, Kalotermitidae, Hodotermitidae, Termopsidae, Rhinotermitidae, and Serritermitidae) and from the Wood-Feeding Roach Cryptocercus (Dictyoptera, Cryptocercidae). Sociobiology 4, 3-119.

Yang, H., Schmitt-Wagner, D., Stingl, U. and Brune, A. (2005). Niche heterogeneity determines bacterial community structure in the termite gut (Reticulitermes santonensis). Environmental Microbiology 7, 916-932.

Zhong, J. H. and Liu, L. L. (2003). Experience with Coptotermes formasanus in China (Isoptera : Rhinotermitidae). Sociobiology 41, 17-26.

81

APPENDIX: LIST OF FULL-LENGTH 16S RIBOSOMAL RNA GENE SEQUENCES (RIBOTYPES), ACCORDING TO DOMINANCE, FOUND IN THE NINE FST CLONE LIBRARIES.

Classification # of Top Match Top Match FST colonies Ribotype Similarity (frequency) Clones Accession (BLAST) LA Fresh LA EtOH China EtOH Number 1 2 3 4 5 6 7 8 9 10

Bacteroidetes Cf2-30 706 AB062769.1 clone:BCf1-03 100% 250 165 118 14 13 56 27 57 6 176

(42.91%) Cf2-32 12 AY571962.1 Bacteroidetes bacterium S1 99% 0 0 1 1 1 3 5 1 0 0

Cf2-10 10 AY571447.1 clone RsaM33 91% 2 0 5 0 1 2 0 0 0 0

Cf2-06 9 AY571449.1 clone RsaHf397 93% 1 2 1 0 1 4 0 0 0 0

Cf2-20 7 AB088942.2 clone:Rs-B45 92% 1 0 1 0 0 3 0 1 1 0

Cf2-21 6 AB088942.2 clone:Rs-B45 92% 0 1 4 0 0 0 1 0 0 0

Cf2-07 5 AY571449.1 clone RsaHf397 91% 0 1 0 0 1 2 0 1 0 0

Cf2-13 5 AB062826.1 clone:BCf5-08 99% 0 0 5 0 0 0 0 0 0 1

Cf2-26 5 AB192208.1 clone: RsTu1-13 92% 2 1 1 1 0 0 0 0 0 0

Cf2-15 4 AB062838.1 clone:BCf7-17 89% 2 0 0 0 0 1 0 1 0 0

Cf2-31 4 AB231051.1 clone:NkW01-014 95% 3 0 0 0 0 0 1 0 0 0

Cf2-37 4 DQ337018.1 clone EV818CFSSAHH218 99% 0 0 0 4 0 0 0 0 0 0

Cf2-02 3 EU252503.1 Dysgonomonas sp. AM15 93% 0 0 0 0 0 0 3 0 0 0

Cf2-19 3 AB088942.2 clone:Rs-B45 92% 0 1 2 0 0 0 0 0 0 0

Cf2-04 2 AY571450.1 clone RsaHf278 89% 0 0 1 0 0 0 0 0 1 0

Cf2-16 2 AB062847.1 clone:BCf9-17 99% 1 1 0 0 0 0 0 0 0 1

Cf2-18 2 AB100460.1 clone: Rs-106 93% 1 1 0 0 0 0 0 0 0 0

Cf2-33 2 AJ319867.2 Dysgonomonas mossii strain CCUG 43457T 96% 0 0 0 0 0 2 0 0 0 0

Cf2-01 1 AB055736.1 Cluster III of the termite bacteroides symbiont 99% 0 0 0 0 0 0 0 1 0 0

Cf2-09 1 AY571428.1 clone RsaHw538 95% 1 0 0 0 0 0 0 0 0 0

Cf2-14 1 AB062835.1 clone:BCf7-02 99% 0 0 1 0 0 0 0 0 0 1

Cf2-17 1 AB100459.1 clone: Rs-096 92% 0 0 1 0 0 0 0 0 0 0

Cf2-22 1 AB088945.1 clone:Rs-F73 94% 1 0 0 0 0 0 0 0 0 0

(Table continued) 82

Cf2-23 1 AB088924.2 clone:Rs-K10 89% 0 0 0 0 0 0 0 1 0 0

Cf2-25 1 AB299528.1 clone: Cc3-055 86% 0 0 1 0 0 0 0 0 0 0

Cf2-28 1 AB198508.1 clone: RsW01-084 92% 1 0 0 0 0 0 0 0 0 0

Cf2-29 1 AB198508.1 clone: RsW01-084 91% 1 0 0 0 0 0 0 0 0 0

Cf2-35 1 AB298723.2 Flavobacteriaceae bacterium WH032 97% 0 0 0 1 0 0 0 0 0 0

Cf2-36 1 EF025213.1 clone B5_B4 99% 0 0 0 0 1 0 0 0 0 0

Cf2-39 1 AY571441.1 clone RsaHw509 94% 0 0 0 1 0 0 0 0 0 0

Cf2-40 1 EF398967.1 clone SJTU_F_09_65 93% 0 0 0 1 0 0 0 0 0 0

Cf2-41 1 AB088938.1 clone:Rs-P65 92% 0 0 0 0 1 0 0 0 0 0

Firmicutes Cf4-32 147 AB062845.1 clone:BCf9-13 99% 3 1 4 7 21 18 43 35 15 1

(30.49%) Cf4-07 125 AY533171.1 Pilibacter termitis strain TI-1 99% 0 0 0 8 15 1 9 36 56 0

Cf4-20 35 AY571396.1 clone RsaHf311 93% 0 0 0 0 4 8 17 4 2 0

Cf4-70 35 AB244437.1 Lactococcus garvieae strain: An1-1 99% 0 0 0 22 5 8 0 0 0 0

Cf4-41 29 AB088952.1 clone:Rs-N86 98% 0 0 0 2 1 0 0 0 26 0

Cf4-05 15 AJ439543.2 Lactovum miscens strain anNAG3 95% 0 0 0 0 0 0 5 10 0 0

Cf4-48 15 AB231039.1 clone:HsW01-031 94% 3 0 1 0 2 6 1 2 0 0

Cf4-03 13 AB198536.1 clone: RsW02-061 95% 0 0 0 0 0 0 12 0 1 0

Cf4-35 11 AB100484.1 clone: Rs-L03 95% 3 1 0 0 2 4 1 0 0 0

Cf4-30 10 AB062842.1 clone:BCf9-05 99% 0 0 0 0 1 1 5 3 0 1

Cf4-14 9 AB088972.1 clone:Rs-H32 92% 0 0 0 0 1 3 0 2 3 0

Cf4-52 9 AB088985.1 clone:Rs-K66 97% 0 2 0 3 1 3 0 0 0 0

Cf4-98 8 AB062774.1 clone:BCf1-20 99% 0 0 0 5 2 1 0 0 0 4

Cf4-29 6 AB062841.1 clone:BCf8-25 99% 0 0 0 0 0 0 3 3 0 0

Cf4-94 5 AB062805.1 clone:BCf10-10 99% 0 0 0 0 0 5 0 0 0 1

Cf4-38 4 AB192177.1 clone: RsTz-92 94% 0 0 0 1 1 0 0 1 1 0

Cf4-57 4 AB088987.2 clone:Rs-E61 96% 0 0 0 0 1 0 2 1 0 0

Cf4-58 4 AB088992.1 clone:Rs-F81 92% 0 1 2 1 0 0 0 0 0 0

Cf4-66 4 DQ394667.1 clone AF17 91% 0 0 1 0 0 1 2 0 0 0

Cf4-81 4 DQ223877.1 Lactococcus sp. RT5 91% 0 0 0 0 2 2 0 0 0 0 (Table continued) 83

Cf4-23 3 AB062775.1 clone:BCf10-04 98% 0 0 0 0 0 0 0 3 0 0

Cf4-39 3 AB088979.1 clone:Rs-M04 96% 0 3 0 0 0 0 0 0 0 0

Cf4-62 3 AB088970.1 clone:Rs-Q64 94% 0 0 0 0 0 0 3 0 0 0

Cf4-64 3 AB198466.1 clone: RsW01-031 98% 0 0 0 0 0 0 0 3 0 0

Cf4-75 3 EF025286.1 clone B5_N17 92% 0 0 0 0 0 3 0 0 0 0

Cf4-91 3 AB089019.1 clone:Rs-J36 96% 0 0 0 1 0 2 0 0 0 0

Cf4-97 3 AB062811.1 clone:BCf2-19 98% 0 0 0 0 0 3 0 0 0 3

Cf4-04 2 AB118592.1 Endosymbiont 'TC1' of Trimyema compressum 93% 0 0 0 0 0 0 1 1 0 0

Cf4-10 2 AF138733.1 Thermoactinomyces dichotomicus KCTC 3667 90% 0 0 0 1 0 0 0 0 1 0

Cf4-34 2 AB100484.1 clone: Rs-L03 96% 0 0 0 0 0 0 2 0 0 0

Cf4-37 2 AB100475.1 clone: Rs-L36 97% 0 1 1 0 0 0 0 0 0 0

Cf4-42 2 AB089040.1 clone:Rs-N94 95% 0 0 0 0 0 1 0 1 0 0

Cf4-45 2 AB198494.1 clone: RsW01-067 91% 0 1 0 0 0 0 1 0 0 0

Cf4-46 2 AB198524.1 clone: RsW02-034 95% 0 0 1 0 1 0 0 0 0 0

Cf4-53 2 AB197849.1 clone:5-1-N 91% 0 0 0 0 0 0 0 0 2 0

Cf4-54 2 AB126235.1 clone: Rs-E65 93% 1 1 0 0 0 0 0 0 0 0

Cf4-55 2 AB126235.1 clone: Rs-E65 95% 0 0 0 0 0 1 1 0 0 0

Cf4-60 2 AB088991.1 clone:Rs-Q18 90% 0 0 2 0 0 0 0 0 0 0

Cf4-67 2 EF436433.1 clone BRC147 91% 0 0 0 0 0 0 2 0 0 0

Cf4-77 2 DQ815751.1 clone aab21g12 89% 0 0 0 0 2 0 0 0 0 0

Cf4-79 2 DQ856510.1 clone C1N 94% 0 0 0 0 2 0 0 0 0 0

Cf4-83 2 AB062841.1 clone:BCf8-25 96% 0 0 0 0 0 2 0 0 0 0

Cf4-85 2 AB100475.1 clone: Rs-L36 96% 0 0 0 0 2 0 0 0 0 0

Cf4-01 1 DQ105968.1 Bacillus megaterium 99% 0 0 1 0 0 0 0 0 0 0

Cf4-02 1 EU057605.1 Clostridium sp. CJT-3 97% 0 0 0 0 0 0 0 1 0 0

Cf4-08 1 AY533171.1 Pilibacter termitis strain TI-1 95% 0 0 0 0 0 0 1 0 0 0

Cf4-09 1 AY533171.1 Pilibacter termitis strain TI-1 94% 0 0 0 0 0 0 0 1 0 0

Cf4-11 1 DQ905456.1 clone 013C-H6 91% 0 0 0 0 0 0 0 1 0 0

Cf4-12 1 AY916298.1 clone MH87 88% 1 0 0 0 0 0 0 0 0 0

Cf4-13 1 AY766465.1 Dehalobacter sp. E1 91% 0 0 0 0 0 0 1 0 0 0 (Table continued) 84

Cf4-16 1 AF371774.1 clone p-878-a5 89% 0 0 0 0 0 0 0 1 0 0

Cf4-17 1 AB288883.1 clone: BOf1-23 88% 1 0 0 0 0 0 0 0 0 0

Cf4-21 1 AF332709.1 clone UASB_brew_B8 87% 0 0 0 0 0 0 0 0 1 0

Cf4-22 1 EF608543.1 clone PCD-27 88% 1 0 0 0 0 0 0 0 0 0

Cf4-24 1 AB062773.1 clone:BCf1-16 98% 1 0 0 0 0 0 0 0 0 1

Cf4-25 1 AB062827.1 clone:BCf5-15 99% 0 0 1 0 0 0 0 0 0 1

Cf4-40 1 AB088979.1 clone:Rs-M04 95% 0 0 1 0 0 0 0 0 0 0

Cf4-44 1 AB088967.1 clone:Rs-P37 97% 0 1 0 0 0 0 0 0 0 0

Cf4-50 1 AB231042.1 clone:HsW01-048 97% 0 0 1 0 0 0 0 0 0 0

Cf4-56 1 AB100465.1 clone: Rs-L24 97% 0 0 1 0 0 0 0 0 0 0

Cf4-59 1 AB088972.1 clone:Rs-H32 96% 0 0 0 0 0 0 0 1 0 0

Cf4-61 1 AB088991.1 clone:Rs-Q18 90% 0 1 0 0 0 0 0 0 0 0

Cf4-63 1 AB198499.1 clone: RsW01-073 90% 1 0 0 0 0 0 0 0 0 0

Cf4-69 1 AF467426.1 clone 87-8 99% 0 0 0 0 0 0 1 0 0 0

Cf4-74 1 DQ279737.1 Clostridium citroniae strain RMA 16102 94% 0 0 0 0 1 0 0 0 0 0

Cf4-86 1 AB234471.1 clone: MgMjD-101 96% 0 0 0 1 0 0 0 0 0 0

Cf4-87 1 AB231039.1 clone:HsW01-031 94% 0 0 0 0 1 0 0 0 0 0

Cf4-88 1 AB231041.1 clone:HsW01-047 95% 0 0 0 0 1 0 0 0 0 0

Cf4-96 1 AB062775.1 clone:BCf10-04 93% 0 0 0 0 0 1 0 0 0 2

Cf10-06 1 AB088988.1 clone:Rs-H81 87% 0 1 0 0 0 0 0 0 0 0

Spirochaetes Cf7-72 35 AB062829.1 clone:BCf5-23 98% 9 9 8 1 8 0 0 0 0 0

(11.30%) Cf7-06 20 AJ419822.1 Spirochaeta sp. clone rs438 93% 2 1 15 0 1 0 0 1 0 0

Cf7-08 16 AJ419822.1 Spirochaeta sp. clone rs443 93% 5 2 1 0 0 1 0 6 1 0

Cf7-01 14 AJ419819.1 Spirochaeta sp. clone nc5 92% 0 3 11 0 0 0 0 0 0 0

Cf7-40 6 AF068347.1 clone CFS121 97% 1 0 4 1 0 0 0 0 0 0

Cf7-46 6 AJ419821.1 Spirochaeta sp. clone rs438 89% 6 0 0 0 0 0 0 0 0 0

Cf7-65 6 AB062768.1 clone:BCf1-01 99% 3 0 3 0 0 0 0 0 0 1

Cf7-13 5 EF454022.2 clone 290cost002-P3L-1474 91% 1 0 3 0 0 0 0 1 0 0

Cf7-31 5 AB062840.1 clone:BCf8-03 99% 1 3 1 0 0 0 0 0 0 1 (Table continued) 85

Cf7-42 5 AJ419821.1 Spirochaeta sp. clone rs438 92% 0 1 2 0 0 2 0 0 0 0

Cf7-18 4 AB062808.1 clone:BCf11-05 95% 0 0 1 0 1 1 1 0 0 0

Cf7-25 4 AB062823.1 clone:BCf4-14 95% 1 2 1 0 0 0 0 0 0 0

Cf7-29 4 AB062834.1 clone:BCf6-24 95% 2 0 2 0 0 0 0 0 0 0

Cf7-34 4 AB062846.1 clone:BCf9-14 96% 0 3 0 0 1 0 0 0 0 0

Cf7-49 4 AB088894.1 clone:Rs-A19 95% 0 0 2 2 0 0 0 0 0 0

Cf7-54 4 X89050.1 clone mpsp2 92% 1 2 0 1 0 0 0 0 0 0

Cf7-03 3 AJ419821.1 Spirochaeta sp. clone rs438 92% 0 1 2 0 0 0 0 0 0 0

Cf7-09 3 AF093251.1 Treponema sp. ZAS-1 91% 1 1 1 0 0 0 0 0 0 0

Cf7-17 3 AB062808.1 clone:BCf11-05 94% 0 3 0 0 0 0 0 0 0 0

Cf7-19 3 AB062808.1 clone:BCf11-05 97% 1 1 1 0 0 0 0 0 0 0

Cf7-22 3 AB062808.1 clone:BCf11-05 94% 2 1 0 0 0 0 0 0 0 0

Cf7-24 3 AB062823.1 clone:BCf4-14 99% 0 1 2 0 0 0 0 0 0 2

Cf7-36 3 AB062846.1 clone:BCf9-14 99% 1 1 1 0 0 0 0 0 0 1

Cf7-41 3 AF068345.1 clone CFS6 95% 1 1 0 1 0 0 0 0 0 0

Cf7-43 3 AB062834.1 clone:BCf6-24 94% 0 1 1 0 0 0 1 0 0 0

Cf7-60 3 EF454909.2 clone 290cost002-P3L-646 93% 0 1 1 1 0 0 0 0 0 0

Cf7-68 3 AB088903.1 clone:Rs-E21 93% 0 2 1 0 0 0 0 0 0 0

Cf7-12 2 AY571482.1 clone RsaHf236 92% 0 1 1 0 0 0 0 0 0 0

Cf7-16 2 AB062808.1 clone:BCf11-05 99% 1 0 0 0 0 1 0 0 0 1

Cf7-27 2 AB062825.1 clone:BCf5-01 94% 0 0 1 0 0 0 0 1 0 0

Cf7-35 2 AB062846.1 clone:BCf9-14 96% 0 0 1 0 0 0 0 1 0 0

Cf7-37 2 AB062846.1 clone:BCf9-14 93% 0 1 1 0 0 0 0 0 0 0

Cf7-39 2 AF068347.1 clone CFS121 100% 0 2 0 0 0 0 0 0 0 0

Cf7-59 2 EF454847.2 clone 290cost002-P3L-517 93% 0 1 0 1 0 0 0 0 0 0

Cf7-61 2 EF454909.2 clone 290cost002-P3L-646 92% 0 1 1 0 0 0 0 0 0 0

Cf7-73 2 EF453867.2 clone 290cost002-P3L-1239 93% 0 0 0 1 0 0 0 1 0 0

Cf7-15 1 AY571483.1 clone RsaHf365 92% 1 0 0 0 0 0 0 0 0 0

Cf7-20 1 AB062808.1 clone:BCf11-05 95% 0 1 0 0 0 0 0 0 0 0

Cf7-21 1 AB062808.1 clone:BCf11-05 94% 1 0 0 0 0 0 0 0 0 0

86 (Table continued)

Cf7-28 1 AB062832.1 clone:BCf6-15 92% 0 1 0 0 0 0 0 0 0 0

Cf7-30 1 AB062839.1 clone:BCf7-24 99% 1 0 0 0 0 0 0 0 0 1

Cf7-32 1 AB062843.1 clone:BCf9-08 99% 0 0 1 0 0 0 0 0 0 1

Cf7-38 1 AB084953.1 clone: NkS4 95% 0 0 1 0 0 0 0 0 0 0

Cf7-47 1 EF453804.1 clone 290cost002-P3L-1157 93% 0 0 1 0 0 0 0 0 0 0

Cf7-48 1 AB032009.1 clone HsDiSp319 91% 0 0 1 0 0 0 0 0 0 0

Cf7-50 1 AY571483.1 clone RsaHf365 92% 1 0 0 0 0 0 0 0 0 0

Cf7-51 1 AM182455.1 Treponema sp. SPIT5 strain SPIT5 92% 0 1 0 0 0 0 0 0 0 0

Cf7-52 1 AY571483.1 clone RsaHf365 92% 0 0 1 0 0 0 0 0 0 0

Cf7-53 1 AB062825.1 clone:BCf5-01 94% 1 0 0 0 0 0 0 0 0 0

Cf7-55 1 AJ458946.1 clone mp4 92% 1 0 0 0 0 0 0 0 0 0

Cf7-56 1 AB088893.2 clone:Rs-E18 92% 1 0 0 0 0 0 0 0 0 0

Cf7-58 1 AY571483.1 clone RsaHf365 90% 0 0 1 0 0 0 0 0 0 0

Cf7-63 1 AB192256.1 clone: RPK-52 93% 0 0 1 0 0 0 0 0 0 0

Cf7-64 1 AB192150.1 clone: RsTz-71 91% 0 0 1 0 0 0 0 0 0 0

Cf7-71 1 AY571483.1 clone RsaHf365 92% 0 0 0 1 0 0 0 0 0 0

Actinobacteria Cf1-12 22 AB062844.1 clone:BCf9-11 99% 2 0 0 1 8 1 3 6 1 1

(5.70%) Cf1-04 16 AY160874.1 clone COB P3-21 94% 0 0 0 1 9 1 1 2 2 0

Cf1-09 15 AB210965.1 Cellulosimicrobium cellulans strain: SSCT73 90% 0 0 0 0 0 0 2 10 3 0

Cf1-15 11 AB089082.1 clone:Rs-Q71 93% 0 0 0 5 4 2 0 0 0 0

Cf1-22 9 AB062818.1 clone:BCf3-22 99% 0 0 0 1 4 0 0 4 0 1

Cf1-06 8 AY160874.1 clone COB P3-21 95% 0 0 0 0 6 0 1 1 0 0

Cf1-11 7 AB062818.1 clone:BCf3-22 92% 0 0 0 0 1 4 2 0 0 0

Cf1-19 5 EU137440.1 clone Oh3137A10B 99% 0 0 0 4 1 0 0 0 0 0

Cf1-02 3 AB089074.1 clone:Rs-J10 96% 0 0 0 0 0 0 0 1 2 0

Cf1-13 3 AM278923.2 clone AP13U.307 90% 1 0 0 0 1 0 0 1 0 0

Cf1-05 2 AY160874.1 clone COB P3-21 96% 0 0 0 0 1 0 0 1 0 0

Cf1-03 1 AB089078.1 clone:Rs-K09 95% 0 0 0 0 0 0 0 1 0 0

Cf1-07 1 AY160874.1 clone COB P3-21 95% 0 0 0 0 0 0 1 0 0 0 (Table continued) 87

Cf1-08 1 DQ303278.1 clone fc3 93% 0 0 0 0 0 0 0 1 0 0

Cf1-10 1 AB062818.1 clone:BCf3-22 94% 0 0 0 0 0 0 0 0 1 0

Cf1-14 1 AB089074.1 clone:Rs-J10 94% 0 0 0 0 1 0 0 0 0 0

Cf1-17 1 AY160874.1 clone COB P3-21 95% 0 0 0 0 1 0 0 0 0 0

Proteobacteria Cf6-10 8 AB062831.1 clone:BCf6-08 99% 1 1 1 0 0 1 0 2 2 2

(2.24%) Cf6-03 6 AF056090.1 Desulfovibrio sp. ABHU2SB 95% 0 0 2 1 1 0 0 1 1 0

Cf6-19 6 AB062810.1 clone:BCf11-19 99% 0 0 4 0 0 0 0 1 1 0

Cf6-06 3 AY160866.1 clone COB P3-26 94% 1 0 2 0 0 0 0 0 0 0

Cf6-07 3 DQ836748.1 clone DB-3 94% 0 2 1 0 0 0 0 0 0 0

Cf6-01 2 AM410705.1 Acinetobacter junii strain ACI289 99% 0 0 0 0 0 1 0 0 1 0

Cf6-05 2 DQ453130.1 Enterobacteriaceae bacterium Eant3-3 98% 0 0 0 1 0 0 0 1 0 0

Cf6-08 2 AY959000.1 clone rRNA227 99% 0 0 0 0 0 0 2 0 0 0

Cf6-11 2 DQ307726.1 clone MTG-93 95% 0 0 1 0 0 0 0 1 0 0

Cf6-02 1 AY528708.1 Burkholderia sp. MPUD4.5 97% 0 0 0 0 0 0 0 1 0 0

Cf6-04 1 EU047556.1 Enterobacter hormaechei strain TMPSB-T10 99% 0 0 1 0 0 0 0 0 0 0

Cf6-12 1 AB089112.2 clone:Rs-P71 95% 0 0 1 0 0 0 0 0 0 0

Cf6-13 1 AM084010.1 Acidovorax sp. R-24667 strain R-24667 99% 0 0 0 1 0 0 0 0 0 0

Cf6-14 1 CP000539.1 Acidovorax sp. JS42 99% 0 0 0 1 0 0 0 0 0 0

Cf6-16 1 EF592491.1 Enterobacter sp. xw 98% 0 0 0 1 0 0 0 0 0 0

Cf6-17 1 EU137383.1 clone Oh_3137A2C 99% 0 0 0 1 0 0 0 0 0 0

Cf6-18 1 AJ292676.1 clone WD2124 99% 0 0 0 0 1 0 0 0 0 0

Tenericutes Cf4-19 24 AY571416.1 clone RsaHf232 95% 3 7 1 0 4 9 0 0 0 0

(1.55%) Cf4-93 4 AB089057.1 clone:Rs-E42 93% 0 0 0 0 0 4 0 0 0 0

Cf4-65 1 AB089057.1 clone:Rs-E42 90% 0 0 0 0 0 0 0 1 0 0

candidate

division Cf10-07 12 AB089050.2 clone:Rs-D43 96% 0 0 12 0 0 0 0 0 0 0

Termite Group 1 Cf10-09 5 AB089050.2 clone:Rs-D43 96% 0 0 4 0 0 1 0 0 0 0 (Table continued) 88

(1.01%) Cf10-08 1 AB126236.1 clone: Rs-E67 89% 1 0 0 0 0 0 0 0 0 0

Cf10-22 1 AB298053.1 clone: CsSn-1 98% 0 0 0 1 0 0 0 0 0 0

candidate

division Cf8-03 8 AB198518.1 clone: RsW02-021 93% 0 0 0 0 0 0 2 6 0 0

TM7 Cf8-02 2 AY571500.1 clone RsaM67 94% 0 0 0 0 0 0 0 2 0 0

(0.64%) Cf8-01 1 AY571500.1 clone RsaM67 96% 0 0 0 0 0 0 1 0 0 0

Cf8-04 1 AB299568.1 clone: Cc3-038 95% 0 0 0 0 0 0 1 0 0 0

Verrucomicrobia Cf9-04 5 U81738.2 clone vadinHA64 92% 1 1 3 0 0 0 0 0 0 0

(0.59%) Cf9-02 3 AB089122.1 clone:Rs-P07 98% 0 0 0 0 0 0 0 3 0 0

Cf9-01 2 AY587231.1 Opitutaceae bacterium TAV1 98% 0 0 1 0 0 1 0 0 0 0

Cf9-05 1 U81738.2 clone vadinHA64 96% 1 0 0 0 0 0 0 0 0 0

Planctomycetes Cf5-01 8 AB062813.1 clone:BCf2-25 99% 3 0 0 0 2 0 1 2 0 2

(0.48%) Cf10-25 1 U81766.2 clone vadinHA49 94% 0 0 1 0 0 0 0 0 0 0

candidate

division Cf4-28 2 AB062836.1 clone:BCf7-04 99% 0 0 1 0 0 0 0 1 0 1

Synergistes Cf4-33 1 AB089066.1 clone:Rs-C88 95% 1 0 0 0 0 0 0 0 0 0

(0.21%) Cf4-18 1 DQ009680.2 clone R38B21 94% 1 0 0 0 0 0 0 0 0 0

candidate

division Cf10-21 1 AB089067.3 clone:Rs-A23 95% 0 0 0 0 1 0 0 0 0 0

ZB3

(0.05%)

Cyanobacteria Cf3-01 1 EU017026.1 Musa acuminata chloroplast 99% 0 0 0 1 0 0 0 0 0 0 (0.05%)

Anomalous/ Cf4-43 8 AB088967.1 clone:Rs-P37 96% 0 0 3 2 0 0 1 2 0 0

chimeric Cf10-12 7 EU233623.1 Cloning vector pEASY-T1 100% 0 0 0 2 5 0 0 0 0 0 (Table continued) 89

(2.77%) Cf10-14 4 EU233623.1 Cloning vector pEASY-T1 100% 0 0 0 4 0 0 0 0 0 0

Cf2-05 4 AY571429.1 clone RsaHf308 87% 1 0 0 1 2 0 0 0 0 0

Cf2-24 4 EF454327.2 clone 290cost002-P3L-1961 88% 2 0 0 1 1 0 0 0 0 0

Cf10-16 3 DQ297764.1 Cloning vector pHP13 99% 0 0 0 1 1 0 0 1 0 0

Cf2-38 3 AY570596.1 clone PL-10B5 87% 0 0 0 3 0 0 0 0 0 0

Cf10-02 2 AJ544074.1 Crassostrea gigas 96% 0 0 0 0 0 0 2 0 0 0

Cf10-05 2 AY570596.1 clone PL-10B5 87% 1 0 0 1 0 0 0 0 0 0

Cf10-01 1 EU233623.1 Cloning vector pEASY-T1 99% 0 0 0 0 0 0 0 1 0 0

Cf10-03 1 AB069651.1 Rhizobium sp. JEYF14 97% 0 0 1 0 0 0 0 0 0 0

Cf10-04 1 AJ419822.1 Spirochaeta sp. clone rs443 88% 1 0 0 0 0 0 0 0 0 0

Cf10-13 1 EU233623.1 Cloning vector pEASY-T1 99% 0 0 0 1 0 0 0 0 0 0

Cf10-15 1 EU233623.1 Cloning vector pEASY-T1 91% 0 0 0 1 0 0 0 0 0 0

Cf10-17 1 DQ297764.1 Cloning vector pHP13 99% 0 0 0 1 0 0 0 0 0 0

Cf10-18 1 DQ882634.1 Coptotermes acinaciformis 18S rRNA 99% 0 0 0 0 1 0 0 0 0 0

Cf10-19 1 DQ882634.1 Coptotermes acinaciformis 18S rRNA 99% 0 0 0 0 1 0 0 0 0 0

Cf10-20 1 AF247461.1 clone Cf2-3 microsatellite 91% 0 0 0 0 1 0 0 0 0 0

Cf10-23 1 EU233623.1 Cloning vector pEASY-T1 99% 0 0 0 0 1 0 0 0 0 0

Cf1-20 1 AY160874.1 clone COB P3-21 90% 0 0 0 0 1 0 0 0 0 0

Cf2-42 1 AF068346.1 clone CFS124 90% 0 1 0 0 0 0 0 0 0 0

Cf4-72 1 EU520326.1 Lactococcus lactis 94% 0 0 0 0 1 0 0 0 0 0

Cf4-84 1 AB100484.1 clone: Rs-L03 88% 0 0 0 0 1 0 0 0 0 0

Cf7-74 1 AB062768.1 clone:BCf1-01 89% 1 0 0 0 0 0 0 0 0 0

90

VITA

Huei-Yang Ho was born in February 1985 in Penang, Malaysia, an island situated off the western side of Peninsular Malaysia bordering Thailand. He was raised on the mainland region of Penang and commuted daily to the island to attend school during his primary and secondary education. Upon completion of his secondary education in December 2001, he proceeded with his college education at KDU college in Penang, with the American University Transfer Program

(AUTP). In Spring 2003, he transferred his credits to Louisiana State University and resumed his pursuit of the Bachelor of Science degree in biological sciences in the United States. He graduated in the Spring 2005 and worked for half a year as a transient worker under Dr Claudia

Husseneder before starting his Master of Science in biological sciences with Dr Claudia

Husseneder as the major professor. Huei-Yang passed his Final Examination in Summer 2008 and officially obtained his master’s degree in Fall 2008.

91