<<

A holobiont characterization of in a

live-bearing , punctata

A dissertation submitted to the

Graduate School of the

University of Cincinnati

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in the Department of Biological Sciences

of the McMicken College of Arts and Sciences

By

Emily C. Jennings

B.S. Neurobiology, University of Cincinnati, April 2013

Committee Chair: Joshua B. Benoit, Ph.D.

Abstract

Viviparous reproduction is characterized by maternal retention of developing offspring within the reproductive tract during , culminating in live . In some cases, a mother will provide nutrition beyond that present in the ; this is known as matrotrophic .

While this phenomenon is best associated with mammals, it is observed in such as the viviparous cockroach, . Female D. punctata carry developing in the brood sac, a reproductive organ that acts as both a and by protecting and providing a nutritive secretion to the intrauterine developing progeny. While the basic physiology and hormonal changes of D. punctata have been characterized, little else is known about this phenomenon. This study attempts to broaden the understanding of D. punctata reproduction by using a multi-omics approach to characterize multiple aspects of the holobiome. First, I utilized RNA-seq analysis to characterize molecular changes associated with

D. punctata reproduction and provides the most complete gene set to date for this . A comparison of four stages of the female reproductive cycle revealed unique gene expression profiles corresponding to each stage. Differentially regulated transcripts of interest include the previously identified family of milk proteins, transcripts associated with juvenile metabolism, and other reproduction-associated transcripts.

I next utilized 16S rRNA gene sequencing to characterize the microbiome of Diploptera punctata during development, from embryonic development to female adulthood. We identified 50 phyla and 121 classes overall and found that mothers and their developing embryos had significantly different microbial communities, with embryos harboring only a single endosymbiont known as Blattabacteria. Our analysis of postnatal development reveals

i that significant amounts of non-Blattabacteria species are not able to colonize newborn D. punctata until melanization in the first instar, after which the microbial community rapidly and dynamically diversifies. This rapid change of the microbial community appears to stabilize after the second juvenile stage with no major changes even through the last molt to adulthood for females.

Lastly, I sought to better understand the role of the cockroach specific endosymbiont

Blattabacteria as the sole member of the D. punctata embryonic microbial community. We leveraged next generation genome sequencing to computationally characterize the metabolic capacity of Blattabacteria isolated from D. punctata. Our analysis produced a 670,273 base pair chromosome with a high degree of similarity to previously sequenced genomes of

Blattabacteria isolated from other cockroach species. Our genome assembly contained 593 putatively protein coding DNA sequences, which included proteins required for nearly all key metabolic processes including DNA replication, transcription, and translation. Additionally, the

D. punctata strain of Blattabacteria also holds the ability to synthesize nearly all 10 essential amino acids and process nitrogenous waste to do so. This suggests that Blattabacteria are likely providing nutritional supplementation to intrauterine developing embryos to accommodate the previously reported lack of methionine and tryptophan in maternally provided nutrition.

Together, the research presented in this dissertation contributes to a more holistic understanding of this human-like reproductive mode in the cockroach Diploptera punctata, broadening our understanding of viviparity throughout the kingdom.

ii

© Copyright

Emily C Jennings

2019

All Rights Reserved

iii Acknowledgements

First and foremost, I want to thank my advisor, Dr. Joshua Benoit, for going out on a limb and taking me on as one of two graduate students the same year he started as a new faculty member at UC. He provided constant support and exhibited unfathomable patience as I stumbled through the last six years in his lab. He encouraged me to think not only critically as a scientist but creatively, encouraging me to take risks on wild experiments like modeling fetal alcohol syndrome in . When some of those risks resulted in failed experiments Josh was always there to help me bounce back, just as he was always there to casually offer kudos when something did manage to work out. Josh helped keep me on track when I started to lose focus and dealt with my constant questions. He never questioned my decisions to enroll in courses outside of traditional biology curriculum, whether it was a course about birth defects, or an undergraduate level python programming course. When I realized that I was interested in at least trying a career outside of academia after graduating, he never wavered in his support even when I felt like a quitter. Josh was always open to and encouraged discussions about my project, which sometimes turned into fun debates about where the experiment was going or needed to go. What wasn’t always fun was when one of us had to admit we were wrong, but

Josh was a good sport and never gloated too much when I was wrong even though I wasn’t always so kind when he was wrong. It’s been a wonderful six years in the lab, and I know I’ll miss it when I’m gone.

I’d also like to acknowledge and thank Dr. Elke Buschbeck, Dr. Joshua Gross, Dr. Mike

Polak, and Dr. Matthew Weirauch, for being part of this crazy journey as members of my research advisory committee. They have offered invaluable guidance and support, helping me

iv to navigate and focus my disjointed project ideas, overcome the roadblocks of failed experiments and helping to know when it was time to let go of a stalled-out project. Through it all, I’ve always been able to count on them to be in my corner of the ring.

In addition to serving on my advisory committee, Dr. Elke Buschbeck has played an especially important role in this journey. I’d like to thank her for gambling on hiring me, a weird bug fanatic in her neuroscience course, as a work study back in 2011. She has consistently served as a mentor and role model, helping me grow as a scientist. I cannot begin to describe the value of her constant encouragement and support, letting me know when I was too hard on myself but also making sure to ask if I had something better to do when she noticed I was slacking.

I want to extend my gratitude to all of the current and former members of the Benoit lab for their constant support, listening to me whine when I needed to, dealing with me talking to myself at the lab bench or in our shared office space, and keeping things fun in and out of lab. This work would not be possible without the undergraduate students who have been part of our lab throughout the last 6 years. LaVeta Burke, Luke Bernhardt, John Cavanaugh, Jake

Hendershot, Matthew Korthauer, Erica McDonald, and Sophie Shemas have all made contributions to my research, running countless experiments. Working with them has taught me so much about mentoring, communication, and leadership. It has been an honor to have worked with them. Christopher Holmes and Elise Didion deserve special recognition. I cannot thank them enough for being wonderful lab mates and two of my best friends. Without them this process would have been much harder and very lonely.

v I am so grateful for the patience and support from my family over the last six years.

They never complained when I missed countless birthdays and holidays because I was working or studying, instead welcoming me home with open arms every chance they got. My parents and step-parents have been supportive, kind, and forgiving beyond any reasonable expectation.

Without them, my education would have been impossible. They believed in me when I couldn’t believe in myself, pushed me to constantly be better and strive for more than I thought I could.

I would like to recognize the immense support of my running family. Not only have they been there for me when I needed a friend, but they have listened to me talk about all aspects of graduate school ad nauseum (sometimes for 26.2 miles). I am grateful for the unique perspective and advice that each one of them has provided me over the last two years. In addition to coaching me through three full marathons, countless training runs, and smaller races, they have also coached me throughout the last two years of my PhD, a very different kind of marathon.

Samantha Isler deserves more recognition than I can give her here, meeting her dancing in the rain back in high school might be one of the greatest things that has ever happened to me. I cannot imagine my life without her, and I cannot thank her enough for being my best friend for all of these years. She has been there for me through every high and every low, my light in the dark.

Lastly, I need to thank my husband, Justin. He has loved me when I was a nightmare, always understood how much my research means to me, and forgiven me when I put him second to science. He has always supported all of my wildest dreams, even when I changed my mind constantly about what those were. Justin is constantly lifting me up, assuring me that I am

vi enough, and that I deserve to be where I am. I’m fairly certain that I wouldn’t be able to accomplish anything without him, including this degree. He is my best friend, an amazing husband, a wonderful dog and cat dad, my number one fan, and also the most resilient person I know. For all this and more, I am eternally grateful.

vii

In loving memory of Dennis Whitehead

Thank you for letting me write a paper about Rosalind Franklin instead of Watson & Crick,

among the many other things you let me get away with.

viii Table of Contents

Abstract ...... i

Acknowledgements ...... iv

Chapter 1 ...... 2

References ...... 11 Chapter 2 ...... 26

Abstract ...... 27 Introduction ...... 28 Methods ...... 29 Results ...... 34 Discussion ...... 38 References ...... 45 Tables ...... 56 Legends and Figures ...... 67 Chapter 3 ...... 76

Introduction ...... 78 Methods ...... 81 Results ...... 85 Discussion ...... 90 References ...... 98 Tables ...... 115 Figures and legends ...... 121 Chapter 4 ...... 129

Abstract ...... 130 Introduction ...... 131 Methods ...... 133 Results ...... 135 Discussion ...... 136 References ...... 142

Tables ...... 150 Figures and legends ...... 151 Chapter 5 ...... 156

References ...... 168 Appendix ...... 184

i

Chapter 1

Introduction and Background

2 Overview of matrotrophic viviparity in insects: Within the animal kingdom there exist three main reproductive strategies: (laying of ), facultative viviparity in which a mother may lay eggs or give birth to live, active offspring, and true viviparity (Hagan, 1948; Meier,

Kotrba, & Ferrar, 1999; Roth & Willis, 1957; Stewart, 2015). Truly viviparous reproduction is characterized by the obligate retention of developing progeny in the maternal reproductive tract for the duration of gestation. This process culminates in the birth of active offspring; early termination of the pregnancy results in stillborn offspring (Hagan, 1948; Kalinka, 2015).

Viviparous reproduction can be further divided into two more specific classifications: lecithotrophic viviparity, in which mothers provide only fluids to developing progeny; and matrotrophic viviparity which is characterized by the provisioning of both fluids and non-yolk nutrients to embryos during intra-uterine development (Blackburn, 2014; Blackburn & Starck,

2015). Females that reproduce via true viviparity must carry, provide fluids, and, in the case of matrotrophic viviparity, supply nutrients to their offspring. Thus, this reproductive strategy is associated with an extremely elevated cost for the mother when compared with the investments of oviparity and facultative viviparity (Bleu, Massot, Haussy, & Meylan, 2012;

Hopkins, Eldridge, & Cech, 1995; Schultz, Webb, & Christian, 2008; Shaffer & Formanowicz,

1996).

Despite the immense investments required, viviparity has independently evolved hundreds of times across the animal kingdom (Blackburn, 2014; Crespi, 1989; Hagan, 1948;

Hussey et al., 2010; Kalinka, 2015; Meier et al., 1999; Roth & Willis, 1957; Shine, 1983;

Tworzydlo, Kisiel, & Bilinski, 2013). Lecithotrophic viviparous include some squamate and , such as the aspic viper, multiple as well as (Blackburn, 1992;

3 Hagan, 1948; Hussey et al., 2010; Roth & Willis, 1957). Present in at least 13 orders of insects

(Clutton-Brock, 1991; Hagan, 1948; Roth & Willis, 1957), viviparity has at least 62 independent origins within the 22 Dipteran families (Kalinka, 2015; Meier et al., 1999).

Matrotrophic viviparity has widespread representation in the animal kingdom (21 of the

34 animal phyla) (Ostrovsky et al., 2016). Squamate , such as the three-toed ; certain anuran amphibians; and fish such as the yellowtail rockfish have documented viviparous species (Blackburn, 1992; Macfarlane & Bowers, 1995). There are 213 families of chordates and 45 families that have been estimated to be matrotrophic viviparous

(Ostrovsky et al., 2016). However, at large have the greatest diversity of progeny incubation and in the animal kingdom (Ostrovsky et al., 2016). Matrotrophic viviparity has been observed in many insects such as dermapterans, multiple species, most notably the Glossinadae (Attardo et al., 2019; Denlinger & Ma, 1974; Ma, Denlinger, Järlfors, &

Smith, 1975; Tobe & Langley, 1978), and the Pacific mimic cockroach, Diploptera punctata (Hagan, 1939; Roth & Stay, 1961; Roth & Willis, 1955; Stay & Coop, 1974; Stay & Roth,

1956). Despite the prevalence of matrotrophic viviparity in insects, little is known about the molecular underpinnings of this phenomenon outside of tsetse (Attardo et al., 2019;

Benoit et al., 2014; Benoit, Attardo, Baumann, Michalkova, & Aksoy, 2015; Guz, Attardo, Wu, &

Aksoy, 2007; Michalkova, Benoit, Attardo, Medlock, & Aksoy, 2014; Michalkova, Benoit, Weiss,

Attardo, & Aksoy, 2014). Increasing the degree to which this human-like pregnancy is understood in a diverse set of animals will broaden the lens through which researchers view the evolution of matrotrophic viviparity.

4

Overview of microbiomes and maternal transmission: It has become common knowledge that animals share their bodies to host a diverse suite of microorganisms known as the microbiome

(Engel & Moran, 2013; The NIH HMP Working Group, 2009). The microbiome has been found to play important roles in a variety of host processes, such as nutrient metabolism to immunity

(Albenberg & Wu, 2014; Chung et al., 2012; Dimmitt et al., 2010; Douglas, 2017; Jašarević,

Rodgers, & Bale, 2015; Michalkova, Benoit, Weiss, et al., 2014; Pais, Lohs, Wu, Wang, & Aksoy,

2008; Snyder & Rio, 2015; Wang, Weiss, & Aksoy, 2013; Weiss, Wang, & Aksoy, 2011). Most insects reproduce by oviparity and consequently acquire the majority of diverse microbial symbionts in their microbiome over the course of development (Abdul Rahman et al., 2015;

Bright & Bulgheresi, 2010; da Costa & Poulsen, 2018; Estes et al., 2013; Funkhouser &

Bordenstein, 2013; Salem, Florez, Gerardo, & Kaltenpoth, 2015; Schwab, Riggs, Newton, &

Moczek, 2016; Shukla, Vogel, Heckel, Vilcinskas, & Kaltenpoth, 2018). In some cases, symbionts can be passed from oviparous mothers to their offspring before birth by incorporating the microbes into the developing eggs (Abdul Rahman et al., 2015; Binkhorst et al., 2012; Estes et al., 2013; Schwab et al., 2016; Shukla et al., 2018). However, the extensive and complex interactions between mother and offspring during pregnancy in matrotrophic viviparous insects provide an opportunity for mothers to vertically transmit microbes to their developing offspring

(Denlinger & Ma, 1975; Funkhouser & Bordenstein, 2013; Morse et al., 2013; Wang et al.,

2013). However, little is known about the role that viviparous reproduction plays in microbial transmission outside of tsetse flies, which are able to transfer important symbiotic to internally developing offspring through nutritive milk-like secretions (Bing et al.,

5 2017; Cheng et al., 2000; De Vooght, Caljon, Van Hees, & Van Den Abbeele, 2015; Douglas,

2017; Snyder, Mclain, & Rio, 2012; Snyder & Rio, 2015; Wang et al., 2013). Considering the ubiquity of viviparity across the animal kingdom and the degree to which the extended interactions between mother and offspring during gestation can impact offspring, further study of microbial inheritance in live-bearing species is warranted.

Nutritional contributions of endosymbionts: Endosymbionts live inside the tissues and cells of their host and are consistently transmitted from mother to offspring during reproduction

(Moran & Wernegreen, 2000; Wernegreen, 2002, 2004). The stability in environment and transmission allows these bacteria to specialize, co-evolving with their host (Hosokawa, Kikuchi,

Nikoh, Shimada, & Fukatsu, 2006; Kinjo et al., 2018; Moran & Wernegreen, 2000; Neef et al.,

2011; Nikoh, Hosokawa, Oshima, Hattori, & Fukatsu, 2011; Sabree et al., 2012; Wernegreen,

2002; Wilson et al., 2010). This co-evolution of endosymbiotic bacteria with their host yields drastic genome reductions over the course of these increasingly mutualistic relationships, often times seen as near complete elimination of some cellular functions in the genome; typical functions lost include cell wall components, cell motility, response to the environment as well as signal transduction (Akman et al., 2002; Jiang et al., 2013; Kinjo et al., 2018; López-Sánchez et al., 2008; McCutcheon & Moran, 2010, 2012; Nakabachi et al., 2006; Neef et al., 2011;

Santos-Garcia et al., 2014; Snyder et al., 2012; van Ham et al., 2003; Wolf & Koonin, 2013).

Despite the drastic reductions seen in previously characterized endosymbiont genomes, obligate endosymbionts often retain genes encoding complex metabolic pathways like those for amino acid or vitamin biosynthesis (Akman et al., 2002; Michalik, Szklarzewicz, Jankowska, &

6 Wieczorek, 2014; Sabree et al., 2012; Sabree, Kambhampati, & Moran, 2009; Snyder & Rio,

2015; Tokuda et al., 2013; Wilson et al., 2010). Wigglesworthia morsitans, an endosymbiont transmitted via milk secretions in the live-bearing , has lost many genes functioning in regulatory processes and transport related functions in its reduced genome (Akman et al.,

2002) (Akman et al 2002). Despite its small size, the W. morsitans genome has retained the ability to synthesize multiple B-vitamins and cofactors (Akman et al., 2002), metabolites that are required for host fitness but absent in the tsetse fly diet of blood (Attardo et al.,

2019; Bing et al., 2017; Michalkova, Benoit, Weiss, et al., 2014; Pais et al., 2008). In phloem feeding aphids, nearly 10% of open reading frames in the genome of its stable endosymbiont,

Buchnera aphidicola, encode the biosynthesis of essential amino acids required by the aphid host (Shigenobu, Watanabe, Hattori, Sakaki, & Ishikawa, 2000; van Ham et al., 2003).

Interestingly, genes encoding non-essential amino acid and cofactor biosynthesis are absent in the B. aphidicola genome meaning that the symbionts must acquire these compounds from the host as many serve as the building blocks for essential amino acid synthesis (Michalik et al.,

2014; van Ham et al., 2003; Wilson et al., 2010).

Cockroaches also have a primary endosymbiont, (Bandi et al., 1994,

1995; Nalepa, Bignell, & Bandi, 2001). Transmitted transovarially, these bacteria reside in specialized bacteriocyte cells in the fat body organ and near the (Bandi et al., 1994,

1995; Nalepa et al., 2001). Genomic surveys of multiple strains of Blattabacterium revealed mechanisms underlying the previously observed effect of these bacteria on the storage of nitrogenous waste as uric acid (Hamilton & Schal, 1988) and sulfur assimilation (Block & Henry,

1961; Henry & Block, 1961), in addition to many other metabolic pathways maintained in a

7 reduced genome (Huang, Sabree, & Moran, 2012; Kambhampati, Alleman, & Park, 2013; López-

Sánchez et al., 2008, 2009; Patino-Navarrete et al., 2014; Patiño-Navarrete, Moya, Latorre, &

Peretó, 2013; Sabree et al., 2009; Tokuda et al., 2013). These bacteria allow cockroaches to recycle uric acid to produce important metabolic products such as essential amino acids which may be lacking in the diet of what are commonly detritivores (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008, 2009; Patino-Navarrete et al., 2014; Patiño-Navarrete et al., 2013; Sabree et al., 2009; Tokuda et al., 2013). These among many other genomic analyses of the relationships between endosymbionts and hosts have revealed likely evolutionary benefits related to diet and niche expansion, but there is still much to be learned about these relationships in contexts such as host reproduction from genomic analyses of endosymbionts.

Diploptera punctata as a model: The beetle mimic cockroach, Diploptera punctata, a Polynesian native species, reproduces by matrotrophic viviparity (Ingram, Stay, & Cain, 1977; Roth & Hahn,

1964; Roth & Stay, 1961; Roth & Willis, 1955; Stay & Coop, 1974). Embryos develop inside the brood sac, a unique organ which functions as both a uterus and pseudo-placenta; embryos are provided with nutrients by a secretion of milk-like components (Roth & Stay, 1961; Roth &

Willis, 1955; Stay & Clark, 1971; Stay & Coop, 1973; Stay & Roth, 1956). This secretion appears in gut contents at 20% of the 60-70-day pregnancy, when the dorsal edge of the body wall is closed. Diploptera milk is a combination of proteins and free amino acids, carbohydrates, and lipids in a water base (Stay & Coop, 1973, 1974; Williford, Stay, & Bhattacharya, 2004). The proteins present include a family of milk proteins derived from 25 unique genes coding for 22

8 different proteins (Stay & Coop, 1973, 1974; Williford et al., 2004). During its gestation period, between nine and thirteen embryos grow from 1.5 mm at the time of to over 6mm at birth, increasing in weight by more than 70-fold. Water content of each increases by 85- fold, and solid dry weight increases over 49-fold (Roth, 1967; Roth & Hahn, 1964; Roth & Stay,

1961; Roth & Willis, 1955; Stay & Coop, 1973, 1974; Stay & Roth, 1956). While the basic physiology of this process has been characterized, studies of molecular and genetic aspects of matrotrophic viviparous reproduction in D. punctata have been limited to early stages in reproduction before ovipositioning into the brood sac (Garside, Koladich, Bendena, & Tobe,

2002; Huang, Hult, Marchal, & Tobe, 2015; Huang et al., 2014; Hult, Huang, Marchal, Lam, &

Tobe, 2015; Liu, Lin, Lin, Yeh, & Chiang, 2005; Marchal et al., 2014; Marchal, Hult, Huang, Stay,

& Tobe, 2013).

Overview of thesis: Matrotrophic viviparity is widespread throughout the animal kingdom and prevalent in the insects. Characterized by the nutrient provisioning to developing offspring within the maternal reproductive tract and culminating in live birth, this type of reproduction is complex and multifaceted. Despite being well represented in the insects, matrotrophic viviparity has not been well characterized at the molecular level outside of the trypanosomiasis vectors in Glossinidae. While matrotrophic viviparity in the Pacific beetle mimic cockroach,

Diploptera punctata, has been well characterized physiologically very little is known about the reproduction of this animal at the molecular level outside the period of and ovipositioning. The work presented in this thesis addresses the knowledge gap in the reproductive biology of this cockroach by taking a holobiont approach leveraging next-

9 generation sequencing techniques. First, I aimed to close the gap in knowledge regarding the genetic underpinnings of pregnancy in D. punctata utilizing transcriptomics and gene knockdowns by RNA interference. Second, I characterized the development of the D. punctata microbiome using 16S rRNA gene sequencing to determine if reproduction by matrotrophic viviparity facilitated the prenatal transmission of symbionts. Lastly, I sequenced and characterized the genome of the primary endosymbiont of D. punctata, a strain of

Blattabacterium. This provided key insights into the metabolic capacity of this bacterium and the potential roles that it plays in the development of its host. The culmination of this work characterizing the holobiome of D. punctata provides novel information about matrotrophic viviparity in this species as well as throughout the animal kingdom, thus broadening our understanding of this extraordinary example of convergent evolution.

10 References:

Abdul Rahman, N., Parks, D. H., Willner, D. L., Engelbrektson, A. L., Goffredi, S. K., Warnecke, F.,

… Hugenholtz, P. (2015). A molecular survey of Australian and North American

genera indicates that vertical inheritance is the primary force shaping termite gut

microbiomes. Microbiome, 3(1), 1–16. http://doi.org/10.1186/s40168-015-0067-8

Akman, L., Yamashita, A., Watanabe, H., Oshima, K., Shiba, T., Hattori, M., & Aksoy, S. (2002).

Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia

glossinidia. Nature Genetics, 32(3), 402–407. http://doi.org/10.1038/ng986

Albenberg, L. G., & Wu, G. D. (2014). Diet and the intestinal microbiome: Associations,

functions, and implications for health and disease. Gastroenterology, 146(6), 1564–1572.

http://doi.org/10.1053/j.gastro.2014.01.058

Attardo, G. M., Abd-Alla, A. M. M., Acosta-Serrano, A., James, E., Bateta, R., Benoit, J. B., …

Aksoy, S. (2019). The Glossina genome cluster: comparative genomic analysis of the

vectors of African trypanosomes. BioRxiv, 531749. http://doi.org/10.1101/531749

Bandi, C., Damiani, G., Magrassi, L., Grigolo, A., Fani, R., & Sacchi, L. (1994). Flavobacteria as

intracellular symbionts in cockroaches. Proceedings of the Royal Society B: Biological

Sciences, 257, 43–48. http://doi.org/10.1098/rspb.1994.0092

Bandi, C., Sironi, M., Damiani, G., Magrassi, L., Nalepa, C. A., Laudant, U., & Sacchi, L. (1995).

The establishment of intracellular symbiosis in an ancestor of cockroaches and .

Proceedings of the Royal Society B: Biological Sciences, 259(1356), 293–299.

http://doi.org/10.1098/rspb.1995.0043

Benoit, J. B., Attardo, G. M., Baumann, A. A., Michalkova, V., & Aksoy, S. (2015). Adenotrophic

11 viviparity in tsetse flies: Potential for population control and as an insect model for

. Annual Review of , 60(1), 351–371. http://doi.org/10.1146/annurev-

ento-010814-020834

Benoit, J. B., Attardo, G. M., Michalkova, V., Krause, T. B., Bohova, J., Zhang, Q., … Aksoy, S.

(2014). A novel highly divergent protein family identified from a viviparous insect by rna-

seq analysis: A potential target for tsetse fly-specific abortifacients. PLoS Genetics, 10(4),

6–10. http://doi.org/10.1371/journal.pgen.1003874

Bing, X., Attardo, G. M., Vigneron, A., Aksoy, E., Scolari, F., Malacrida, A., … Aksoy, S. (2017).

Unravelling the relationship between the tsetse fly and its obligate symbiont

Wigglesworthia : transcriptomic and metabolomic landscapes reveal highly integrated

physiological networks. Proceedings of the Royal Society B: Biological Sciences, 284(1857),

20170360. http://doi.org/10.1098/rspb.2017.0360

Binkhorst, M., Wortmann, S. B., Funke, S., Kozicz, T., Wevers, R. a., & Morava, E. (2012).

Glycosylation defects underlying fetal alcohol spectrum disorder: A novel pathogenetic

model. Journal of Inherited Metabolic Disease, 35, 399–405.

http://doi.org/10.1007/s10545-011-9425-2

Blackburn, D. G. (1992). Convergent Evolution of Viviparity, Matrotrophy, and Specializations

for Fetal Nutrition in Reptiles and Other . American Zoologist, 32(2), 313–321.

http://doi.org/10.1093/icb/32.2.313

Blackburn, D. G. (2014). Evolution of vertebrate viviparity and specializations for fetal nutrition:

a quantitative and qualitative analysis. Journal of Morphology, 276(8) 961-990.

http://doi.org/10.1002/jmor.20272

12 Blackburn, D. G., & Starck, J. M. (2015). Morphological specializations for fetal maintenance in

viviparous vertebrates: An introduction and historical retrospective. Journal of

Morphology. http://doi.org/10.1002/jmor.20410

Bleu, J., Massot, M., Haussy, C., & Meylan, S. (2012). Experimental litter size reduction reveals

costs of gestation and delayed effects on offspring in a viviparous . Proceedings of the

Royal Society B: Biological Sciences, 279(1728), 489–498.

http://doi.org/10.1098/rspb.2011.0966

Block, R. J., & Henry, S. M. (1961). Metabolism of the sulphur amino-acids and of sulphate in

Blattella germanica. Nature, 191(4786), 392–393. http://doi.org/10.1038/191392a0

Bright, M., & Bulgheresi, S. (2010). A complex journey: Transmission of microbial symbionts.

Nature Reviews Microbiology, 8(3), 218–230. http://doi.org/10.1038/nrmicro2262

Cheng, Q., Ruel, T. D., Zhou, W., Moloo, S. K., Majiwa, P., O’Neill, S. L., & Aksoy, S. (2000). Tissue

distribution and prevalence of Wolbachia infections in tsetse flies, Glossina spp. Medical

and Veterinary Entomology, 14(1), 44–50. http://doi.org/10.1046/j.1365-

2915.2000.00202.x

Chung, H., Pamp, S. J., Hill, J. A., Surana, N. K., Edelman, S. M., Troy, E. B., … Kasper, D. L. (2012).

Gut immune maturation depends on colonization with a host-specific microbiota. Cell,

149(7), 1578–1593. http://doi.org/10.1016/j.cell.2012.04.037

Clutton-Brock, T. H. (1991). The Evolution of . (J. R. Krebs & T. H. Clutton-Brock,

Eds.). Princeton: Princeton University Press.

Crespi, B. J. (1989). Facultative viviparity in a thrips. Nature, 337(6205), 357–358. da Costa, R. R., & Poulsen, M. (2018). Mixed-mode transmission shapes termite gut community

13 assemblies. Trends in Microbiology, 26(7), 557–559.

http://doi.org/10.1016/j.tim.2018.04.005

De Vooght, L., Caljon, G., Van Hees, J., & Van Den Abbeele, J. (2015). Paternal transmission of a

secondary symbiont during mating in the viviparous tsetse fly. Molecular Biology and

Evolution, 32(8), 1977–1980. http://doi.org/10.1093/molbev/msv077

Denlinger, D. L., & Ma, W. C. (1975). Maternal nutritive secretions as possible channels for

vertical transmission of microorganisms in insects: The tsetse fly example. Annals of the

New York Academy of Sciences, 266(1), 162–165. http://doi.org/10.1111/j.1749-

6632.1975.tb35097.x

Denlinger, D. L., & Ma, W. C. (1974). Dynamics of the pregnancy cycle in the tsetse Glossina

morsitans. Journal of Insect Physiology, 20(6). http://doi.org/10.1016/0022-

1910(74)90143-7

Dimmitt, R. A., Staley, E. M., Chuang, G., Tanner, S. M., Soltau, T. D., & Lorenz, R. G. (2010). Role

of postnatal acquisition of the intestinal microbiome in the early development of immune

function. Journal of Pediatric Gastroenterology and Nutrition, 51(3), 1.

http://doi.org/10.1097/MPG.0b013e3181e1a114

Douglas, A. E. (2017). The B vitamin nutrition of insects: the contributions of diet, microbiome

and horizontally acquired genes. Current Opinion in Insect Science, 23, 65–69.

http://doi.org/10.1016/j.cois.2017.07.012

Engel, P., & Moran, N. A. (2013). The gut microbiota of insects-diversity in structure and

function. FEMS Microbiology Reviews, 37, 699–735.

Estes, A. M., Hearn, D. J., Snell-Rood, E. C., Feindler, M., Feeser, K., Abebe, T., … Moczek, A. P.

14 (2013). Brood ball-mediated transmission of microbiome members in the dung beetle,

Onthophagus taurus (Coleoptera: ). PLoS ONE, 8(11), 1–15.

http://doi.org/10.1371/journal.pone.0079061

Funkhouser, L. J., & Bordenstein, S. R. (2013). Mom Knows Best: The Universality of Maternal

Microbial Transmission. PLoS Biology, 11(8), 1–9.

http://doi.org/10.1371/journal.pbio.1001631

Garside, C. S., Koladich, P. M., Bendena, W. G., & Tobe, S. S. (2002). Expression of allatostatin in

the of the cockroach Diploptera punctata. Insect Biochemistry and Molecular

Biology, 32(9), 1089–1099. http://doi.org/10.1016/S0965-1748(02)00046-2

Guz, N., Attardo, G. M., Wu, Y., & Aksoy, S. (2007). Molecular aspects of transferrin expression

in the tsetse fly (Glossina morsitans morsitans). Journal of Insect Physiology, 53(7), 715–

723. http://doi.org/10.1016/j.jinsphys.2007.03.013

Hagan, H. R. (1939). Diploptera dytiscoides, a viviparous roach with elongate pleuropodia.

Annals of the Entomological Society of America, 32, 264–266.

Hagan, H. R. (1948). A Brief Analysis of Viviparity in Insects. Journal of the New York

Entomological Society, 56(1), 63–68.

Hamilton, R. L., & Schal, C. (1988). Effects of dietary protein levels on reproduction and food

consumption in the German cockroach (Dictyoptera: Blattellidae). Annals of the

Entomological Society of America, 81, 969–976.

Henry, S. M., & Block, R. J. (1961). Sulfur metabolism of insects. 6. Metabolism of sulfur amino

acids and related compounds in German cockroach, Blattella germanica (L.). Contrib.

Boyce Thompson Inst., 21, 129–145.

15 Hopkins, T. E., Eldridge, M. B., & Cech, J. J. (1995). Metabolic costs of viviparity in yellowtail

rockfish, Sebastes flavidus. Environmental Biology of , 43, 77–84.

http://doi.org/10.1007/BF00001819

Hosokawa, T., Kikuchi, Y., Nikoh, N., Shimada, M., & Fukatsu, T. (2006). Strict host-symbiont

cospeciation and reductive genome evolution in insect gut bacteria. PLoS Biology, 4(10),

1841–1851. http://doi.org/10.1371/journal.pbio.0040337

Huang, C. Y., Sabree, Z. L., & Moran, N. A. (2012). Genome sequence of Blattabacterium sp.

strain BGIGA, endosymbiont of the giganteus cockroach. Journal of Bacteriology,

194(16), 4450–4451. http://doi.org/10.1128/JB.00789-12

Huang, J., Hult, E. F., Marchal, E., & Tobe, S. S. (2015). Identification and characterization of the

NMDA receptor and its role in regulating reproduction in the cockroach Diploptera

punctata. Journal of Experimental Biology, 218, 983–990.

http://doi.org/10.1242/jeb.115154

Huang, J., Marchal, E., Hult, E. F., Zels, S., Vanden Broeck, J., & Tobe, S. S. (2014). Mode of

action of allatostatins in the regulation of juvenile hormone biosynthesis in the cockroach,

Diploptera punctata. Insect Biochemistry and Molecular Biology, 54, 61–68.

http://doi.org/10.1016/j.ibmb.2014.09.001

Hult, E. F., Huang, J., Marchal, E., Lam, J., & Tobe, S. S. (2015). RXR/USP and EcR are critical for

the regulation of reproduction and the control of JH biosynthesis in Diploptera punctata.

Journal of Insect Physiology, 80, 48–60. http://doi.org/10.1016/j.jinsphys.2015.04.006

Hussey, N. E., Wintner, S. P., Dudley, S. F. J., Cliff, G., Cocks, D. T., & Aaron MacNeil, M. (2010).

Maternal investment and size-specific reproductive output in carcharhinid . Journal

16 of Animal Ecology, 79(1), 184–193. http://doi.org/10.1111/j.1365-2656.2009.01623.x

Ingram, M. J., Stay, B., & Cain, G. D. (1977). Composition of milk from the viviparous cockroach,

Diploptera punctata. Insect Biochemistry, 7(3), 257–267. http://doi.org/10.1016/0020-

1790(77)90023-3

Jašarević, E., Rodgers, A. B., & Bale, T. L. (2015). A novel role for maternal stress and microbial

transmission in early life programming and neurodevelopment. Neurobiology of Stress,

1(1), 81–88. http://doi.org/10.1016/j.ynstr.2014.10.005

Jiang, Z. F., Xia, F., Johnson, K. W., Brown, C. D., Bartom, E., Tuteja, J. H., … Ghanim, M. (2013).

Comparison of the genome sequences of “Candidatus portiera aleyrodidarum” primary

endosymbionts of the whitefly Bemisia tabaci B and Q biotypes. Applied and

Environmental Microbiology, 79, 1757–1759. http://doi.org/10.1128/AEM.02976-12

Kalinka, A. T. (2015). How did viviparity originate and evolve? Of conflict, co-option, and cryptic

choice. BioEssays, 37(7), 721-731. http://doi.org/10.1002/bies.201400200

Kambhampati, S., Alleman, A., & Park, Y. (2013). Complete genome sequence of the

endosymbiont Blattabacterium from the cockroach Nauphoeta cinerea (:

Blaberidae). Genomics, 102(2013), 479–483. http://doi.org/10.1016/j.ygeno.2013.09.003

Kinjo, Y., Bourguignon, T., Tong, K. J., Kuwahara, H., Lim, S. J., Yoon, K. B., … Tokuda, G. (2018).

Parallel and gradual genome erosion in the Blattabacterium endosymbionts of

Mastotermes darwiniensis and wood roaches. Genome Biology and Evolution,

10(6), 1622–1630. http://doi.org/10.1093/gbe/evy110

Liu, H. P., Lin, S. C., Lin, C. Y., Yeh, S. R., & Chiang, A. S. (2005). Glutamate-gated chloride

channels inhibit juvenile hormone biosynthesis in the cockroach, Diploptera punctata.

17 Insect Biochemistry and Molecular Biology, 35(11), 1260–1268.

http://doi.org/10.1016/j.ibmb.2005.06.004

López-Sánchez, M. J., Neef, A., Patiño-Navarrete, R., Navarro, L., Jiménez, R., Latorre, A., &

Moya, A. (2008). Blattabacteria, the endosymbionts of cockroaches, have small genome

sizes and high genome copy numbers. Environmental Microbiology, 10(12), 3417–3422.

http://doi.org/10.1111/j.1462-2920.2008.01776.x

López-Sánchez, M. J., Neef, A., Peretó, J., Patiño-Navarrete, R., Pignatelli, M., Latorre, A., &

Moya, A. (2009). Evolutionary convergence and nitrogen metabolism in Blattabacterium

strain Bge, primary endosymbiont of the cockroach Blattella germanica. PLoS Genetics,

5(11), e1000721. http://doi.org/10.1371/journal.pgen.1000721

Ma, W. C., Denlinger, D. L., Järlfors, U., & Smith, D. S. (1975). Structural modulations in the

tsetse fly milk gland during a pregnancy cycle. Tissue and Cell, 7(2), 319–330.

http://doi.org/10.1016/0040-8166(75)90008-7

Macfarlane, R., & Bowers, M. (1995). Matrotrophic viviparity in the yellowtail rockfish Sebastes

flavidus. The Journal of Experimental Biology, 198, 1197–206.

Marchal, E., Hult, E. F., Huang, J., Pang, Z., Stay, B., & Tobe, S. S. (2014). Methoprene-tolerant

(Met) knockdown in the adult female cockroach, Diploptera punctata completely inhibits

ovarian development. PLoS ONE, 9(9), e106737.

http://doi.org/10.1371/journal.pone.0106737

Marchal, E., Hult, E. F., Huang, J., Stay, B., & Tobe, S. S. (2013). Diploptera punctata as a model

for studying the endocrinology of arthropod reproduction and development. General and

Comparative Endocrinology, 188, 85–93.

18 http://doi.org/http://dx.doi.org/10.1016/j.ygcen.2013.04.018

McCutcheon, J. P., & Moran, N. A. (2010). Functional Convergence in Reduced Genomes of

Bacterial Symbionts Spanning 200 My of Evolution. Genome Biology and Evolution, 2(1),

708–718. http://doi.org/10.1093/gbe/evq055

McCutcheon, J. P., & Moran, N. A. (2012). Extreme genome reduction in symbiotic bacteria.

Nature Reviews Microbiology, 10(1), 13–26. http://doi.org/10.1038/nrmicro2670

Meier, R., Kotrba, M., & Ferrar, P. (1999). and viviparity in the Diptera. Biological

Reviews, 74, 199–258. http://doi.org/10.1111/j.1469-185X.1999.tb00186.x

Michalik, A., Szklarzewicz, T., Jankowska, W., & Wieczorek, K. (2014). Endosymbiotic

microorganisms of aphids (Hemiptera: Sternorrhyncha: Aphidoidea): Ultrastructure,

distribution and transovarial transmission. European Journal of Entomology, 111(1), 91–

104. http://doi.org/10.14411/eje.2014.011

Michalkova, V., Benoit, J. B., Attardo, G. M., Medlock, J., & Aksoy, S. (2014). Amelioration of

reproduction-associated oxidative stress in a viviparous insect is critical to prevent

reproductive senescence. PloS One, 9(4), e87554.

http://doi.org/10.1371/journal.pone.0087554

Michalkova, V., Benoit, J. B., Weiss, B. L., Attardo, G. M., & Aksoy, S. (2014). Vitamin B6

generated by obligate symbionts is critical for maintaining proline homeostasis and

fecundity in tsetse flies. Applied and Environmental Microbiology, 80(18), 5844–5853.

http://doi.org/10.1128/AEM.01150-14

Moran, N. A., & Wernegreen, J. J. (2000). Lifestyle evolution in symbiotic bacteria: insights from

genomics. Trends in Ecology & Evolution, 15(8), 321–326. http://doi.org/10.1016/S0169-

19 5347(00)01902-9

Morse, S. F., Bush, S. E., Patterson, B. D., Dick, C. W., Gruwell, M. E., & Dittmar, K. (2013).

Evolution, multiple acquisition, and localization of endosymbionts in bat flies (Diptera:

Hippoboscoidea: Streblidae and Nycteribiidae). Applied and Environmental Microbiology,

79(9), 2952–2961. http://doi.org/10.1128/AEM.03814-12

Nakabachi, A., Yamashita, A., Toh, H., Ishikawa, H., Dunbar, H. E., Moran, N. A., & Hattori, M.

(2006). The 160-Kilobase genome of the bacterial endosymbiont Carsonella. Science,

314(5797), 267–267. http://doi.org/10.1126/science.1134196

Nalepa, C. A., Bignell, D. E., & Bandi, C. (2001). Detritivory, coprophagy, and the evolution of

digestive mutualisms in Dictyoptera. Insectes Sociaux, 48(3), 194–201.

http://doi.org/10.1007/PL00001767

Neef, A., Latorre, A., Peretó, J., Silva, F. J., Pignatelli, M., & Moya, A. (2011). Genome

economization in the endosymbiont of the wood roach due to

drastic loss of amino acid synthesis capabilities. Genome Biology and Evolution, 3(1), 1437–

1448. http://doi.org/10.1093/gbe/evr118

Nikoh, N., Hosokawa, T., Oshima, K., Hattori, M., & Fukatsu, T. (2011). Reductive Evolution of

Bacterial Genome in Insect Gut Environment. Genome Biology and Evolution, 3(1), 702–

714. http://doi.org/10.1093/gbe/evr064

Ostrovsky, A. N., Lidgard, S., Gordon, D. P., Schwaha, T., Genikhovich, G., & Ereskovsky, A. V.

(2016). Matrotrophy and placentation in invertebrates: a new paradigm. Biological

Reviews of the Cambridge Philosophical Society, 91(3), 673–711.

http://doi.org/10.1111/brv.12189

20 Pais, R., Lohs, C., Wu, Y., Wang, J., & Aksoy, S. (2008). The obligate mutualist Wigglesworthia

glossinidia influences reproduction, digestion, and immunity processes of its host, the

tsetse fly. Applied and Environmental Microbiology, 74(19), 5965–5974.

http://doi.org/10.1128/AEM.00741-08

Patiño-Navarrete, R., Moya, A., Latorre, A., & Peretó, J. (2013). Comparative genomics of

Blattabacterium cuenoti: The frozen legacy of an ancient endosymbiont genome. Genome

Biology and Evolution, 5(2), 351–361. http://doi.org/10.1093/gbe/evt011

Patino-Navarrete, R., Piulachs, M.-D., Belles, X., Moya, A., Latorre, A., & Pereto, J. (2014). The

cockroach Blattella germanica obtains nitrogen from uric acid through a metabolic

pathway shared with its bacterial endosymbiont. Biology Letters, 10(7), 20140407–

20140407. http://doi.org/10.1098/rsbl.2014.0407

Roth, L. M. (1967). Water changes in cockroach oothecae in relation to the evolution of

ovoviviparity and viviparity. Annals of the Entomological Society of America, 60(5), 928–

946. http://doi.org/10.1093/aesa/60.5.928

Roth, L. M., & Hahn, W. (1964). Size of new-born larvae of cockroaches incubating eggs

internally. Journal of Insect Physiology. 10, 65-72. http://doi.org/10.1016/0022-

1910(64)90094-0

Roth, L. M., & Stay, B. (1961). Oöcyte development in Diploptera punctata (Eschscholtz)

(Blattaria). Journal of Insect Physiology. 7, 186-202. http://doi.org/10.1016/0022-

1910(61)90071-3

Roth, L. M., & Willis, E. R. (1955). Intra-uterine nutrition of the “beetle-roach” Diploptera

dytiscoides (serv.) during embryogenesis, with notes on its biology in the laboratory

21 (Blattaria: Diplopteridae). Psyche: A Journal of Entomology. 62, 55-68.

http://doi.org/10.1155/1955/12542

Roth, L. M., & Willis, E. R. (1957). An analysis of oviparity and viviparity in the Blattaria.

American Entomological Society, 83(4), 221–238.

Sabree, Z. L., Huang, C. Y., Arakawa, G., Tokuda, G., Lo, N., Watanabe, H., & Moran, N. A. (2012).

Genome shrinkage and loss of nutrient-providing potential in the obligate symbiont of the

primitive termite . Applied and Environmental Microbiology,

78(1), 204–210. http://doi.org/10.1128/AEM.06540-11

Sabree, Z. L., Kambhampati, S., & Moran, N. a. (2009). Nitrogen recycling and nutritional

provisioning by Blattabacterium, the cockroach endosymbiont. Proceedings of the National

Academy of Sciences of the United States of America, 106(46), 19521–19526.

http://doi.org/10.1073/pnas.0907504106

Salem, H., Florez, L., Gerardo, N., & Kaltenpoth, M. (2015). An out-of-body experience: the

extracellular dimension for the transmission of mutualistic bacteria in insects. Proceedings

of the Royal Society B: Biological Sciences, 282(1804), 20142957–20142957.

http://doi.org/10.1098/rspb.2014.2957

Santos-Garcia, D., Latorre, A., Moya, A., Gibbs, G., Hartung, V., Dettner, K., … Silva, F. J. (2014).

Small but powerful, the primary endosymbiont of moss bugs, Candidatus evansia muelleri,

holds a reduced genome with large biosynthetic capabilities. Genome Biology and

Evolution, 6(7), 1875–1893. http://doi.org/10.1093/gbe/evu149

Schultz, T. J., Webb, J. K., & Christian, K. A. (2008). The physiological cost of pregnancy in a

tropical viviparous . Copeia, 2008(3), 637–642. http://doi.org/10.1643/CP-06-182

22 Schwab, D. B., Riggs, H. E., Newton, I. L. G., & Moczek, A. P. (2016). Developmental and

ecological benefits of the maternally transmitted microbiota in a dung beetle. The

American Naturalist, 188(6), 679–692. http://doi.org/10.1086/688926

Shaffer, L. R., & Formanowicz, D. R. (1996). A cost of viviparity and parental care in :

Reduced sprint speed and behavioural compensation. Animal Behaviour, 51 (5), 1017–

1023. http://doi.org/DOI 10.1006/anbe.1996.0104

Shigenobu, S., Watanabe, H., Hattori, M., Sakaki, Y., & Ishikawa, H. (2000). Genome sequence of

the endocellular bacterial symbiont of aphids Buchnera sp. APS. Nature, 407(6800), 81–86.

http://doi.org/10.1038/35024074

Shine, R. (1983). Reptilian reproductive modes: The oviparity viviparity continuum.

Herpetologica, 39(1), 1–8.

Shukla, S. P., Vogel, H., Heckel, D. G., Vilcinskas, A., & Kaltenpoth, M. (2018). Burying

regulate the microbiome of carcasses and use it to transmit a core microbiota to their

offspring. Molecular Ecology, 27(8), 1980–1991. http://doi.org/10.1111/mec.14269

Snyder, A. K., Mclain, C., & Rio, R. V. M. (2012). The tsetse fly obligate mutualist Wigglesworthia

morsitans alters gene expression and population density via exogenous nutrient

provisioning. Applied and Environmental Microbiology, 78(21), 7792–7797.

http://doi.org/10.1128/AEM.02052-12

Snyder, A. K., & Rio, R. V. M. (2015). “Wigglesworthia morsitans” folate (vitamin B9)

biosynthesis contributes to tsetse host fitness. Applied and Environmental Microbiology,

81(16), 5375–5386. http://doi.org/10.1128/AEM.00553-15

Stay, B., & Clark, J. K. (1971). Fluctuation of protein granules in the fat body of the viviparous

23 cockroach, Diploptera punctata, during the reproductive cycle. Journal of Insect

Physiology, 17, 1747–1762. http://doi.org/10.1016/0022-1910(71)90072-2

Stay, B., & Coop, A. (1973). Developmental stages and chemical composition in embryos of the

cockroach, Diploptera punctata, with observations on the effect of diet. Journal of Insect

Physiology, 19(1), 147–171. http://doi.org/10.1016/0022-1910(73)90230-8

Stay, B., & Coop, A. C. (1974). “Milk” secretion for embryogenesis in a viviparous cockroach.

Tissue and Cell, 6(4), 669–693. http://doi.org/10.1016/0040-8166(74)90009-3

Stay, B., & Roth, L. M. (1956). The reproductive behavior of Diploptera punctata (Blattaria:

Diplopterinae). Proceedings of the International Congress of Entomology. 6(4), 669-693.

Stewart, J. R. (2015). A universal model for the evolution of viviparity. BioEssays, 37, 1.

http://doi.org/10.1002/bies.201400200

The NIH HMP Working Group. (2009). The NIH Human Microbiome Project. Genome Research,

19(12), 2317–2323. http://doi.org/10.1101/gr.096651.109

Tobe, S. S., & Langley, P. A. (1978). Reproductive physiology of Glossina. Annual Review of

Entomology, 23(1), 283–307. http://doi.org/10.1146/annurev.en.23.010178.001435

Tokuda, G., Elbourne, L. D. H., Kinjo, Y., Saitoh, S., Sabree, Z. L., Hojo, M., … Lo, N. (2013).

Maintenance of essential amino acid synthesis pathways in the Blattabacterium cuenoti

symbiont of a wood-feeding cockroach. Biology Letters, 9, 20121153.

http://doi.org/10.1098/rsbl.2012.1153

Tworzydlo, W., Kisiel, E., & Bilinski, S. M. (2013). Embryos of the viviparous dermapteran,

Arixenia esau develop sequentially in two compartments: terminal ovarian follicles and the

uterus. PLoS ONE, 8(5), e64087. http://doi.org/10.1371/journal.pone.0064087

24 van Ham, R. C. H. J., Kamerbeek, J., Palacios, C., Rausell, C., Abascal, F., Bastolla, U., … Moya, A.

(2003). Reductive genome evolution in Buchnera aphidicola. Proceedings of the National

Academy of Sciences, 100(2), 581–586. http://doi.org/10.1073/pnas.0235981100

Wang, J., Weiss, B. L., & Aksoy, S. (2013). Tsetse fly microbiota: form and function. Frontiers in

Cellular and Infection Microbiology, 3, 1–6. http://doi.org/10.3389/fcimb.2013.00069

Weiss, B. L., Wang, J., & Aksoy, S. (2011). Tsetse immune system maturation requires the

presence of obligate symbionts in larvae. PLoS Biology, 9(5), e1000619.

http://doi.org/10.1371/journal.pbio.1000619

Wernegreen, J. J. (2002). Genome evolution in bacterial endosymbionts of insects. Nature

Reviews Genetics, 3(11), 850–861. http://doi.org/10.1038/nrg931

Wernegreen, J. J. (2004). Endosymbiosis: Lessons in Conflict Resolution. PLoS Biology, 2(3), e68.

http://doi.org/10.1371/journal.pbio.0020068

Williford, A., Stay, B., & Bhattacharya, D. (2004). Evolution of a novel function: Nutritive milk in

the viviparous cockroach, Diploptera punctata. Evolution and Development, 6, 67–77.

http://doi.org/10.1111/j.1525-142X.2004.04012.x

Wilson, A. C. C., Ashton, P. D., Calevro, F., Charles, H., Colella, S., Febvay, G., … Douglas, A. E.

(2010). Genomic insight into the amino acid relations of the pea aphid, Acyrthosiphon

pisum, with its symbiotic bacterium Buchnera aphidicola. Insect Molecular Biology, 19,

249–258. http://doi.org/10.1111/j.1365-2583.2009.00942.x

Wolf, Y. I., & Koonin, E. V. (2013). Genome reduction as the dominant mode of evolution.

BioEssays, 35(9), 829–837. http://doi.org/10.1002/bies.201300037

25

Chapter 2

RNA-seq analysis sheds light on the molecular mechanisms underlying

the control of live birth in the cockroach, Diploptera punctata

Emily C. Jennings 1 *

Jacob M. Hendershot1, Matthew T. Weirauch2, Jose M. C. Ribeiro3, and Joshua B. Benoit1

1 Department of Biological Sciences, University of Cincinnati

2 Center for Autoimmune Genomics and Etiology and Divisions of Biomedical Informatics and

Developmental Biology, Cincinnati Children’s Hospital Medical Center

3 Laboratory of Malaria and Vector Research, National Institute of Allergy and Infectious

Diseases

* The work presented here was principally composed by Emily C Jennings with assistance

from the listed co-authors.

26 Abstract

Viviparous reproduction is characterized by maternal retention of developing offspring within her reproductive tract during gestation, culminating in live birth. In some cases, a mother will provide nutrition beyond that present in the yolk; this is known as matrotrophic viviparity.

While this phenomenon is best associated with mammals, it is observed in insects such as the viviparous cockroach, Diploptera punctata. Female D. punctata carry developing embryos in the brood sac, a reproductive organ that acts as both a uterus and placenta by protecting and providing a nutritive secretion to the intrauterine developing progeny. While the basic physiology of D. punctata pregnancy has been characterized, little is known about the molecular mechanisms underlying this phenomenon. This study combines RNA-seq analysis,

RNA interference, and other assays to characterize molecular and physiological changes associated with D. punctata reproduction, providing the most complete gene set to date for this species. A comparison of four stages of the female reproductive cycle revealed unique gene expression profiles corresponding to each stage. Differentially regulated transcripts of interest include the previously identified family of milk proteins and transcripts associated with juvenile hormone metabolism. RNA interference experiments reveal potential impacts of juvenile hormone breakdown in maintaining pregnancy in D. punctata.

27 Introduction

Within the animal kingdom there exist three main reproductive strategies: oviparity or the laying of eggs, facultative viviparity (encompassing ovoviviparity to aplacental viviparity) in which a mother may lay eggs or give birth to live, active offspring, and viviparity (Hagan, 1948;

Meier, Kotrba, & Ferrar, 1999; Roth & Willis, 1957; Stewart, 2015). True viviparous reproduction is characterized by the obligate retention of and nutritional provisioning to developing progeny in the maternal reproductive tract for the duration of gestation, which culminates in the birth of active offspring; early termination of the pregnancy results in stillborn offspring (Hagan, 1948; Kalinka, 2015). While most insects are oviparous, facultative or true viviparity has been observed in at least 13 orders of insects (Clutton-Brock, 1991; Hagan, 1948;

Roth & Willis, 1957). One such viviparous insect is the pacific beetle mimic cockroach,

Diploptera punctata (Ingram, Stay, & Cain, 1977; Roth & Hahn, 1964; Roth & Stay, 1961; Roth &

Willis, 1955; Stay & Coop, 1974).

Diploptera punctata, a Polynesian native species, reproduces by matrotrophic viviparity

(Figure 1A and B) (Ingram et al., 1977; Roth & Hahn, 1964; Roth & Stay, 1961; Roth & Willis,

1955; Stay & Coop, 1974). Embryos develop inside the brood sac, a unique organ which functions as both a uterus and pseudo-placenta (Figure 1B); embryos are provided with nutrients by a secretion of milk-like components (Roth & Stay, 1961; Roth & Willis, 1955; Stay &

Clark, 1971; Stay & Coop, 1973; Stay & Roth, 1956). Crystalized milk secretion accumulates in embryo gut contents at 20% of the 60-70-day gestation period, at which point the dorsal edge of the body wall has closed. Diploptera milk is a combination of proteins and free amino acids, carbohydrates, and lipids in a water base (Stay & Coop, 1973, 1974; Williford, Stay, &

28 Bhattacharya, 2004). The proteins present include a family of milk proteins derived from 25 unique mRNAs coding for 22 different proteins (Stay & Coop, 1973, 1974; Williford et al., 2004).

During its gestation period, between nine and thirteen embryos grow from 1.5 mm at ovulation to over 6mm at birth, increasing in weight by more than 70-fold. Water content of each egg increases by 85-fold, and solid dry weight increases over 49-fold (Roth, 1967; Roth & Hahn,

1964; Roth & Stay, 1961; Roth & Willis, 1955; Stay & Coop, 1973, 1974; Stay & Roth, 1956).

While the basic physiology of D. punctata pregnancy has been characterized, little is known about the molecular mechanisms underlying this phenomenon. This study combines RNA-seq analysis, RNA interference, and other assays to characterize molecular changes associated with

D. punctata reproduction. We present a transcriptome containing 11,987 coding DNA sequences, 2,474 of which are differentially regulated across pregnancy in D. punctata.

Additionally, gene knockdown of juvenile hormone esterase indicates a key role for this enzyme in maintaining milk protein production but does not act alone for complete maintenance of pregnancy.

Methods

Animals

Colonies were reared at the University of Cincinnati in a climate-controlled facility. Ambient temperature was held between 24-28°C and relative humidity (RH) was held between 70-80%.

A 12:12 hour light-dark photoperiod was maintained for the duration of the experiment.

Animals were provided water and fed Old Roy Complete Nutrition brand dog food (Mars, Inc.) ad libitum.

29

Sample collection

Females used in transcriptome analysis are divided into four categories: mated but not pregnant, pre-lactation, early lactation, and late lactation (Table 1). Pregnant females harboring embryos less than 1.6 mm in length are considered pre-lactation; those with embryos between

1.6 and 2.5 mm are early lactation; and those with embryos greater than 2.5 mm are considered late lactation. Lastly, mated but not pregnant females are characterized by the presence of a spermatophore in the bursa copulatrix and absence of embryos in the brood sac.

These parameters are based on those reported by Stay and Coop (1973). Females were randomly chosen from the colony and embryos were carefully dissected out of the brood sac in sterile PBS and immediately measured. Once the samples were categorized, embryos were discarded, and due to their size, females were cut in half at the junction between the thorax and abdomen and stored in Trizol (Invitrogen) at -80°C until processed.

RNA extraction and library preparation

Total RNA was extracted from whole animals or females without developing embryos using

Trizol and then treated with DNase I (Thermo Scientific) to remove genomic DNA. Samples were then treated with a GeneJET RNA Cleanup and Concentration Micro Kit (Thermo Scientific) to further remove contamination. For pregnant female samples, equal parts of RNA extracted from the anterior and posterior were combined before DNase I treatment. Total RNA was pooled from 3 individuals – each extracted individually – for each replicate; a second set of samples was prepared for qPCR validation of select transcripts. In addition, three male samples

30 each consisting of a pool of three individuals and one sample containing equal amounts of RNA from a male, first instar nymph, pregnant female, and non-pregnant female were collected for sequencing. RNA concentration and quality were examined via A260/A280 and A260/A230 with a NanoDrop 2000 (Thermo Scientific).

Poly(A) libraries were prepared by the DNA Sequencing and Genotyping Core at

Cincinnati Children’s Hospital Medical Center. RNA was quantified using a Qubit 3.0

Fluorometer (Life Technologies). Total RNA (150-300 ng) was poly(A) selected for reverse transcription using a TruSeq Stranded mRNA Library Preparation Kit (Illumina). Multiplexing was conducted by ligating an 8-base molecular barcode to sequences before 15 cycles of PCR amplification and HiSeq 2500 (Illumina) Rapid Mode library sequencing. Sequencing resulted in

30-40 million single end strand specific reads per sample at either 75bp or 100 bp long.

Bioinformatic analysis

Initial bioinformatic analyses (e.g. de novo assembly of contigs, identification of coding sequences, and basic characterization of coding sequences) were conducted as previously described (Ribeiro, Genta, et al., 2014; Ribeiro, Chagas, Pham, Lounibos, & Calvo, 2014). Briefly, low quality regions and reads of low quality are removed using RTA 1.12.4.2 (Illumina) and

CASAVA 1.8.2 (Illumina). The remaining high-quality reads are used to generate de novo assemblies with both ABySS software (Birol et al., 2009; Simpson et al., 2009) and the

SOAPdenovo-Trans assembler (Luo et al., 2012), which were joined by an iterative BLAST and cap3 assembler. Coding DNA sequences were determined based on similarity to known proteins or by obtaining CDS containing a signal peptide. Annotation utilized multiple databases

31 including Swissprot, Gene Ontology, KOG, Pfam, and SMART as well as the non-redundant protein database of the National Center for Biotechnology Information. Additional manual annotation was conducted as needed. Further description of this procedure can be found in previous publications (Budachetri & Karim, 2015; Ribeiro, Chagas, et al., 2014). Assessment of differential expression was conducted using the EDGE method with default settings in CLC

Genomics Workbench (QIAGEN) to generate normalized expression values as the number of transcripts per kilobase per million (TPM); false detection rate (FDR) corrected p-values less than 0.05 were considered significant. Differential enrichment of functional annotations was conducted using annotation and Fisher’s Exact test functions in Blast2GO5 Basic (BioBam), and the Gene Set Enrichment Analysis (GSEA) function of CLC Genomics Workbench.

RNA interference

Dicer-substrate short interfering RNAs (siRNA) consisting of two duplex sequences were designed for the D. punctata juvenile hormone esterase transcript (jhe) and green florescent protein (gfp) as a control using the IDT custom siRNA design tool (Integrated DNA Technologies)

(Table 2). The siRNA was reconstituted in nuclease free water according to manufacturer specifications and diluted to 1µg/µl. Pregnant females (status was confirmed by the presence of embryos in the brood sac) were collected from the colony and grouped based on weight.

Cockroaches were injected with 2 µl of either siGFP or siJHE according to the following schedule based on previous studies in D. punctata (Hult, Huang, Marchal, Lam, & Tobe, 2015). The first injection was considered as day 0 and subsequent injections were given on day 2, day 4, day 5,

32 and day 6. Embryos were removed from cold anesthetized females on day 7 and female tissue was stored in Trizol (Invitrogen) at -80° C until RNA was extracted.

cDNA synthesis and quantitative PCR

Extracted RNA was diluted to a concentration of 200 ng/µl for use in reverse transcription reactions. Complimentary DNA (cDNA) was synthesized using a DyNAmo cDNA Synthesis Kid

(Thermo Scientific) from 1µg of RNA for gene specific expression validation and 250 ng of RNA for RNAi knockdown experiments. KiCqStart SYBR Green qPCR ReadyMix (Sigma Aldrich) was utilized in all reactions with gene specific primers designed using Primer3 (Hancock & Zvelebil,

2014) (Table 2). Quantitative PCR was conducted in an Illumina Eco quantitative PCR system; reactions were run according to a previous study (Rosendale, Romick-Rosendale, Watanabe,

Dunlevy, & Benoit, 2016). Relative expression of genes of interest was calculated with the DDCq method (Schmittgen & Livak, 2008) using Elongation factor 1 alpha (Ef1a) for normalization. For

RNA-seq expression validation, fold change in the genes of interest was calculated relative to non-pregnant females and the logarithmic fold change was plotted against the corresponding value from the RNA-seq analysis to calculate a Pearson correlation coefficient (r). Relative mRNA expression levels for each gene of interest, calculated as described for RNA-seq validation, were compared between treatment groups in RStudio (R Core Team, 2017; RStudio

Team, 2015) and were compared between controls and knockdowns using a Wilcox test.

33 Data processing and visualization

Data processing was conducted in Microsoft Excel (v.16.22) and R (v.3.3.3) (R Core

Team, 2017) using RStudio (v1.1.423) (RStudio Team, 2015). Additional statistics and graphical representations of data were also performed in R using RStudio. Packages utilized include dplyr

(Wickham, Francois, Henry, & Müller, 2017), ggplot2 (Wickham, 2016), reshape2 (Wickham,

2007), RColorBrewer (Neuwirth, 2014), Rmisc (Hope, 2013), wesanderson (Ram & Wickham,

2018), and yarrrr (Phillips, 2017).

Results

General assembly characteristics

Quality control processing of the raw sequencing data produced 460,755,803 single end reads which were assembled into 102,880 contigs (Table 3). Representing 10,820,347 bases, 11,987 of those contigs were determined to be coding sequences (Table 3) and 3,289 of them had signatures indicative of signaling peptides (i.e., exported from the cell). BUSCO analysis revealed that 80.5% of the benchmarking universal single copy orthologs were represented as complete genes in our gene set with another 2.3% present as fragments (Kriventseva et al.

2015), indicating that the assembly is of good quality (Figure 2A). Contigs were annotated by searching NCBI’s arthropod non-redundant database and Swissprot using BLASTx; 77.8% of extracted CDS matched a protein in the nr arthropod database with an e-value ≤ 0.001. The termite Zootermopsis nevadensis appeared as the species with the highest similarity for 57.30% of BLAST annotations followed by Tribolium castaneum, representing only 2.85% of BLAST top- hits (Figure 2B).

34 Our analysis revealed 2,474 contigs that are differentially expressed after mating across

D. punctata pregnancy. In these 2,474 contigs, 75 Gene Ontology (GO) terms were overrepresented (Figure 3A). Overrepresented terms included biological processes such as cytoskeleton organization, cell cycle processes, morphogenesis of epithelium, multiple metabolic and biosynthetic processes; cellular components like ribosomes; and molecular functions associated with cuticle components and the binding of protein and carbohydrates.

Each stage of pregnancy had a visibly and statistically unique profile of differentially regulated genes associated with an equally unique sets of GO terms (Table 4). Using an FDR corrected p- value cut off of 0.01 and a minimum absolute value of fold change of 1.5, we detected 163 pre- lactation-specific genes (38 upregulated and 125 downregulated), 414 early lactation-specific genes (123 upregulated and 291 downregulated), and 287 late lactation-specific (41 upregulated and 246 downregulated) (Figure 3B). Pregnant females producing milk secretion show additional enrichment of GO terms for protein metabolic process, cuticle development, transposition, and RNA splicing relative to non-pregnant females while pre lactation females have increased representation of terms for inorganic ion transport, generation of precursor metabolites and energy, cell differentiation, nucleic acid metabolism, and mitochondria related energy production (Table 4).

To identify specific male expressed genes, we selected CDS with five-fold or greater expression in males when compared to both late lactation females and mated but not pregnant females with an FDR p-value cutoff ≤ 0.01. This analysis produced 286 CDS with male specific expression (Figure 4A). When annotated using BLASTx against the Zootermopsis nevadensis nr database subset and arthropod nr database, only 64 of the 286 CDS had blast matches with an

35 e-value ≤ 0.001. GO analysis revealed an enrichment of serine-type endopeptidase activity, serine-type peptidase activity, serine hydrolase activity and general hydrolase and catalytic activities; also enriched were the biological processes for proteolysis and protein metabolic processes (Figure 4B).

Differential expression of transcriptional regulators

We also sought to identify transcription factors that could be associated with or driving reproduction in D. punctata. Bioinformatic analyses identified 329 putative transcription factors in our assembly, using methods previously described (Schoville et al., 2018). These transcription factors belong to 35 structural families, with C2H2 zinc fingers being the most abundant followed by homeodomains, bHLH, bZIP and then Myb domains (Figure 5A). Of the corresponding 329 transcripts, ten are differentially regulated, including krüppel homolog 1 and ftz-f1, which are known to interact with juvenile hormone as well as ecdysone, another important insect hormone (Figure 5B).

Reproduction specific gene expression

Previous research of Diploptera punctata reproduction resulted in the identification of 25 mRNA sequences that appear to encode the proteins that constitute the main nutritional content of the milky secretion provided to intrauterine developing embryos during pregnancy

(Evans & Stay, 1994; Ingram et al., 1977; Stay & Coop, 1974; Williford et al., 2004). These proteins have been shown to increase in concentration across pregnancy (Evans & Stay, 1994;

Ingram et al., 1977; Stay & Coop, 1974; Williford et al., 2004). Using the available mRNA clone

36 sequences from the NCBI database as a reference for mapping our RNA-seq reads, we calculated their expression in the four reproductive stages surveyed. The pattern of expression identified in our analysis combined with the previously conducted mRNA clone sequencing and protein sequence analysis confirms that these sequences encode the milk-proteins and that their transcription mirrors the secretion patterns (Figure 6A). Additionally, we identified and characterized the expression of transcripts for vitellogenin, the main yolk protein precursor

(Figure 6B) and juvenile hormone esterase (jhe), the enzyme primarily responsible for breaking down and suppressing levels of juvenile hormone to allow the maintenance of pregnancy (Rotin

& Tobe, 1983). Both were differentially regulated across pregnancy and as expected inversely mirrored the known titer of juvenile hormone (Figure 6C).

Knockdown of JHE

In our RNA-seq data set we identified a putative transcript encoding JH esterase, which was differentially regulated across female reproduction (Figure 6C). To determine the degree to which the increased expression of the JH esterase transcript contributes to the low JH titer controlling milk production, we utilized short interfering RNA (siRNA) to transiently knock down the expression of JH esterase. We measured relative expression of a milk protein clone MP13Z, ecdysone receptor (EcR) and transcripts known to respond to juvenile hormone levels, vitellogenin (Vg) and krüppel homolog 1 (Krh1). Relative to the siGFP RNAi injected females, the siJHE individuals demonstrated a significant reduction in JHE expression, indicating a successful knockdown of the transcript (Figure 7). There was also a significant decrease in levels of the milk protein clone MP13Z, indicating that JHE contributes to maintaining low circulating levels

37 of juvenile hormone in pregnant D. punctata. However, we did not see a corresponding increase in vitellogenin; rather, the levels of Vg transcripts did not vary between the siGPF controls or the siJHE knockdowns, but both groups had high variability in the expression of this gene. While we had predicted that Krh1 would increase in expression due to an increase in circulating JH, we instead saw a reduction of this transcript in siJHE knockdowns relative to siGFP controls. Additionally, there was no change in the expression of EcR (Figure 7).

Discussion

Utilizing RNA-seq analysis we assembled an 102,880 contig de novo transcriptome of D. punctata. BLAST analysis of the extracted 11,987 CDS revealed high sequence similarity with the termite Zootermopsis nevadensis; until recently, Z. nevadensis represented the most complete and well annotated Dictyopteran genome available at the time of analyses (Terrapon et al., 2014). The size of our transcriptome is comparable to recent transcriptomes for the

American cockroach, americana, containing 85,984 contigs with 17,744 annotated

(Kim et al., 2016), as well as for the lobster cockroach, Nauphoeta cinerea, which had 57,928 assembled contigs (Segatto, Diesel, Loreto, & da Rocha, 2018).

Differential expression analysis revealed 2,474 transcripts with significantly different expression relative to non-pregnant females enriched in predicted functions such as organonitrogen compound metabolism and protein metabolic processes. This heavy investment into nitrogen and protein metabolism is unsurprising considering the well documented maternally synthesized protein component of the milk-like secretion consumed by intrauterine developing embryos (Banerjee et al., 2016; Ingram et al., 1977; Marchal, Hult, Huang, Stay, &

38 Tobe, 2013; Stay & Coop, 1974; Williford et al., 2004). Through our male-female comparisons, we also identified a set of 286 genes with male specific expression. These genes generally lacked similarity to existing annotations in available databases but were enriched for processes such as metal ion binding, peptidase activity, proteolysis, and phosphatase activities among others. Commonly seen enriched in transcriptomes of male reproductive organs, these enzymatic activities have been linked to a variety of post-mating changes in female reproductive physiology such as ovulation and storage (Bretãs et al., 2012; Gotoh et al.,

2018; McGraw, Gibson, Clark, & Wolfner, 2004; Sirot, Rubinstein, Avila, Wolfner, & LaFlamme,

2010). Male specific transcriptome libraries often contain sequences with little to no homology to known genes and functions (Attardo et al., 2019; Bretãs et al., 2012; Gotoh et al., 2018;

Meibers et al., 2019; Wei et al., 2016); this is attributed to what are known as male accessory gland or seminal proteins since they are undergoing rapid evolution (Attardo et al., 2019; Bretãs et al., 2012; Gotoh et al., 2018; Meibers et al., 2019; Sirot et al., 2010; Wei et al., 2016), explaining the lack of annotations associated with our male enriched gene set.

Our data also support the previously characterized increase in milk protein synthesis and secretion (Evans & Stay, 1994; Ingram et al., 1977; Stay & Coop, 1974; Williford et al.,

2004), confirming that this increase is controlled at the transcriptional level. It is exceptionally notable that while milk proteins are synthesized by the cells of the brood sac epithelium

(Hagan, 1941; Ingram et al., 1977; Stay & Coop, 1974; Williford et al., 2004), the magnitude of milk cDNA expression is so great that it is not masked by the sequencing of a whole-body RNA sample, despite the fact that this organ makes up a relatively small amount of the body mass of a female D. punctata. Juvenile hormone (JH), produced in the corpora allata (CA) of insects is

39 known for its roles in development and reproduction (de Kort & Granger, 1996; Engelmann &

Mala, 2000; Rankin & Stay, 1984; Rotin & Tobe, 1983; Stoltzman, Stocker, Borst, & Stay, 2000).

The fluctuations of this hormone have been extensively characterized across female reproduction in insects (de Kort & Granger, 1996; Rankin & Stay, 1984; Rotin & Tobe, 1983;

Stoltzman, Stocker, Borst, & Stay, 2000). During D. punctata reproduction, JH gradually increases in the first several days post mating, corresponding to development and , peaking at roughly five days post mating at peak yolk deposition (de Kort &

Granger, 1996; Rotin & Tobe, 1983; Stoltzman et al., 2000; Tobe et al., 1985). Shortly after, JH levels drop indicative of vitellogenesis and ovipositioning into the brood sac, remaining low until the time of parturition (de Kort & Granger, 1996; Rotin & Tobe, 1983; Stoltzman et al.,

2000; Tobe et al., 1985). JH titers are modulated in part by juvenile hormone esterase (JHE) activity; the activity of this enzyme shows an inverse relationship with JH levels across reproduction indicating that combined with fluctuating biosynthesis (Marchal et al., 2013;

Paulson & Stay, 1987; Tobe et al., 1985), JHE maintains the low JH titers required for maintenance of pregnancy (de Kort & Granger, 1996; Rotin & Tobe, 1983; Stoltzman et al.,

2000; Tobe et al., 1985). Our RNA-seq analysis suggests that this increase in enzymatic activity is being controlled at the transcriptional level through increased expression of jhe, similar to what has been seen in the distantly related viviparous tsetse fly (Baumann et al., 2013). Of note, we did not see a change in the expression of transcripts encoding juvenile hormone biosynthesis corresponding to the JH titer as previously described in D. punctata (Couillaud &

Feyereisen, 1991; Hult et al., 2015; Marchal et al., 2013; Paulson & Stay, 1987; Stay, Tobe,

Mundall, & Rankin, 1983; Tobe et al., 1985) or other insects (Baumann et al., 2013; Borras-

40 Castells, Nieva, Maestro, Maestro, Belles and Martín, 2017; Dominguez & Maestro, 2017;

Ishikawa et al., 2012; Naghdi, Maestro, Belles, & Bandani, 2016; Qu, Bendena, Tobe, & Hui,

2018; Rankin, Palmer, Yagi, Scott, & Tobe, 1995). Although it is possible the synthesis of juvenile hormone is not downregulated during pregnancy, the production of juvenile hormone occurs in the corpora allata (Marchal et al., 2013; Paulson & Stay, 1987; Tobe et al., 1985), a part of the insect central nervous system that is relatively small compared to the amount of tissue from which RNA was extracted for this study. Consequently, it is probable that changes in expression of these transcripts are being masked by our sampling technique and tissue specific sampling would reveal differential expression of JH biosynthesis.

To further characterize the role of the differential expression of jhe transcripts, we performed an RNA interference knockdown of jhe. Our knockdown experiments suggest that this transcriptional regulation of jhe is a key component in maintaining milk production.

Interestingly, our knockdown was able to inhibit milk production within seven days of treatment but did not cause termination of any in that time. Additionally, our knockdown produced unexpected patterns of expression of Khr1 and vitellogenin. Krh1 and Vg are known for their strong expression in response to JH (Baumann et al., 2013; Hult et al., 2015;

Konopova, Smykal, & Jindra, 2011; Marchal et al., 2013; Minakuchi, Namiki, & Shinoda, 2009;

Stay & Clark, 1971; Wilson, Landers, & Happ, 1983); however while our knockdown increased

JH enough to suppress milk protein production, it did not change the expression levels of Vg and in fact decreased expression of Krh1, which usually increases as JH levels rise. It is possible that the change in JH levels was high enough to prevent milk production but was not maintained at high enough level for sufficient time to trigger Vg transcription. As for Krh1, a

41 previous study increasing JH production by knocking down the retinoid X receptor/ultraspiracle

(RXR/USP) complex and ecdysone receptor (EcR) found organ-specific changes in expression of

Krh1 (Hult et al., 2015). Specifically, increased JH biosynthesis decreased Krh1 expression in corpora allata, increased it in the , and expression in the fat body was not impacted by changes in JH alone (Hult et al., 2015). Additionally, we saw that during pregnancy, Krh1 levels are dramatically lowered after ovipositioning in D. punctata and are maintained low throughout pregnancy in our transcriptome. Consequently, we conclude that our knockdown did not produce or sufficiently sustain a significant enough increase in JH to trigger the expression of

Krh1 or Vg.

It is important to note, however, that viviparity and its associated physiology are not unique to D. punctata. Many other insects are live-bearing and even matrotrophic (Attardo et al., 2019; Clutton-Brock, 1991; Denlinger & Ma, 1974; Hagan, 1948; Ma, Denlinger, Järlfors, &

Smith, 1975; Roth & Willis, 1957; Tobe & Langley, 1978) including some dermapterans, aphids, and most notably tsetse flies (Glossinidae) and other Hippoboscoidea (Attardo et al., 2019;

Clutton-Brock, 1991; Denlinger & Ma, 1974; Hagan, 1948; Ma et al., 1975; Roth & Willis, 1957;

Tobe & Langley, 1978). Despite the phylogenetic distance separating Diploptera and Glossina, these genera have converged upon matrotrophic viviparity and employ similar regulatory mechanisms for the process (Attardo et al., 2019; Baumann et al., 2013; Denlinger & Ma, 1974;

Ejezie & Davey, 1976; Evans & Stay, 1989, 1995; Ingram et al., 1977; Langley & Pimley, 1986;

Marchal et al., 2013; Stay & Coop, 1974; Stay, Ostedgaard, Tobe, Strambi, & Spaziani, 1984;

Tobe et al., 1985; Tobe, Davey, & Huebner, 1973; Tobe & Langley, 1978; Williford et al., 2004).

Further underscoring this mechanistic convergence are the shared adverse reactions to

42 perturbations in juvenile hormone titers during pregnancy; JH inhibits production and secretion of milk proteins and can act as an abortifacient in both D. punctata and tsetse flies (Baumann et al., 2013; Denlinger, 1975; Evans & Stay, 1995; Langley & Pimley, 1986; Stay & Lin, 1981; Terr

Wee & Stay, 1987).

While viviparous reproduction is widespread throughout the animal kingdom, little is known about the molecular mechanisms underlying live-birth and pregnancy in insects. Despite numerous studies characterizing its reproductive endocrinology, little is known about other aspects of matrotrophic viviparity in Diploptera punctata, with most molecular research targeting hormonal processes occurring before fertilized eggs are ever deposited into the brood sac. The goal of this study was to expand our understanding of the molecular changes associated with this unique and complex reproductive process using RNA-seq analysis. In this study we report the first transcriptome analysis of D. punctata, focusing on characterizing changes in gene expression across pregnancy. We identified 11,987 putative protein coding sequences with 2,474 differentially regulated across pregnancy. Three hundred twenty-nine transcription factors were identified in this gene set, corresponding to 35 structural families, 10 of which are differentially regulated across pregnancy. The 25 previously sequenced milk protein cDNA clones were differentially expressed across all stages of pregnancy, being most highly expressed in the late lactation stages of pregnancy. In addition, we identified a significant pattern in the expression of juvenile hormone esterase (jhe), an enzyme previously shown to break down juvenile hormone in D. punctata that is associated with pregnancy maintenance. To better characterize the role of this differential expression, we knocked down jhe using RNAi, finding a significant reduction in milk protein transcript expression but no increase in

43 vitellogenin or krüppel homolog 1. In addition to providing new molecular resources for future studies into the biology of D. punctata, our study contributes to broadening the understanding of transcriptional changes associated with viviparity and enables further comparative research to be done among live bearing insects.

44 References:

Attardo, G. M., Abd-Alla, A. M. M., Acosta-Serrano, A., James, E., Bateta, R., Benoit, J. B., …

Aksoy, S. (2019). The Glossina genome cluster: Comparative genomic analysis of the

vectors of African Trypanosomes. BioRxiv, 531749. http://doi.org/10.1101/531749

Banerjee, S., Coussens, N. P., Gallat, F. X., Sathyanarayanan, N., Srikanth, J., Yagi, K. J., …

Ramaswamy, S. (2016). Structure of a heterogeneous, glycosylated, lipid-bound, in vivo-

grown protein crystal at atomic resolution from the viviparous cockroach Diploptera

punctata. IUCrJ, 3, 282–293. http://doi.org/10.1107/S2052252516008903

Baumann, A. A., Benoit, J. B., Michalkova, V., Mireji, P. O., Attardo, G. M., Moulton, J. K., …

Aksoy, S. (2013). Juvenile hormone and insulin suppress lipolysis between periods of

lactation during tsetse fly pregnancy. Molecular and Cellular Endocrinology, 372(1–2), 30–

41. http://doi.org/10.1016/j.mce.2013.02.019

Birol, I., Jackman, S. D., Nielsen, C. B., Qian, J. Q., Varhol, R., Stazyk, G., … Jones, S. J. M. (2009).

De novo transcriptome assembly with ABySS. Bioinformatics. 25(21), 2872-2877.

http://doi.org/10.1093/bioinformatics/btp367

Borras-Castells, F., Nieva, C., Maestro, J.L., Maestro, O., Belles, X. and Martín, D. (2017).

Juvenile hormone biosynthesis in adult Blattella germanica requires nuclear receptors

Seven-up and FTZ-F1. Scientific Reports, 7(1), 40234. http://doi.org/10.1038/srep40234

Bretãs, J. A. C., Peixoto, A. A., Mazzoni, C. J., Davila, A. M. R., Dias, D. B. S., Wagner, G., …

Azevedo, R. V. D. M. (2012). The transcriptome of Lutzomyia longipalpis (Diptera:

Psychodidae) male reproductive organs. PLoS ONE, 7(4), e34495.

http://doi.org/10.1371/journal.pone.0034495

45 Budachetri, K., & Karim, S. (2015). An insight into the functional role of thioredoxin reductase, a

selenoprotein, in maintaining normal native microbiota in the Gulf Coast tick (Amblyomma

maculatum). Insect Molecular Biology. 24(5), 570-581. http://doi.org/10.1111/imb.12184

Clutton-Brock, T. H. (1991). The Evolution of Parental Care. (J. R. Krebs & T. H. Clutton-Brock,

Eds.). Princeton: Princeton University Press.

Couillaud, F., & Feyereisen, R. (1991). Assay of HMG-CoA synthase in Diploptera punctata

corpora allata. Insect Biochemistry, 21(2), 131–135. http://doi.org/10.1016/0020-

1790(91)90042-D de Kort, C. A. D., & Granger, N. A. (1996). Regulation of JH titers: The relevance of degradative

enzymes and binding proteins. Archives of Insect Biochemistry and Physiology, 33(1), 1–26.

http://doi.org/10.1002/(SICI)1520-6327(1996)33:1<1::AID-ARCH1>3.0.CO;2-2

Denlinger, D. L. (1975). Insect as tsetse abortifacients. Nature, 253(5490), 347–8.

http://doi.org/10.0.4.14/253347a0

Denlinger, D. L., & Ma, W. C. (1974). Dynamics of the pregnancy cycle in the tsetse Glossina

morsitans. Journal of Insect Physiology, 20(6), 1015-1026. http://doi.org/10.1016/0022-

1910(74)90143-7

Dominguez, C. V, & Maestro, J. L. (2017). Expression of juvenile hormone acid O -

methyltransferase and juvenile hormone synthesis in Blattella germanica. Insect Science,

25(5), 1–10. http://doi.org/10.1111/1744-7917.12467

Ejezie, G. C., & Davey, K. G. (1976). Some effects of allatectomy in the female tsetse, Glossina

austeni. Journal of Insect Physiology, 22, 1743–1749.

Engelmann, F., & Mala, J. (2000). The interactions between juvenile hormone (JH), lipophorin,

46 vitellogenin, and JH esterases in two cockroach species. Insect Biochemistry and Molecular

Biology, 30(8–9), 793–803. http://doi.org/10.1016/S0965-1748(00)00051-5

Evans, L. D., & Stay, B. (1989). Humoral induction of milk synthesis in the viviparous cockroach

Diploptera punctata. Invertebrate Reproduction and Development, 15(3), 171–176.

Evans, L. D., & Stay, B. (1994). Development of competence for milk production in a viviparous

cockroach and its association with changes in brood sac glycogen. Invertebrate

Reproduction and Development, 25(2), 111–122.

http://doi.org/10.1080/07924259.1994.9672376

Evans, L. D., & Stay, B. (1995). Regulation of competence for milk production in Diploptera

punctata: interaction between mating, ovaries and the corpus allatum. Invertebrate

Reproduction and Development, 28(3), 161–170.

Gotoh, A., Shigenobu, S., Yamaguchi, K., Kobayashi, S., Ito, F., & Tsuji, K. (2018). Transcriptome

characterization of male accessory glands in to identify molecules involved in their

reproductive success. Insect Molecular Biology, 27(2), 212–220.

http://doi.org/10.1111/imb.12364

Hagan, H. R. (1941). The general morphology of the female of a viviparous

roach, Diploptera dytiscoides (Serville). Psyche: A Journal of Entomology, 48(1), 1–8.

http://doi.org/10.1155/1941/45020

Hagan, H. R. (1948). A brief analysis of viviparity in insects. Journal of the New York

Entomological Society, 56(1), 63–68.

Hancock, J. M., & Zvelebil, M. J. (2014). PRIMER3. In Dictionary of Bioinformatics and

Computational Biology. http://doi.org/10.1002/9780471650126.dob0560.pub2

47 Hope, R. M. (2013). Rmisc: Ryan miscellaneous.

Hult, E. F., Huang, J., Marchal, E., Lam, J., & Tobe, S. S. (2015). RXR/USP and EcR are critical for

the regulation of reproduction and the control of JH biosynthesis in Diploptera punctata.

Journal of Insect Physiology, 80, 48–60. http://doi.org/10.1016/j.jinsphys.2015.04.006

Ingram, M. J., Stay, B., & Cain, G. D. (1977). Composition of milk from the viviparous cockroach,

Diploptera punctata. Insect Biochemistry, 7(3), 257–267. http://doi.org/10.1016/0020-

1790(77)90023-3

Ishikawa, A., Ogawa, K., Gotoh, H., Walsh, T. K., Tagu, D., Brisson, J. A., … Miura, T. (2012).

Juvenile hormone titre and related gene expression during the change of reproductive

modes in the pea aphid. Insect Molecular Biology, 21(1), 49–60.

http://doi.org/10.1111/j.1365-2583.2011.01111.x

Kalinka, A. T. (2015). How did viviparity originate and evolve? Of conflict, co-option, and cryptic

choice. BioEssays, 37(7), 721-731. http://doi.org/10.1002/bies.201400200

Kim, I.-W., Lee, J. H., Subramaniyam, S., Yun, E.-Y., Kim, I., Park, J., & Hwang, J. S. (2016). De

novo transcriptome analysis and detection of antimicrobial peptides of the American

cockroach Periplaneta americana (Linnaeus). PLOS ONE, 11(5), e0155304.

http://doi.org/10.1371/journal.pone.0155304

Konopova, B., Smykal, V., & Jindra, M. (2011). Common and distinct roles of juvenile hormone

signaling genes in of holometabolous and hemimetabolous insects. PLoS

ONE, 6(12), 19–23. http://doi.org/10.1371/journal.pone.0028728

Langley, P. A., & Pimley, R. W. (1986). A role for juvenile hormone and the effects of so-called

anti-juvenile hormones in Glossina morsitans. Journal of Insect Physiology, 32(8), 727–734.

48 http://doi.org/10.1016/0022-1910(86)90115-0

Luo, R., Liu, B., Xie, Y., Li, Z., Huang, W., Yuan, J., … Wang, J. (2012). SOAPdenovo2: An

empirically improved memory-efficient short-read de novo assembler. GigaScience. 1(1),

18. http://doi.org/10.1186/2047-217X-1-18

Ma, W. C., Denlinger, D. L., Järlfors, U., & Smith, D. S. (1975). Structural modulations in the

tsetse fly milk gland during a pregnancy cycle. Tissue and Cell, 7(2), 319–330.

http://doi.org/10.1016/0040-8166(75)90008-7

Marchal, E., Hult, E. F., Huang, J., Stay, B., & Tobe, S. S. (2013). Diploptera punctata as a model

for studying the endocrinology of arthropod reproduction and development. General and

Comparative Endocrinology, 188, 85–93.

http://doi.org/http://dx.doi.org/10.1016/j.ygcen.2013.04.018

McGraw, L. A., Gibson, G., Clark, A. G., & Wolfner, M. F. (2004). Genes regulated by mating,

sperm, or seminal proteins in mated female Drosophila melanogaster. Current Biology,

14(16), 1509–1514. http://doi.org/10.1016/j.cub.2004.08.028

Meibers, H. E., Finch, G., Gregg, R. T., Glenn, S., Assani, K. D., Jennings, E. C., … Benoit, J. B.

(2019). - and developmental-specific transcriptomic analyses of the Antarctic mite,

Alaskozetes antarcticus, reveal transcriptional shifts underlying oribatid mite reproduction.

Polar Biology, 42(2), 357–370. http://doi.org/10.1007/s00300-018-2427-x

Meier, R., Kotrba, M., & Ferrar, P. (1999). Ovoviviparity and viviparity in the Diptera. Biological

Reviews, 74, 199–258. http://doi.org/10.1111/j.1469-185X.1999.tb00186.x

Minakuchi, C., Namiki, T., & Shinoda, T. (2009). Krüppel homolog 1, an early juvenile hormone-

response gene downstream of Methoprene-tolerant, mediates its anti-metamorphic action

49 in the red flour beetle Tribolium castaneum. Developmental Biology, 325(2), 341–350.

http://doi.org/10.1016/j.ydbio.2008.10.016

Naghdi, M., Maestro, J., Belles, X., & Bandani, A. (2016). Transduction of the vitellogenic signal

of juvenile hormone by Methoprene-tolerant in the cockroach Blattella germanica ( L .) (

Dictyoptera , Blattellidae ). Arthropods, 5(4), 130–137.

Neuwirth, E. (2014). RColorBrewer: ColorBrewer Palettes.

Paulson, C. R., & Stay, B. (1987). Humoral inhibition of the corpora allata in larvae of Diploptera

punctata: Role of the brain and ecdysteroids. Journal of Insect Physiology, 33(9), 613–622.

http://doi.org/10.1016/0022-1910(87)90129-6

Phillips, N. (2017). yarrr: A Companion to the e-Book “YaRrr!: The Pirate’s Guide to R.”

Qu, Z., Bendena, W. G., Tobe, S. S., & Hui, J. H. L. (2018). Juvenile hormone and

sesquiterpenoids in arthropods: Biosynthesis, signaling, and role of microRNA. Journal of

Steroid Biochemistry and Molecular Biology, 184, 69-76.

http://doi.org/10.1016/j.jsbmb.2018.01.013

R Core Team. (2017). R: A language and environment for statistical computing. Vienna, Austria:

R Foundation for Statistical Computing.

Ram, K., & Wickham, H. (2018). wesanderson: A Wes Anderson palette generator.

Rankin, S. M., Palmer, J. O., Yagi, K. J., Scott, G. L., & Tobe, S. S. (1995). Biosynthesis and release

of juvenile hormone during the reproductive cycle of the ring-legged earwig.

Comp.Biochem.Physiol., 110(3), 241–251. http://doi.org/10.1016/0742-8413(95)00016-H

Rankin, S. M., & Stay, B. (1984). The changing effect of the ovary on rates of juvenile hormone

synthesis in Diploptera punctata. General and Comparative Endocrinology, 54(3), 382–388.

50 http://doi.org/10.1016/0016-6480(84)90151-5

Ribeiro, J. M. C., Chagas, A. C., Pham, V. M., Lounibos, L. P., & Calvo, E. (2014). An insight into

the sialome of the biting fly, Corethrella appendiculata. Insect Biochemistry and

Molecular Biology, 44, 23–32. http://doi.org/10.1016/j.ibmb.2013.10.006

Ribeiro, J. M. C., Genta, F. A., Sorgine, M. H. F., Logullo, R., Mesquita, R. D., Paiva-Silva, G. O., …

Oliveira, P. L. (2014). An insight into the transcriptome of the digestive tract of the

bloodsucking bug, Rhodnius prolixus. PLoS Neglected Tropical Diseases, 8(1), e2594.

http://doi.org/10.1371/journal.pntd.0002594

Rosendale, A. J., Romick-Rosendale, L. E., Watanabe, M., Dunlevy, M. E., & Benoit, J. B. (2016).

Mechanistic underpinnings of dehydration stress in the American dog tick revealed

through RNA-Seq and metabolomics. The Journal of Experimental Biology, 219(12), 1808–

1819. http://doi.org/10.1242/jeb.137315

Roth, L. M. (1967). Water changes in cockroach oothecae in relation to the evolution of

ovoviviparity and viviparity. Annals of the Entomological Society of America, 60(5), 928–

946. http://doi.org/10.1093/aesa/60.5.928

Roth, L. M., & Hahn, W. (1964). Size of new-born larvae of cockroaches incubating eggs

internally. Journal of Insect Physiology. 10, 65-72. http://doi.org/10.1016/0022-

1910(64)90094-0

Roth, L. M., & Stay, B. (1961). Oöcyte development in Diploptera punctata (Eschscholtz)

(Blattaria). Journal of Insect Physiology. 7, 186-202. http://doi.org/10.1016/0022-

1910(61)90071-3

Roth, L. M., & Willis, E. R. (1955). Intra-uterine nutrition of the “beetle-roach” Diploptera

51 Dytiscoides (Serv.) during embryogenesis, with notes on its biology in the laboratory

(Blattaria: Diplopteridae). Psyche: A Journal of Entomology. 62, 55-68.

http://doi.org/10.1155/1955/12542

Roth, L. M., & Willis, E. R. (1957). An analysis of oviparity and viviparity in the Blattaria.

American Entomological Society, 83(4), 221–238.

Rotin, D., & Tobe, S. S. (1983). The possible role of juvenile hormone esterase in the regulation

of juvenile hormone titre in the female cockroach Diploptera punctata. Canadian Journal

of Biochemistry and Cell Biology, 61(7), 811–817. http://doi.org/10.1139/o83-103

RStudio Team. (2015). RStudio: Integrated development for R. Boston, MA: RStudio, Inc.

Schmittgen, T. D., & Livak, K. J. (2008). Analyzing real-time PCR data by the comparative CT

method. Nature Protocols, 3(6), 1101–1108. http://doi.org/10.1038/nprot.2008.73

Schoville, S. D., Chen, Y. H., Andersson, M. N., Benoit, J. B., Bhandari, A., Bowsher, J. H., …

Richards, S. (2018). A model species for agricultural genomics: the genome of the

Colorado potato beetle, Leptinotarsa decemlineata (Coleoptera: Chrysomelidae). Scientific

Reports, 8(1), 1931. http://doi.org/10.1038/s41598-018-20154-1

Segatto, A. L. A., Diesel, J. F., Loreto, E. L. S., & da Rocha, J. B. T. (2018). De novo transcriptome

assembly of the lobster cockroach Nauphoeta cinerea (). Genetics and

Molecular Biology, 41(3), 713–721. http://doi.org/10.1590/1678-4685-gmb-2017-0264

Simpson, J. T., Wong, K., Jackman, S. D., Schein, J. E., Jones, S. J. M., & Birol, I. (2009). ABySS: A

parallel assembler for short read sequence data. Genome Research. 19(6), 1117-1123.

http://doi.org/10.1101/gr.089532.108

Sirot, L. K., Rubinstein, C. D., Avila, F. W., Wolfner, M. F., & LaFlamme, B. A. (2010). Insect

52 seminal fluid proteins: Identification and function. Annual Review of Entomology, 56(1),

21–40. http://doi.org/10.1146/annurev-ento-120709-144823

Stay, B., & Clark, J. K. (1971). Fluctuation of protein granules in the fat body of the viviparous

cockroach, Diploptera punctata, during the reproductive cycle. Journal of Insect

Physiology, 17, 1747–1762. http://doi.org/10.1016/0022-1910(71)90072-2

Stay, B., & Coop, A. (1973). Developmental stages and chemical composition in embryos of the

cockroach, Diploptera punctata, with observations on the effect of diet. Journal of Insect

Physiology, 19(1), 147–171. http://doi.org/10.1016/0022-1910(73)90230-8

Stay, B., & Coop, A. C. (1974). “Milk” secretion for embryogenesis in a viviparous cockroach.

Tissue and Cell, 6(4), 669–693. http://doi.org/10.1016/0040-8166(74)90009-3

Stay, B., & Lin, H. L. (1981). The inhibition of milk synthesis by juvenile hormone in the

viviparous cockroach, Diploptera punctata. Journal of Insect Physiology, 27(8), 551–557.

http://doi.org/10.1016/0022-1910(81)90043-3

Stay, B., Ostedgaard, L. S., Tobe, S. S., Strambi, A., & Spaziani, E. (1984). Ovarian and

haemolymph titres of ecdysteroid during the gonadotrophic cycle in Diploptera punctata.

Journal of Insect Physiology, 30(8), 643–651. http://doi.org/10.1016/0022-1910(84)90049-

0

Stay, B., & Roth, L. M. (1956). The reproductive behavior of Diploptera punctata (Blattaria:

Diplopterinae). Proceedings of the International Congress of Entomology. 2, 547-552.

Stay, B., Tobe, S. S., Mundall, E. C., & Rankin, S. (1983). Ovarian stimulation of juvenile hormone

biosynthesis in the viviparous cockroach, Diploptera punctata. General and Comparative

Endocrinology, 52(3), 341–349. http://doi.org/10.1016/0016-6480(83)90171-5

53 Stewart, J. R. (2015). A universal model for the evolution of viviparity. BioEssays, 37, 1.

http://doi.org/10.1002/bies.201400200

Stoltzman, C. A., Stocker, C., Borst, D., & Stay, B. (2000). Stage-specific production and release

of juvenile hormone esterase from the ovary of Diploptera punctata. Journal of Insect

Physiology, 46(5), 771–782. http://doi.org/10.1016/S0022-1910(99)00166-3

Terr Wee, J., & Stay, B. (1987). Termination of pregnancy and milk synthesis in the viviparous

cockroach, Diploptera Punctata: a role of juvenile hormone. International Journal of

Invertebrate Reproduction and Development, 11(1), 59–75.

Terrapon, N., Li, C., Robertson, H. M., Ji, L., Meng, X., Booth, W., … Liebig, J. (2014). Molecular

traces of alternative social organization in a termite genome. Nature Communications,

5(1), 3636. http://doi.org/10.1038/ncomms4636

Tobe, S. S., Davey, K. G., & Huebner, E. (1973). Nutrient transfer during the reproductive cycle in

Glossina austeni newst.: Histology and histochemistry of the milk gland, fat body, and

oenocytes. Tissue and Cell, 5(4), 633–650. http://doi.org/10.1016/S0040-8166(73)80050-3

Tobe, S. S., & Langley, P. A. (1978). Reproductive physiology of Glossina. Annual Review of

Entomology, 23(1), 283–307. http://doi.org/10.1146/annurev.en.23.010178.001435

Tobe, S. S., Ruegg, R. P., Stay, B., Baker, F. C., Miller, C. A., & Schooley, D. A. (1985). Juvenile

hormone titre and regulation in the cockroach Diploptera punctata. Experientia, 41(8),

1028–34.

Wei, D., Tian, C. B., Liu, S. H., Wang, T., Smagghe, G., Jia, F. X., … Wang, J. J. (2016).

Transcriptome analysis to identify genes for peptides and proteins involved in immunity

and reproduction from male accessory glands and ejaculatory duct of Bactrocera dorsalis.

54 Peptides, 80, 48–60. http://doi.org/10.1016/j.peptides.2015.08.007

Wickham, H. (2007). Reshaping data with the reshape Package. Journal of Statistical Software,

21(12), 1–20. http://doi.org/10.18637/jss.v021.i12

Wickham, H. (2016). ggplot2: Elegant graphics for data analysis. Springer-Verlag New York.

Wickham, H., Francois, R., Henry, L., & Müller, K. (2017). dplyr: A grammar of data

manipulation.

Williford, A., Stay, B., & Bhattacharya, D. (2004). Evolution of a novel function: Nutritive milk in

the viviparous cockroach, Diploptera punctata. Evolution and Development, 6, 67–77.

http://doi.org/10.1111/j.1525-142X.2004.04012.x

Wilson, T. G., Landers, M. H., & Happ, G. M. (1983). Precocene I and II inhibition of vitellogenic

oöcyte development in Drosophila melanogaster. Journal of Insect Physiology, 29(3), 249–

254. http://doi.org/10.1016/0022-1910(83)90091-4

55 Tables:

Table 1. Pregnancy stage determination. This table describes the measurements utilized to determine pregnancy stage based on a previous study by Stay & Coop (1973).

Reproductive Stage Embryo Length Estimated Embryo Age

Not Pregnant (NPF) Not present n/a

Pre Lactation (PreL) <1.6 mm 0-11 days

Early Lactation (EarL) 1.6-2.5 mm 12-27 days

Late Lactation (LateL) >2.5 mm 28-55+ days

56 Table 2. Primer and siRNA sequences utilized in this study

Gene Assay Sequence

elongation factor 1 a qPCR F: 5’—CAAGATTGGAGGTATTGGAACAGTG—3’

R: 5’—GACTTTACTTCAGTGGTCAAGTTGG—3’

vitellogenin qPCR F: 5’—AAAGGTGTCCTCAGCCAGC—3’

R: 5’—TCCTCCATCTCGGATTGGGA—3’

juvenile hormone esterase qPCR F: 5’—CCTGGACAAGGATGTTGTTATG—3’

R: 5’—CACCTCCGAAACTTGCTATG—3’

milk protein 13Y qPCR F: 5’—CAATATGGACAAGAGACACATCGTG—3’

R: 5’—CTGCAAGTATCCGACTTTCTGAATC—3’

krüppel homolog 1 qPCR F: 5’—ACACAGCGGCAAGTTACA—3’

R: 5’—AAGTTGACCGCTCTGGATAAA—3’

57 ecdysone receptor qPCR F: 5’—ATCAGTGAACGGAGTAAAACCTGTA—3’

R: 5’—TTGAGGTCATCATCAGAAGGTGATT—3’ juvenile hormone esterase siRNA F: 5’—rCrArArGrGrArUrGrUrUrGrUrUrArUrGrGrUrArArCrArATC—3’

R: 5’—rGrArUrUrGrUrUrArCrCrArUrArArCrArArCrArUrCrCrUrUrGrUrC—3’ green fluorescent protein siRNA F: 5’—rCrUrUrGrArCrUrUrCrArGrCrArCrGrUrGrUrCrUrUrGrUrArGrUrU—3’

R: 5’—rCrUrArCrArArGrArCrArCrGrUrGrCrUrGrArArGrUrCrArArG—3’

58 Table 3. General assembly information for both the full assembly and extracted CDS

All Contigs Extracted CDS

Number of contigs 102,880 11,987

Total size of contigs 56,694,784 10,820,347

Longest contig 24,290 23,307

Shortest contig 150 150

Mean contig size 551 903

Median contig size 314 693

N50 contig length 866 1326

%A 31.88 31.4

%C 18.22 19.74

%G 17.88 22.41

%T 32.01 26.45

59 Table 4. GO Term enrichment as generated by GSEA in CLC Genomics Workbench, t-statstics and P-values are from comparisons against non-pregnant females

Pre Lactation Early Lactation Late Lactation

Description GO ID Gene count T-Statistic P-value T-Statistic P-value T-Statistic P-value cuticle development 42335 26 -5.95 0.0087 -5.97 0.0018 -5.16 0.0009 transposition, DNA-mediated 6313 50 -1.99 0.4758 -6.97 0.0004 -5.54 0.0001 -based cuticle development 40003 24 -6.17 0.0071 -6.69 0.0005 -5.90 0.0003 nucleic acid phosphodiester bond hydrolysis 90305 246 -14.60 <0.00001 -7.06 0.0009 -2.92 0.001 transposition 32196 50 -1.99 0.4758 -6.97 0.0004 -5.54 0.0001

DNA biosynthetic process 71897 187 -19.78 <0.00001 -10.49 <0.00001 -8.04 <0.00001

RNA-dependent DNA replication 6278 165 -19.41 <0.00001 -9.53 <0.00001 -8.25 <0.00001

DNA metabolic process 6259 415 -15.20 <0.00001 -8.86 <0.00001 -4.92 <0.00001 regulation of protein metabolic process 51246 185 -0.92 0.0506 2.41 0.0114 8.15 0.0004 translation 6412 401 6.84 <0.00001 13.00 <0.00001 11.67 <0.00001 single-organism cellular process 44763 1468 -5.89 0.0017 1.19 0.0002 11.77 0.0006 organic substance metabolic process 71704 3315 -13.01 0.0151 0.67 <0.00001 14.14 <0.00001

60 peptide biosynthetic process 43043 411 7.04 <0.00001 13.44 <0.00001 11.91 <0.00001 biosynthetic process 9058 1274 -5.95 0.0078 6.13 <0.00001 10.66 0.0009 regulation of cellular protein metabolic process 32268 179 -0.91 0.0532 2.39 0.0128 8.05 0.0006 organonitrogen compound metabolic process 1901564 944 2.56 <0.00001 7.72 <0.00001 11.26 <0.00001 organonitrogen compound biosynthetic process 1901566 601 5.79 <0.00001 11.73 <0.00001 12.10 <0.00001 cellular component assembly 22607 199 1.36 0.0008 5.56 <0.00001 8.75 0.0002 peptide metabolic process 6518 427 7.28 <0.00001 13.40 <0.00001 12.13 <0.00001 mRNA splicing, via spliceosome 398 59 2.61 0.0014 8.13 <0.00001 8.55 0.0006

RNA splicing, via transesterification reactions with bulged adenosine as nucleophile 377 59 2.61 0.0014 8.13 <0.00001 8.55 0.0006 cellular metabolic process 44237 2935 -12.78 0.0721 0.33 <0.00001 13.62 0.0002 primary metabolic process 44238 3065 -13.50 0.1141 0.56 <0.00001 13.84 <0.00001

RNA splicing, via transesterification 375 59 2.61 0.0014 8.13 <0.00001 8.55 0.0006

61 reactions cellular biosynthetic process 44249 1210 -5.45 0.0056 6.43 <0.00001 10.59 0.0006 small molecule biosynthetic process 44283 125 1.09 0.0069 3.37 0.004 8.17 0.0004 cellular process 9987 3531 -15.72 0.3736 -1.41 <0.00001 13.44 0.0004 cellular amide metabolic process 43603 456 7.18 <0.00001 12.98 <0.00001 11.97 <0.00001 amide biosynthetic process 43604 420 7.51 <0.00001 13.76 <0.00001 12.21 <0.00001 macromolecule modification 43412 658 -12.45 0.0003 -4.57 0.146 2.84 0.259 organic substance transport 71702 452 -11.10 0.0008 -4.29 0.1324 0.56 0.0664 protein modification process 36211 602 -12.37 0.0001 -5.28 0.0602 2.16 0.1788 nucleobase-containing compound biosynthetic process 34654 486 -15.27 <0.00001 -5.48 0.0312 -1.63 0.0023 heterocycle biosynthetic process 18130 548 -14.20 <0.00001 -5.03 0.069 -0.31 0.0118 organic cyclic compound biosynthetic process 1901362 557 -14.15 <0.00001 -4.95 0.077 -0.14 0.0151 aromatic compound biosynthetic process 19438 525 -14.76 <0.00001 -5.31 0.0466 -1.00 0.0052

RNA phosphodiester bond hydrolysis, endonucleolytic 90502 46 -9.47 0.0003 -2.97 0.1006 -1.66 0.0651

62 nucleic acid metabolic process 90304 942 -15.18 <0.00001 -3.41 0.4912 2.91 0.1429 signal transduction 7165 539 -12.03 0.0003 -7.86 0.0007 1.17 0.0898 regulation of biological process 50789 1227 -14.48 0.0008 -6.93 0.0249 4.92 0.3858 cellular protein modification process 6464 602 -12.37 0.0001 -5.28 0.0602 2.16 0.1788 carboxylic acid metabolic process 19752 344 0.55 0.0001 2.84 0.0022 8.55 0.0015 single-organism metabolic process 44710 1534 -0.40 <0.00001 1.28 0.0001 11.00 0.0018 electron transport chain 22900 57 7.91 <0.00001 2.62 0.0238 5.80 0.0116 respiratory electron transport chain 22904 51 7.44 <0.00001 2.47 0.0296 5.23 0.0186 oxoacid metabolic process 43436 349 -0.03 0.001 2.40 0.0051 8.06 0.0027 mitotic cell cycle process 1903047 167 1.58 0.001 6.37 <0.00001 5.94 0.018 mitochondrial transport 6839 27 3.48 0.001 5.80 0.0004 5.56 0.0105 metabolic process 8152 4921 -15.84 0.0001 -4.73 0.0003 13.39 0.0022 inorganic ion transmembrane transport 98660 162 1.81 0.0005 3.27 0.0025 3.16 0.2502 inorganic cation transmembrane transport 98662 151 2.17 0.0004 3.74 0.0015 3.72 0.1555 mitochondrial electron transport, NADH to ubiquinone 6120 26 6.06 <0.00001 1.12 0.1634 2.89 0.1241 single-organism process 44699 2456 -6.68 <0.00001 -1.34 0.0011 11.28 0.0082

63 oxidation-reduction process 55114 813 3.62 <0.00001 1.78 0.0014 7.81 0.029 small molecule metabolic process 44281 600 0.81 <0.00001 2.92 0.0004 9.20 0.0016 generation of precursor metabolites and energy 6091 115 5.42 <0.00001 3.54 0.0028 7.05 0.0041 actomyosin structure organization 31032 24 3.99 0.0007 4.57 0.0027 4.98 0.0173 organic acid metabolic process 6082 350 0.02 0.0008 2.42 0.0048 8.08 0.0027 cell differentiation 30154 110 2.23 0.0005 7.71 <0.00001 7.46 0.0015 digestion 7586 10 5.13 0.0001 4.38 0.0056 5.05 0.0125 single-organism biosynthetic process 44711 318 -0.35 0.0033 3.33 0.0004 7.56 0.0063 nitrogen compound metabolic process 6807 1842 -8.68 0.0211 2.82 <0.00001 9.82 0.0237 synapse organization 50808 28 3.11 0.0022 5.38 0.0006 5.80 0.007 macromolecule biosynthetic process 9059 879 -6.93 0.2452 5.46 <0.00001 7.47 0.0513 neurogenesis 22008 89 0.99 0.0172 6.44 <0.00001 5.66 0.0154

RNA splicing 8380 74 1.74 0.0063 7.38 <0.00001 7.37 0.0023 cellular nitrogen compound metabolic process 34641 1620 -7.66 0.0131 4.92 <0.00001 10.45 0.0047 cellular macromolecule biosynthetic 34645 833 -6.51 0.2124 5.60 <0.00001 6.62 0.1186

64 process organic substance biosynthetic process 1901576 1220 -6.15 0.0149 6.10 <0.00001 10.33 0.0017 mRNA processing 6397 99 -0.02 0.0534 6.30 <0.00001 6.17 0.0086

RNA processing 6396 224 -3.33 0.332 4.76 0.0003 6.02 0.0277 cell cycle process 22402 204 0.68 0.0019 5.78 <0.00001 6.24 0.0165 ribonucleoprotein complex assembly 22618 49 0.94 0.0403 5.12 0.0005 5.23 0.0164 organelle organization 6996 311 -1.20 0.016 3.96 0.0003 6.52 0.0198 organelle assembly 70925 29 3.25 0.0016 5.43 0.0003 6.64 0.0047 ribonucleoprotein complex subunit organization 71826 50 0.80 0.0477 5.17 0.0005 5.27 0.0146 cellular component organization or biogenesis 71840 629 -4.13 0.0463 5.08 <0.00001 8.81 0.0022 macromolecule metabolic process 43170 2570 -13.89 0.4511 0.11 <0.00001 10.97 0.0193 cellular macromolecule metabolic process 44260 2015 -13.84 0.1242 0.69 <0.00001 9.33 0.0715 cellular protein metabolic process 44267 1009 -5.26 0.0182 4.11 <0.00001 9.15 0.0091 cellular developmental process 48869 229 -0.05 0.0069 6.38 <0.00001 7.61 0.0035 cellular nitrogen compound biosynthetic 44271 938 -6.16 0.0895 5.07 <0.00001 7.32 0.0707

65 process regulation of developmental process 50793 97 0.68 0.0217 4.95 0.0003 4.32 0.066 mRNA metabolic process 16071 115 -0.75 0.0971 6.11 <0.00001 6.72 0.0045 single-organism organelle organization 1902589 124 1.64 0.0016 4.19 0.0005 7.40 0.0018 cellular component organization 16043 596 -4.36 0.0783 4.34 <0.00001 7.99 0.0073 protein metabolic process 19538 1543 -7.30 0.0115 2.69 <0.00001 11.12 0.0016

66 Legends and Figures

Legends

Figure 1. (a) Female D. punctata giving birth (b) Micro CT scan of pregnant D. punctata

Figure 2. Quality metrics of the de novo assembly from Diploptera punctata (A) BUSCO representation in the complete combined assembly. (B) Top BLAST hits from the extracted CDS against the nr arthropod database

Figure 3. Functional annotation of genes differentially expressed across pregnancy. (A) GO terms enriched in the 2,474 transcripts differentially regulated across pregnancy. (B) Analysis of biological function GO term composition of significantly up and down regulated contigs in female D. punctata reveals unique transcriptional profiles for each stage of pregnancy.

Figure 4. Transcripts with male specific expression patterns (A) Venn diagram of genes with male specific expression relative to not pregnant and late lactation females as well as the core male gene set (B) GO term enrichment of the 284 male specific genes

Figure 5. Transcription factor identification and expression (A) 329 transcription factors belonging to 35 structural families were identified in our assembly. C2H2 zinc finger domains are the most abundant family, followed by homeodomain, bHLH, bZIP, and Myb domains (B)

TPM expression of ftz-f1 beta and Krh1 transcription factors in the four female stages surveyed.

* FDR p ≤ 0.01; ** FDR p ≤ 0.001

67 Figure 6. Differential expression of select reproduction associated genes in female D. punctata

(A) Milk protein mRNA clone expression (TPM) abundance across the four reproductive stages.

(B) Vitellogenin expression across pregnancy represented as TPM. (C) Expression of Juvenile hormone esterase across pregnancy represented as TPM. Significance is denoted by different letter combinations (FDR p-value < 0.05)

Figure 7. Relative mRNA expression of key transcripts associated with pregnancy after RNAi knockdown of JHE. * denotes a Wilcox test p-value <0.05

68 Figures

69 A B

Species Pediculus humanus corporis Zootermopsis nevadensisZootermopsis Stegodyphus mimosarum Coptotermes formosanus Coptotermes Metaseiulus occidentalis 0% Camponotus floridanus Culex quinquefasciatus Harpegnathos saltator Harpegnathos Acyrthosiphon pisum Acyrthosiphon Tribolium castaneum Megachile rotundata Blattella germanica Nasonia vitripennis Bombus impatiens Ixodes scapularis Fopius arisanus Diaphorina citri Diaphorina Daphnia pulex Aedes aegypti 0 0.80% 0.81% 0.81% 0.82% 0.84% 0.86% 0.91% 0.93% 0.96% 1.01% 1.05% 1.12% 1.33% 1.47% 1.68% 1.79% 2.20% 2.25% 2.85% 25% 2000 B USCOs USCOs % 50% No. Top BLAST Hits 75% 4000 100% 57.30% 6000 Missing F CompleteSingle Co Duplicated Complete B r agmented USCO status p y

70 GO Name B A nucleobase-containing compound biosynthetic process

organonitrogen compound biosynthetic process cellular component organization or biogenesis organonitrogen compound metabolic process anatomical structure morphogenesis multicellular organism development anatomical structure development small molecule metabolic process carbohydrate derivative binding cellular amide metabolic process cellular component organization nucleic acid metabolic process peptide biosynthetic process amide biosynthetic process structural molecule activity peptide metabolic process protein metabolic process cytoskeleton organization developmental process organelle organization nucleic acid binding system development extracellular region transferase activity protein binding anion binding drug binding DNA binding translation cell cycle 0 1 2 3 4 5 6 Enriched Bar Chart 7 Female DEGs 8 9 10 11 12 13 % Seqs 14 15 16 17 18 19 20 21 22 23 24 25 26 27

71 GO Name oxidoreductase activity, or reduction of molecular oxygen acting on paired donors, with incorporation

B A

protein tyrosine/serine/threonine phosphatase activity protein tyrosine/serine/threonine peptidase activity, acting on L organonitrogen compound metabolic process catalytic activity, acting on a protein serine serine − type endopeptidase activity transition metal ion binding protein metabolic process serine hydrolaseserine activity − endopeptidase activity type peptidase activity − amino acid peptides peptidase activity hydrolase activity Pregnant catalytic activity proteolysis 1454 Male Male 546 Not vs vs 0

1

2

3

4 5 286 15 6 7

8

9 Lactation 10 11 Male Male Late 12 12 89 vs 13 14

15

16 Sequence count

17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37

72 A

TPM B

TPM 10000 20000 30000 40000 10000 20000 30000 40000 0 0 ftz − ftz f1 beta * − f1 beta

Transcription factor Transcription factor Krh1 ** Krh1

reorder(Group, order) LateL EarL PreL NPF reorder(Group, order) LateL EarL PreL NPF

73 A

A B B B C AB

1000 100 C A

reorder(Group, Order) reorder(Group, Order) 100 NPF NPF

PreL PreL EarL EarL

B LateL10 LateL

Vitellogenin expression (TPM) Vitellogenin expression 10

B Juvenile hormone esterase expression (TPM) hormone expression esterase Juvenile

1 1 NPF PreL EarL LateL NPF PreL EarL LateL Group Group

74

2.0

1.5

group siGFP siJHE 1.0 Relative mRNA expression Relative * 0.5

* * 0.0 JHE Krh1 MP13 Vg EcR Gene of interest

75

Chapter 3

Matrotrophic viviparity as a constraint on microbiome acquisition during

gestation in a live-bearing cockroach, Diploptera punctata

Emily C Jennings 1*

Matthew W Korthauer 1, Trinity L Hamilton 2, Joshua B Benoit 1

1 Department of Biological Sciences University of Cincinnati. Cincinnati, OH 45221

2 Plant and Microbial Biology and the BioTechnology Institute, College of Biological Sciences

University of Minnesota. St. Paul, MN 55108

* The work presented here was principally composed by Emily C Jennings with assistance from the listed co-authors.

76 Abstract

The vertical transmission of microbes from mother to offspring is critical to the survival, development, and health of animals. Invertebrate systems offer unique opportunities to conduct studies on microbiome-development-reproduction dynamics since reproductive modes ranging from oviparity to multiple types of viviparity are found in these animals. One such invertebrate is the live-bearing cockroach, Diploptera punctata. Females carry embryos in their brood sac, which acts as the functional equivalent of the uterus and placenta. In this study, 16S rRNA sequencing was used to characterize maternal and embryonic microbiomes as well as the development of the whole-body microbiome across nymphal development. We identified 50 phyla and 121 classes overall and found that mothers and their developing embryos had significantly different microbial communities. Of particular interest is the notable lack of diversity in the embryonic microbiome, which is comprised exclusively of Blattabacteriaceae, indicating microbial transmission of only this symbiont during gestation. Analysis of postnatal development reveals that significant amounts of non-Blattabacteria species are not able to colonize newborn D. punctata until melanization, after which the microbial community rapidly and dynamically diversifies. While the role of these microbes during development has not been characterized, Blattabacteria must serve a critical role providing specific micronutrients lacking in milk secretions to the embryos during gestation. Collectively, this research provides insight into the microbiome development, specifically with relation to viviparity, provisioning of milk- like secretions, and mother-offspring interactions during pregnancy.

77 Introduction

Animals share their bodies with a diverse suite of microorganisms known as the microbiome (Engel & Moran, 2013; The NIH HMP Working Group, 2009). These microbes have important roles in a variety of processes benefiting their host, ranging from nutrient metabolism to immunity {Formatting Citation}. For most animals, their microbial community is established over development through interactions with the environment, through diet, as well as interactions with other organisms (Abdul Rahman et al., 2015; Blaser & Dominguez-Bello,

2016; Carrasco et al., 2014; da Costa & Poulsen, 2018; Estes et al., 2013; Funkhouser &

Bordenstein, 2013; Gilbert, 2014; Korpela et al., 2018; Kostic et al., 2015; Morse et al., 2013;

Mueller, Bakacs, Combellick, Grigoryan, & Dominguez-Bello, 2015; Perez-Muñoz, Arrieta,

Ramer-Tait, & Walter, 2017; Schwab, Riggs, Newton, & Moczek, 2016; Shukla, Vogel, Heckel,

Vilcinskas, & Kaltenpoth, 2018; Torrazza & Neu, 2011; Wang & Rozen, 2017). Of interest is the role that parent-offspring interactions play in the microbial acquisition during early development, specifically from mother to her offspring (Adair & Douglas, 2017; Dimmitt et al.,

2010; Duranti et al., 2017; Fox & Eichelberger, 2015; Funkhouser & Bordenstein, 2013; Gilbert,

2014; Jašarević, Rodgers, et al., 2015; Korpela et al., 2018; Kostic et al., 2015; Perez-Muñoz et al., 2017; Schwab et al., 2016; Torrazza & Neu, 2011; Wade, 2014; Walker, Clemente, Peter, &

Loos, 2017).

The animal’s reproductive mode, in part, mediates the types of interactions mothers have with their offspring. Egg laying (oviparous) organisms have limited opportunity to pass microbes to offspring before they are born through hatching (Abdul Rahman et al., 2015; Bright

& Bulgheresi, 2010; da Costa & Poulsen, 2018; Estes et al., 2013; Funkhouser & Bordenstein,

78 2013; Salem, Florez, Gerardo, & Kaltenpoth, 2015; Schwab et al., 2016; Shukla et al., 2018). This forces vertical symbiont transmission to occur through incorporation during oogenesis or by inoculating the external egg surface for consumption immediately upon juvenile emergence

(Abdul Rahman et al., 2015; Estes et al., 2013; Funkhouser & Bordenstein, 2013; Schwab et al.,

2016; Shukla et al., 2018). Viviparous (live-bearing) animals can have extensive and complex interactions between mother and offspring during gestation and birth, the impacts of which can last for a few days to years (Cao-Lei et al., 2014, 2017; Duranti et al., 2017; Funkhouser &

Bordenstein, 2013; Jašarević, Rodgers, et al., 2015; Jiménez-Chillarón et al., 2015; Ma et al.,

2014; Ogawa & Miura, 2014; Poulin & Thomas, 2008; Stein & Lumey, 2000; Torrazza & Neu,

2011; Weiss et al., 2011). These prolonged interactions provide means for multiple routes of vertical transmission of microbes from mother to her progeny (Funkhouser & Bordenstein,

2013; Ma et al., 2014; Mueller et al., 2015). In humans, while placental transmission of microbes is debated (Aagaard et al., 2014; Blaser & Dominguez-Bello, 2016; Fardini, Chung,

Dumm, Joshi, & Han, 2010; Perez-Muñoz et al., 2017; Walker et al., 2017), mother to newborn transfer can occur during passage through the birth canal, feeding and throughout early postnatal development (Ballard & Morrow, 2013; Dahlen, Downe, Kennedy, & Foureur, 2014;

Duranti et al., 2017; Funkhouser & Bordenstein, 2013; Jašarević, Howerton, Howard, & Bale,

2015; Jašarević, Rodgers, et al., 2015; Korpela et al., 2018; Ma et al., 2014; Mueller et al., 2015).

Other live-bearing animals and their symbionts have evolved to utilize the extended gestation as a time to inoculate progeny with bacteria (Denlinger & Ma, 1975; Funkhouser & Bordenstein,

2013; Ma et al., 2014; Morse et al., 2013; Mueller et al., 2015; Wang et al., 2013). This is exemplified in tsetse flies and other members of Hippoboscoidea, where mothers utilize

79 nutritive secretions as a mechanism to transfer required symbiotic bacteria to their intra- uterine developing larvae (Douglas, 2017; Morse et al., 2013; Snyder & Rio, 2015; Wang et al.,

2013; Weiss et al., 2011). For tsetse flies, symbiotic bacteria, specifically Wigglesworthia, provide key B vitamins that are low in their food source (blood) or within milk transferred to the developing intrauterine (Akman et al., 2002; Attardo et al., 2019; Rio et al., 2012).

Here, we examine shifts in the microbiome of the live bearing cockroach, Diploptera punctata, during pregnancy and development.

D. punctata reproduces by matrotrophic viviparity, in which embryos develop inside the brood sac, a unique organ which functions as both a uterus and pseudo-placenta, and are provided with nutrients by a secretion of milk-like components (Hagan, 1939, 1941, Roth &

Willis, 1955, 1957; Stay & Coop, 1973). This secretion appears in embryo gut contents at 20% of the 60-70-day pregnancy, when the dorsal edge of the body wall is closed (Ingram, Stay, & Cain,

1977; Roth & Willis, 1955; Stay & Coop, 1973, 1974). Diploptera milk is a combination of proteins and free amino acids, carbohydrates, and lipids in a water base (Ingram et al., 1977;

Stay & Coop, 1974; Williford, Stay, & Bhattacharya, 2004; Youngsteadt, Fan, Stay, & Schal,

2005). The proteins present include a unique family of lipocalin-like milk proteins (Ingram et al.,

1977; Stay & Coop, 1974; Williford et al., 2004). While this milky secretion provides vital nutrients to developing embryo, it is deficient in two essential amino acids, methionine and tryptophan (Ingram et al., 1977; Williford et al., 2004). It has been proposed that bacterial endosymbionts provide these two nutrients (Williford et al., 2004), however in oviparous cockroaches the only bacterium transmitted from mother to embryo belongs to the

Flavobacteria family Blattabacteriaceae (Bandi et al., 1994, 1995; Giorgi & Nordin, 1994). Most,

80 but not all, strains of Blattabacteria have an incomplete biosynthetic pathway for methionine

(Huang, Sabree, & Moran, 2012; Kambhampati, Alleman, & Park, 2013; López-Sánchez et al.,

2008, 2009; Patiño-Navarrete, Moya, Latorre, & Peretó, 2013; Sabree, Kambhampati, & Moran,

2009; Tokuda et al., 2013).

This study determined the microbiome of Diploptera punctata throughout development, characterizing the microbial communities inhabiting female D. punctata and their offspring across development using 16S rRNA gene sequencing. The information generated by this study will provide the first step in developing D. punctata as a model system to elucidate how intrauterine development and the prenatal microbiome affect later acquisition of microbial endosymbionts. Developing a new model system understanding microbial shifts during invertebrate matrotrophic viviparity will widen the evolutionary lens through which we view reproduction and the microbiome in viviparous animals.

Methods

Animals

Colonies reared at the University of Cincinnati (UC) Department of Biological Sciences

(Cincinnati, OH) were housed in a climate-controlled facility. Ambient temperature was held between 24-28°C and relative humidity (RH) was held between 70-80%. A 12:12 hour light-dark photoperiod was maintained for the duration of the experiment. Animals were provided water and fed Old Roy Complete Nutrition brand dog food (Mars, Inc.) ad libitum. A second group of

D. punctata were obtained from The Ohio State University (OSU) Biological Sciences

Greenhouse (Columbus, OH) insect collection where they were reared in similar conditions with

81 the exception of being fed a diet of Tetramin fish food (Spectrum Brands Pet). Upon retrieval, this sampling of the OSU colony was housed separately under identical conditions from the UC colony and provided the same food and water sources as the UC colonies.

Sample collection

Visibly pregnant females were selected from the colony for use in mother-embryo comparisons.

Females were surface sterilized by rinsing for 1 minute in each of the following solutions: 70% ethanol and 2% sodium hypochlorite. This was followed by four rinses in sterile phosphate- buffered saline (PBS; 81mM Na2HPO4, 19mM NaH2PO4, 150mM NaCl, pH 7.4). Embryo broods were then dissected from the brood sac in sterile PBS by making two small incisions at the opening of the brood sac, one on each side, and removed using ethanol sterilized forceps. To determine the developmental stage of the embryos, a single embryo from the center of each brood was measured on a bleach sterilized ruler and designated as pre-lactation, early lactation, or late lactation based on its length (Chapter 1, Table 1) (Stay & Coop, 1973). Entire broods of embryos and individual mothers were then placed into separate 1.5 mL centrifuge tubes with silica beads and stored at -80°C until processing. While mother-embryo pairs were collected for all three trimesters, only late lactation pairs were utilized in this study.

To characterize the postnatal development of the microbiome, visibly pregnant females were again selected from the colony and housed in individual containers with food and water ad libitum and monitored for active birthing. Nymphs were collected as neonates (identified by lack of cuticular melanization), first instars (identified by melanization within 12 hours of birth).

Second, third, and fourth instar nymphs were sampled and identified by size and the presence

82 of molts in the living quarters. Postnatal samples were collected only from the UC colonies.

Upon collection, nymphs were surface sterilized using the methods described above and then stored in 1.5 mL centrifuge tubes with silica beads at -80°C until processing.

Genomic DNA preparation

Samples were homogenized in 1µl of sterile 1X PBS and DNA was extracted using a QIAGEN

DNeasy Blood and Tissue Kit (QIAGEN). The homogenate (200µl) was incubated with proteinase

K (QIAGEN) over night before continuing the provided protocol. DNA concentration and quality were measured via A260/A280 and A260/A230 using a Nanodrop 2000 (Thermo Scientific). All samples were diluted to 20 ng/µl for sequencing.

16S rRNA Sequencing and Bioinformatic Analyses

The V4 hypervariable region of the bacterial 16S rRNA gene was PCR amplified using the 515f

(GTGYCAGCMGCCGCGGTAA) and 806r (GGACTACNVGGGTWTCTAAT) universal primers (Apprill,

Mcnally, Parsons, & Weber, 2015; Caporaso et al., 2011). Amplicon sequencing using the MiSeq

Illumina 2x300bp chemistry was conducted at the Miami University Center for Bioinformatics &

Functional Genomics (Oxford, OH, USA) as well as the University of Minnesota Genomics Center

(Minneapolis, MN, USA).

Using the Ohio Supercomputer Center resources (Ohio Supercomputer Center, 1987), sequence reads were processed in mothur (v.1.39.5) (Schloss et al., 2009) based on the published MiSeq SOP (Kozich, Westcott, Baxter, Highlander, & Schloss, 2013). Briefly, the make.contigs command was used to extract quality data from the reads and only reads

83 possessing a quality score greater than 25 were joined to make the contigs for further analysis.

Screen.seqs was utilized to remove contigs containing ambiguous bases, contigs longer than

275 bp, and those containing homopolymers longer than 8 bp. Unique.seqs and count.seqs were utilized to remove duplicate sequences and generate count tables. Taxonomic assignment of sequences was conducted using align.seqs to compare the contigs to the SILVA database

(v.123) (Quast et al., 2013) containing only the V4 region aligning with the primers used.

Filter.seqs was used to remove sequences that have large gaps in the alignments. Chimeric sequences were removed using the UCHIME (Edgar, Haas, Clemente, Quince, & Knight, 2011) algorithm using the chimers.uchime and remove.seqs commands. Non 16S rRNA gene sequences were removed using the classify.seqs and remove.lineage commands. Sequences were clustered using the cluster.split command at the taxonomic level 4, representing order. All further analyses were conducted using operational taxonomic unit (OTU) assignments generated in the above steps. Rarefaction curves were generated using the rarefaction.single and the number of observed OTUs (sobs), demonstrating adequate sequencing depth. Alpha diversity was assessed using the inverse Simpson, and Shannon diversity metrics. NMDS and

PCOA analyses were conducted using mothur. For mother-embryo comparisons, community composition was manually assessed for visualization at taxonomic level 5, representing bacterial families. In addition to mothur we performed a second analysis of our data for validation purposes utilizing QIIME (v. 1.9.1) (Caporaso et al., 2010) as implemented by the

Nephele pipeline (v. 2.2.2) (Weber et al., 2018) using the default settings, referencing the SILVA database (v.128 SSU REF 99) (Quast et al., 2013). When relative abundances calculated at the class level by both methods were compared, they were found to be significantly correlated

84 (Figure 1); consequently, results from mothur were reported. Additional results from the QIIME analysis can be found in supplemental data files 1 and 2.

Data processing was conducted in Microsoft Excel (v.16.22) and R (v.3.3.3) (R Core

Team, 2017) using RStudio (v1.1.423) (RStudio Team, 2015). Additional statistics and graphical representations of data were also performed in R using RStudio. Packages utilized include dplyr

(Wickham, Francois, Henry, & Müller, 2017), dunn.test (Dinno, 2017), ggplot2 (Wickham, 2016), reshape2 (Wickham, 2007), RColorBrewer (Neuwirth, 2014), Rmisc (Hope, 2013), and wesanderson (Ram & Wickham, 2018).

Results

Maternal and embryonic microbiomes

Amplicons from the 16S rRNA generated 2,180,632 paired end reads from both OSU and UC colony mothers and embryos of D. punctata, assembled into 2,170,187 contigs when joined. Of those, 1,759,259 total sequences passed quality control and were classified as Archaea (8,320 reads; 0.473%), Bacteria (1,750,772 reads; 99.518%), or unknown (167 reads; 0.009%). Removal of unwanted classifications (archeae, chloroplast, eukaryote, mitochondria, and unknown) yielded 1,749,921 merged reads, ultimately generating 38,969 bacterial operational taxonomic units (OTUs) corresponding to 44 phyla, 108 classes, 204 orders, 386 families, and 710 genera.

Overall, was the most prominent (21,099 OTUs; 54.143%), followed by

Firmicutes (5,513 OTUs; 14.147%), Proteobacteria (4,783 OTUs; 12.274%), and unclassified bacteria (4,286 OTUs; 10.998%) (Figure 2). Flavobacteria, a class of Bacteroidetes, was the most prominent class overall (14,636 OTUs; 37.558%), with unclassified bacteria (4,286 OTUs;

85 10.998%) and Bacteroidia (4,042 OTUs; 10.372%) as the second and third most represented.

Orders followed suit with (14,636 OTUs; 37.558%), unclassified bacteria (4,286

OTUs; 10.998%) and Bacteroidales (3,897; 10.000%) as the only orders representing 10% or more of the total OTUs. At the family level, Blattabacteriaceae, a family of Flavobacteria, was the most represented overall in both OTUs (14,426 OTUs; 37.019%) and reads (1,038,785 reads;

59.047% of all reads including non-bacterial) with unclassified bacteria being the next most abundant family (4,286 OTUs; 10.998%) followed by unclassified Bacteroidetes (2,260; 5.799%) and Ruminococcaceae (1,890; 4.850%) (Figure 2).

In mothers, OTUs were distributed among the same top four phyla (Bacteroidetes,

35.354%; Firmicutes, 27.714%; Proteobacteria, 14.138%; unclassified bacteria, 11.609%), with a similar distribution among mothers of both the OSU and UC colonies. At the family level, OTUs derived from D. punctata mothers were most represented in Blattabacteriaceae (6,734;

13.842%), Ruminococcaceae (5,781; 11.883%) and unclassified bacteria (5,648; 11.609%)

(Figure 2). Mothers from the OSU and UC colonies had similar distributions of OTUs among families. Additionally, there was no significant difference between the two colonies in community diversity or evenness (Figure 3). We identified a core community of 2,314 OTUs shared between mothers of both colonies, composed of 25 phyla with Firmicutes and

Bacteroidetes representing more than 60% of OTUs (Figure 4). Clostridia and Clostridiales were the most abundant of the 47 classes and orders respectively, each making up just over 30% of the core community. No individual family represented more than 16% of the core OTUs, with

Ruminococcaceae (16%) being the most abundant of the top eight families (52%).

86 Approximately 89% of OTUs and 99% of sequencing reads from embryos of both colonies belonged to the family Blattabacteriaceae, while all other families individually represented 1% or less of OTUs and 0.08% of embryo derived sequencing reads (Figure 2).

These findings were corroborated by secondary analyses completed using the Nephele implementation of QIIME, despite inherent differences in computational methods (Figure 1).

Embryos of both UC and OSU colonies did not differ significantly in diversity, evenness, and species richness (Figure 3). However, microbial communities of embryos were less diverse and less so than mothers across both colonies (Figure 3). Analysis of molecular variance and homogeneity of molecular variance in mothur revealed that despite our four sampling groups consisting of mothers and embryos from two distinct colonies (Table 1 and Table 2), there exist three distinct sub communities corresponding to UC mothers, OSU mothers, and all embryos

(Figure 4).

While the transmission of the cockroach specific endosymbiont Blattabacteria is known to occur during oogenesis (Sacchi et al., 1996), surface sterilization of oothecae and hatching into a sterile environment results in a microbiome exclusively composed of Blattabacteria, indicating any other bacteria must be acquired from food or feces (Pietri, Tiffany, & Liang,

2018). Such is the case in the intergenerational transfer of microbiota via proctodeal trophallaxis in Cryptocercus punctulatus and Mastotermes darwiniensis (McMahan, 1969).

Because D. punctata harbor their developing embryos for their gestational period, it is possible other bacteria may be transmitted via the brood sac. The low diversity and overall OTUs present in embryonic samples, however, suggest that if other bacteria are transmitted during gestation the number is very low and is not likely of significance to D. punctata embryos. This

87 indicates that Blattabacteria are the main endosymbiont during intrauterine development in D. punctata and that any additional constituents of the microbiome colonize after birth.

Postnatal microbiome development

We next sought to determine the progression of the microbiome over postnatal development.

Because we found no significant differences between the OSU and UC colonies of D. punctata the samples were recategorized for subsequent analyses and denoted simply as mothers and embryos. To determine the succession of the microbial communities inhabiting D. punctata from embryo to adulthood, we surveyed the microbiome of neonate nymphs and each of the following nymphal instars (one through four).

A total of 6,453,793 paired reads from mothers, embryos, and juvenile instars were used to generate 6,443,348 contigs in mothur. Of these, 4,752,552 passed quality control and were able to be taxonomically classified as either Archeae (14,141; 0.298%), Bacteria

(4,737,007; 99.673%), or unknown (1,402; 0.029%). There were also two Eukaryotic reads

(<0.001%). Removing unwanted reads as before, 4,734,605 remained and were utilized to generate 209,554 bacterial operational taxonomic units (OTUs) including 50 phyla, 122 classes,

252 orders, 485 families, and 1008 genera. As expected, Bacteroidetes was again the most abundant phylum (122,945 OTUs; 58.670%) when all samples were combined, followed by

Firmicutes (29,705 OTUs; 14.175%), unclassified bacteria (24,777 OTUs; 11.824%), and

Proteobacteria (20,068 OTUs; 9.577%). Flavobacteria and unclassified bacterial classes comprised 54.979% of class-level OTUs, a trend that holds true at the order level as well. At the family level Blattabacteriaceae (42.877%) was again the most prominent taxon followed by

88 unclassified bacteria (11.824%), unclassified Bacteroidetes (5.515%), and Porphyromonadaceae

(4.506%) (Figure 5).

Bacteroidetes (94.985%), and more specifically Blattabacteriaceae (93.387%), again were the primary constituent of the embryonic microbial community, while other families each represented 0.716% or less of the OTUs present (Figure 5). The dominance of the microbial community by Blattabacteriaceae persisted after birth during the neonate stage (93.741%) with each other family representing less than 1% of the community (Figure 5). Post-melanization first instars, however, have a more diverse microbial community. While Bacteroidetes remains the major phylum represented, it represents a much smaller portion of the OTUs (53.563% down from the roughly 95% of embryos and neonates). Firmicutes and Proteobacteria are the next most abundant phyla. Blattabacteriaceae is still the most abundant family (40.442%), a significant portion of the community (a combined 23.494% of OTUs) is made up by unclassified bacteria (6.281%), Enterobacteriaceae (6.125%), unclassified Lactobacillales (5.737%), and

Porphyromonadaceae (5.351%), while all other families individually represented less than 4% of the first instar microbial community (Figure 5). Second instars had the same major phyla represented, but had more families represented in high levels. Blattabacteriaceae represented only 28.303% of the community, while unclassified bacteria (12.978%), Porphyromonadaceae

(7.728%), Ruminococcaceae (7.705%), and unclassified Bacteroidetes (5.854%) increased in representation and together make up 34.265% of the OTUs. This redistribution of abundance from Blattabacteriacieae is maintained after the second instar stage, with abundances in third and fourth instars remaining around 30% (Figure 5). Adult females had even lower abundances of Blattabacteriaceae, although it was still the most abundant family (18.826%). All families

89 represented less than 20% of the OTUs present, with Ruminococcaceae and unclassified bacteria being the only two with abundances greater than 10% (Figure 5).

Multiple measures of diversity varied across the life stages of D. punctata. While embryos and neonates did not differ in either the Shannon index or Inverse Simpson, all other instars and mothers were significantly different from embryos in both measures (Figure 6).

Neonates also did not differ from first instars but showed significant differences in both diversity metrics compared to second, third, and fourth instars as well as adult females. Second, third, and fourth instars, however, did not differ from each other or mothers in any diversity measure (Figure 6). While amova and homova analyses revealed slightly different relationships between the samples (Table 3 and Table 4), the analyses consistently showed that embryos and neonates differed from the other juvenile stages and adult females. These results further support our hypothesis that D. punctata acquire microbial endosymbionts (outside of

Blattabacteria), not through direct maternal transfer during gestation but in the days and weeks after birth, primarily during and after initial melanization during the first nymphal instar.

Discussion

We identified 50 phyla, 122 classes, 252 orders, 485 families, and 1008 genera as part of the overall Diploptera punctata microbial community. Our analyses revealed that Bacteroidetes,

Firmicutes, and Proteobacteria were the dominant phyla in addition to unclassified bacteria.

Previous studies have characterized microbial communities of cockroaches, primarily the gut microbiome. Consistent with our findings, Bacteroidetes, Firmicutes, Proteobacteria and unclassified bacteria are repeatedly found to be prominent members of adult cockroach

90 endosymbiont communities (Bauer et al., 2015; Bertino-Grimaldi et al., 2013; Carrasco et al.,

2014; Gontang et al., 2017; Kakumanu, Maritz, Carlton, & Schal, 2018; Pérez-Cobas et al., 2015;

Schauer, Thompson, & Brune, 2014; Tinker & Ottesen, 2016). Similar to our adult female samples, other studies have shown Porphyromonadaceae, Rikenellaceae, and Bacteroidaceae to be the most abundant families of Bacteroidetes; while Lachnospiraceae, Ruminococcaceae,

Clostridiaceae, and Lactobacillaceae are commonly represented Firmicutes (Bauer et al., 2015;

Bertino-Grimaldi et al., 2013; Carrasco et al., 2014; Gontang et al., 2017; Kakumanu et al., 2018;

Pérez-Cobas et al., 2015; Sabree & Moran, 2014; Schauer et al., 2014; Tinker & Ottesen, 2016).

Proteobacteria present in cockroach microbiomes often belong to the families

Desulfobacteraceae, Enterobacteriaceae, Desulfovibrionaceae (Bauer et al., 2015; Bertino-

Grimaldi et al., 2013; Carrasco et al., 2014; Gontang et al., 2017; Kakumanu et al., 2018; Pérez-

Cobas et al., 2015; Sabree & Moran, 2014; Schauer et al., 2014; Tinker & Ottesen, 2016). Most previous cockroach microbiome studies found extremely low representation of Blattabacteria or do not report on its abundance due to the specific sampling of gut tissue; Blattabacteria reside in the fat body and ovaries and thus will be lacking in studies focus on the gut microbiome (Bauer et al., 2015; Bertino-Grimaldi et al., 2013; Carrasco et al., 2014; Gontang et al., 2017; Kakumanu et al., 2018; Pérez-Cobas et al., 2015; Sabree & Moran, 2014; Schauer et al., 2014; Tinker & Ottesen, 2016). The few studies that performed microbiome analyses on whole bodies or carcasses without guts, however, report Blattabacteriaceae abundances ranging from 8% to 90% depending on the habitat sampled, although carcasses without guts were generally found to contain predominantly Blattabacteria (Carrasco et al., 2014; Kakumanu et al., 2018).

91 Investigations of developmental acquisition of the cockroach microbiome are rare, however one study characterized the succession of the microbiota in the oviparous German cockroach, Blattella germanica (Carrasco et al., 2014). Contents of surface sterilized oothecae contain exclusively Blattabacteria and whole bodies of first instar nymphs that hatched from unsterilized oothecae contain predominantly Blattabacteria, but have begun to acquire other gut symbionts (Carrasco et al., 2014). Despite the difference in reproductive mode, we found similar results in the intrauterine developing embryos and neonatal D. punctata.

One previous study has attempted to characterize the microbiome of D. punctata mothers and embryos, concluding that there are significant amounts of non-Blattabacteria microbes in embryos (Ayayee, Keeney, Sabree, & Muñoz-Garcia, 2017). In direct contrast, our embryo samples from two independent colonies, including the colony used in the previous study, produced sequencing reads that were 99.5% assigned to Blattabacteriaceae. Two taxa identified to be significantly enriched in the embryonic microbiome by this previous study were

Halomonadaceae and Shewanellaceae (Ayayee et al., 2017), neither of which were present in our maternal, embryo, or postnatal development samples. While our analysis using mothur did identify non-Blattabacteria sequences in embryonic samples, the extremely low abundances

(less than 0.5% of total raw reads combined) suggest they are sequencing artifacts or misidentified and are not likely critical for embryos during gestation. This is further supported by our secondary analysis using the Nephele implementation of QIIME, which identified no taxa representing more than 0.2% of the embryonic community other than Blattabacteria. Because of our robust sampling method, including two separately housed colonies of D. punctata from separate institutional origins and use of two independent pipelines for analysis, we conclude

92 that no bacterial transmission occurs after oogenesis during intrauterine development in D. punctata. Thus, Blattabacteria is the only bacterial component of the microbiome during intrauterine development. This is further supported by the lack of additional bacterial components in first instar nymphs collected immediately after birth (=neonate). While we cannot eliminate rearing differences, our study indicates that other bacteria, beyond

Blattabacteria, are not required for D. punctata development.

After determining that there was no significant gestational transmission of endosymbionts, we sought to characterize the microbial community across nymphal development. D. punctata juveniles have a minimum of three nymphal instars with females molting an additional time to a fourth instar stage. Newborn, unmelanized first instar nymphs did not differ in bacterial community from intrauterine developing embryos suggesting that significant bacterial transmission does not occur during the birthing process, unlike humans.

However, by the time first instars fully develop a hardened cuticle they have developed a more diverse microbial community where Blattabacteria represents only 35% of the OTUs. This substantial increase is likely the results of food and water consumption that occurs following melanization. Across the remaining instars, the community continues to become more diverse, however the changes become much less dramatic after the second instar stage. These findings are again consistent with a previous study investigating the juvenile microbiome of B. germanica as well as in other egg laying organisms such as burying beetle Nicrophorus vespilloides (Carrasco et al., 2014; Wang & Rozen, 2017). Consequently, we conclude that the microbial community is largely acquired during the first and second instar stages, likely from their environment where they cohabitate with both adult and other juvenile cockroaches, after

93 they have started to feed and drink There are continually changes to the microbiome throughout the life of the animal, but these are minor compared to the initial acquisition in early developmental stages.

This initial acquisition period of the microbiome is extremely important to animal development (Albenberg & Wu, 2014; Ballou et al., 2016; Breznak & Kane, 1990; Brownlie &

Johnson, 2009; Chung et al., 2012; Colston, 2017; Coon, Brown, & Strand, 2016; Coon, Vogel,

Brown, & Strand, 2014; Diaz Heijtz, 2016; Dimmitt et al., 2010; Hamdi et al., 2011; Kostic et al.,

2015; Lee & Brey, 2013; Ma et al., 2014; Malmuthuge, Griebel, & Guan, 2015; McFall-Ngai,

2014; Michalkova et al., 2014; Pais et al., 2008; Pietri et al., 2018; Schwab et al., 2016; Snyder &

Rio, 2015; Thompson, Rivera, Closek, & Medina, 2015; Torrazza & Neu, 2011; Wade, 2014; Yang et al., 2016). Studies in insect systems have demonstrated this by ablating the microbiome of juvenile animals and observing the phenotypes. Consistently, these experiments find that animals unable to acquire microbes from their environment or mothers face severe disadvantages, often failing to progress from one instar to the next, unable to molt to adulthood or undergo pupation, or dying. One example of this is the inability of axenic larvae to reach adulthood (Coon et al., 2016, 2014). In the dung beetle Onthophagus gazella, removal of a maternally provided fecal secretion, known as the pedestal, significantly reduces bacterial load in larvae hatched from surface sterilized eggs (Schwab et al., 2016).

While preventing microbiome acquisition in O. gazelle larvae doesn’t result in mortality as in mosquitoes, it is associated with reduced larval mass, increased time to adulthood, smaller adult body size, and impaired dehydration tolerance (Schwab et al., 2016). In tsetse flies,

Wigglesworthia glossinidia transmission via milk gland secretions is not only essential for B

94 vitamin provisioning, but also immune function (Weiss et al., 2011). Targeted elimination of this symbiont decreased the population of phagocytic hemocytes and reduced melanization ability

(Weiss et al., 2011). Symbiont community composition has also been implicated in resistance in the German cockroach (Pietri et al., 2018). Elimination of all bacteria from the cockroach except for Blattabacteria throughout development suggests that insecticide resistance are due to changes in non-Blattabacteria bacteria which are acquired after hatching

(Pietri et al., 2018). These studies underscore the importance of developing a diverse and robust microbial community during early nymphal development, which we have found primarily occurs during the first instar of D. punctata.

The embryonic microbiome comprised exclusively of Blattabacteria is of interest relative to the intra-uterine development of D. punctata embryos, as the milk-like secretion provided by mothers as the sole form of nutrition during development is largely devoid of two essential amino acids, methionine and tryptophan (Stay & Coop, 1974; Williford et al., 2004).

Consequently, it has been suggested that these amino acids are acquired from bacterial endosymbionts (Williford et al., 2004). Bacterial symbionts commonly serve to supplement nutrients that may be lacking in the diet (Bermingham & Wilkinson, 2009; Douglas, 2017; Engel

& Moran, 2013; Funkhouser & Bordenstein, 2013; Michalik, Szklarzewicz, Jankowska, &

Wieczorek, 2014; Michalkova et al., 2014). Viviparous insects, such as tsetse flies, take advantage of endosymbionts to fill such nutritional gaps during development, mostly through the provisioning of B vitamins (Douglas, 2017; Snyder, Mclain, & Rio, 2012; Snyder & Rio, 2015;

Wang et al., 2013). However, while Wolbachia is transmitted through the germ line before nutrient provisioning (Wang et al., 2013), other symbionts in these flies, such as Wigglesworthia

95 and Sodalis, have been shown to be transmitted from mother to offspring during their extended gestation period (Denlinger & Ma, 1975; Douglas, 2017; Snyder et al., 2012; Snyder &

Rio, 2015; Wang et al., 2013). The exclusively Blattabacterial composition of the microbiome in embryos suggests that this symbiont must be the source of these essential nutrients. However, previous studies characterizing the genome of Blattabacteria inhabiting other species of cockroaches have shown that only the strain belonging to the German cockroach (Blattella germanica) possesses the capability to synthesize methionine, one of the amino acids lacking in

D. punctata milk, in any capacity (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008, 2009; Neef et al., 2011; Patiño-Navarrete et al., 2013; Sabree et al., 2012, 2009;

Tokuda et al., 2013). Consequently, further investigation of this symbiotic relationship is required to understand the role of Blattabacteria during intrauterine development. Sequencing the genome of the D. punctata strain of Blattabacteria may reveal the presence of biosynthetic pathways that can provide amino acids required for prenatal development.

In conclusion, we provide a comprehensive survey of the microbial communities of mothers and their developing embryos along with succession of the microbiome community across postnatal development in D. punctata. This study provides evidence that, unlike other viviparous insects, there is no transmission of bacteria from mothers to offspring during their

63+ day pregnancy. Surprisingly, we also found no evidence that there is significant bacterial colonization of D. punctata during birth or within the few hours immediately following birth.

Rather, a majority of the microbiome components are acquired, likely from their environment, throughout the full duration of the first instar and melanization period. Further investigation will be required to further elucidate the specific mechanisms underlying nutrient provisioning

96 by Blattabacteria during embryonic development in D. punctata, as well as the role of the microbiome during nymphal development.

Data Accessibility

Sequence data have been added to the NCBI Sequence Read Archive (SRA) database

(PRJNA522760).

Author Contributions

E.C.J. and J.B.B. conceived the study. E.C.J. designed the experiments and, with guidance from

T.L.H., analyzed all data. E.C.J. and M.W.K. collected samples, performed DNA extractions and prepared samples for sequencing. T.L.H. coordinated sample transportation and sequencing.

E.C.J. and J.B.B. wrote the paper and T.L.H. edited the paper. E.C.J., M.W.K., T.L.H. and J.B.B. contributed substantially to interpreting the data and developing the manuscript and take full responsibility for the content of the paper.

97 References:

Aagaard, K., Ma, J., Antony, K. M., Ganu, R., Petrosino, J., & Versalovic, J. (2014). The placenta

harbors a unique microbiome. Science Translational Medicine, 6(237), 237ra65.

http://doi.org/10.1126/scitranslmed.3009864

Abdul Rahman, N., Parks, D. H., Willner, D. L., Engelbrektson, A. L., Goffredi, S. K., Warnecke, F.,

… Hugenholtz, P. (2015). A molecular survey of Australian and North American termite

genera indicates that vertical inheritance is the primary force shaping termite gut

microbiomes. Microbiome, 3(1), 1–16. http://doi.org/10.1186/s40168-015-0067-8

Adair, K. L., & Douglas, A. E. (2017). Making a microbiome: the many determinants of host-

associated microbial community composition. Current Opinion in Microbiology, 35, 23–29.

http://doi.org/10.1016/j.mib.2016.11.002

Akman, L., Yamashita, A., Watanabe, H., Oshima, K., Shiba, T., Hattori, M., & Aksoy, S. (2002).

Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia

glossinidia. Nature Genetics, 32(3), 402–407. http://doi.org/10.1038/ng986

Albenberg, L. G., & Wu, G. D. (2014). Diet and the intestinal microbiome: Associations,

functions, and implications for health and disease. Gastroenterology, 146(6), 1564–1572.

http://doi.org/10.1053/j.gastro.2014.01.058

Apprill, A., Mcnally, S., Parsons, R., & Weber, L. (2015). Minor revision to V4 region SSU rRNA

806R gene primer greatly increases detection of SAR11 bacterioplankton. Aquatic

Microbial Ecology, 75(2), 129–137. http://doi.org/10.3354/ame01753

Attardo, G. M., Abd-Alla, A. M. M., Acosta-Serrano, A., James, E., Bateta, R., Benoit, J. B., …

Aksoy, S. (2019). The Glossina genome cluster: Comparative genomic analysis of the

98 vectors of African Trypanosomes. BioRxiv, 531749. http://doi.org/10.1101/531749

Ayayee, P. A., Keeney, G., Sabree, Z. L., & Muñoz-Garcia, A. (2017). Compositional differences

among female-associated and embryo-associated microbiota of the viviparous Pacific

Beetle cockroach, Diploptera punctata. FEMS Microbiology Ecology, 93(6) fix052.

http://doi.org/10.1093/femsec/fix052

Ballard, O., & Morrow, A. L. (2013). Human Milk Composition. Nutrients and Bioactive Factors.

Pediatric Clinics of North America, 60(1), 49–74. http://doi.org/10.1016/j.pcl.2012.10.002

Ballou, A. L., Ali, R. A., Mendoza, M. A., Ellis, J. C., Hassan, H. M., Croom, W. J., & Koci, M. D.

(2016). Development of the chick microbiome: How early exposure influences future

microbial diversity. Frontiers in Veterinary Science, 3(January), 1–12.

http://doi.org/10.3389/fvets.2016.00002

Bandi, C., Damiani, G., Magrassi, L., Grigolo, A., Fani, R., & Sacchi, L. (1994). Flavobacteria as

intracellular symbionts in cockroaches. Proceedings of the Royal Society B: Biological

Sciences, 257, 43–48. http://doi.org/10.1098/rspb.1994.0092

Bandi, C., Sironi, M., Damiani, G., Magrassi, L., Nalepa, C. A., Laudant, U., & Sacchi, L. (1995).

The establishment of intracellular symbiosis in an ancestor of cockroaches and termites.

Proceedings of the Royal Society B: Biological Sciences, 259(1356), 293–299.

http://doi.org/10.1098/rspb.1995.0043

Bauer, E., Lampert, N., Mikaelyan, A., Köhler, T., Maekawa, K., & Brune, A. (2015).

Physicochemical conditions, metabolites and community structure of the bacterial

microbiota in the gut of wood-feeding cockroaches (Blaberidae: Panesthiinae). FEMS

Microbiology Ecology, 91(2), 1–14. http://doi.org/10.1093/femsec/fiu028

99 Bermingham, J., & Wilkinson, T. L. (2009). Embryo nutrition in parthenogenetic viviparous

aphids. Physiological Entomology, 34(2), 103–109. http://doi.org/10.1111/j.1365-

3032.2008.00669.x

Bertino-Grimaldi, D., Medeiros, M. N., Vieira, R. P., Cardoso, A. M., Turque, A. S., Silveira, C. B.,

… Machado, E. A. (2013). Bacterial community composition shifts in the gut of Periplaneta

americana fed on different lignocellulosic materials. SpringerPlus, 2(1), 1–11.

http://doi.org/10.1186/2193-1801-2-609

Blaser, M. J., & Dominguez-Bello, M. G. (2016). The Human Microbiome before Birth. Cell Host

and Microbe, 20(5), 558–560. http://doi.org/10.1016/j.chom.2016.10.014

Breznak, J. A., & Kane, M. D. (1990). Microbial H2/CO2 acetogenesis in animal guts: nature and

nutritional significance. FEMS Microbiology Letters, 87(3–4), 309–313.

http://doi.org/10.1016/0378-1097(90)90471-2

Bright, M., & Bulgheresi, S. (2010). A complex journey: Transmission of microbial symbionts.

Nature Reviews Microbiology, 8(3), 218–230. http://doi.org/10.1038/nrmicro2262

Brownlie, J. C., & Johnson, K. N. (2009). Symbiont-mediated protection in insect hosts. Trends in

Microbiology, 17(8), 348–354. http://doi.org/10.1016/j.tim.2009.05.005

Cao-Lei, L., de Rooij, S. R., King, S., Matthews, S. G., Metz, G. A. S., Roseboom, T. J., & Szyf, M.

(2017). Prenatal stress and epigenetics. Neuroscience & Biobehavioral Reviews, Pre-print.

http://doi.org/10.1016/j.neubiorev.2017.05.016

Cao-Lei, L., Massart, R., Suderman, M. J., Machnes, Z., Elgbeili, G., Laplante, D. P., … King, S.

(2014). DNA methylation signatures triggered by prenatal maternal stress exposure to a

natural disaster: Project Ice Storm. PLoS ONE, 9(9), e107653.

100 http://doi.org/10.1371/journal.pone.0107653

Caporaso, J. G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F. D., Costello, E. K., …

Knight, R. (2010). QIIME allows analysis of high-throughput community sequencing data.

Nature Methods, 7(5), 335–336. http://doi.org/10.1038/nmeth.f.303

Caporaso, J. G., Lauber, C. L., Walters, W. A., Berg-Lyons, D., Lozupone, C. A., Turnbaugh, P. J., …

Knight, R. (2011). Global patterns of 16S rRNA diversity at a depth of millions of sequences

per sample. Proceedings of the National Academy of Sciences, 108, 4516–4522.

http://doi.org/10.1073/pnas.1000080107

Carrasco, P., Pérez-Cobas, A. E., van de Pol, C., Baixeras, J., Moya, A., & Latorre, A. (2014).

Succession of the gut microbiota in the cockroach Blattella germanica. International

Microbiology, 17, 99–109. http://doi.org/10.2436/20.1501.01.212

Chung, H., Pamp, S. J., Hill, J. A., Surana, N. K., Edelman, S. M., Troy, E. B., … Kasper, D. L. (2012).

Gut immune maturation depends on colonization with a host-specific microbiota. Cell,

149(7), 1578–1593. http://doi.org/10.1016/j.cell.2012.04.037

Colston, T. J. (2017). Gut microbiome transmission in lizards. Molecular Ecology, 26(4), 972–

974. http://doi.org/10.1111/mec.13987

Coon, K. L., Brown, M. R., & Strand, M. R. (2016). Mosquitoes host communities of bacteria that

are essential for development but vary greatly between local habitats. Molecular Ecology,

25(22), 5806–5826. http://doi.org/10.1111/mec.13877

Coon, K. L., Vogel, K. J., Brown, M. R., & Strand, M. R. (2014). Mosquitoes rely on their gut

microbiota for development. Molecular Ecology, 23(11), 2727–2739.

http://doi.org/10.1111/mec.12771

101 da Costa, R. R., & Poulsen, M. (2018). Mixed-mode transmission shapes termite gut community

assemblies. Trends in Microbiology, 26(7), 557–559.

http://doi.org/10.1016/j.tim.2018.04.005

Dahlen, H. G., Downe, S., Kennedy, H. P., & Foureur, M. (2014). Is society being reshaped on a

microbiological and epigenetic level by the way women give birth? Midwifery, 1–3.

http://doi.org/10.1016/j.midw.2014.07.007

Denlinger, D. L., & Ma, W. -C. (1975). Maternal nutritive secretions as possible channels for

vertical transmission of microorganisms in insects: the tsetse fly example. Annals of the

New York Academy of Sciences, 266(1), 162–165. http://doi.org/10.1111/j.1749-

6632.1975.tb35097.x

Diaz Heijtz, R. (2016). Fetal, neonatal, and infant microbiome: Perturbations and subsequent

effects on brain development and behavior. Seminars in Fetal and Neonatal Medicine,

21(6), 410–417. http://doi.org/10.1016/j.siny.2016.04.012

Dimmitt, R. A., Staley, E. M., Chuang, G., Tanner, S. M., Soltau, T. D., & Lorenz, R. G. (2010). Role

of postnatal acquisition of the intestinal microbiome in the early development of immune

function. Journal of Pediatric Gastroenterology and Nutrition, 51(3), 1.

http://doi.org/10.1097/MPG.0b013e3181e1a114

Dinno, A. (2017). dunn.test: Dunn’s test of multiple comparisons using rank sums.

Douglas, A. E. (2017). The B vitamin nutrition of insects: the contributions of diet, microbiome

and horizontally acquired genes. Current Opinion in Insect Science, 23, 65–69.

http://doi.org/10.1016/j.cois.2017.07.012

Duranti, S., Lugli, G. A., Mancabelli, L., Armanini, F., Turroni, F., James, K., … Ventura, M. (2017).

102 Maternal inheritance of bifidobacterial communities and bifidophages in infants through

vertical transmission. Microbiome, 5(1), 1–13. http://doi.org/10.1186/s40168-017-0282-6

Edgar, R. C., Haas, B. J., Clemente, J. C., Quince, C., & Knight, R. (2011). UCHIME improves

sensitivity and speed of chimera detection. Bioinformatics, 27(16), 2194–2200.

http://doi.org/10.1093/bioinformatics/btr381

Engel, P., & Moran, N. A. (2013). The gut microbiota of insects-diversity in structure and

function. FEMS Microbiology Reviews, 37, 699–735.

Estes, A. M., Hearn, D. J., Snell-Rood, E. C., Feindler, M., Feeser, K., Abebe, T., … Moczek, A. P.

(2013). Brood ball-mediated transmission of microbiome members in the dung beetle,

Onthophagus taurus (Coleoptera: Scarabaeidae). PLoS ONE, 8(11), 1–15.

http://doi.org/10.1371/journal.pone.0079061

Fardini, Y., Chung, P., Dumm, R., Joshi, N., & Han, Y. W. (2010). Transmission of diverse oral

bacteria to murine placenta: Evidence for the oral microbiome as a potential source of

intrauterine infection. Infection and Immunity, 78(4), 1789–1796.

http://doi.org/10.1128/IAI.01395-09

Fox, C., & Eichelberger, K. (2015). Maternal microbiome and pregnancy outcomes. and

Sterility, 104(6), 1358–1363. http://doi.org/10.1016/j.fertnstert.2015.09.037

Funkhouser, L. J., & Bordenstein, S. R. (2013). Mom knows best: the universality of maternal

microbial transmission. PLoS Biology, 11(8), 1–9.

http://doi.org/10.1371/journal.pbio.1001631

Gilbert, S. F. (2014). A holobiont birth narrative: The epigenetic transmission of the human

microbiome. Frontiers in Genetics, 5, 1–7. http://doi.org/10.3389/fgene.2014.00282

103 Giorgi, F., & Nordin, J. H. (1994). Structure of yolk granules in and eggs of Blattella

germanica and their interaction with vitellophages and endosymbiotic bacteria during

granule degradation. Journal of Insect Physiology, 40(12), 1077–1092.

http://doi.org/10.1016/0022-1910(94)90061-2

Gontang, E. A., Aylward, F. O., Carlos, C., Glavina del Rio, T., Chovatia, M., Fern, A., … Kolter, R.

(2017). Major changes in microbial diversity and community composition across gut

sections of a juvenile cockroach. PLOS ONE, 12(5), e0177189.

http://doi.org/10.1371/journal.pone.0177189

Hagan, H. R. (1939). Diploptera dytiscoides, a viviparous roach with elongate pleuropodia.

Annals of the Entomological Society of America, 32, 264–266.

Hagan, H. R. (1941). The General Morphology of the Female Reproductive System of a

Viviparous Roach, Diploptera Dytiscoides (Serville). Psyche: A Journal of Entomology, 48(1),

1–8. http://doi.org/10.1155/1941/45020

Hamdi, C., Balloi, A., Essanaa, J., Crotti, E., Gonella, E., Raddadi, N., … Cherif, A. (2011). Gut

microbiome dysbiosis and honeybee health. Journal of Applied Entomology, 135(7), 524–

533. http://doi.org/10.1111/j.1439-0418.2010.01609.x

Hope, R. M. (2013). Rmisc: Ryan miscellaneous.

Huang, C. Y., Sabree, Z. L., & Moran, N. a. (2012). Genome sequence of Blattabacterium sp.

strain BGIGA, endosymbiont of the cockroach. Journal of Bacteriology,

194(16), 4450–4451. http://doi.org/10.1128/JB.00789-12

Ingram, M. J., Stay, B., & Cain, G. D. (1977). Composition of milk from the viviparous cockroach,

Diploptera punctata. Insect Biochemistry, 7(3), 257–267. http://doi.org/10.1016/0020-

104 1790(77)90023-3

Jašarević, E., Howerton, C. L., Howard, C. D., & Bale, T. L. (2015). Alterations in the vaginal

microbiome by maternal stress are associated with metabolic reprogramming of the

offspring gut and brain. Endocrinology, 156(9), 3265–3276.

http://doi.org/10.1210/en.2015-1177

Jašarević, E., Rodgers, A. B., & Bale, T. L. (2015). A novel role for maternal stress and microbial

transmission in early life programming and neurodevelopment. Neurobiology of Stress,

1(2015), 81–88. http://doi.org/10.1016/j.ynstr.2014.10.005

Jiménez-Chillarón, J. C., Nijland, M. J., Ascensão, A. A., Sardão, V. A., Magalhães, J., Hitchler, M.

J., … P.J., O. (2015). Back to the future: Transgenerational transmission of xenobiotic-

induced epigenetic remodeling. Epigenetics, 10(4), 259–273.

http://doi.org/10.1080/15592294.2015.1020267

Kakumanu, M. L., Maritz, J. M., Carlton, J. M., & Schal, C. (2018). Overlapping community

compositions of gut and fecal microbiomes in lab-reared and field-collected German

cockroaches. Applied and Environmental Microbiology, 84(17), 1–17.

http://doi.org/10.1128/AEM.01037-18

Kambhampati, S., Alleman, A., & Park, Y. (2013). Complete genome sequence of the

endosymbiont Blattabacterium from the cockroach Nauphoeta cinerea (Blattodea:

Blaberidae). Genomics, 102(5–6), 479–483. http://doi.org/10.1016/j.ygeno.2013.09.003

Korpela, K., Costea, P., Coelho, L. P., Kandels-Lewis, S., Willemsen, G., Boomsma, D. I., … Bork, P.

(2018). Selective maternal seeding and environment shape the human gut microbiome.

Genome Research, 28(4), 561–568. http://doi.org/10.1101/gr.233940.117

105 Kostic, A. D., Gevers, D., Siljander, H., Vatanen, T., Hyötyläinen, T., Hämäläinen, A.-M., … Xavier,

R. J. (2015). The dynamics of the human infant gut microbiome in development and in

progression toward type 1 diabetes. Cell Host & Microbe, 17(2), 260–273.

http://doi.org/10.1016/j.chom.2015.01.001

Kozich, J. J., Westcott, S. L., Baxter, N. T., Highlander, S. K., & Schloss, P. D. (2013). Development

of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence

data on the miseq illumina sequencing platform. Applied and Environmental Microbiology,

79(17), 5112–5120. http://doi.org/10.1128/AEM.01043-13

Lee, W.-J., & Brey, P. T. (2013). How microbiomes influence Metazoan development:Insights

from history and Drosophila modeling of gut-microbe interactions. Annual Review of Cell

and Developmental Biology, 29(1), 571–592. http://doi.org/10.1146/annurev-cellbio-

101512-122333

López-Sánchez, M. J., Neef, A., Patiño-Navarrete, R., Navarro, L., Jiménez, R., Latorre, A., &

Moya, A. (2008). Blattabacteria, the endosymbionts of cockroaches, have small genome

sizes and high genome copy numbers. Environmental Microbiology, 10(12), 3417–3422.

http://doi.org/10.1111/j.1462-2920.2008.01776.x

López-Sánchez, M. J., Neef, A., Peretó, J., Patiño-Navarrete, R., Pignatelli, M., Latorre, A., &

Moya, A. (2009). Evolutionary convergence and nitrogen metabolism in Blattabacterium

strain Bge, primary endosymbiont of the cockroach Blattella germanica. PLoS Genetics,

5(11). http://doi.org/10.1371/journal.pgen.1000721

Ma, J., Prince, A. L., Bader, D., Hu, M., Ganu, R., Baquero, K., … Aagaard, K. M. (2014). High-fat

maternal diet during pregnancy persistently alters the offspring microbiome in a primate

106 model. Nature Communications, 5(May), 3889. http://doi.org/10.1038/ncomms4889

Malmuthuge, N., Griebel, P. J., & Guan, L. L. (2015). The gut microbiome and its potential role in

the development and function of newborn calf gastrointestinal tract. Frontiers in

Veterinary Science, 2, 1–10. http://doi.org/10.3389/fvets.2015.00036

McFall-Ngai, M. J. (2014). The importance of microbes in animal development: lessons from the

squid-vibrio symbiosis. Annual Review of Microbiology, 68(1), 177–194.

http://doi.org/10.1146/annurev-micro-091313-103654

McMahan, E. A. (1969). Feeding relationships and radioisotope techniques. In K. Krishna & F. M.

Weesner (Eds.), Biology of Termites (1st ed., pp. 387–406). New York: Academic Press.

http://doi.org/10.1016/B978-0-12-395529-6.50016-7

Michalik, A., Szklarzewicz, T., Jankowska, W., & Wieczorek, K. (2014). Endosymbiotic

microorganisms of aphids (Hemiptera: Sternorrhyncha: Aphidoidea): ultrastructure,

distribution and transovarial transmission. European Journal of Entomology, 111(1), 91–

104. http://doi.org/10.14411/eje.2014.011

Michalkova, V., Benoit, J. B., Weiss, B. L., Attardo, G. M., & Aksoy, S. (2014). Vitamin B6

generated by obligate symbionts is critical for maintaining proline homeostasis and

fecundity in tsetse flies. Applied and Environmental Microbiology, 80(18), 5844–5853.

http://doi.org/10.1128/AEM.01150-14

Morse, S. F., Bush, S. E., Patterson, B. D., Dick, C. W., Gruwell, M. E., & Dittmar, K. (2013).

Evolution, multiple acquisition, and localization of endosymbionts in bat flies (Diptera:

Hippoboscoidea: Streblidae and Nycteribiidae). Applied and Environmental Microbiology,

79(9), 2952–2961. http://doi.org/10.1128/AEM.03814-12

107 Mueller, N. T., Bakacs, E., Combellick, J., Grigoryan, Z., & Dominguez-Bello, M. G. (2015). The

infant microbiome development: mom matters. Trends in Molecular Medicine, 21(2), 109–

117. http://doi.org/10.1016/j.molmed.2014.12.002

Neef, A., Latorre, A., Peretó, J., Silva, F. J., Pignatelli, M., & Moya, A. (2011). Genome

economization in the endosymbiont of the wood roach Cryptocercus punctulatus due to

drastic loss of amino acid synthesis capabilities. Genome Biology and Evolution, 3(1), 1437–

1448. http://doi.org/10.1093/gbe/evr118

Neuwirth, E. (2014). RColorBrewer: ColorBrewer Palettes.

Ogawa, K., & Miura, T. (2014). Aphid polyphenisms: Trans-generational developmental

regulation through viviparity. Frontiers in Physiology, 5, 1–11.

http://doi.org/10.3389/fphys.2014.00001

Ohio Supercomputer Center. (1987). Ohio Supercomputer Center.

Pais, R., Lohs, C., Wu, Y., Wang, J., & Aksoy, S. (2008). The obligate mutualist Wigglesworthia

glossinidia influences reproduction, digestion, and immunity processes of its host, the

tsetse fly. Applied and Environmental Microbiology, 74(19), 5965–5974.

http://doi.org/10.1128/AEM.00741-08

Patiño-Navarrete, R., Moya, A., Latorre, A., & Peretó, J. (2013). Comparative genomics of

Blattabacterium cuenoti: The frozen legacy of an ancient endosymbiont genome. Genome

Biology and Evolution, 5(2), 351–361. http://doi.org/10.1093/gbe/evt011

Pérez-Cobas, A. E., Maiques, E., Angelova, A., Carrasco, P., Moya, A., & Latorre, A. (2015). Diet

shapes the gut microbiota of the omnivorous cockroach Blatella germanica. FEMS

Microbiology Ecology, 91(4), 1–14. http://doi.org/10.1093/femsec/fiv022

108 Perez-Muñoz, M. E., Arrieta, M. C., Ramer-Tait, A. E., & Walter, J. (2017). A critical assessment

of the “sterile womb” and “in utero colonization” hypotheses: Implications for research on

the pioneer infant microbiome. Microbiome. 5(1), 48. http://doi.org/10.1186/s40168-017-

0268-4

Pietri, J. E., Tiffany, C., & Liang, D. (2018). Disruption of the microbiota affects physiological and

evolutionary aspects of insecticide resistance in the German cockroach, an important

urban pest. PLOS ONE, 13(12), e0207985. http://doi.org/10.1371/journal.pone.0207985

Poulin, R., & Thomas, F. (2008). Epigenetic effects of infection on the phenotype of host

offspring: parasites reaching across host generations. Oikos, 117(3), 331–335.

http://doi.org/10.1111/j.2007.0030-1299.16435.x

Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., … Glöckner, F. O. (2013). The

SILVA ribosomal RNA gene database project: improved data processing and web-based

tools. Nucleic Acids Research, 41(Database issue), D590-6.

http://doi.org/10.1093/nar/gks1219

R Core Team. (2017). R: A language and environment for statistical computing. Vienna, Austria:

R Foundation for Statistical Computing.

Ram, K., & Wickham, H. (2018). wesanderson: A Wes Anderson palette generator.

Rio, R. V. M., Symula, R. E., Wang, J., Lohs, C., Wu, Y.-N., Snyder, A. K., … Aksoy, S. (2012).

Insight into the transmission biology and species-specific functional capabilities of tsetse

(Diptera: Glossinidae) obligate symbiont Wigglesworthia. MBio, 3(1), 1–13.

http://doi.org/10.1128/mBio.00240-11

Roth, L. M., & Willis, E. R. (1955). Intra-uterine nutrition of the “beetle-roach” Diploptera

109 Dytiscoides (Serv.) during embryogenesis, with notes on its biology in the laboratory

(Blattaria: Diplopteridae). Psyche: A Journal of Entomology. 62, 55-68

http://doi.org/10.1155/1955/12542

Roth, L. M., & Willis, E. R. (1957). An analysis of oviparity and viviparity in the Blattaria.

American Entomological Society, 83(4), 221–238.

RStudio Team. (2015). RStudio: Integrated development for R. Boston, MA: RStudio, Inc.

Sabree, Z. L., Huang, C. Y., Arakawa, G., Tokuda, G., Lo, N., Watanabe, H., & Moran, N. A. (2012).

Genome shrinkage and loss of nutrient-providing potential in the obligate symbiont of the

primitive termite Mastotermes darwiniensis. Applied and Environmental Microbiology,

78(1), 204–210. http://doi.org/10.1128/AEM.06540-11

Sabree, Z. L., Kambhampati, S., & Moran, N. a. (2009). Nitrogen recycling and nutritional

provisioning by Blattabacterium, the cockroach endosymbiont. Proceedings of the National

Academy of Sciences of the United States of America, 106(46), 19521–19526.

http://doi.org/10.1073/pnas.0907504106

Sabree, Z. L., & Moran, N. A. (2014). Host-specific assemblages typify gut microbial communities

of related insect species. SpringerPlus, 3(1), 138. http://doi.org/10.1186/2193-1801-3-138

Sacchi, L., Corona, S., Grigolo, A., Laudani, U., Selmi, M. G., & Bigliardi, E. (1996). The fate of the

endocytobionts of Blattella germanica (Blattaria: Blattellidae) and Periplaneta americana

(Blattaria: ) during embryo development. Italian Journal of Zoology, 63(1), 1–11.

http://doi.org/10.1080/11250009609356100

Salem, H., Florez, L., Gerardo, N., & Kaltenpoth, M. (2015). An out-of-body experience: the

extracellular dimension for the transmission of mutualistic bacteria in insects. Proceedings

110 of the Royal Society B: Biological Sciences, 282(1804), 20142957–20142957.

http://doi.org/10.1098/rspb.2014.2957

Schauer, C., Thompson, C., & Brune, A. (2014). Pyrotag Sequencing of the Gut Microbiota of the

Cockroach Shelfordella lateralis Reveals a Highly Dynamic Core but Only Limited Effects of

Diet on Community Structure. PLoS ONE, 9(1), e85861.

http://doi.org/10.1371/journal.pone.0085861

Schloss, P. D., Westcott, S. L., Ryabin, T., Hall, J. R., Hartmann, M., Hollister, E. B., … Weber, C. F.

(2009). Introducing mothur: Open-source, platform-independent, community-supported

software for describing and comparing microbial communities. Applied and Environmental

Microbiology, 75(23), 7537–7541. http://doi.org/10.1128/AEM.01541-09

Schwab, D. B., Riggs, H. E., Newton, I. L. G., & Moczek, A. P. (2016). Developmental and

ecological benefits of the maternally transmitted microbiota in a dung beetle. The

American Naturalist, 188(6), 679–692. http://doi.org/10.1086/688926

Shukla, S. P., Vogel, H., Heckel, D. G., Vilcinskas, A., & Kaltenpoth, M. (2018). Burying beetles

regulate the microbiome of carcasses and use it to transmit a core microbiota to their

offspring. Molecular Ecology, 27(8), 1980–1991. http://doi.org/10.1111/mec.14269

Snyder, A. K., Mclain, C., & Rio, R. V. M. (2012). The tsetse fly obligate mutualist Wigglesworthia

morsitans alters gene expression and population density via exogenous nutrient

provisioning. Applied and Environmental Microbiology, 78(21), 7792–7797.

http://doi.org/10.1128/AEM.02052-12

Snyder, A. K., & Rio, R. V. M. (2015). “Wigglesworthia morsitans” folate (vitamin B9)

biosynthesis contributes to tsetse host fitness. Applied and Environmental Microbiology,

111 81(16), 5375–5386. http://doi.org/10.1128/AEM.00553-15

Stay, B., & Coop, A. (1973). Developmental stages and chemical composition in embryos of the

cockroach, Diploptera punctata, with observations on the effect of diet. Journal of Insect

Physiology, 19(1), 147–71.

Stay, B., & Coop, A. C. (1974). “Milk” secretion for embryogenesis in a viviparous cockroach.

Tissue and Cell, 6(4), 669–693. http://doi.org/10.1016/0040-8166(74)90009-3

Stein, A. D., & Lumey, L. H. (2000). The relationship between maternal and offspring birth

weights after maternal prenatal famine exposure: the Dutch famine birth cohort study.

Human Biology; an International Record of Research, 72(4), 641–654.

The NIH HMP Working Group. (2009). The NIH Human Microbiome Project. Genome Research,

19(12), 2317–2323. http://doi.org/10.1101/gr.096651.109

Thompson, J. R., Rivera, H. E., Closek, C. J., & Medina, M. (2015). Microbes in the coral

holobiont: partners through evolution, development, and ecological interactions. Frontiers

in Cellular and Infection Microbiology, 4, 1–20. http://doi.org/10.3389/fcimb.2014.00176

Tinker, K. A., & Ottesen, E. A. (2016). The core gut microbiome of the American cockroach,

Periplaneta americana, is stable and resilient to dietary shifts. Applied and Environmental

Microbiology, 82(22), 6603–6610. http://doi.org/10.1128/AEM.01837-16

Tokuda, G., Elbourne, L. D. H., Kinjo, Y., Saitoh, S., Sabree, Z. L., Hojo, M., … Lo, N. (2013).

Maintenance of essential amino acid synthesis pathways in the Blattabacterium cuenoti

symbiont of a wood-feeding cockroach. Biology Letters, 9, 20121153.

http://doi.org/10.1098/rsbl.2012.1153

Torrazza, R. M., & Neu, J. (2011). The developing intestinal microbiome and its relationship to

112 health and disease in the neonate. Journal of Perinatology, 31(S1), S29–S34.

http://doi.org/10.1038/jp.2010.172

Wade, M. J. (2014). Paradox of mother’s curse and the maternally provisioned offspring

microbiome. Cold Spring Harbor Perspectives in Biology, 6, a017541.

http://doi.org/10.1177/00220345950740020901

Walker, R. W., Clemente, J. C., Peter, I., & Loos, R. J. F. (2017). The prenatal gut microbiome: are

we colonized with bacteria in utero? Pediatric Obesity, 12(August), 3–17.

http://doi.org/10.1111/ijpo.12217

Wang, J., Weiss, B. L., & Aksoy, S. (2013). Tsetse fly microbiota: form and function. Frontiers in

Cellular and Infection Microbiology, 3, 1–6. http://doi.org/10.3389/fcimb.2013.00069

Wang, Y., & Rozen, D. E. (2017). Gut microbiota colonization and transmission Nicrophorus

vespilloides throughout development. Applied and Environmental Microbiology, 83(9),

e03250-16. http://doi.org/10.1128/AEM.03250-16

Weber, N., Liou, D., Dommer, J., Macmenamin, P., Quiñones, M., Misner, I., … Hurt, D. E. (2018).

Nephele: A cloud platform for simplified, standardized and reproducible microbiome data

analysis. Bioinformatics, 34(8), 1411–1413. http://doi.org/10.1093/bioinformatics/btx617

Weiss, B. L., Wang, J., & Aksoy, S. (2011). Tsetse immune system maturation requires the

presence of obligate symbionts in larvae. PLoS Biology, 9(5), e1000619s.

http://doi.org/10.1371/journal.pbio.1000619

Wickham, H. (2007). Reshaping data with the reshape Package. Journal of Statistical Software,

21(12), 1–20. http://doi.org/10.18637/jss.v021.i12

Wickham, H. (2016). ggplot2: Elegant graphics for data analysis. Springer-Verlag New York.

113 Wickham, H., Francois, R., Henry, L., & Müller, K. (2017). dplyr: A grammar of data

manipulation.

Williford, A., Stay, B., & Bhattacharya, D. (2004). Evolution of a novel function: Nutritive milk in

the viviparous cockroach, Diploptera punctata. Evolution and Development, 6, 67–77.

http://doi.org/10.1111/j.1525-142X.2004.04012.x

Yang, I., Corwin, E. J., Brennan, P. A., Jordan, S., Murphy, J. R., & Dunlop, A. (2016). The Infant

Microbiome: Implications for Infant Health and Neurocognitive Development. Nursing

Research, 65(1), 76–88. http://doi.org/10.1097/NNR.0000000000000133

Youngsteadt, E., Fan, Y., Stay, B., & Schal, C. (2005). Cuticular hydrocarbon synthesis and its

maternal provisioning to embryos in the viviparous cockroach Diploptera punctata. Journal

of Insect Physiology, 51, 803–809. http://doi.org/10.1016/j.jinsphys.2005.03.008

114 Tables

Table 1. HOMOVA comparisons of thetayc distances calculated from mothers and embryos of both colonies

Groups Compared Bvalue Pvalue Sum of Squares for each group

UC Embryo-UC Mother- 232.761 <0.001* 1.34E-07 0.317 2.20E-07 0.038 OSU Embryo-OSU Mother

UC Embryo-UC Mother 100.043 <0.001* 1.34E-07 0.317

UC Embryo-OSU Embryo 0.517 0.654 1.34E-07 2.20E-07

UC Embryo-OSU Mother 85.427 <0.001* 1.34E-07 0.038

UC Mother-OSU Embryo 132.335 <0.001* 3.17E-01 2.20E-07

UC Mother- OSU Mother 9.247 0.014 3.17E-01 0.038

OSU Embryo- OSU Mother 112.345 <0.001* 2.20E-07 0.038

115 Table 2. AMOVA comparisons of thetayc distances calculated from mothers and embryos of both colonies

Sum of Squares Degrees of Freedom Mean Squares

Groups Compared Among Within Total Among Within Total Among Within F-statistic p-value

UC Embryo-UC Mother-OSU 4.187 2.955 7.143 3 38 41 1.396 0.078 17.949 <0.001

Embryo-OSU Mother

UC Embryo-UC Mother 1.935 2.535 4.470 1 16 17 1.935 0.158 12.214 <0.001

UC Embryo-OSU Embryo 2.58E-06 3.50E-06 6.08E-06 1 19 20 0.000 0.000 14.015 0.011

UC Embryo-OSU Mother 1.634 0.420 2.054 1 19 20 1.634 0.022 73.862 <0.001

UC Mother-OSU Embryo 2.192 2.535 4.726 1 19 20 2.192 0.133 16.427 <0.001

UC Mother- OSU Mother 0.761 2.955 3.716 1 19 20 0.761 0.156 4.892 0.004

OSU Embryo- OSU Mother 1.864 0.420 2.284 1 22 23 1.864 0.019 97.593 <0.001

116 Table 3. HOMOVA comparisons of thetayc distances calculated from all developmental stages

Groups Compared Bvalue Pvalue Sum of Squares for each group

Embryo-First_instar-Fourth_instar-

Mother-Neonate-Second_instar-

Third_instar 252.199 <0.001* 4.41E-06 0.006 0.009 0.187 1.84E-06 0.004 0.013

Embryo-First_instar 100.589 <0.001* 4.41E-06 0.006

Embryo-Fourth_instar 105.298 <0.001* 4.41E-06 0.009

Embryo-Mother 180.880 <0.001* 4.41E-06 0.187

Embryo-Neonate 0.962 0.739 4.41E-06 1.84E-06

Embryo-Second_instar 98.711 <0.001* 4.41E-06 0.004

Embryo-Third_instar 126.723 <0.001* 4.41E-06 0.013

First_instar-Fourth_instar 0.213 0.547 6.09E-03 0.009

First_instar-Mother 13.197 0.001* 6.09E-03 0.1871

First_instar-Neonate 24.708 0.002* 6.09E-03 1.84E-06

First_instar-Second_instar 0.161 0.602 6.09E-03 0.004

First_instar-Third_instar 1.064 0.21 6.09E-03 0.013

117 Fourth_instar-Mother 9.153 0.014 9.19E-03 0.187

Fourth_instar-Neonate 25.859 0.004 9.19E-03 1.84E-06

Fourth_instar-Second_instar 0.779 0.332 9.19E-03 0.004

Fourth_instar-Third_instar 0.214 0.625 9.19E-03 0.013

Mother-Neonate 38.437 <0.001* 1.87E-01 1.84E-06

Mother-Second_instar 19.852 0.001* 1.87E-01 0.004

Mother-Third_instar 21.371 0.005 1.87E-01 0.013

Neonate-Second_instar 23.965 <0.001* 1.84E-06 0.004

Neonate-Third_instar 28.514 <0.001* 1.84E-06 0.013

Second_instar-Third_instar 2.552 0.12 4.44E-03 0.013

118 Table 4. AMOVA comparisons of thetayc distances calculated from from all developmental stages

Sum of Squares Degrees of Freedom Mean Squares

Groups Compared Among Within Total Among Within Total Among Within F-statistic p-value

Embryo-First_instar-Fourth_instar-Mother- 5.790 4.058 9.849 6 78 84 0.965 0.052 18.548 <0.001 Neonate-Second_instar-Third_instar

Embryo-First_instar 1.040 0.037 1.077 1 26 27 1.040 0.001 738.15 <0.001

Embryo-Fourth_instar 1.585 0.046 1.631 1 25 26 1.585 0.002 860.55 <0.001

Embryo-Mother 3.417 3.742 7.159 1 40 41 3.417 0.094 36.520 <0.001

Embryo-Neonate 1.15E-04 9.5E-05 2.10E-04 1 24 25 0.000 4.0E-06 28.878 <0.001

Embryo-Second_instar 2.023 0.036 2.059 1 28 29 2.023 0.001 1591.1 <0.001

Embryo-Third_instar 2.362 0.198 2.560 1 35 36 2.362 0.006 417.61 <0.001

First_instar-Fourth_instar 0.107 0.082 0.190 1 11 12 0.107 0.007 14.319 0.001

First_instar-Mother 0.342 3.779 4.121 1 26 27 0.342 0.145 2.355 0.067

First_instar-Neonate 0.500 3.7E-02 0.536 1 10 11 0.500 0.004 136.74 0.001

First_instar-Second_instar 0.092 7.2E-02 0.165 1 14 15 0.092 0.005 17.963 0.001

119 First_instar-Third_instar 0.059 0.234 0.294 1 21 22 0.059 0.011 5.319 0.025

Fourth_instar-Mother 0.245 3.788 4.0 1 25 26 0.245 0.152 1.617 0.124

Fourth_instar-Neonate 0.848 0.046 0.894 1 9 10 0.848 0.005 166.12 0.001

Fourth_instar-Second_instar 0.006 0.081 0.087 1 13 14 0.006 0.006 0.918 0.374

Fourth_instar-Third_instar 0.029 2.4E-01 0.272 1 20 21 0.029 0.012 2.348 0.134

Mother-Neonate 1.234 3.742 4.976 1 24 25 1.234 0.156 7.911 0.001

Mother-Second_instar 0.326 3.778 4.104 1 28 29 0.326 0.135 2.415 0.061

Mother-Third_instar 0.493 3.940 4.433 1 35 36 0.493 0.113 4.380 0.005

Neonate-Second_instar 0.939 0.036 0.974 1 12 13 0.939 0.003 317.10 <0.001

Neonate-Third_instar 0.888 0.198 1.086 1 19 20 0.888 0.010 85.270 0.001

Second_instar-Third_instar 0.012 0.233 0.246 1 23 24 0.012 0.010 1.224 0.292

120 Figures and legends

Legends

Figure 1. Scatter plot of the relative abundances of the ten most abundant classes and all other classes combined. Point color indicates class designation while point shape corresponds to the treatment group. The linear model describing the relationship between the two methods is represented by the black trendline (y = 0.0092 + 1.1x). r2 = 0.943.

Figure 2. Relative abundances of the 19 most abundant bacterial families in D. punctata mothers and embryos. The remaining families are cumulatively represented as “other”. The y- axis represents the percent of total OTUs present in each sample for each family. Each bar represents an individual mother or brood of embryos.

Figure 3. Measure of diversity and evenness calculated using mothur for mothers and embryos of both UC and OSU colonies. (A) Inverse Simpson measure of alpha diversity (B) Shannon’s diversity index (C) Shannon’s evenness index. Error bars signify standard error. Significance determined by Kruskall-Wallace and Dunn’s test, alpha=0.025.

Figure 4. Community comparisons between D. punctata mothers and embryos of both colonies

(A) Number of OTUs recovered for mothers of the UC and OSU colonies. A 2314 OTU core component of the maternal microbiome was identified using mothur (B) Principle coordinate analysis [PCOA] of mothers and embryos from both colonies (C) Non-metric multi-dimensional

121 scaling [NMDS] of mothers and embryos from both colonies. In B and C embryos cluster so closely that the samples are indistinguishable.

Figure 5. Relative abundances of the 19 most abundant bacterial families in D. punctata embryos, nymphs, and adult females. The y-axis represents the percent of total OTUs present in each sample for each family. Each bar represents an individual sequencing replicate; nymphs and mothers were individual animals while embryos were whole broods.

Figure 6. Measures of diversity and evenness generated by mothur for embryos, all nymphs, and adult females of D. punctata (A) Inverse Simpson (B) Shannon’s diversity index. Error bars signify standard error. Significance determined by Kruskall-Wallace and Dunn’s test, alpha=0.025.

122 Figures

1.00

Bacterial Class 0.75 Bacteroidales Clostridiales Desulfobacterales Desulfovibrionales Enterobacteriales Flavobacteriales Fusobacteriales

TUs in QIIME Lactobacillales O 0.50 other Rs−K70_termite_group tion of r Synergistales Propo Development Stage First Instar Fourth Instar Neonate OSU Embryo 0.25 OSU Mother Second Instar Third Instar Total UC Embryo UC Mother

0.00

0.00 0.25 0.50 0.75 1.00 Proportion of OTUs in mothur

123 Relativeabundance (%OTUs) 100% 25% 50% 75% 0%

RelativeaRelativebundance abundance (%O (%)TUs) 100% 25% 50% 75% 0% Total

Embryos UC UC Mothers UC OSU Embryos OSU Mothers Synergistaceae Ruminococcaceae Rs−K70_te Ri Proteobacte P Other_ Lactobacillales_unclassifed Lachnospi Gammaproteobacte Fusobacte Fusobacte Fi Enterobacte Desul Clost Blattabacte Bacteroidetes_unclassifed Bacteroidales_unclassifed Bacte o r k r micutes_unclassifed enellaceae p h r f r yromonadaceae idiales_unclassifed obacte ia_unclassifed f amilies r r r aceae r r iales_unclassifed iaceae mite_group_unclassifed iaceae r r ia_unclassifed iaceae r aceae Synergistaceae Ruminococcaceae Rs−K70_te Ri Proteobacte P Other_ Lactobacillales_unclassifed Lachnospi Gammaproteobacte Fusobacte Fusobacte Fi Enterobacte Desul Clost Blattabacte Bacteroidetes_unclassifed Bacteroidales_unclassifed Bacte r ia_unclassifed o r k r micutes_unclassifed enellaceae p h r f r yromonadaceae idiales_unclassifed obacte ia_unclassifed f amilies r r r aceae r r iales_unclassifed iaceae mite_group_unclassifed iaceae r r ia_unclassifed iaceae r aceae r ia_unclassifed

124 A

Inverse SimpsonInverse measure Simpson of alpha diversity 10 20 30 0 OSU_Embryo a OSU_Mother b Group UC_Embryo a UC_Mother b group UC_Mother UC_Embryo OSU_Mother OSU_Embryo B

ShannonShannon diversity Index index (H) 0 1 2 3 4 5 OSU_Embryo a OSU_Mother b Group UC_Embryo a UC_Mother b group UC_Mother UC_Embryo OSU_Mother OSU_Embryo C

Shannon Evenness 0.0 0.2 0.4 0.6 0.8 OSU_Embryo a OSU_Mother b Group UC_Embryo a UC_Mother b

group UC_Mother UC_Embryo OSU_Mother OSU_Embryo

125

A

B C

OSU Mothers

0.2 0.50 UC Mothers 0.0 group group 0.25 OSU_Embryo Embryos OSU_Embryo OSU_Mother −0.2 OSU_Mother UC_Embryo UC_Embryo Embryos UC_Mother NMDS 2 UC_Mother UC Mothers −0.4 0.00 PCOA 2 (18.73%) PCOA

−0.6

−0.25 OSU Mothers −0.25 0.00 0.25 0.50 −0.4 −0.2 0.0 0.2

PCOA 1 (38.17%) NMDS 1

126 100% 25% 50% 75% 0%

Relative abundance (%) 100% 25% 50% 75% Total 0%

Embryo Neonate 1 st instar 2 nd instar 3 rd instar 4 th instar Mother Synergistaceae Ruminococcaceae Rs−K70_te Ri Proteobacte P O Lactobacillales_unclassifed Lactobacillaceae Lachnospi Gammaproteobacte Fi E Enterobacte Desul Clost Blattabacte Bacteroidetes_unclassifed Bacteroidales_unclassifed Bacte o r THER r k ysipelot r micutes_unclassifed enellaceae p h r f r yromonadaceae idiales_unclassifed obacte ia_unclassifed r r ichaceae aceae r r mite_group_unclassifed iaceae r r ia_unclassifed iaceae r aceae r ia_unclassifed Synergistaceae Ruminococcaceae Rs−K70_te Ri Proteobacte P O Lactobacillales_unclassifed Lactobacillaceae Lachnospi Gammaproteobacte Fi E Enterobacte Desul Clost Blattabacte Bacteroidetes_unclassifed Bacteroidales_unclassifed Bacte o r THER r k ysipelot r micutes_unclassifed enellaceae p h r f r yromonadaceae idiales_unclassifed obacte ia_unclassifed r r ichaceae aceae r r mite_group_unclassifed iaceae r r ia_unclassifed iaceae r aceae r ia_unclassifed

127

A

30 C

20 BC BC BC

B Inverse Simpson Inverse 10 AB A

0 Embryo Neonate First_instar Second_instar Third_instar Fourth_instar Mother

Group

C C BC B BC 5 B

4

group Embryo 3 First_instar Fourth_instar Mother Neonate Second_instar

Shannon Index Third_instar

2 AB Shannon Shannon diversity index (H) A

1

Embryo Neonate First_instar Second_instar Third_instar Fourth_instar Mother

Group

128

Chapter 4

Genome analysis of obligate endosymbiont Blattabacterium sp. reveals potential metabolic contributions to intrauterine developing embryos of

Diploptera punctata

Emily C Jennings 1*

Matthew W Korthauer 1, Joshua B Benoit 1

1 Department of Biological Sciences University of Cincinnati. Cincinnati, OH 45221

* The work presented here was principally composed by Emily C Jennings with assistance from the listed co-authors.

129 Abstract

Bacterial endosymbionts of insects have often co-evolved with their host species to play important roles in the host’s fitness. Such is the case with tsetse flies and aphids and their mutualistic symbionts, in which these bacteria provide their hosts with vital nutrients that are lacking in their diet. Cockroaches universally have a common transovarially inherited primary endosymbiont, Blattabacterium, that resides in bacteriocytes present in the fat body organ where it functions in recycling uric acid into ammonia for host use. In addition, it is known that these symbionts are capable of synthesizing many essential amino acids, although there is variation between strains. Previously, we reported the embryonic microbiome of the live bearing cockroach, Diploptera punctata, consisting only of Blattabacterium. The intrauterine developing embryos of this species are provided with milk-like secretions from the mother during their gestation in the brood-sac, however this milk is deficient in two essential amino acids—methionine and tryptophan. To determine if Blattabacterium as the only member of the embryonic microbiome is capable of supplementing the diet of developing D. punctata we present the genome of Blattabacterium isolated from the fat body of D. punctata. Our analyses indicate that in addition to the previously known function in nitrogen cycling, this strain of

Blattabacterium is likely capable of providing all ten essential amino acids to the host, supporting the intrauterine development of D. puncata embryos.

130 Introduction

Despite the diversity in habitat and diet, all species of cockroach except one harbor an obligate

Flavobacterial endosymbiotic bacteria (Bandi et al., 1994, 1995). These bacteria, belonging to the Blattabacterium, live in specialized bacteriocytes in the fat body and are maternally inherited through incorporation into developing oocytes through a phagocytosis like process

(Bandi et al., 1994, 1995). All other bacterial components of the microbiome appear to be inherited after hatching from the environment and oothecal surface (Chapter 3). The genomes of Blattabacteria endosymbionts from several species of cockroach have been sequenced, revealing their role in recycling of nitrogen from uric acid stores to produce various amino acids with the exception of methionine in all strains but that of Blattella germanica (Huang, Sabree, &

Moran, 2012; Kambhampati, Alleman, & Park, 2013; López-Sánchez et al., 2008, 2009; Patino-

Navarrete et al., 2014; Sabree et al., 2012; Sabree, Kambhampati, & Moran, 2009; Tokuda et al.,

2013).

Cockroaches also display extensive diversity in reproductive strategy with species that reproduce through oviparity (laying eggs), facultative viviparity or ovoviviparity (may have live birth or lay eggs but do not provide embryos with non-yolk nutrients), and some species that reproduce by true matrotrophic viviparity (harboring developing offspring inside the reproductive tract and providing non-yolk nutrition to developing offspring) (Hagan, 1948; Roth

& Willis, 1957). Only one species of cockroach, Diploptera punctata, has been documented to use matrotrophic viviparity as a reproductive strategy (Roth & Willis, 1955; Stay & Coop, 1973,

1974; Stay & Roth, 1956).

131 D. punctata embryos develop inside the brood sac, a unique organ which functions as both a uterus and pseudo-placenta; embryos are provided with nutrients by a secretion of milk- like components (Marchal, Hult, Huang, Stay, & Tobe, 2013; Roth & Willis, 1955; Stay & Coop,

1974; Williford, Stay, & Bhattacharya, 2004). This secretion appears in embryo gut contents at

20% of the 60-70-day pregnancy, when the dorsal edge of the body wall is closed. Diploptera milk is a combination of proteins and free amino acids, carbohydrates, and lipids in a water base (Ingram, Stay, & Cain, 1977; Roth & Stay, 1961; Stay & Coop, 1973; Williford et al., 2004).

The proteins present include a family of milk proteins derived from 25 unique mRNAs coding for

22 different proteins (Williford et al., 2004). During its gestation period, between nine and thirteen embryos grow from 1.5 mm at the time of ovulation to over 6mm at birth, increasing in weight by more than 70-fold. Water content of each egg increases by 85-fold, and solid dry weight increases over 49-fold (Roth, 1967; Roth & Willis, 1955; Stay & Coop, 1973). Of importance is the observation the D. punctata milk is deficient in two essential amino acids, tryptophan and methionine (Ingram et al., 1977; Williford et al., 2004). It has been hypothesized that this dietary deficiency is remediated with the assistance of endosymbiont metabolism. Previously we demonstrated that, as in other cockroach species, Blattabacteria are the only bacteria maternally inherited and present in embryos during intrauterine gestation

(Chapter 3). In this study, we investigate the genome of Blattabacterium derived from D. punctata (= Blattabacterium str. DPU) to determine the role of that this bacterial endosymbiont potentially has during embryonic development of D. punctata.

132 Methods

Animals and Sample Collection

Colonies reared at the University of Cincinnati were housed in a climate-controlled facility.

Ambient temperature was held between 24-28°C and relative humidity (RH) was held between

70-80%. A 12:12 hour light-dark photoperiod was maintained for the duration of the experiment. Animals were provided water and fed dog food ad libitum.

Visibly pregnant females were selected from the colony for use in mother-embryo comparisons. Females were surface sterilized by rinsing in 70% ethanol for 10 seconds and then sterile PBS prior to dissection. Embryo broods were then dissected from the brood sac in sterile

PBS by making two incisions at the opening of the brood sac, one on each side, and removed using ethanol sterilized forceps and discarded. The ventral portion of the abdominal cuticle was then removed to expose abdominal organs, specifically the fat body. Fat body tissue was then dissected and placed into a 1.5 mL centrifuge tube. Each replicate was comprised of fat body tissue from five female D. punctata.

Genomic DNA preparation

Bacterial DNA was collected using a modified version of the protocols described by Sabree,

Kambhampati and Moran (2009) and López-Sánchez et al. (2009) with the use of a QIAGEN

DNeasy Blood & Tissue kit. Briefly, samples were homogenized using sterile plastic pestles with a battery-operated pestle in 200 µl of sterile 1X PBS. This extract was then passed through a 20

µm glass syringe filter (Millipore) and centrifuged for 10 minutes at 8000 x g and 4°C to reduce the amount of host tissue in the sample. Genomic DNA was extracted from the pellet

133 resuspended in the lysis buffer provided in the extraction kit after which the manufacturer protocol was followed. Extracted DNA was examined for quality and quantified via A260/A280 and A260/A230 using a Nanodrop 2000 (Thermo Scientific).

Sequencing and bioinformatic analysis

Illumina Nextera library preparation and paired end Illumina HiSeq analysis of genomic DNA was conducted at the Cincinnati Children’s Hospital Medical Center (CCHMC) Genetic Variation and Gene Discovery Core Facility. Sequence reads were trimmed and filtered to remove adapter sequences and low-quality reads with ambiguous bases using Trimmomatic (Bolger,

Lohse, & Usadel, 2014). Paired reads produced by Trimmomatic were used in the assembly. In anticipation of host sequences being present in the samples, reads were assembled using metaSPAdes (Nurk et al., 2013) as implemented on KBASE (Arkin et al., 2018). Genome quality was assessed using QUAST (Gurevich, Saveliev, Vyahhi, & Tesler, 2013) also implemented on

KBASE. Contigs from the metSPAdes assembly were searched for sequence similarity to the

Blattabacterium strain bbge genome (López-Sánchez et al., 2009) using BLASTn as implemented on CLC Genomics Workbench (QIAGEN). Coding DNA sequences (CDS) were identified and annotated using Prokka (Seemann, 2014) as implemented on Galaxy (Afgan et al., 2018) as well as BLASTx in CLC Genomic Workbench. Predicted metabolic pathways were determined by assigning KEGG orthology terms to the putative genes produced by Prokka using - mapper and visualized using iPATH3 (Darzi, Letunic, Bork, & Yamada, 2018). Metabolic pathway completeness was determined using the online KEGG module mapping tool (Kanehisa, Sato,

Furumichi, Morishima, & Tanabe, 2019). BRIG software (Alikhan, Petty, Ben Zakour, & Beatson,

134 2011) was utilized to visualize the genome structure and features such as predicted CDS, GC content and GC skew.

The Bbge genome, plasmid and coding sequences as well as the coding DNA sequences from the seven other strains of Blattabacterium were downloaded from the NCBI GenBank database.

Results

Illumina Hiseq paired end sequencing produced 6,778,349 paired end reads of 75bp length and

4,444,306 of 125bp length, less than 1% of which were lost during quality control using

Trimmomatic. MetaSPAdes generated 187 total contigs of varying lengths with an N50 of

625,590 base pairs, which is also the length of the largest contig generated. BLASTn comparison of these contigs to the Blattabacterium strain Bbge genome identified this largest contig as a candidate genome sequence (E<0.0001).

BLASTn comparison of all metaSPAdes generated contigs to the Blattabacterium strain

Bbge plasmid revealed that the 2,852 bp plasmid sequence had been assembled as part of the genome; consequently, we removed this sequence from the contig for further analyses. After removal of the plasmid sequence a 623,008 bp contig remained with 28.03% GC content.

Prokka identified 618 ORFs, 580 of which were predicted to be protein coding as well as 34 transfer RNAs, 3 ribosomal RNAs and one transfer messenger RNA (Table 1, Figure 1). Prokka annotation of the plasmid sequence identified 4 ORFs. Similar to previously sequenced strains, almost all genes required for DNA replication, RNA transcription, and mRNA translational machinery were identified in the assembly. The plasmid genes were annotated as deoxyuridine

5’ -triphosphate nucleotidohydrolase, ribonucleoside-diphosphate reductase subunit beta, and

135 two hypothetical proteins. Assignment of orthology terms (COG) revealed that, like other sequenced Blattabacterium genomes, the majority of coding genes serve functions in translation and ribosome formation (Figure 2). Aside from terms associated with unknown functions, the next most prominent genome function is amino acid metabolism and transport followed by energy production and conversion (Figure 2). Our analyses assigned no genes functions associated with cell motility or signal transduction to any of the genomes we searched.

In addition to presence of enzymes required for central carbohydrate metabolism and nitrogen salvage, metabolic pathway prediction revealed that BDIPPU possesses complete biosynthetic pathways for nearly all ten essential amino acids. The traditional bacterial biosynthetic pathway for methionine is, however, incomplete. Genes for all enzymatic reactions required to produce methionine are present with the exception of the gene metA which facilitates the conversion of homoserine and succinyl coenzyme A to O-succinylhomoserine and the alternative metX which produces O-acetylhomoserine. The ability to synthesize selenomethionine, however, is retained. Blattabacterium str. DPU also retains the ability to synthesize the non-essential amino acids alanine, arginine, cysteine, glutamate, and glycine.

Discussion

We utilized genome sequencing and assembly to characterize the potential metabolic contributions of Blattabacterium str. DPU to intrauterine developing embryos of its host. We assembled and characterized the 623,008 bp genome and 2,852 bp plasmid. The small size of the genome is consistent with the reduced genomes found in previously sequenced strains. We

136 identified 580 and 4 coding sequences in the genome and plasmid respectively, including genes coding for basic cellular processes such as DNA replication and transcription, ribosomes and other translational machinery. Also consistent with previously sequenced strains was our identification of 3 rRNAs, 34 tRNAs, 1 tmRNA. Compared to previously characterized strains of

Blattabacterium, the DPU strain is similar in composition of functional terms represented in the genome. Genes functioning in cell motility, signal transduction mechanisms, and secondary metabolite processes are virtually unrepresented in the genomes. The most represented functional category is that of translation, ribosomal structures and biogenesis. Of interest is the also high representation of amino acid transport and metabolism in coding genes.

This genomic profile is not unique to Blattabacterium sp., in fact it is a trend that has been seen across the genomes of multiple obligate endosymbiont bacteria (Akman et al., 2002;

Gil et al., 2003; Jiang et al., 2013; Nakabachi et al., 2006; Santos-Garcia et al., 2014; Shigenobu,

Watanabe, Hattori, Sakaki, & Ishikawa, 2000). Obligate endosymbionts often have genome sizes less than 1 Mbp with low GC content, and few coding sequences (McCutcheon & Moran, 2010,

2012; Moran & Wernegreen, 2000; Nikoh, Hosokawa, Oshima, Hattori, & Fukatsu, 2011;

Wernegreen, 2002; Wolf & Koonin, 2013). Relying on the stable environment of host tissue, these bacteria are able to shed unnecessary functions over time resulting in this reduced genome such as the loss of cell motility and signal transduction seen across Blattabacteria, likely associated with their direct vertical transmission pattern in which these bacteria are incorporated into developing oocytes (Gil et al., 2003; Huang et al., 2012; Kambhampati et al.,

2013; López-Sánchez et al., 2008, 2009; McCutcheon & Moran, 2010; Patino-Navarrete et al.,

2014; Sabree et al., 2012, 2009; Tokuda et al., 2013; van Ham et al., 2003; Wernegreen, 2002).

137 Maintenance of metabolic pathways producing nutrients, especially essential amino acids or vitamins, is also common in obligate endosymbionts (Akman et al., 2002; Wernegreen, 2002;

Wilson et al., 2010). In the genomes of the obligate endosymbionts of aphids, Buchnera aphidicola, and of carpenter ants, Blochmania floridanus, amino acid metabolism is similarly represented in coding genes (Gil et al., 2003; Shigenobu et al., 2000). While the genomes of tsetse fly symbionts Wigglesworthia sp. do not share the over-representation of amino acid biosynthesis, genes associated with synthesis and transport of cofactors and vitamins are abundant, allowing these bacteria to supplement the B-vitamin depleted blood-based diet of their host (Akman et al., 2002; Gil et al., 2003; Rio et al., 2012).

Blattabacteria are universal endosymbionts among cockroaches, residing in bacteriocytes in the fat body of their host (Bandi et al., 1994, 1995). Consequently, genomes of many strains have been sequenced, revealing a consistent pattern of metabolic capabilities.

Specifically, these bacteria have been implicated in recycling nitrogen waste which is stored in this organ as uric acid, synthesizing multiple amino acids in the process including nearly all ten essential amino acids (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008,

2009; Patino-Navarrete et al., 2014; Sabree et al., 2012, 2009; Tokuda et al., 2013). The amino acid biosynthetic capacity of this endosymbiont is of particular interest in the case of the

Diploptera punctata symbiosis due to the proposed role of Blattabacterium in intrauterine developing embryos (Williford et al., 2004; Chapter 3). Previous research has documented that the milk like secretion provided by D. punctata mothers to their developing embryos during pregnancy is deficient in methionine and tryptophan, two essential amino acids (Ingram et al.,

1977; Williford et al., 2004). Williford et al. (2004) proposed that bacterial symbionts

138 supplement the milk-like secretion, providing these amino acids to embryos. We previously demonstrated that the only bacteria present in the embryonic microbiome is the maternally transmitted Blattabacteria, further underscoring the importance of this symbiont to developing

D. punctata. While there is some variation between the strains in their biosynthetic abilities for essential amino acids, consistently the metabolic pathway for methionine biosynthesis is incomplete in the genomes and plasmids of Blattabacteria, including the DPU strain. All strains lack the genes metA and metX (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2009; Patino-Navarrete et al., 2014; Sabree et al., 2009; Tokuda et al., 2013), which perform the first steps in methionine biosynthesis by cysteine transsulfuration and acetylation respectively (Ferla & Patrick, 2014). Of note, previous research has demonstrated the ability of cultured fat body tissues, and presumably the associated bacteria, to synthesize methionine by cysteine transsulfuration (Block & Henry, 1961; Henry & Block, 1961). Additionally, most strains, including the DPU strain, maintain the remainder of the transsulfuration based metabolic pathway, metB, metC, and metE (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2009; Patino-Navarrete et al., 2014; Sabree et al., 2009; Tokuda et al., 2013). Because the fat body tissue of B. germanica has been shown capable of methionine biosynthesis, an alternative synthetic pathway has been suggested by Sabree et al (2009) where cystathionine-g- synthase produces homocysteine and O-phosphohomoserine, a product of threonine biosynthesis. Although there is little evidence of this pathway being employed by bacteria and is best characterized in Arabidopsis thaliana (Bartlem et al., 2000; Ferla & Patrick, 2014), flexibility in the affinity of cystathionine-g-synthase was observed in bacteria (Thangavelu &

Viola, 2016) supporting a possible implementation for methionine production in Blattabacteria.

139 Alternatively, Blattabacterium may utilize their close association with the host as a source for alternative substrates, as has been proposed in the B. aphidicola (Wilson et al 2010). B. aphidicola is proposed to utilize homocysteine synthesized by the aphid host from endosymbiont produced cysteine (Wilson et al 2010). Importantly, transcripts encoding cystathionine-g-lyases and cystathionine-b-synthase, the enzymes catalyzing the conversion of cysteine to homocysteine, can be found in the transcriptome generated in Chapter 1.

Additionally, all sequenced Blattabacterium strain genomes, with the exception of that of

Mastotermes darwinienses and Cryptocercus punctulatus, encode the full synthetic pathway for tryptophan synthesis (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008,

2009; Patino-Navarrete et al., 2014; Sabree et al., 2012, 2009; Tokuda et al., 2013) which is also lacking in D. punctata milk secretions acids (Ingram et al., 1977; Williford et al., 2004).

In conclusion, we present the genome sequence of a new Blattabacterium strain belonging to the live-bearing cockroach, D. punctata. This strain is genetically and functionally similar to that harbored by other cockroach species, possessing a reduced genome and the ability to recycle nitrogenous waste, such as urea and ammonia. As in many other insect endosymbionts, Blattabacterium has lost genes related to functions such as cell mobility, signal transduction as well as vitamin and cofactor metabolism, an indicator of adaptation to the specialized environment of the host. This specialization is also seen in the retention of ability to biosynthesize nearly all essential amino acids including tryptophan and likely methionine. The metabolic provisioning to the cockroach host by Blattabacterium has been suggested to be a key factor in allowing cockroaches to expand into such numerous and diverse ecological niches where these nutrients may have been scarce. Such provisioning may also have been a key

140 factor in allowing D. punctata embryos to survive on maternal nutrient secretion alone despite being deficient in two essential amino acids. However, to fully determine the ability of

Blattabacterium to synthesize methionine for D. punctata, especially intrauterine developing embryos, further physiological investigation will be required.

141 References:

Afgan, E., Baker, D., Batut, B., van den Beek, M., Bouvier, D., Čech, M., … Blankenberg, D.

(2018). The Galaxy platform for accessible, reproducible and collaborative biomedical

analyses: 2018 update. Nucleic Acids Research, 46(W1), W537–W544.

http://doi.org/10.1093/nar/gky379

Akman, L., Yamashita, A., Watanabe, H., Oshima, K., Shiba, T., Hattori, M., & Aksoy, S. (2002).

Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia

glossinidia. Nature Genetics, 32(3), 402–407. http://doi.org/10.1038/ng986

Alikhan, N. F., Petty, N. K., Ben Zakour, N. L., & Beatson, S. A. (2011). BLAST Ring Image

Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics, 12(1), 402.

http://doi.org/10.1186/1471-2164-12-402

Arkin, A. P., Cottingham, R. W., Henry, C. S., Harris, N. L., Stevens, R. L., Maslov, S., … Yu, D.

(2018). KBase: The United States Department of Energy Systems Biology Knowledgebase.

Nature Biotechnology, 36(7), 566–569. http://doi.org/10.1038/nbt.4163

Bandi, C., Damiani, G., Magrassi, L., Grigolo, A., Fani, R., & Sacchi, L. (1994). Flavobacteria as

intracellular symbionts in cockroaches. Proceedings of the Royal Society B: Biological

Sciences, 257, 43–48. http://doi.org/10.1098/rspb.1994.0092

Bandi, C., Sironi, M., Damiani, G., Magrassi, L., Nalepa, C. A., Laudant, U., & Sacchi, L. (1995).

The establishment of intracellular symbiosis in an ancestor of cockroaches and termites.

Proceedings of the Royal Society B: Biological Sciences, 259(1356), 293–299.

http://doi.org/10.1098/rspb.1995.0043

Bartlem, D., Lambein, I., Okamoto, T., Itaya, A., Uda, Y., Kijima, F., … Naito, S. (2000). Mutation

142 in the threonine synthase gene results in an over-accumulation of soluble methionine in

Arabidopsis. Plant Physiology, 123(1), 101–110. http://doi.org/10.1104/pp.123.1.101

Block, R. J., & Henry, S. M. (1961). Metabolism of the sulphur amino-acids and of sulphate in

Blattella germanica. Nature, 191(4786), 392–393. http://doi.org/10.1038/191392a0

Bolger, A. M., Lohse, M., & Usadel, B. (2014). Trimmomatic: a flexible trimmer for Illumina

sequence data. Bioinformatics, 30(15), 2114–2120.

http://doi.org/10.1093/bioinformatics/btu170

Darzi, Y., Letunic, I., Bork, P., & Yamada, T. (2018). iPath3.0: interactive pathways explorer v3.

Nucleic Acids Research, 46(W1), W510–W513. http://doi.org/10.1093/nar/gky299

Ferla, M. P., & Patrick, W. M. (2014). Bacterial methionine biosynthesis. Microbiology (United

Kingdom), 160, 1571–1584. http://doi.org/10.1099/mic.0.077826-0

Gil, R., Silva, F. J., Zientz, E., Delmotte, F., Gonzalez-Candelas, F., Latorre, A., … Moya, A. (2003).

The genome sequence of Blochmannia floridanus: Comparative analysis of reduced

genomes. Proceedings of the National Academy of Sciences, 100(16), 9388–9393.

http://doi.org/10.1073/pnas.1533499100

Gurevich, A., Saveliev, V., Vyahhi, N., & Tesler, G. (2013). QUAST: quality assessment tool for

genome assemblies. Bioinformatics, 29(8), 1072–1075.

http://doi.org/10.1093/bioinformatics/btt086

Hagan, H. R. (1948). A brief analysis of viviparity in insects. Journal of the New York

Entomological Society, 56(1), 63–68.

Henry, S. M., & Block, R. J. (1961). Sulfur metabolism of insects. 6. Metabolism of sulfur amino

acids and related compounds in German cockroach, Blattella germanica (L.). Contrib.

143 Boyce Thompson Inst., 21, 129–145.

Huang, C. Y., Sabree, Z. L., & Moran, N. a. (2012). Genome sequence of Blattabacterium sp.

strain BGIGA, endosymbiont of the Blaberus giganteus cockroach. Journal of Bacteriology,

194(16), 4450–4451. http://doi.org/10.1128/JB.00789-12

Ingram, M. J., Stay, B., & Cain, G. D. (1977). Composition of milk from the viviparous cockroach,

Diploptera punctata. Insect Biochemistry, 7(3), 257–267. http://doi.org/10.1016/0020-

1790(77)90023-3

Jiang, Z. F., Xia, F., Johnson, K. W., Brown, C. D., Bartom, E., Tuteja, J. H., … Ghanim, M. (2013).

Comparison of the genome sequences of “Candidatus portiera aleyrodidarum” primary

endosymbionts of the whitefly Bemisia tabaci B and Q biotypes. Applied and

Environmental Microbiology, 79, 1757–1759. http://doi.org/10.1128/AEM.02976-12

Kambhampati, S., Alleman, A., & Park, Y. (2013). Complete genome sequence of the

endosymbiont Blattabacterium from the cockroach Nauphoeta cinerea (Blattodea:

Blaberidae). Genomics, 102(5–6), 479–483. http://doi.org/10.1016/j.ygeno.2013.09.003

Kanehisa, M., Sato, Y., Furumichi, M., Morishima, K., & Tanabe, M. (2019). New approach for

understanding genome variations in KEGG. Nucleic Acids Research, 47(D1), D590–D595.

http://doi.org/10.1093/nar/gky962

López-Sánchez, M. J., Neef, A., Patiño-Navarrete, R., Navarro, L., Jiménez, R., Latorre, A., &

Moya, A. (2008). Blattabacteria, the endosymbionts of cockroaches, have small genome

sizes and high genome copy numbers. Environmental Microbiology, 10(12), 3417–3422.

http://doi.org/10.1111/j.1462-2920.2008.01776.x

López-Sánchez, M. J., Neef, A., Peretó, J., Patiño-Navarrete, R., Pignatelli, M., Latorre, A., &

144 Moya, A. (2009). Evolutionary convergence and nitrogen metabolism in Blattabacterium

strain Bge, primary endosymbiont of the cockroach Blattella germanica. PLoS Genetics,

5(11). http://doi.org/10.1371/journal.pgen.1000721

Marchal, E., Hult, E. F., Huang, J., Stay, B., & Tobe, S. S. (2013). Diploptera punctata as a model

for studying the endocrinology of arthropod reproduction and development. General and

Comparative Endocrinology, 188, 85–93.

http://doi.org/http://dx.doi.org/10.1016/j.ygcen.2013.04.018

McCutcheon, J. P., & Moran, N. A. (2010). Functional convergence in reduced genomes of

bacterial symbionts spanning 200 my of evolution. Genome Biology and Evolution, 2(1),

708–718. http://doi.org/10.1093/gbe/evq055

McCutcheon, J. P., & Moran, N. A. (2012). Extreme genome reduction in symbiotic bacteria.

Nature Reviews Microbiology, 10(1), 13–26. http://doi.org/10.1038/nrmicro2670

Moran, N. A., & Wernegreen, J. J. (2000). Lifestyle evolution in symbiotic bacteria: insights from

genomics. Trends in Ecology & Evolution, 15(8), 321–326. http://doi.org/10.1016/S0169-

5347(00)01902-9

Nakabachi, A., Yamashita, A., Toh, H., Ishikawa, H., Dunbar, H. E., Moran, N. A., & Hattori, M.

(2006). The 160-kilobase genome of the bacterial endosymbiont Carsonella. Science,

314(5797), 267–267. http://doi.org/10.1126/science.1134196

Nikoh, N., Hosokawa, T., Oshima, K., Hattori, M., & Fukatsu, T. (2011). Reductive evolution of

bacterial genome in insect gut environment. Genome Biology and Evolution, 3(1), 702–714.

http://doi.org/10.1093/gbe/evr064

Nurk, S., Bankevich, A., Antipov, D., Gurevich, A. A., Korobeynikov, A., Lapidus, A., … Pevzner, P.

145 A. (2013). Assembling single-cell genomes and mini-metagenomes from chimeric mda

products. Journal of Computational Biology, 20(10), 714–737.

http://doi.org/10.1089/cmb.2013.0084

Patino-Navarrete, R., Piulachs, M.-D., Belles, X., Moya, A., Latorre, A., & Pereto, J. (2014). The

cockroach Blattella germanica obtains nitrogen from uric acid through a metabolic

pathway shared with its bacterial endosymbiont. Biology Letters, 10(7), 20140407–

20140407. http://doi.org/10.1098/rsbl.2014.0407

Rio, R. V. M., Symula, R. E., Wang, J., Lohs, C., Wu, Y.-N., Snyder, A. K., … Aksoy, S. (2012).

Insight into the transmission biology and species-specific functional capabilities of tsetse

(Diptera: Glossinidae) obligate symbiont Wigglesworthia. MBio, 3(1), 1–13.

http://doi.org/10.1128/mBio.00240-11

Roth, L. M. (1967). Water changes in cockroach oothecae in relation to the evolution of

ovoviviparity and viviparity. Annals of the Entomological Society of America, 60(5), 928–

946. http://doi.org/10.1093/aesa/60.5.928

Roth, L. M., & Stay, B. (1961). Oöcyte development in Diploptera punctata (Eschscholtz)

(Blattaria). Journal of Insect Physiology. http://doi.org/10.1016/0022-1910(61)90071-3

Roth, L. M., & Willis, E. R. (1955). Intra-uterine nutrition of the “beetle-roach” Diploptera

dytiscoides (Serv.) during embryogenesis, with notes on its biology in the laboratory

(Blattaria: Diplopteridae). Psyche: A Journal of Entomology.

http://doi.org/10.1155/1955/12542

Roth, L. M., & Willis, E. R. (1957). An analysis of oviparity and viviparity in the Blattaria.

American Entomological Society, 83(4), 221–238.

146 Sabree, Z. L., Huang, C. Y., Arakawa, G., Tokuda, G., Lo, N., Watanabe, H., & Moran, N. A. (2012).

Genome shrinkage and loss of nutrient-providing potential in the obligate symbiont of the

primitive termite Mastotermes darwiniensis. Applied and Environmental Microbiology,

78(1), 204–210. http://doi.org/10.1128/AEM.06540-11

Sabree, Z. L., Kambhampati, S., & Moran, N. a. (2009). Nitrogen recycling and nutritional

provisioning by Blattabacterium, the cockroach endosymbiont. Proceedings of the National

Academy of Sciences of the United States of America, 106(46), 19521–19526.

http://doi.org/10.1073/pnas.0907504106

Santos-Garcia, D., Latorre, A., Moya, A., Gibbs, G., Hartung, V., Dettner, K., … Silva, F. J. (2014).

Small but powerful, the primary endosymbiont of moss bugs, Candidatus evansia muelleri,

holds a reduced genome with large biosynthetic capabilities. Genome Biology and

Evolution, 6(7), 1875–1893. http://doi.org/10.1093/gbe/evu149

Seemann, T. (2014). Prokka: rapid prokaryotic genome annotation. Bioinformatics, 30(14),

2068–2069. http://doi.org/10.1093/bioinformatics/btu153

Shigenobu, S., Watanabe, H., Hattori, M., Sakaki, Y., & Ishikawa, H. (2000). Genome sequence of

the endocellular bacterial symbiont of aphids Buchnera sp. APS. Nature, 407(6800), 81–86.

http://doi.org/10.1038/35024074

Stay, B., & Coop, A. (1973). Developmental stages and chemical composition in embryos of the

cockroach, Diploptera punctata, with observations on the effect of diet. Journal of Insect

Physiology, 19(1), 147–171. http://doi.org/10.1016/0022-1910(73)90230-8

Stay, B., & Coop, A. C. (1974). “Milk” secretion for embryogenesis in a viviparous cockroach.

Tissue and Cell, 6(4), 669–693. http://doi.org/10.1016/0040-8166(74)90009-3

147 Stay, B., & Roth, L. M. (1956). The reproductive behavior of Diploptera punctata (Blattaria:

Diplopterinae). Proceedings of the International Congress of Entomology.

Thangavelu, B., & Viola, R. (2016). Alternative microbial routes to homoserine activation and

sulfuration: critical steps in methionine biosynthesis. International Journal of Biochemistry

Research & Review, 11(3), 1–22. http://doi.org/10.9734/ijbcrr/2016/25074

Tokuda, G., Elbourne, L. D. H., Kinjo, Y., Saitoh, S., Sabree, Z. L., Hojo, M., … Lo, N. (2013).

Maintenance of essential amino acid synthesis pathways in the Blattabacterium cuenoti

symbiont of a wood-feeding cockroach. Biology Letters, 9, 20121153.

http://doi.org/10.1098/rsbl.2012.1153 van Ham, R. C. H. J., Kamerbeek, J., Palacios, C., Rausell, C., Abascal, F., Bastolla, U., … Moya, A.

(2003). Reductive genome evolution in Buchnera aphidicola. Proceedings of the National

Academy of Sciences, 100(2), 581–586. http://doi.org/10.1073/pnas.0235981100

Wernegreen, J. J. (2002). Genome evolution in bacterial endosymbionts of insects. Nature

Reviews Genetics, 3(11), 850–861. http://doi.org/10.1038/nrg931

Williford, A., Stay, B., & Bhattacharya, D. (2004). Evolution of a novel function: nutritive milk in

the viviparous cockroach, Diploptera punctata. Evolution and Development, 6, 67–77.

http://doi.org/10.1111/j.1525-142X.2004.04012.x

Wilson, A. C. C., Ashton, P. D., Calevro, F., Charles, H., Colella, S., Febvay, G., … Douglas, A. E.

(2010). Genomic insight into the amino acid relations of the pea aphid, Acyrthosiphon

pisum, with its symbiotic bacterium Buchnera aphidicola. Insect Molecular Biology,

19(SUPPL. 2), 249–258. http://doi.org/10.1111/j.1365-2583.2009.00942.x

Wolf, Y. I., & Koonin, E. V. (2013). Genome reduction as the dominant mode of evolution.

148 BioEssays, 35(9), 829–837. http://doi.org/10.1002/bies.201300037

149 Tables

Table 1. Genome characteristics of Blattabacterium from D. punctata and 6 other published strains. Total CDS is represented as the number of CDS on the chromosome + the number found on the plasmid.

D. punctata B. germanica C. punctulatus P. americana N. cinerea B. orientalis B. giganteus Chromosome size (bp) 623,008 636,850 605,745 636,994 622,952 634,449 629,165 G + C content (%) 28.0 27.1 23.8 28.2 26.2 28.2 25.7 Total CDS 580 + 4 586 + 4 545 + 3 587 + 4 581 + 5 572 + 7 573 + 4 rRNAs 3 3 3 3 3 3 3 tRNAs 34 34 32 33 32 33 34 Other RNAs 1 3 3 1 1 3 1 Plasmid size (bp) 2,852 4,085 3,816 3,448 3,675 3,735 3,423

150 Figures and legends

Legends

Figure 1. Genome presentation of Blattabacterium str. Dpu. Outward to inward rings represent

GC Skew (purple and green bars are negative and positive respectively), GC content of each strand, RNA genes including tRNAs, rRNA, and tmRNA (Blue), and lastly predicted CDS in teal.

Figure 2. COG composition of the DPU and 8 previously sequenced strains of Blattabacterium.

Letters refer to COG functional categories (C-Energy production and conversion; D- Cell division and chromosome partitioning; E- Amino acid transport and metabolism; F- Nucleotide transport and metabolism; G- Carbohydrate transport and metabolism; H- Coenzyme metabolism; I- Lipid metabolism; J- Translation, ribosomal structure and biogenesis; K- Transcription; L- DNA replication, recombination and repair; M- Cell envelope biogenesis, outer membrane; O-Post- translational modification, protein turnover, chaperones; P- Inorganic ion transport and metabolism; Q- Secondary metabolites biosynthesis, transport and catabolism; S- COG of unknown function; N/A- not assigned). Host abbreviations are as follows: Bbge, Blatella germanica; Bbgiga, Blaberus giganteus; Bbor, Blatella orientalis; Bcpu, Cryptocercus punctulatus; Bdpu, D. punctata; Bncin, Nauphoeta cinerea; Bpaa, Paraethria angustipennis;

Bplan, Periplaneta americana; Madar, Mastotermes darwiniensis.

Figure 3. Biosynthetic ability of 7 strains of Blattabacterium for essential amino acids. Presence or absence (filled and white boxes, respectively) of genes encoding key enzymes associated

151 with biosynthesis of the ten essential amino acids for D. punctata (Blue filled) and 6 additional strains of Blattabacterium (Green filled).

152 Figures

Blattabacterium str. Dpu 623,008 bp

153 Bbge Bbgiga Bbor Bcpu Bdpu Bncin Bpaa Bplan Madar Strain NA V U S Q P O M L K COG Term J I H G F E D C 0.25 0.20 0.15 0.10 0.05 0.00

Frequency in genome in Frequency 154 Met Lys Thr Ile/Val/Leu Arg His Phe Trp asd ilvA ilvC ilvE hisl lysA hisC trpF thrA thrB thrC leuA leuB argF hisG hisA hisD tyrB trpE trpC trpA trpB dapF argA argB argC argD argE argH metB metE dapD dapE ilvBH argG pheA metA DapA DapB hisHF Host Species leuDC trpGD

D. punctata

B. giganteus

P. angustipennis

B. germanica

P. americana

C. punctulatus

M. darwiniensis

155

Chapter 5

Discussion and Conclusions

156 The principle aim of my dissertation research was to further characterize the reproductive biology of the Pacific beetle mimic cockroach, Diploptera punctata, at the molecular levels and in relation to potential microbial symbionts. First, I sought to examine molecular aspects of pregnancy in D. punctata, utilizing transcriptomics to measure transcript changes as embryo development progresses in the brood sac (Chapter 2). I then leveraged 16S rRNA sequencing to characterize the developmental trajectory of the microbiome in this cockroach, with a focus on determining if the viviparous reproduction of D. punctata facilitates prenatal acquisition of microbes by developing embryos (Chapter 3). Lastly, I utilized genome sequencing and annotation to elucidate the potential metabolic interactions between D. punctata and their primary endosymbionts, Blattabacterium (Chapter 4). The work presented in this dissertation represents the successful completion of each individual aim, together creating a holobiome characterization of D. punctata which will facilitate future research in the molecular mechanisms underlying the reproductive biology of this insect system as well as in the broader scope of viviparity in the animal kingdom.

In Chapter 2, I describe the transcriptome of D. punctata across pregnancy, generating the largest set of sequenced genes and first reported transcriptome of this species. I identified

11, 987 putative protein coding sequences containing over 90% of the benchmark universal single copy ortholog genes as fragmented or complete sequences, indicating a high quality and complete gene set. Of these, 2,474 were differentially regulated across pregnancy including the previously characterized milk protein cDNAs. I identified differential regulation of juvenile hormone esterase (jhe), which is known to suppress circulating levels of juvenile hormone (JH) and differential activity during oogenesis and ovipositioning (de Kort & Granger, 1996;

157 Engelmann & Mala, 2000; Rankin & Stay, 1984; Rotin & Tobe, 1983; Stoltzman, Stocker, Borst,

& Stay, 2000). JH has been shown to play an important role in both development and reproduction in most insects, including D. punctata, with high levels of this hormone inducing vitellogenic oocyte development (Luo et al., 2017; Mundall, Tobe, & Stay, 1981; Naghdi,

Maestro, Belles, & Bandani, 2016; Stoltzman & Stay, 1997; Tufail, Nagaba, Elgendy, & Takeda,

2014). In D. punctata circulating JH and JH biosynthesis increases in the days after mating, peaking at 4-5 days post mating when yolk formation is at its peak; also, during that time juvenile hormone esterase enzymatic (JHE) activity is at its lowest (de Kort & Granger, 1996;

Rotin & Tobe, 1983; Stoltzman et al., 2000; Tobe et al., 1985). Shortly after yolk generation, JHE activity increases and circulating JH levels rapidly drop as females oviposit their eggs into the brood sac; JH remains low until near the end of the pregnancy when the next cycle of oogenesis birth (de Kort & Granger, 1996; Rotin & Tobe, 1983; Stoltzman et al., 2000; Tobe et al., 1985).

The suppression of JH levels has been demonstrated as imperative in the matrotrophic viviparity of D. punctata, as artificial elevation of JH through application of JH analogs or implantation of additional corpora allata inhibits milk production and may even result in the termination of pregnancy altogether while stimulating new egg development (Evans & Stay,

1995; Stay & Lin, 1981; Stoltzman & Stay, 1997; Terr Wee & Stay, 1987). Utilizing RNA- interference to suppress transcript levels, I identified jhe expression as a key component in the maintenance of milk production. This effect is likely due to known role of JHE in suppressing circulating juvenile hormone titers. While JHE activity was known to be associated with the low

JH titer across the gonadotropic cycle of D. punctata (de Kort & Granger, 1996; Rotin & Tobe,

1983; Stoltzman et al., 2000; Tobe et al., 1985), previous studies had also shown a general

158 decrease in the biosynthesis and release of JH in the corpora allata (Marchal, Hult, Huang, Stay,

& Tobe, 2013; Paulson & Stay, 1987; Tobe et al., 1985). The relative contribution of each in suppressing JH to maintain pregnancy was previously unknown. Our findings determined that suppressing jhe significantly reduce the expression of milk protein transcripts, which likely will either delay embryo development or lead to abortion.

Importantly, neither JH as a regulator of reproduction nor matrotrophic viviparity are unique to D. punctata. While many other insects are live-bearing, tsetse flies (Glossina spp.) are possibly the most well studied due to their capacity as vectors for trypanosomiasis (Langley,

1994; Rio et al., 2012). Phylogenetically these two species are extremely distant, yet they have both strikingly converged upon not only milk production (Attardo et al., 2019; Denlinger & Ma,

1974; Ingram, Stay, & Cain, 1977; Stay & Coop, 1974; Tobe, Davey, & Huebner, 1973; Tobe &

Langley, 1978; Williford, Stay, & Bhattacharya, 2004), but similar hormonal controls of milk production and pregnancy despite the cooption of different organs both for milk production and organs that act as uterus-like structures (Baumann et al., 2013; Ejezie & Davey, 1976; Evans

& Stay, 1989, 1995; Langley & Pimley, 1986; Marchal et al., 2013; Stay, Ostedgaard, Tobe,

Strambi, & Spaziani, 1984; Tobe et al., 1985). While the tsetse fly JH titer has not been directly characterized, one study utilizing expression of JH biosynthesis enzymes and downstream gene targets of JH as a proxy revealed that the biosynthesis and circulating levels of JH are high during vitellogenesis and low during pregnancy in the tsetse fly, Glossina morsitans (Baumann et al., 2013; Robert, Strambi, & Strambi, 1986). This convergence is further underscored by the shared negative responses to increased JH levels or JH mimics; JH inhibits milk protein production and may induce premature parturition (Baumann et al., 2013; Denlinger, 1975;

159 Evans & Stay, 1995; Langley & Pimley, 1986; Stay & Lin, 1981; Terr Wee & Stay, 1987).

Interestingly, in tsetse flies there appears to be an interaction between insulin signaling and JH signaling that has not yet been examined in D. punctata (Baumann et al., 2013). The structural and morphological differences of reproduction in D. punctata and in tsetse flies combined with the degree of convergence upon matrotrophic viviparity and the hormonal regulation thereof underscore the importance of comparative molecular studies to understand the evolution of the diversity in matrotrophic viviparous insects. The research presented in Chapter 2 identifies transcriptomic patterns that may be key to the evolution of viviparity, such as differential regulation of JH breakdown genes, and demonstrates the need for genetic and transcriptomic studies in more diverse matrotrophic viviparous insect systems.

In Chapter 3 I utilized 16S rRNA gene sequencing to characterize the microbiome of pregnant female D. punctata and their embryos, as well as a developmental trajectory from birth to adulthood of this microbial community. This study revealed that despite the immense interactions between mothers and their offspring during pregnancy, there are no microbial endosymbionts present in embryos outside of Blattabacterium, the single endosymbiont known to be vertically transmitted from mother to offspring in cockroaches (Bandi et al., 1994, 1995;

Giorgi & Nordin, 1994). This is a novel finding considering little is known regarding the impact of reproductive mode on microbe transmission for invertebrate systems. The literature, and logic, suggests that egg laying insects have restricted opportunity to stably transmit symbionts with high fidelity due to the short period in which mothers interact with their eggs (essentially during oogenesis and then ovipositioning) (Abdul Rahman et al., 2015; Bright & Bulgheresi,

2010; da Costa & Poulsen, 2018; Estes et al., 2013; Funkhouser & Bordenstein, 2013; Salem,

160 Florez, Gerardo, & Kaltenpoth, 2015; Schwab, Riggs, Newton, & Moczek, 2016; Shukla, Vogel,

Heckel, Vilcinskas, & Kaltenpoth, 2018). Thus, it has been found that microbes are only able to be directly transmitted to offspring transovarially during oogenesis and all other symbionts must be acquired after hatching; though, these species have made behavioral adaptations such as oviposition substrate curation, fecal smearing of eggs, or trophallaxis, to ensure their offspring encounter and internalize symbiotic microbes (Abdul Rahman et al., 2015; Bright &

Bulgheresi, 2010; da Costa & Poulsen, 2018; Estes et al., 2013; Funkhouser & Bordenstein,

2013; Salem et al., 2015; Schwab et al., 2016; Shukla et al., 2018).

Intrauterine gestation associated with live-bearing animals provides ample time for mothers to exchange microbes with their offspring. Yet, vertical transmission of symbionts has not been well studied in viviparous species outside of mammals and select insect systems.

Placental transmission of symbionts during mammalian pregnancy is debated, although the current consensus appears that it is unlikely (Aagaard et al., 2014; Blaser & Dominguez-Bello,

2016; Fardini, Chung, Dumm, Joshi, & Han, 2010; Perez-Muñoz, Arrieta, Ramer-Tait, & Walter,

2017; Walker, Clemente, Peter, & Loos, 2017). Infants will acquire important microbes from their mothers when passing through the birth canal and through during early development (Ballard & Morrow, 2013; Dahlen, Downe, Kennedy, & Foureur, 2014; Duranti et al., 2017; Funkhouser & Bordenstein, 2013; Jašarević, Rodgers, & Bale, 2015; Mueller, Bakacs,

Combellick, Grigoryan, & Dominguez-Bello, 2015). Tsetse flies and other Hippoboscoidea, transfer vital symbionts to offspring during pregnancy through their nutritive secretions

(Denlinger & Ma, 1975; Funkhouser & Bordenstein, 2013; Morse et al., 2013; Mueller et al.,

2015; Wang, Weiss, & Aksoy, 2013). In contrast, viviparous aphids exclusively employ

161 transovarial symbiont transmission before parturition (Michalik, Szklarzewicz, Jankowska, &

Wieczorek, 2014); although, while intrauterine developing aphids do acquire nutrients from the maternal hemocoel, the ovariole sheathe acts as a semipermeable barrier surrounding the embryos and it is unknown the degree to which nutrients are synthesized by the mother

(Bermingham & Wilkinson, 2009; Couchman & King, 1980; Davis, 2012; Michalik et al., 2014).

Additionally, female aphids are able to dynamically transition from oviparous to viviparous reproduction (Bermingham & Wilkinson, 2009; Bickel et al., 2013; Davis, 2012; Miura et al.,

2003; Ogawa & Miura, 2014), The shift in reproductive mode may limit their capacity for non- transovarial symbiont transfer. My findings that D. punctata fail to transmit symbionts outside of Blattabacterium to their offspring further highlights viviparity diversity and vertical symbiont transmission across insects and the animal kingdom.

In addition, I found that no microbes are acquired or internalized as symbionts by D. punctata during the birthing process, in contrast to the diverse suite of bacteria that colonize human infants during birth. Rather, first instar D. punctata nymphs do not have a significant symbiont community until melanization. The microbiome continues to diversify across the remaining three instar stages until adulthood. Overall, we identified 50 phyla, 122 classes, 252 orders, 485 families and 1008 genera of bacteria in these samples. Over the course of nymphal development Blattabacteria proportion decreases relative to the general bacterial community, ultimately representing only 20% of the microbiome. The increase in bacterial components are consistent with previous studies on microbiome development in cockroaches as well as studies on adult microbial communities (Bauer et al., 2015; Bertino-Grimaldi et al., 2013; Carrasco et

162 al., 2014; Gontang et al., 2017; Kakumanu, Maritz, Carlton, & Schal, 2018; Pérez-Cobas et al.,

2015; Sabree & Moran, 2014; Schauer, Thompson, & Brune, 2014; Tinker & Ottesen, 2016).

My findings are of particular interest in this species because the embryonic microbiome has been proposed to serve an important role in providing vital essential amino acids that have been found to be lacking in milk secretions, tryptophan and methionine (Ingram et al., 1977;

Williford et al., 2004). The minimal microbiome of D. punctata embryos indicates that

Blattabacterium must be the source of supplemental nutrients during gestation. It has been well established that endosymbionts can serve to supplement a nutrient deficient diet across the animal kingdom (Bermingham & Wilkinson, 2009; Douglas, 2017; Engel & Moran, 2013;

Funkhouser & Bordenstein, 2013; Michalik et al., 2014; Michalkova, Benoit, Weiss, Attardo, &

Aksoy, 2014). Viviparous insects such as tsetse flies and aphids both have endosymbionts that act as nutritional complements during intrauterine development (Wilson et al 2010; Michalik et al 2014; Bermingham and Wilkinson 2009; Davis 2012; Douglas, 2017; Snyder, Mclain, & Rio,

2012; Snyder & Rio, 2015; J. Wang et al., 2013).

In Chapter 4 I report the genome of Blattabacterium isolated from D. punctata. I assembled and annotated the 623,000 bp chromosome and 2,852 bp plasmid from this strain of

Blattabacterium, containing 580 and 4 coding DNA sequences respectively. This genome also contained 3 ribosomal RNAs, 34 transfer RNAs, and 1 transfer messenger RNA. These genes represent capacity for DNA replication and transcription, translational machinery as well as ability to biosynthesize essential and non-essential amino acids; however, cell motility, signal transduction and secondary metabolite synthesis are all virtually absent in this genome. These values are within the range of previously sequenced Blattabacterium strain genomes (Huang,

163 Sabree, & Moran, 2012; Kambhampati, Alleman, & Park, 2013; López-Sánchez et al., 2008,

2009; Patino-Navarrete et al., 2014; Sabree et al., 2012; Sabree, Kambhampati, & Moran, 2009;

Tokuda et al., 2013). The reduction in genome size and selective metabolic process retention is characteristic of obligate bacterial endosymbionts and can be seen in endosymbionts like

Buchnera aphidicola, Wigglesworthia spp., Blochmania floridanus and more (Akman et al.,

2002; Gil et al., 2003; Jiang et al., 2013; McCutcheon & Moran, 2010, 2012; Moran &

Wernegreen, 2000; Nakabachi et al., 2006; Nikoh, Hosokawa, Oshima, Hattori, & Fukatsu, 2011;

Santos-Garcia et al., 2014; Shigenobu, Watanabe, Hattori, Sakaki, & Ishikawa, 2000;

Wernegreen, 2002; Wolf & Koonin, 2013).

This bacterium is a universal constituent of the cockroach microbiome; it lives in specialized bacteriocytes in the fat body and ovaries where it contributes to nitrogen cycling and essential amino acid synthesis for the cockroach host (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008, 2009; Patino-Navarrete et al., 2014; Sabree et al., 2012,

2009; Tokuda et al., 2013). Interestingly, fat body tissue of the German cockroach has been reported to be capable of methionine biosynthesis (Block & Henry, 1961; Henry & Block, 1961) but sequenced genomes of the Blattella germanica and 9 other strains revealed incomplete methionine biosynthetic pathways encoded in the genome and plasmids, although most strains retain complete biosynthetic pathways for all other essential amino acids including tryptophan

(Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008, 2009; Patino-

Navarrete et al., 2014; Sabree et al., 2012, 2009; Tokuda et al., 2013). Similarly, Blattabacterium from D. punctata is able to produce nearly all essential amino acids, but lacks the enzyme responsible for homoserine activation, the first step in methionine biosynthesis. The

164 incomplete biosynthetic pathway for methionine is of great importance relative as supplementation of this amino acid during embryonic development is critical since it is lacking in milk (Ingram et al., 1977; Williford et al., 2004; Chapter 3).

Bacteria biosynthesize methionine through cysteine transsulfuration or acetylation, a process initiated with homoserine activation by O-succinylhomoserine transferase (metA) or O- acetylhomoserine transferase (metX) (Ferla & Patrick, 2014). Despite lacking either enzyme,

Blattabacterium genomes typically contain the remainder of the transsulfuration based pathway (metB, metC, metE) (Huang et al., 2012; Kambhampati et al., 2013; López-Sánchez et al., 2008, 2009; Patino-Navarrete et al., 2014; Sabree et al., 2012, 2009; Tokuda et al., 2013).

While an alternative homoserine activation pathway has yet to be seen in bacteria, Arabidopsis thaliana is capable of synthesizing methionine without metA or metX using O- phosphohomoserine (thrB), a biproduct of threonine biosynthesis, as a substrate of cystathionine-g-synthase (metB) (Bartlem et al., 2000; Ferla & Patrick, 2014). Blattabacterium from D. punctata does contain thrB and metB, making this pathway feasible. Although the ability of bacterial metB to bind O-phosphohomoserine has yet to be observed (Ferla & Patrick,

2014), cystathionine-g-synthase is notoriously flexible (Thangavelu & Viola, 2016). Alternatively, the symbiont may acquire the intermediate substrate or an alternate substrate, likely from the host cockroach as has been proposed in the B. aphidicola—aphid symbiosis (Wilson et al.,

2010). It has been suggested that B. aphidicola exchanges cysteine and homocysteine with its host to produce methionine despite lacking enzymes for these intermediate steps (Wilson et al., 2010); transcripts encoding cystathionine-g-lyases and cystathionine-b-synthase, the enzymes catalyzing the conversion of cysteine to homocysteine, importantly can be found in

165 the host transcriptome generated in Chapter 1. Although a likely product of the extensive co- evolution occurring between Blattabacterium and D. punctata, additional experiments employing techniques such as radiolabeling would be required to confirm if and how

Blattabacterium and the cockroach host exchange these metabolites. The results of Chapters 3 and 4 suggest that despite lacking a complete biosynthetic pathway for methionine,

Blattabacterium is likely still capable of supplementing missing nutrients in D. punctata embryos since it is the only symbiont. While extensive and unique symbioses have allowed insects such as cockroaches to invade new niche spaces by reducing dietary needs (Douglas,

2017; Peterson & Scharf, 2016; Rosenberg, Sharon, & Zilber-Rosenberg, 2009; Rosenberg &

Zilber-Rosenberg, 2011; Sabree et al., 2009), our findings suggest that Blattabacterium and D. punctata interactions have allowed this species to evolve matrotrophic viviparity that allows D. punctata to utilize lipocalins, which can act as both nutrients and carrier for hydrophobic molecules (Banerjee et al., 2016; Williford et al., 2004), as milk proteins despite a deficiency in both methionine and tryptophan in these proteins.

In conclusion, the work presented in this thesis provides a holistic characterization of

Diploptera punctata reproductive biology. The transcriptome generated by Chapter 1, combined with the microbiome description in Chapter 2, and genome sequence of the primary

D. punctata endosymbiont, Blattabacterium, in Chapter 3 has initiated the development of D. punctata as a model system to study matrotrophic viviparity. The utilization of my transcriptome and Blattabacterium genome in the ongoing sequencing and assembly of the D. punctata genome (in progress and should be completed in 2019/2020) serves as evidence of the utility of this work. Additionally, my findings have begun to identify trends of shared

166 evolutionary pathways to matrotrophic viviparity, as evidenced in Chapter 1; as well as ways in which viviparous insects differ such as in the vertical transmission of symbionts that I describe in Chapter 2; and reinforced the understanding of the intricate co-evolution between cockroaches and Blattabacterium that we see in the likely sharing of nutrients between host and symbiont proposed in Chapter 3. These patterns of convergence and diversity I show in the

D. punctata system demonstrate the necessity of continued comparative studies on matrotrophic viviparity and development of molecular resources for new and diverse study systems.

167 References:

Aagaard, K., Ma, J., Antony, K. M., Ganu, R., Petrosino, J., & Versalovic, J. (2014). The placenta

harbors a unique microbiome. Science Translational Medicine, 6(237), 237ra65.

http://doi.org/10.1126/scitranslmed.3009864

Abdul Rahman, N., Parks, D. H., Willner, D. L., Engelbrektson, A. L., Goffredi, S. K., Warnecke, F.,

… Hugenholtz, P. (2015). A molecular survey of Australian and North American termite

genera indicates that vertical inheritance is the primary force shaping termite gut

microbiomes. Microbiome, 3(1), 1–16. http://doi.org/10.1186/s40168-015-0067-8

Akman, L., Yamashita, A., Watanabe, H., Oshima, K., Shiba, T., Hattori, M., & Aksoy, S. (2002).

Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia

glossinidia. Nature Genetics, 32(3), 402–407. http://doi.org/10.1038/ng986

Attardo, G. M., Abd-Alla, A. M. M., Acosta-Serrano, A., James, E., Bateta, R., Benoit, J. B., …

Aksoy, S. (2019). The Glossina genome cluster: comparative genomic analysis of the

vectors of African Trypanosomes. BioRxiv, 531749. http://doi.org/10.1101/531749

Ballard, O., & Morrow, A. L. (2013). Human Milk Composition. Nutrients and Bioactive Factors.

Pediatric Clinics of North America, 60(1), 49–74. http://doi.org/10.1016/j.pcl.2012.10.002

Bandi, C., Damiani, G., Magrassi, L., Grigolo, A., Fani, R., & Sacchi, L. (1994). Flavobacteria as

intracellular symbionts in cockroaches. Proceedings of the Royal Society B: Biological

Sciences, 257, 43–48. http://doi.org/10.1098/rspb.1994.0092

Bandi, C., Sironi, M., Damiani, G., Magrassi, L., Nalepa, C. A., Laudant, U., & Sacchi, L. (1995).

The establishment of intracellular symbiosis in an ancestor of cockroaches and termites.

Proceedings of the Royal Society B: Biological Sciences, 259(1356), 293–299.

168 http://doi.org/10.1098/rspb.1995.0043

Banerjee, S., Coussens, N. P., Gallat, F. X., Sathyanarayanan, N., Srikanth, J., Yagi, K. J., …

Ramaswamy, S. (2016). Structure of a heterogeneous, glycosylated, lipid-bound, in vivo-

grown protein crystal at atomic resolution from the viviparous cockroach Diploptera

punctata. IUCrJ, 3, 282–293. http://doi.org/10.1107/S2052252516008903

Bartlem, D., Lambein, I., Okamoto, T., Itaya, A., Uda, Y., Kijima, F., … Naito, S. (2000). Mutation

in the threonine synthase gene results in an over-accumulation of soluble methionine in

Arabidopsis. Plant Physiology, 123(1), 101–110. http://doi.org/10.1104/pp.123.1.101

Bauer, E., Lampert, N., Mikaelyan, A., Köhler, T., Maekawa, K., & Brune, A. (2015).

Physicochemical conditions, metabolites and community structure of the bacterial

microbiota in the gut of wood-feeding cockroaches (Blaberidae: Panesthiinae). FEMS

Microbiology Ecology, 91(2), 1–14. http://doi.org/10.1093/femsec/fiu028

Baumann, A. A., Benoit, J. B., Michalkova, V., Mireji, P. O., Attardo, G. M., Moulton, J. K., …

Aksoy, S. (2013). Juvenile hormone and insulin suppress lipolysis between periods of

lactation during tsetse fly pregnancy. Molecular and Cellular Endocrinology, 372(1–2), 30–

41. http://doi.org/10.1016/j.mce.2013.02.019

Bermingham, J., & Wilkinson, T. L. (2009). Embryo nutrition in parthenogenetic viviparous

aphids. Physiological Entomology, 34(2), 103–109. http://doi.org/10.1111/j.1365-

3032.2008.00669.x

Bertino-Grimaldi, D., Medeiros, M. N., Vieira, R. P., Cardoso, A. M., Turque, A. S., Silveira, C. B.,

… Machado, E. A. (2013). Bacterial community composition shifts in the gut of Periplaneta

americana fed on different lignocellulosic materials. SpringerPlus, 2(1), 1–11.

169 http://doi.org/10.1186/2193-1801-2-609

Bickel, R. D., Cleveland, H. C., Barkas, J., Jeschke, C. C., Raz, A. A., Stern, D. L., & Davis, G. K.

(2013). The pea aphid uses a version of the terminal system during oviparous, but not

viviparous, development. EvoDevo, 4(1), 10. http://doi.org/10.1186/2041-9139-4-10

Blaser, M. J., & Dominguez-Bello, M. G. (2016). The Human Microbiome before Birth. Cell Host

and Microbe, 20(5), 558–560. http://doi.org/10.1016/j.chom.2016.10.014

Block, R. J., & Henry, S. M. (1961). Metabolism of the sulphur amino-acids and of sulphate in

Blattella germanica. Nature, 191(4786), 392–393. http://doi.org/10.1038/191392a0

Bright, M., & Bulgheresi, S. (2010). A complex journey: Transmission of microbial symbionts.

Nature Reviews Microbiology, 8(3), 218–230. http://doi.org/10.1038/nrmicro2262

Carrasco, P., Pérez-Cobas, A. E., van de Pol, C., Baixeras, J., Moya, A., & Latorre, A. (2014).

Succession of the gut microbiota in the cockroach Blattella germanica. International

Microbiology, 17, 99–109. http://doi.org/10.2436/20.1501.01.212

Couchman, J. R., & King, P. E. (1980). Ovariole sheath structure and its relationship with

developing embryos in a parthenogenetic viviparous aphid. Acta Zoologica, 61(3), 147–

155. http://doi.org/10.1111/j.1463-6395.1980.tb01302.x da Costa, R. R., & Poulsen, M. (2018). Mixed-mode transmission shapes termite gut community

assemblies. Trends in Microbiology, 26(7), 557–559.

http://doi.org/10.1016/j.tim.2018.04.005

Dahlen, H. G., Downe, S., Kennedy, H. P., & Foureur, M. (2014). Is society being reshaped on a

microbiological and epigenetic level by the way women give birth? Midwifery, 1–3.

http://doi.org/10.1016/j.midw.2014.07.007

170 Davis, G. K. (2012). Cyclical parthenogenesis and viviparity in aphids as evolutionary novelties.

Journal of Experimental Zoology Part B: Molecular and Developmental Evolution, 318B,

448–459. http://doi.org/10.1002/jez.b.22441 de Kort, C. A. D., & Granger, N. A. (1996). Regulation of JH titers: The relevance of degradative

enzymes and binding proteins. Archives of Insect Biochemistry and Physiology, 33(1), 1–26.

http://doi.org/10.1002/(SICI)1520-6327(1996)33:1<1::AID-ARCH1>3.0.CO;2-2

Denlinger, D. L. (1975). Insect hormones as tsetse abortifacients. Nature, 253(5490), 347–8.

http://doi.org/10.0.4.14/253347a0

Denlinger, D. L., & Ma, W. -C. (1975). Maternal nutritive secretions as possible channels for

vertical transmission of microorganisms in insects: the tsetse fly example. Annals of the

New York Academy of Sciences, 266(1), 162–165. http://doi.org/10.1111/j.1749-

6632.1975.tb35097.x

Denlinger, D. L., & Ma, W. C. (1974). Dynamics of the pregnancy cycle in the tsetse Glossina

morsitans. Journal of Insect Physiology, 20(6). http://doi.org/10.1016/0022-

1910(74)90143-7

Douglas, A. E. (2017). The B vitamin nutrition of insects: the contributions of diet, microbiome

and horizontally acquired genes. Current Opinion in Insect Science, 23, 65–69.

http://doi.org/10.1016/j.cois.2017.07.012

Duranti, S., Lugli, G. A., Mancabelli, L., Armanini, F., Turroni, F., James, K., … Ventura, M. (2017).

Maternal inheritance of bifidobacterial communities and bifidophages in infants through

vertical transmission. Microbiome, 5(1), 1–13. http://doi.org/10.1186/s40168-017-0282-6

Ejezie, G. C., & Davey, K. G. (1976). Some effects of allatectomy in the female tsetse, Glossina

171 austeni. Journal of Insect Physiology, 22, 1743–1749.

Engel, P., & Moran, N. A. (2013). The gut microbiota of insects-diversity in structure and

function. FEMS Microbiology Reviews, 37, 699–735.

Engelmann, F., & Mala, J. (2000). The interactions between juvenile hormone (JH), lipophorin,

vitellogenin, and JH esterases in two cockroach species. Insect Biochemistry and Molecular

Biology, 30(8–9), 793–803. http://doi.org/10.1016/S0965-1748(00)00051-5

Estes, A. M., Hearn, D. J., Snell-Rood, E. C., Feindler, M., Feeser, K., Abebe, T., … Moczek, A. P.

(2013). Brood ball-mediated transmission of microbiome members in the dung beetle,

Onthophagus taurus (Coleoptera: Scarabaeidae). PLoS ONE, 8(11), 1–15.

http://doi.org/10.1371/journal.pone.0079061

Evans, L. D., & Stay, B. (1989). Humoral induction of milk synthesis in the viviparous cockroach

Diploptera punctata. Invertebrate Reproduction and Development, 15(3), 171–176.

Evans, L. D., & Stay, B. (1995). Regulation of competence for milk production in Diploptera

punctata: interaction between mating, ovaries and the corpus allatum. Invertebrate

Reproduction and Development, 28(3), 161–170.

Fardini, Y., Chung, P., Dumm, R., Joshi, N., & Han, Y. W. (2010). Transmission of diverse oral

bacteria to murine placenta: Evidence for the oral microbiome as a potential source of

intrauterine infection. Infection and Immunity, 78(4), 1789–1796.

http://doi.org/10.1128/IAI.01395-09

Ferla, M. P., & Patrick, W. M. (2014). Bacterial methionine biosynthesis. Microbiology (United

Kingdom), 160, 1571–1584. http://doi.org/10.1099/mic.0.077826-0

Funkhouser, L. J., & Bordenstein, S. R. (2013). Mom knows best: the universality of maternal

172 microbial transmission. PLoS Biology, 11(8), 1–9.

http://doi.org/10.1371/journal.pbio.1001631

Gil, R., Silva, F. J., Zientz, E., Delmotte, F., Gonzalez-Candelas, F., Latorre, A., … Moya, A. (2003).

The genome sequence of Blochmannia floridanus: Comparative analysis of reduced

genomes. Proceedings of the National Academy of Sciences, 100(16), 9388–9393.

http://doi.org/10.1073/pnas.1533499100

Giorgi, F., & Nordin, J. H. (1994). Structure of yolk granules in oocytes and eggs of Blattella

germanica and their interaction with vitellophages and endosymbiotic bacteria during

granule degradation. Journal of Insect Physiology, 40(12), 1077–1092.

http://doi.org/10.1016/0022-1910(94)90061-2

Gontang, E. A., Aylward, F. O., Carlos, C., Glavina del Rio, T., Chovatia, M., Fern, A., … Kolter, R.

(2017). Major changes in microbial diversity and community composition across gut

sections of a juvenile Panchlora cockroach. PLOS ONE, 12(5), e0177189.

http://doi.org/10.1371/journal.pone.0177189

Henry, S. M., & Block, R. J. (1961). Sulfur metabolism of insects. 6. Metabolism of sulfur amino

acids and related compounds in German cockroach, Blattella germanica (L.). Contrib.

Boyce Thompson Inst., 21, 129–145.

Huang, C. Y., Sabree, Z. L., & Moran, N. a. (2012). Genome sequence of Blattabacterium sp.

strain BGIGA, endosymbiont of the Blaberus giganteus cockroach. Journal of Bacteriology,

194(16), 4450–4451. http://doi.org/10.1128/JB.00789-12

Ingram, M. J., Stay, B., & Cain, G. D. (1977). Composition of milk from the viviparous cockroach,

Diploptera punctata. Insect Biochemistry, 7(3), 257–267. http://doi.org/10.1016/0020-

173 1790(77)90023-3

Jašarević, E., Rodgers, A. B., & Bale, T. L. (2015). A novel role for maternal stress and microbial

transmission in early life programming and neurodevelopment. Neurobiology of Stress,

1(1), 81–88. http://doi.org/10.1016/j.ynstr.2014.10.005

Jiang, Z. F., Xia, F., Johnson, K. W., Brown, C. D., Bartom, E., Tuteja, J. H., … Ghanim, M. (2013).

Comparison of the genome sequences of “Candidatus portiera aleyrodidarum” primary

endosymbionts of the whitefly Bemisia tabaci B and Q biotypes. Applied and

Environmental Microbiology, 79, 1757–1759. http://doi.org/10.1128/AEM.02976-12

Kakumanu, M. L., Maritz, J. M., Carlton, J. M., & Schal, C. (2018). Overlapping community

compositions of gut and fecal microbiomes in lab-reared and field-collected German

cockroaches. Applied and Environmental Microbiology, 84(17), 1–17.

http://doi.org/10.1128/AEM.01037-18

Kambhampati, S., Alleman, A., & Park, Y. (2013). Complete genome sequence of the

endosymbiont Blattabacterium from the cockroach Nauphoeta cinerea (Blattodea:

Blaberidae). Genomics, 102, 479–483. http://doi.org/10.1016/j.ygeno.2013.09.003

Langley, P. A. (1994). Understanding tsetse flies. Onderstepoort J Vet Res, 61(4), 361–367.

Langley, P. A., & Pimley, R. W. (1986). A role for juvenile hormone and the effects of so-called

anti-juvenile hormones in Glossina morsitans. Journal of Insect Physiology, 32(8), 727–734.

http://doi.org/10.1016/0022-1910(86)90115-0

López-Sánchez, M. J., Neef, A., Patiño-Navarrete, R., Navarro, L., Jiménez, R., Latorre, A., &

Moya, A. (2008). Blattabacteria, the endosymbionts of cockroaches, have small genome

sizes and high genome copy numbers. Environmental Microbiology, 10(12), 3417–3422.

174 http://doi.org/10.1111/j.1462-2920.2008.01776.x

López-Sánchez, M. J., Neef, A., Peretó, J., Patiño-Navarrete, R., Pignatelli, M., Latorre, A., &

Moya, A. (2009). Evolutionary convergence and nitrogen metabolism in Blattabacterium

strain Bge, primary endosymbiont of the cockroach Blattella germanica. PLoS Genetics,

5(11). http://doi.org/10.1371/journal.pgen.1000721

Luo, M., Li, D., Wang, Z., Guo, W., Kang, L., & Zhou, S. (2017). Juvenile hormone differentially

regulates two Grp78 genes encoding protein chaperones required for insect fat body cell

homeostasis and vitellogenesis. Journal of Biological Chemistry, 292(21), 8823–8834.

http://doi.org/10.1074/jbc.M117.780957

Marchal, E., Hult, E. F., Huang, J., Stay, B., & Tobe, S. S. (2013). Diploptera punctata as a model

for studying the endocrinology of arthropod reproduction and development. General and

Comparative Endocrinology, 188, 85–93.

http://doi.org/http://dx.doi.org/10.1016/j.ygcen.2013.04.018

McCutcheon, J. P., & Moran, N. A. (2010). Functional convergence in reduced genomes of

bacterial symbionts spanning 200 my of evolution. Genome Biology and Evolution, 2(1),

708–718. http://doi.org/10.1093/gbe/evq055

McCutcheon, J. P., & Moran, N. A. (2012). Extreme genome reduction in symbiotic bacteria.

Nature Reviews Microbiology, 10(1), 13–26. http://doi.org/10.1038/nrmicro2670

Michalik, A., Szklarzewicz, T., Jankowska, W., & Wieczorek, K. (2014). Endosymbiotic

microorganisms of aphids (Hemiptera: Sternorrhyncha: Aphidoidea): ultrastructure,

distribution and transovarial transmission. European Journal of Entomology, 111(1), 91–

104. http://doi.org/10.14411/eje.2014.011

175 Michalkova, V., Benoit, J. B., Weiss, B. L., Attardo, G. M., & Aksoy, S. (2014). Vitamin B6

generated by obligate symbionts is critical for maintaining proline homeostasis and

fecundity in tsetse flies. Applied and Environmental Microbiology, 80(18), 5844–5853.

http://doi.org/10.1128/AEM.01150-14

Miura, T., Braendle, C., Shingleton, A., Sisk, G., Kambhampati, S., & Stern, D. L. (2003). A

comparison of parthenogenetic and sexual embryogenesis of the pea aphid Acyrthosiphon

pisum (Hemiptera: Aphidoidea). Journal of Experimental Zoology (Mol Dev Evol), 295B(1),

59–81. http://doi.org/10.1613/jair.301

Moran, N. A., & Wernegreen, J. J. (2000). Lifestyle evolution in symbiotic bacteria: insights from

genomics. Trends in Ecology & Evolution, 15(8), 321–326. http://doi.org/10.1016/S0169-

5347(00)01902-9

Morse, S. F., Bush, S. E., Patterson, B. D., Dick, C. W., Gruwell, M. E., & Dittmar, K. (2013).

Evolution, multiple acquisition, and localization of endosymbionts in bat flies (Diptera:

Hippoboscoidea: Streblidae and nycteribiidae). Applied and Environmental Microbiology,

79(9), 2952–2961. http://doi.org/10.1128/AEM.03814-12

Mueller, N. T., Bakacs, E., Combellick, J., Grigoryan, Z., & Dominguez-Bello, M. G. (2015). The

infant microbiome development: mom matters. Trends in Molecular Medicine, 21(2), 109–

117. http://doi.org/10.1016/j.molmed.2014.12.002

Mundall, E. C., Tobe, S. S., & Stay, B. (1981). Vitellogenin fluctuations in haemolymph and fat

body and dynamics of uptake into oöcytes during the reproductive cycle of Diploptera

punctata. Journal of Insect Physiology, 27(12), 821–827. http://doi.org/10.1016/0022-

1910(81)90083-4

176 Naghdi, M., Maestro, J., Belles, X., & Bandani, A. (2016). Transduction of the vitellogenic signal

of juvenile hormone by Methoprene-tolerant in the cockroach Blattella germanica ( L .) (

Dictyoptera , Blattellidae ). Arthropods, 5(4), 130–137.

Nakabachi, A., Yamashita, A., Toh, H., Ishikawa, H., Dunbar, H. E., Moran, N. A., & Hattori, M.

(2006). The 160-kilobase genome of the bacterial endosymbiont Carsonella. Science,

314(5797), 267–267. http://doi.org/10.1126/science.1134196

Nikoh, N., Hosokawa, T., Oshima, K., Hattori, M., & Fukatsu, T. (2011). Reductive evolution of

bacterial genome in insect gut environment. Genome Biology and Evolution, 3(1), 702–714.

http://doi.org/10.1093/gbe/evr064

Ogawa, K., & Miura, T. (2014). Aphid polyphenisms: Trans-generational developmental

regulation through viviparity. Frontiers in Physiology, 5, 1–11.

http://doi.org/10.3389/fphys.2014.00001

Patino-Navarrete, R., Piulachs, M.-D., Belles, X., Moya, A., Latorre, A., & Pereto, J. (2014). The

cockroach Blattella germanica obtains nitrogen from uric acid through a metabolic

pathway shared with its bacterial endosymbiont. Biology Letters, 10(7), 20140407–

20140407. http://doi.org/10.1098/rsbl.2014.0407

Paulson, C. R., & Stay, B. (1987). Humoral inhibition of the corpora allata in larvae of Diploptera

punctata: Role of the brain and ecdysteroids. Journal of Insect Physiology, 33(9), 613–622.

http://doi.org/10.1016/0022-1910(87)90129-6

Pérez-Cobas, A. E., Maiques, E., Angelova, A., Carrasco, P., Moya, A., & Latorre, A. (2015). Diet

shapes the gut microbiota of the omnivorous cockroach Blatella germanica. FEMS

Microbiology Ecology, 91(4), 1–14. http://doi.org/10.1093/femsec/fiv022

177 Perez-Muñoz, M. E., Arrieta, M. C., Ramer-Tait, A. E., & Walter, J. (2017). A critical assessment

of the “sterile womb” and “in utero colonization” hypotheses: Implications for research on

the pioneer infant microbiome. Microbiome. 5(1), 48. http://doi.org/10.1186/s40168-017-

0268-4

Peterson, B. F., & Scharf, M. E. (2016). Metatranscriptome analysis reveals bacterial symbiont

contributions to lower termite physiology and potential immune functions. BMC

Genomics, 17(1), 772. http://doi.org/10.1186/s12864-016-3126-z

Rankin, S. M., & Stay, B. (1984). The changing effect of the ovary on rates of juvenile hormone

synthesis in Diploptera punctata. General and Comparative Endocrinology, 54(3), 382–388.

http://doi.org/10.1016/0016-6480(84)90151-5

Rio, R. V. M., Symula, R. E., Wang, J., Lohs, C., Wu, Y.-N., Snyder, A. K., … Aksoy, S. (2012).

Insight into the transmission biology and species-specific functional capabilities of tsetse

(Diptera: Glossinidae) obligate symbiont Wigglesworthia. MBio, 3(1), 1–13.

http://doi.org/10.1128/mBio.00240-11

Robert, A., Strambi, A., & Strambi, C. (1986). Haemolymph ecdysteroid levels in female tsetse

fly Glossina fuscipes (Diptera) during the first reproductive cycle: a comparison between

virgin and mated females. Journal of Insect Physiology, 32(8), 665–671.

http://doi.org/10.1016/0022-1910(86)90107-1

Rosenberg, E., Sharon, G., & Zilber-Rosenberg, I. (2009). The hologenome theory of evolution

contains Lamarckian aspects within a Darwinian framework. Environmental Microbiology,

11(12), 2959–2962. http://doi.org/10.1111/j.1462-2920.2009.01995.x

Rosenberg, E., & Zilber-Rosenberg, I. (2011). Symbiosis and development: The hologenome

178 concept. Birth Defects Research Part C - Embryo Today: Reviews, 93(1), 56–66.

http://doi.org/10.1002/bdrc.20196

Rotin, D., & Tobe, S. S. (1983). The possible role of juvenile hormone esterase in the regulation

of juvenile hormone titre in the female cockroach Diploptera punctata. Canadian Journal

of Biochemistry and Cell Biology, 61(7), 811–817. http://doi.org/10.1139/o83-103

Sabree, Z. L., Huang, C. Y., Arakawa, G., Tokuda, G., Lo, N., Watanabe, H., & Moran, N. A. (2012).

Genome shrinkage and loss of nutrient-providing potential in the obligate symbiont of the

primitive termite Mastotermes darwiniensis. Applied and Environmental Microbiology,

78(1), 204–210. http://doi.org/10.1128/AEM.06540-11

Sabree, Z. L., Kambhampati, S., & Moran, N. a. (2009). Nitrogen recycling and nutritional

provisioning by Blattabacterium, the cockroach endosymbiont. Proceedings of the National

Academy of Sciences of the United States of America, 106(46), 19521–19526.

http://doi.org/10.1073/pnas.0907504106

Sabree, Z. L., & Moran, N. A. (2014). Host-specific assemblages typify gut microbial communities

of related insect species. SpringerPlus, 3(1), 138. http://doi.org/10.1186/2193-1801-3-138

Salem, H., Florez, L., Gerardo, N., & Kaltenpoth, M. (2015). An out-of-body experience: the

extracellular dimension for the transmission of mutualistic bacteria in insects. Proceedings

of the Royal Society B: Biological Sciences, 282(1804), 20142957–20142957.

http://doi.org/10.1098/rspb.2014.2957

Santos-Garcia, D., Latorre, A., Moya, A., Gibbs, G., Hartung, V., Dettner, K., … Silva, F. J. (2014).

Small but powerful, the primary endosymbiont of moss bugs, Candidatus evansia muelleri,

holds a reduced genome with large biosynthetic capabilities. Genome Biology and

179 Evolution, 6(7), 1875–1893. http://doi.org/10.1093/gbe/evu149

Schauer, C., Thompson, C., & Brune, A. (2014). Pyrotag sequencing of the gut microbiota of the

cockroach Shelfordella lateralis reveals a highly dynamic core but only limited effects of

diet on community structure. PLoS ONE, 9(1), e85861.

http://doi.org/10.1371/journal.pone.0085861

Schwab, D. B., Riggs, H. E., Newton, I. L. G., & Moczek, A. P. (2016). Developmental and

ecological benefits of the maternally transmitted microbiota in a dung beetle. The

American Naturalist, 188(6), 679–692. http://doi.org/10.1086/688926

Shigenobu, S., Watanabe, H., Hattori, M., Sakaki, Y., & Ishikawa, H. (2000). Genome sequence of

the endocellular bacterial symbiont of aphids Buchnera sp. APS. Nature, 407(6800), 81–86.

http://doi.org/10.1038/35024074

Shukla, S. P., Vogel, H., Heckel, D. G., Vilcinskas, A., & Kaltenpoth, M. (2018). Burying beetles

regulate the microbiome of carcasses and use it to transmit a core microbiota to their

offspring. Molecular Ecology, 27(8), 1980–1991. http://doi.org/10.1111/mec.14269

Stay, B., & Coop, A. C. (1974). “Milk” secretion for embryogenesis in a viviparous cockroach.

Tissue and Cell, 6(4), 669–693. http://doi.org/10.1016/0040-8166(74)90009-3

Stay, B., & Lin, H. L. (1981). The inhibition of milk synthesis by juvenile hormone in the

viviparous cockroach, Diploptera punctata. Journal of Insect Physiology, 27(8), 551–557.

http://doi.org/10.1016/0022-1910(81)90043-3

Stay, B., Ostedgaard, L. S., Tobe, S. S., Strambi, A., & Spaziani, E. (1984). Ovarian and

haemolymph titres of ecdysteroid during the gonadotrophic cycle in Diploptera punctata.

Journal of Insect Physiology, 30(8), 643–651. http://doi.org/10.1016/0022-1910(84)90049-

180 0

Stoltzman, C. A., & Stay, B. (1997). Gonadotrophic and morphogenetic effects of a juvenile

hormone analog treatment and ovary presence on last instar male and female Diploptera

punctata (Blattaria: Blaberidae). European Journal of Entomology, 94(3), 335–348.

Stoltzman, C. A., Stocker, C., Borst, D., & Stay, B. (2000). Stage-specific production and release

of juvenile hormone esterase from the ovary of Diploptera punctata. Journal of Insect

Physiology, 46(5), 771–782. http://doi.org/10.1016/S0022-1910(99)00166-3

Terr Wee, J., & Stay, B. (1987). Termination of pregnancy and milk synthesis in the viviparous

cockroach, Diploptera Punctata: a role of juvenile hormone. International Journal of

Invertebrate Reproduction and Development, 11(1), 59–75.

Thangavelu, B., & Viola, R. (2016). Alternative microbial routes to homoserine activation and

sulfuration: critical steps in methionine biosynthesis. International Journal of Biochemistry

Research & Review, 11(3), 1–22. http://doi.org/10.9734/ijbcrr/2016/25074

Tinker, K. A., & Ottesen, E. A. (2016). The core gut microbiome of the American cockroach,

Periplaneta americana, is stable and resilient to dietary shifts. Applied and Environmental

Microbiology, 82(22), 6603–6610. http://doi.org/10.1128/AEM.01837-16

Tobe, S. S., Davey, K. G., & Huebner, E. (1973). Nutrient transfer during the reproductive cycle in

Glossina austeni newst.: Histology and histochemistry of the milk gland, fat body, and

oenocytes. Tissue and Cell, 5(4), 633–650. http://doi.org/10.1016/S0040-8166(73)80050-3

Tobe, S. S., & Langley, P. A. (1978). Reproductive physiology of Glossina. Annual Review of

Entomology, 23(1), 283–307. http://doi.org/10.1146/annurev.en.23.010178.001435

Tobe, S. S., Ruegg, R. P., Stay, B., Baker, F. C., Miller, C. A., & Schooley, D. A. (1985). Juvenile

181 hormone titre and regulation in the cockroach Diploptera punctata. Experientia, 41(8),

1028–34.

Tokuda, G., Elbourne, L. D. H., Kinjo, Y., Saitoh, S., Sabree, Z. L., Hojo, M., … Lo, N. (2013).

Maintenance of essential amino acid synthesis pathways in the Blattabacterium cuenoti

symbiont of a wood-feeding cockroach. Biology Letters, 9, 20121153.

http://doi.org/10.1098/rsbl.2012.1153

Tufail, M., Nagaba, Y., Elgendy, A. M., & Takeda, M. (2014). Regulation of vitellogenin genes in

insects. Entomological Science, 17(3), 269–282. http://doi.org/10.1111/ens.12086

Walker, R. W., Clemente, J. C., Peter, I., & Loos, R. J. F. (2017). The prenatal gut microbiome: are

we colonized with bacteria in utero? Pediatric Obesity, 12, 3–17.

http://doi.org/10.1111/ijpo.12217

Wang, J., Weiss, B. L., & Aksoy, S. (2013). Tsetse fly microbiota: form and function. Frontiers in

Cellular and Infection Microbiology, 3, 1–6. http://doi.org/10.3389/fcimb.2013.00069

Wernegreen, J. J. (2002). Genome evolution in bacterial endosymbionts of insects. Nature

Reviews Genetics, 3(11), 850–861. http://doi.org/10.1038/nrg931

Williford, A., Stay, B., & Bhattacharya, D. (2004). Evolution of a novel function: Nutritive milk in

the viviparous cockroach, Diploptera punctata. Evolution and Development, 6, 67–77.

http://doi.org/10.1111/j.1525-142X.2004.04012.x

Wilson, A. C. C., Ashton, P. D., Calevro, F., Charles, H., Colella, S., Febvay, G., … Douglas, A. E.

(2010). Genomic insight into the amino acid relations of the pea aphid, Acyrthosiphon

pisum, with its symbiotic bacterium Buchnera aphidicola. Insect Molecular Biology, 19,

249–258. http://doi.org/10.1111/j.1365-2583.2009.00942.x

182 Wolf, Y. I., & Koonin, E. V. (2013). Genome reduction as the dominant mode of evolution.

BioEssays, 35(9), 829–837. http://doi.org/10.1002/bies.201300037

183

Appendix

Non-thesis publication list

184 Attardo, G. M. et al. (2019). The Glossina genome cluster: Comparative Genomic Analysis of the

Vectors of African Trypanosomes. BioRxiv, 531749. http://doi.org/10.1101/531749

Genome Biology. In review

Background

Tsetse flies (Glossina sp.) are the sole vectors of human and animal trypanosomiasis throughout sub-Saharan Africa. Tsetse are distinguished from other Diptera by unique adaptations, including lactation and the birthing of live young (obligate viviparity), a vertebrate blood specific diet by both and obligate bacterial symbiosis. This work describes comparative analysis of six Glossina genomes representing three sub-genera: Morsitans (G. morsitans morsitans (G.m. morsitans), G. pallidipes, G. austeni), Palpalis (G. palpalis, G. fuscipes) and

Fusca (G. brevipalpis) which represent different habitats, host preferences and vectorial capacity.

Results

Genomic analyses validate established evolutionary relationships and sub-genera. Syntenic analysis of Glossina relative to Drosophila melanogaster shows reduced structural conservation across the sex-linked X chromosome. Sex linked scaffolds show increased rates of female specific gene expression and lower evolutionary rates relative to autosome associated genes.

Tsetse specific genes are enriched in protease, odorant binding and helicase activities. Lactation associated genes are conserved across all Glossina species while male seminal proteins are rapidly evolving. Olfactory and gustatory genes are reduced across the genus relative to other characterized insects. Vision associated Rhodopsin genes show conservation of motion

185 detection/tracking functions and significant variance in the Rhodopsin detecting colors in the blue wavelength ranges.

Conclusions

Expanded genomic discoveries reveal the genetics underlying Glossina biology and provide a rich body of knowledge for basic science and disease control. They also provide insight into the evolutionary biology underlying novel adaptations and are relevant to applied aspects of vector control such as trap design and discovery of novel pest and disease control strategies.

186 Holmes, C., Jennings, E. C., et al. (2019). Comprehensive microbiome analysis of the Antarctic

mite, Alaskozetes antarcticus, reveals a highly diverse but common bacterial community

shared across life stages and sexes. Polar Biology. In review

Abstract

The Antarctic mite (Alaskozetes antarcticus) is widely distributed on sub-Antarctic islands and throughout the Antarctic Peninsula, making it one of the most abundant terrestrial arthropods in the region. Despite the impressive ability of A. antarcticus to thrive in harsh Antarctic conditions, little is known about the biology of this species. In this study, we performed 16S rRNA gene sequencing to examine the microbiome of the final immature instar (tritonymph) and both male and female adults. The microbiome included a limited number of microbial classes and genera, with few differences noted among the different stages. Five classes —

Actinobacteria, , Sphingobacteriia, Gammaproteobacteria, and

Betaproteobacteria — comprised ~82.0% of the microbial composition, and five (identified) genera — Dermacoccus, Pedobacter, Chryseobacterium, Pseudomonas, and Flavobacterium — accounted for ~68.0% of the total composition. The core microbiome present in all surveyed A. antarcticus was dominated by the families Flavobacteriaceae, Comamonadaceae,

Sphingobacteriaceae, Chitinophagaceae and Cytophagaceae, but the majority of the core consisted of operational taxonomic units of low abundance. This comprehensive analysis reveals a diverse microbiome among individuals of different stages, with overlap likely due to their shared habitat and common feeding preferences as herbivores and detritivores. The microbiome of the Antarctic mite shows considerably more diversity than observed in mite species from lower latitudes.

187

Martinsen, E. O., Peyton, J. et al. (2019) Genome and ontogenetic-based transcriptomic

analyses of the flesh fly, Sarcophaga bullata, provides a platform for probing

development and host- interactions. G3 Genes|Genomes|Genetics. In press

Abstract

The flesh fly, Sarcophaga bullata, is a widely-used model for probing the physiology of insect , development, stress tolerance, neurobiology, and host-parasitoid interactions. Flies in this taxon are implicated in myiasis and, as carrion feeders, are important in forensic studies.

Here we present the genome along with developmental- and reproduction-based RNA-seq analyses. We predicted 15,768 protein coding genes, identified orthology in relation to closely related flies, and established sex and developmental-specific gene sets based on our RNA-seq analyses. Genomic sequences, predicted genes, and sequencing data sets have been deposited at the National Center for Biotechnology Information. Our results provide groundwork for genomic-based studies on this flesh fly that can expand its utility as a fly model.

188 Oeyen, J. P. et al. (2018) Draft genomes of two sawflies reveal evolutionary acquisitions that

fostered the megaradiation of parasitoid and eusocial Hymenoptera. Current Biology. In

review

Summary

The evolution of parasitoidism, a life style in which a larva develops by feeding upon and thereby killing a single host specimen, is considered the most important factor having fostered the diversification of Hymenoptera (sawflies, , ants, and ) [1,2]. The transition from an ancestral phytophagous life style, retained in most sawflies ("Symphyta"), to parasitoidism occurred only once in Hymenoptera: in the last common ancestor of the parasitoid sawfly family Orussidae and the -waisted Hymenoptera (Apocrita) [1]. However, parasitoidism alone is not a reliable predictor of species richness, as Orussidae and Apocrita differ dramatically in their species diversity [3]. Speciation in Apocrita was thus likely promoted by the evolutionary acquisition of additional traits. Here we present draft genomes of the phytophagous sawfly Athalia rosae and the parasitoid sawfly Orussus abietinus to shed light on these acquired traits in a comparative genomic analysis. Evolution of parasitoidism in

Hymenoptera was accompanied by an increase of nitrogen-rich storage proteins, possibly having facilitated diet change, but seemingly had no major impact on the metabolism-related gene repertoire. Intriguingly, parasitoid and eusocial Apocrita evolved a significantly larger array of odorant receptors than phytophagous and parasitoid sawflies, which may have promoted diversification of the two former groups. Finally, our analysis revealed that the gene repertoire of the last common ancestor of Hymenoptera was wider than previously thought and encompassed genes thought to be absent in Hymenoptera (e.g., genes for CO2 and light

189 detection) and genomic features previously only known from eusocial Apocrita (e.g., small immune gene repertoire, low transposable element content and activity).

190 Panfilio, K. A. et al. (2019) Molecular evolutionary trends and feeding ecology diversification in

the Hemiptera, anchored by the milkweed bug genome. Genome Biology. In Press

Background

The Hemiptera (aphids, , and true bugs) are a key insect order, with high diversity for feeding ecology and excellent experimental tractability for molecular genetics. Building upon recent sequencing of hemipteran pests such as phloem-feeding aphids and blood-feeding bed bugs, we present the genome sequence and comparative analyses centered on the milkweed bug Oncopeltus fasciatus, a seed feeder of the family Lygaeidae.

Results: The 926-Mb Oncopeltus genome is well represented by the current assembly and official gene set. We use our genomic and RNA-seq data not only to characterize the protein- coding gene repertoire and perform isoform-specific RNAi, but also to elucidate the patterns of molecular evolution and physiology. We find ongoing, lineage-specific expansion and diversification of repressive C2H2 zinc finger proteins. The discovery of intron gain and turnover specific to the Hemiptera also prompted the evaluation of lineage and genome size as predictors of gene structure evolution. Furthermore, we identify enzymatic gains and losses that correlate with feeding biology, particularly for the reductions associated with derived, fluid nutrition feeding.

Conclusions

With the milkweed bug, we now have a critical mass of sequenced species for hemimetabolous insect order and close outgroup to the Holometabola, substantially improving the diversity of insect genomics. We thereby define the commonalities among the Hemiptera and delve into how hemipteran genomes reflect distinct feeding ecologies. Given Oncopeltus’s strength as an

191 experimental model, these new sequence resources bolster the foundation for molecular research and highlight technical considerations for the analysis of medium-sized invertebrate genomes.

192 Armisen, D. et al. (2018) The genome of the water strider Gerris buenoi reveals expansions of

gene repertoires associated with adaptations to life on the water. BMC Genomics.

19:832 DOI: 10.1186/s12864-018-5163-2

Background

Having conquered water surfaces worldwide, the semi-aquatic bugs occupy ponds, streams, lakes, mangroves, and even open oceans. The diversity of this group has inspired a range of scientific studies from ecology and evolution to developmental genetics and hydrodynamics of fluid locomotion. However, the lack of a representative water strider genome hinders our ability to more thoroughly investigate the molecular mechanisms underlying the processes of adaptation and diversification within this group.

Results

Here we report the sequencing and manual annotation of the Gerris buenoi (G. buenoi) genome; the first water strider genome to be sequenced thus far. The size of the G. buenoi genome is approximately 1,000 Mb, and this sequencing effort has recovered 20,949 predicted protein-coding genes. Manual annotation uncovered a number of local (tandem and proximal) gene duplications and expansions of gene families known for their importance in a variety of processes associated with morphological and physiological adaptations to a water surface lifestyle. These expansions may affect key processes associated with growth, vision, desiccation resistance, detoxification, olfaction and epigenetic regulation. Strikingly, the G. buenoi genome contains three insulin receptors, suggesting key changes in the rewiring and function of the insulin pathway. Other genomic changes affecting with opsin genes may be

193 associated with wavelength sensitivity shifts in opsins, which is likely to be key in facilitating specific adaptations in vision for diverse water habitats.

Conclusions

Our findings suggest that local gene duplications might have played an important role during the evolution of water striders. Along with these findings, the sequencing of the G. buenoi genome now provides us the opportunity to pursue exciting research opportunities to further understand the genomic underpinnings of traits associated with the extreme body plan and life history of water striders.

194 Meibers, H. E. et al. (2018) Sex-specific RNA-seq analysis of the Antarctic mite, Alaskozetes

antarcticus, reveal transcriptional shifts underlying orabatid mite reproduction. Polar

Biology. DOI: 10.1007/s00300-018-2427-x

Abstract

The oribatid mite Alaskozetes antarcticus, one of the most abundant terrestrial invertebrates in

Antarctica, survives extreme temperature fluctuation and desiccation, and thrives in the short growing season characteristic of this polar environment. Several aspects of the mite’s ecology and physiology are well studied, but little is known about its reproduction. In this study, we utilize sex- and development-specific next-generation RNA-sequencing (RNA-seq) analyses to identify differentially regulated transcripts underlying reproduction of A. antarcticus. Pairwise comparisons between males, females, and tritonymphs revealed more than 4000 enriched transcripts based on different transcriptional levels among sexes and developmental stages.

More than 500 of these enriched transcripts were differentially upregulated over 1000-fold.

Many of the highly enriched and sex-specific transcripts were previously uncharacterized or have no known orthology. Of the transcripts identified, gene ontology-based analyses linked the transcriptional distinctions to differences in reproduction, chemosensation, and stress response. Our comparative approach allowed us to determine sexually dimorphic transcript expression in A. antarcticus. We anticipate that this study will provide a baseline to better understand the mechanisms that underlie reproduction in both polar and non-polar oribatid mites.

195 Hagan, R. H. et al. (2018) Dehydration prompts increased activity and blood feeding by

mosquitoes. Scientific Reports. 8:16804 DOI: 10.1038/s41598-018-24893-z

Abstract

Current insights into the mosquito dehydration response rely on studies that examine specific responses but ultimately fail to provide an encompassing view of mosquito biology.

Here, we examined underlying changes in the biology of mosquitoes associated with dehydration. Specifically, we show that dehydration increases blood feeding in the northern house mosquito, Culex pipiens, which was the result of both higher activity and a greater tendency to land on a host. Similar observations were noted for Aedes aegypti and Anopheles quadrimaculatus. RNA-seq and metabolome analyses in C. pipiens following dehydration revealed that factors associated with carbohydrate metabolism are altered, specifically the breakdown of trehalose. Suppression of trehalose breakdown in C. pipiens by RNA interference reduced phenotypes associated with lower hydration levels. Lastly, mesocosm studies for C. pipiens confirmed that dehydrated mosquitoes were more likely to host feed under ecologically relevant conditions. Disease modeling indicates dehydration bouts will likely enhance viral transmission. This dehydration-induced increase in blood feeding is therefore likely to occur regularly and intensify during periods when availability of water is low.

196 Schoville, S. D. et al. (2018) A model species for agricultural pest genomics: the genome of the

Colorado potato beetle, Leptinotarsa decemlineata (Coleoptera: Chrysomelidae).

Scientific Reports. 8: 1931 DOI:10.1038/s41598-018-20154-1

Abstract

The Colorado potato beetle is one of the most challenging agricultural pests to manage. It has shown a spectacular ability to adapt to a variety of solanaceaeous plants and variable climates during its global invasion, and, notably, to rapidly evolve insecticide resistance. To examine evidence of rapid evolutionary change, and to understand the genetic basis of herbivory and insecticide resistance, we tested for structural and functional genomic changes relative to other arthropod species using genome sequencing, transcriptomics, and community annotation. Two factors that might facilitate rapid evolutionary change include transposable elements, which comprise at least 17% of the genome and are rapidly evolving compared to other Coleoptera, and high levels of nucleotide diversity in rapidly growing pest populations. Adaptations to plant feeding are evident in gene expansions and differential expression of digestive enzymes in gut tissues, as well as expansions of gustatory receptors for bitter tasting. Surprisingly, the suite of genes involved in insecticide resistance is similar to other beetles. Finally, duplications in the

RNAi pathway might explain why Leptinotarsa decemlineata has high sensitivity to dsRNA.

The L. decemlineata genome provides opportunities to investigate a broad range of phenotypes and to develop sustainable methods to control this widely successful pest.

197 Benoit, J.B., et al. (2016) Unique features of the bed bug, a global human ectoparasite,

identified through genome sequencing. Nature Communications. 7:10165 DOI:

10.1038/ncomms10165

Abstract

The bed bug, Cimex lectularius, has re-established itself as a ubiquitous human ectoparasite throughout much of the world during the past two decades. This global resurgence is likely linked to increased international travel and commerce in addition to widespread insecticide resistance. Analyses of the C. lectularius sequenced genome (650 Mb) and 14,220 predicted protein-coding genes provide a comprehensive representation of genes that are linked to traumatic , a reduced chemosensory repertoire of genes related to obligate hematophagy, host–symbiont interactions, and several mechanisms of insecticide resistance. In addition, we document the presence of multiple putative lateral gene transfer events. Genome sequencing and annotation establish a solid foundation for future research on mechanisms of insecticide resistance, human–bed bug and symbiont–bed bug associations, and unique features of bed bug biology that contribute to the unprecedented success of C. lectularius as a human ectoparasite.

198 Benoit, J.B., Zhang, Q., Jennings, E.C., Rosendale, A.J. & Denlinger, D.L. (2015) Dramatic

suppression of net transpiration conserves water resources during pupal diapause in the

corn earworm, Helicoverpa zea. Physiological Entomology. 40:4 336-342

DOI: 10.1111/phen.12119

Abstract

One critical aspect of an insect's ability to overwinter successfully is the effective management of its water resources. Maintenance of adequate water levels during winter is challenging because of the prevailing low relative humidity at that time of year and the short supply of environmental water that is not in the form of ice. These issues are further exacerbated for insects overwintering as pupae, comprising an immobile stage that is unable to move to new microhabitats if conditions deteriorate. The present study compares the water balance attributes of diapausing and nondiapausing pupae of the corn earworm Helicoverpa zea Boddie, aiming to identify the mechanisms used by diapausing pupae to maintain water balance during winter. Diapausing pupae are 10% larger than nondiapausing individuals. Water loss rates for nondiapausing pupae are low (0.21 mg h−1) and are suppressed (0.01 mg h−1) in diapausing pupae. Cuticular lipids, which serve to waterproof the cuticle and thus suppress cuticular water loss, are more than two-fold more abundant on the surface of diapausing pupae, and oxygen consumption rates during diapause drop to almost one-third the rate observed in nondiapausing pupae. Water gain can be accomplished only when atmospheric water content is near saturation or during contact with free water. At moderate relative humidities (20–40%), water loss rates are very low for diapausing pupae, suggesting that these pupae have robust mechanisms for combating water loss. The exceptional ability of H. zea to suppress

199 water loss during diapause is probably a result of the combined effects of increased size, more abundant cuticular lipids and decreased metabolic rates.

200