Juvenile hormone biosynthesis in the cockroach, …...the JH biosynthetic pathway in CA of day 6...

195
Juvenile hormone biosynthesis in the cockroach, Diploptera punctata: the characterization of the biosynthetic pathway and the regulatory roles of allatostatins and NMDA receptor by Juan Huang A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Cell and Systems Biology University of Toronto © Copyright by Juan Huang (2015)

Transcript of Juvenile hormone biosynthesis in the cockroach, …...the JH biosynthetic pathway in CA of day 6...

Juvenile hormone biosynthesis in the cockroach, Diploptera punctata: the characterization of the

biosynthetic pathway and the regulatory roles of allatostatins and NMDA receptor

by

Juan Huang

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Department of Cell and Systems Biology University of Toronto

© Copyright by Juan Huang (2015)

ii

Juvenile hormone biosynthesis in the cockroach, Diploptera

punctata: the characterization of the biosynthetic pathway and the

regulatory roles of allatostatins and NMDA receptor

Juan Huang

Doctor of Philosophy (2015), Department of Cell and Systems Biology, University of Toronto

Abstract

The juvenile hormones (JH) play essential roles in regulating growth, development,

metamorphosis, ageing, caste differentiation and reproduction in insects. Diploptera punctata,

the only truly viviparous cockroach is a well-known model system in the study of JH

biosynthesis and its regulation. The physiology of this animal is characterized by very stable and

high rates of JH biosynthesis and precise and predictable reproductive events that correlate well

with rates of JH production. Many studies have been performed on D. punctata to determine the

function of JH. However, the pathway of JH biosynthesis has not been identified. In addition,

although many factors are known to regulate JH biosynthesis, the exact mechanisms remain

unclear. The aim of my research was to elucidate the JH biosynthetic pathway in D. punctata and

study the mechanisms by which allatostatin (AST) and N-methyl-D-aspartate (NMDA) receptor

regulate JH production. I have (1) identified genes in the JH biosynthetic pathway, and

determined their roles in JH biosynthesis; (2) investigated the mode of action of AST by

determining the signaling pathway of AstR and the target of AST action; (3) determined the role

of the NMDA receptor in JH biosynthesis using RNA interference and treatment with an NMDA

receptor antagonist. To validate the application value of my research, AST analogs with high JH

inhibitory activity were designed and their activities on JH biosynthesis were measured by in

vitro and in vivo bioassays.

iii

Acknowledgements

It has been four and a half years since I began my studies at the University of Toronto. I still

remember the first day in Toronto. I was excited and nervous. Starting from the first day I went

to the lab, I was surrounded by friendly faces and help. My supervisor Stephen S. Tobe and his

wife Martha Tobe helped me to get used to the culture on another continent. Jinrui Zhang helped

me to find a place to live. And all the documents were handled in one day with the help of

Ekaterina F. Hult and Jane Linley. They successfully took away all my nervousness and made

me feel that my life in Toronto would be exciting.

Before I came to Canada, I barely spoke English. Language has been troublesome for me. I was

worried that no one would like to talk to me because of my poor English. But again, people in

my lab, Ekaterina F. Hult, Jinrui Zhang, Koichiro J. Yagi, Shirley H. Tiu, Elisabeth Marchal and

Ilke van Hazel helped me to get over my trouble. They have been super patient, and always

encouraged me to speak. My English was greatly improved with their help.

In regard to research, there are many people to thank. First, my supervisor Stephen S. Tobe; he

has an interesting way to train his students. To start my project, Steve asked me to read papers

and find a project I am interested in, instead of assigning me one. It was difficult in the

beginning, but when I look back now, I find it was great training. Now I am able to start and

complete a project independently, thanks to him. In addition, Steve always provided great

suggestions for my projects, and he always encouraged me to try new things. He taught me:

never to be afraid of failure, because that is part of PhD training. I would never have finished my

PhD without his support and supervision.

iv

And I would also like to thank Ekaterina F. Hult and Elisabeth Marchal for their great help

during my PhD study. I was a chemist before I came to Toronto. I knew very little biology.

Ekaterina F. Hult not only helped me to start my project, but also taught me many techniques in

biology. Most importantly, she has an ability to make me feel good about myself. She

encouraged me many times when I was frustrated with my failed experiments. Elisabeth Marchal

is the most kind and sweet person I have ever met. She is always very thoughtful and nice, and

she always came up with great ideas. We were a great team and worked on several projects

together. I learned a lot from her, not only her knowledge in biology, but also her attitude to

research and life.

I would also like to acknowledge my other collaborators who contributed to my research

projects: Jinrui Zhang, who taught me Radiochemical assay, and cockroach dissection; Koichiro

J. Yagi, who helped me set up HPLC and gave suggestions for my projects; Prof Barbara Stay,

who taught me cockroach dissection and provided suggestions for my projects; Prof Jozef

Vanden Broeck, who provided the equipment and reagents to run the AstR functional assays;

Sven Zels, who taught me the technique of receptor functional assays; Ilke van Hazel, who

helped me with the cell culture and the expression of NMDA receptors.

I would like to express my gratitude to my committee members: Belinda Chang, Ian Orchard,

William G. Bendena, David Lovejoy and Les Buck. I thank them for their great suggestions for

my projects and my thesis.

I would like to thank all my friends in Toronto and in China. They brought so much happiness

and joy to my life, and made my life in Toronto pleasant and colorful. Lastly, I would like to

thank my parents and my brother for their support in the past four and a half years.

v

Table of Contents

Abstract ……………………………………………………………………………………….ii

Acknowledgement …………………………………………………………………………….iii

Table of Contents……………………………………………………………………………...v

Table of Figures………………………………………………………………………………..vii

Table of Tables…………………………………………………………………………………ix

Abbreviations…………………………………………………………………………………..x

Chapter 1: General Introduction

1.1 Juvenile hormones………………………………………………………………… 1

1.2 JH biosynthetic pathway…………………………………………………………...7

1.3 JH signaling pathway……………………………………………………………... 26

1.4 Diploptera punctata………………………………………………………………. 32

1.5 Regulation of JH titre……………………………………………………………... 38

1.6 Rational and objectives of my study……………………………………………... 49

1.7 References………………………………………………………………………….. 52

Chapter 2: Characterization of the Juvenile Hormone pathway in the viviparous

cockroach, Diploptera punctata

2.1 Summary…………………………………………………………………………... 69

2.2 Introduction……………………………………………………………………….. 69

2.3 Materials and Methods…………………………………………………………… 73

2.4 Results…………………………………………………………………………….... 80

2.5 Discussion………………………………………………………………………….. 90

2.6 Supplementary data………………………………………………………………. 96

2.7 References………………………………………………………………………… 100

Chapter 3: Mode of action of allatostatins in the regulation of juvenile hormone

biosynthesis in the cockroach, Diploptera punctata

3.1 Summary………………………………………………………………………….. 105

vi

3.2 Introduction………………………………………………………………………. 105

3.3 Materials and Methods…………………………………………………………... 108

3.4 Results……………………………………………………………………………...114

3.5 Discussion………………………………………………………………………….126

3.6 Supplementary data……………………………………………………………….131

3.7 References………………………………………………………………………… 134

Chapter 4: Identification and characterization of the NMDA receptor and its role in

regulating reproduction in the cockroach, Diploptera punctata

4.1 Summary………………………………………………………………………….. 138

4.2 Introduction………………………………………………………………………. 138

4.3 Materials and Methods…………………………………………………………... 141

4.4 Results……………………………………………………………………………...144

4.5 Discussion………………………………………………………………………….155

4.6 Supplementary data……………………………………………………………….160

4.7 References………………………………………………………………………… 166

Chapter 5: General discussion

5.1 Function of JH in reproduction...……………………………………………….. 169

5.2 Evolution of the JH biosynthetic pathway...……………………………………. 170

5.3 Regulation of JH biosynthesis….………………………………………………... 172

5.4 The value of my study in insect control…..……………………………………...175

5.5 Future perspective……………………………………………………………….. 176

5.6 References………………………………………………………………………….177

Chapter 6: Appendices……………………………………………………………………… 180

vii

Table of Figures

Figure 1.1 Structures of the JH homologues in insects 2

Figure 1.2 The JH biosynthetic pathway 8

Figure 1.3 Dynamics of JH metabolism throughout the life cycle of D. punctata 35

Figure 2.1 Scheme of JH biosynthetic pathway 71

Figure 2.2 Tissue specific expression of genes encoding JH biosynthetic enzymes 82

Figure 2.3 Developmental expression of genes encoding JH biosynthetic enzymes

during the first gonadotrophic

cycle of D. punctata

84

Figure 2.4 The effect of JH precursors on JH biosynthesis by CA from mated female

D. punctata

86

Figure 2.5 Efficiency of HMGR-JHAMT RNAi-mediated knockdown and the effect

of silencing on the transcription of the other genes encoding enzymes in the

JH biosynthetic pathway in day 4 mated female D. punctata.

86

Figure 2.6 JH regulates ovarian development 88

Figure 2.7 Transverse sections of the basal oocytes from day 4 control and HMGR-

JHAMT

dsRNA-treated animals.

89

Figure 3.1 Relative expression levels of Dippu-AstR and Dippu-AST mRNA in tissues

of day 4 males and mated females.

115

Figure 3.2 Relative expression levels of Dippu-AstR and Dippu-AST mRNA during the

first

gonadotrophic cycle

116

Figure 3.3 The effect of Dippu-AST dsRNA on JH biosynthesis by the CA. 120

Figure 3.4 Dose-response curves for ASTs in CHO-WTA11 cells expressing Dippu-

AstR.

121

Figure 3.5 Dose-response curves for the bioluminescence response induced in

CHOPAM28 and HEK293 cells expressing Dippu-AstR

122

Figure 3.6 The effect of Dippu-AstR dsRNA on JH biosynthesis by the CA and on the

expression of genes encoding enzymes in the JH biosynthetic pathway of

D. punctata

123

Figure 3.7 The effect of AST on the expression levels of genes encoding enzymes in

the JH biosynthetic pathway in CA of day 6 mated female D. punctata

124

Figure 3.8 JH precursors rescue the AST-induced JH inhibition. 124

Figure 4.1 Amino acid sequence alignment of the two Diploptera NR1 subunit

(DpNR1A, DpNR1B), and homologous receptors from D. melanogaster

and T. castaneum

145

Figure 4.2 Phylogram depicting the relationship between the NR1 subunits from

Diploptera and orthologues of this receptor from other insects.

146

Figure 4.3 Molecular characterization of DpNR2. 149

viii

Figure 4.4 Graphic representation of the relative tissue distribution of (A) DpNR1A

transcript levels, (B) DpNR1B transcript levels and (C) DpNR2 transcript

levels in tissues of day 4 adult male and mated female D. punctata.

150

Figure 4.5 Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in brains of

mated female D. punctata from day 0-day 7 after ecdysis.

151

Figure 4.6 Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in CA of of

mated female D. punctata from day 0-day 7 after ecdysis

151

Figure 4.7 Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in testes of

differentages of male D. punctata.

152

Figure 4.8 The effect of DpNR2 dsRNA treatment on JH biosynthesis and basal

oocyte growth, and the interactions among these genes in mated female D.

punctata.

152

Figure 4.9 In vivo effect of MK-801 on JH biosynthesis, basal oocyte growth and

relative Vg mRNA levels.

154

Figure 6.1 The effect of topical application of K15 and W206 on JH biosynthesis and

oocyte growth

183

ix

Table of Tables

Table 2.1 q-RT-PCR primer sequences and reaction efficiencies and correlation

coefficients in the q-RT-PCR assay

77

Table 2.2 Primers for dsRNA construction 79

Table 3.1 Potency of Dippu-ASTs a: activation of AstR in CHO-WTA11 cells

(EC50) or inhibitory effect on JH release (IC50)

125

Table 6.1 Structure of AST analogs 181

Table 6.2 Potency of Dippu-AST analogs a: inhibitory effect on JH release (IC50)

and activation of AstR in CHO-WTA11 cells (EC50)

182

x

Abbreviations

20E 20-hydroxyecdysone

AC adenylate cyclase

AST allatostatin

AstR Allatostatin receptor

AT Allatotropin

CA corpora allata

CHO Chinese hamster ovary

CRE cAMP responsive element

DMMP diphosphomevalonate

DMPP dimethylallyl pyrophosphate

FALD Farnesal dehydrogenase

FOLD Farnesol dehydrogenase

FPP farnesyl diphosphate

FPPP Farnesyl diphosphate pyrophosphatase

FPPS Farnesyl diphosphate synthase

GnRH gonadotropin releasing hormone

GPP Geranyl pyrophosphate

HEK human embryonic kidney

HMGR 3-hydroxy-3-methylglutaryl-CoA reductase

HMGS 3-hydroxy-3-methylglutaryl-CoA synthase

IPP isopentenyl pyrophosphate

IPPI Isopentenyl diphosphate isomerase

JH juvenile hormones

JHMAT Juvenile hormone acid O-methyltransferase

Kr-h1 Krüpel-homologues 1

LH luteinizing hormone

MA mevalonic acid

Met Methoprene-tolerant

MF methyl farnesoate

MK Mevalonate kinase

NMDAR N-methyl-D-aspartate receptor

PMK Phosphomevalonate kinase

PPMD Diphosphomevalonate decarboxylase

PTX pertussis toxin

RXR retinoid X receptor

Thiol Acetoacetyl-CoA thiolase

USP Ultraspiracle

Vg vitellogenin

Vn vitellin

1

Chapter 1

General Introduction

1 Juvenile hormones

The juvenile hormones (JH), a family of acyclic sesquiterpenoids, play essential roles in

regulating growth, development, metamorphosis, aging, caste differentiation and reproduction in

insects. This family of hormones has been extensively studied because of its central role in insect

development and reproduction and their potential value in pest control. This section will review

the current knowledge of JHs, including the JH homologues, enzymes in the JH biosynthetic

pathway and signal pathways of JH in insects.

1.1 JH homologues

JHs are synthesized and secreted by specialized, paired endocrine glands, the corpora allata (CA).

As early as 1934, Wigglesworth pointed out that insect metamorphosis was controlled by a

hormone produced by CA, a gland near the insect brain (Wigglesworth, 1934). In 1956, a highly

active extract, which produced anomalies in metamorphosis, was obtained from Cecropia Moth

Hyalophora cecropia (Williams, 1956). The structure of the first JH homologue was later

elucidated by Röller et al (Röller et al., 1967), as methyl (2E,6E,10-cis)-10,11-epoxy-7-ethyl-3,

11-dimethyl-2,6-tridecadienoate. The structure was further confirmed as the 2E,6E,10-cis isomer

(Dahm et al., 1968), and the absolute configuration of the chiral centers (C10 and C11) was

determined to be 10R,11S (Faulkner and Petersen, 1971; Meyer et al., 1971; Nakanishi et al.,

1971). This JH homologue was known as JH I (Fig. 1.1) (Goodman and Cusson, 2012).

2

COOCH3

O

COOCH3

O

JH I

JH II

COOCH3

O

JH III

COOCH3

O

JH 0

COOCH3

O

iso-JH 0

COOCH3

O O

JHB3

COOCH3

O

JHSB3

O

COOCH3

O

4'-Hydroxy JH III

COOCH3

O

HO

HO

COOCH3

O

8'-Hydroxy JH III

12'-Hydroxy JH III

HO

COOCH3

MF

23

4

5

67

8

9

1011

1

Figure 1.1 Structures of the JH homologues in insects. Figure adapted from Goodman and

Cusson (2012)

3

JH I, which has only been identified in the Lepidoptera, not only plays important roles in

regulating development, morphogenesis and reproduction in the Lepidoptera, but also has an

effect on the development of other insects (Fisher and Mayer, 1982; Granger et al., 1979;

Granger et al., 1982; Shalaby et al., 1990; Steiner et al., 1999). In larval development of Corcyra,

JH I treatment on ligated early-last instar resulted in a stimulation on DNA synthesis with a

consequent increase in DNA content and DNA concentration (Lakshmi and Dutta-Gupta, 1990).

During the last half of the larval molt of the tobacco hornworm, M. sexta, the presence of JH I at

the peak of the ecdysteroid titer is important in inducing dopa decarboxylase (DDC), an enzyme

which converts dopa to dopamine (Hiruma and Riddiford, 1985). In other insects, topical

application of the synthetic JH I to adult Musca domestica vicina Macq resulted in a shortened

gonoadotrophic cycle, decreased number of eggs and reduced hatching rate (Shalaby et al., 1990).

In addition, addition of JH I to the culture medium improved the development of single two-cell-

stage embryos of a polyembryonic wasp Copidosoma floridanum (Iwabuchi, 1995).

A second JH homologue, JH II (methyl (2E, 6E, 10-cis)-10,11-epoxy-3,7,11-trimethyl-2,6-

tridecadienoate) was identified in H. cecropia extracts (Meyer et al., 1970; Meyer et al., 1968).

JH II is the 2E, 6E, 10-cis isomer, as in JH I, but differs from JH I by a methyl group at C7 (Fig.

1.1). The absolute configuration of natural JH II at the C10, C11 positions has not yet been

determined (Goodman and Cusson, 2012). Same as JH I, JH II has only been identified in the

Lepidoptera. Nevertheless, relatively little research has been performed on JH II. In Trichoplusia

ni, the JH (JH I and JH II) and ecdysteroid titres were determined from the egg to the pupal molt

(Grossniklaus-Burgin and Lanzrein, 1990). Very little JH was detected in the freshly laid eggs of

T. ni, while in larval stages, JH II appeared to be the predominant or exclusive juvenile hormone

to interact with ecdysteroids to regulate the larval development. JH II, which is the most

4

abundant JH in Sesamia nonagrioides, is involved both in diapause programming and diapause

manifestation in this animal (Eizaguirre et al., 2005). In addition, JH II appears to be able to

initiate male production followed by sexual reproduction in the water flea Daphnia magna

(Cladocera, Crustacea). Exposure of D. magna to either JH I or JH II reduces the reproduction

rate, and induces parthenogenetically reproducing D. magna to produce male neonates (Oda et

al., 2005).

JH III was first identified from organ cultures of CA of the tobacco hornworm moth, Manduca

sexta (Judy et al., 1973). JH III displays the same E, E configuration at C2, C3 and C6, C7;

however, it differs from other JH homologues, with methyl groups at the C3, C7, and C11

positions. This hormone only contains one chiral carbon (C10), which displays the 10R

configuration in insects (Fig. 1.1). Of the juvenile hormone family, JH III is the most ubiquitous

JH homologue since it is the only JH biosynthesized and released in Orthoptera, Coleoptera,

Diptera, Hymenoptera, Dictyoptera, Lepidoptera, and the primitive ametamorphic Thysanura

(Baker et al., 1984; Tobe and Stay, 1985b). In larvae and adults of many insects, JH III is the

principal or only JH homologue identified, such as in the cockroach, Nauphoeta cinerea,

Diploptera punctata and the firebrat, Thermobia domestica (Baker et al., 1984; Tobe et al., 1985).

JH 0 and its isomer 4-methyl JH I (iso-JH 0) were identified in M. sexta eggs (Bergot et al.,

1981). Differing from JH III, JH 0 contains ethyl groups at the C3, C7, and C11 positions (Fig.

1.1). To date, JH 0 and its isomer (iso-JH 0) have been identified only in the Lepidoptera and

their functions in insects were unclear. JH III bisepoxide (JHB3), which contains a second

epoxide substitution at C6, C7, was first identified in Drosophila melanogaster (Richard et al.,

1989b). JHB3 was determined to be the major in vitro JH product of larval ring glands and of

adult CA-corpus cardiacum (CC) complexes of D. melanogaster. Later study identified JHB3 in

5

various dipteran species, such as Ceratitis capitata, Lucilia cuprina, Phormia regina,

Sarcophaga bullata (Bylemans et al., 1998; Lefevere et al., 1993; Moshitzky and Applebaum,

1995; Moshitzky et al., 2003). It has been demonstrated that the higher cyclorrhaphous Diptera

produce JHB3 predominantly, and JHB3 is believed to be restricted to the higher Diptera

(Richard et al., 1989a; Richard et al., 1989b). In L. cuprina, JHB3 is the only juvenile hormone

biosynthesized in vitro (Lefevere et al., 1993). Although it appeared that JHB3 production was

restricted to higher Diptera, JHB3 was reported to be synthesized by CA and the male accessory

glands of the mosquitoes, A. aegypti in vitro (Borovsky et al., 1994). However, recent work by Li

et al. (Li et al., 2003) was unable to detect any JHB3 synthesized by CA complex of A. aegypti.

The existence of JHB3 in other orders has yet to be confirmed.

Even though JHs were identified and characterized in various species of insects, the structure of

the JH in order Hemiptera has been a matter of controversy (Kotaki, 1993, 1996). Although JH

III and methyl farnesoate (MF) were reported as the products of CA in vitro in Dysdercus

fasciatus (Bowers et al., 1983; Feldlaufer et al., 1982) and the presence of JH I in the

hemolymph of Riptortus clavatus (Numata et al., 1992), there were no significant levels of the

known JH or related compounds in the milkweed bug, Oncopeltus fasciatus (Baker et al., 1988).

The presence of an unknown Heteropteran JH was suggested (Miyawaki et al., 2006). The

mystery regarding the JH in Hemiptera was not resolved until 2009, when Kotaki et al (2009)

identified a new JH homologue, JH III skipped bisepoxide (JHSB3; Fig. 1.1) from Plautia stali, a

member of the family Pentatomidae, suborder Heteroptera, order Hemiptera using a novel

approach. The term “skipped” refers to a second epoxide substitution switching from C6, C7 as

in JHB3 to C2, C3. The absolute chemical structure of the novel skipped bisepoxide JH was

characterized by the screening of a JH molecular library, and the juvenilizing activity of JHSB3

6

with different configurations on C2 and C3 and chirality on C10 was determined. Their result

shows that the (2R, 3S) configuration is more important for biological activity than the chirality

of C10, C11. JHSB3 with the 2R, 3S-configuration was more potent than those with the 2S, 3R-

configuration and 2,3-double bond (Kotaki et al., 2011). The function of JHSB3 was determined

in the last instars and adults of P. stali (Kotaki et al., 2011). Topical application of JHSB3 to last

instar nymphs inhibited their metamorphosis, and JHSB3 application in allatectomized and

diapausing adults stimulated the development of ovaries and ectadenia in females and males,

respectively.

Another family of JH homologues, hydroxylated JHs (HJHs; Fig. 1.1) was identified from the

African locust Locusta migratoria (Mauchamp et al., 1999). JH III was identified as the main

product released by the CA in vitro of L. migratoria (Mauchamp et al., 1985), while later studies

discovered three different hydroxylated forms of JH III (4-OH, 8-OH, and 12-OH JH III)

exhibited JH-like biological effects (Darrouzet et al., 1997; Mauchamp et al., 1999), in which 12-

OH JH III was found to be 100-fold more active than JH III.

Methyl farnesoate (MF; Fig. 1.1), a JH precursor without a C10, C11 epoxidation as in JH III,

was first isolated to function as a JH from the hemolymph of the spider crab Libinia emarginata

(Laufer et al., 1987). Recent studies have shown multifunctional roles of MF in crustaceans,

including reproduction, molting, larval development, morphogenesis, behaviour and general

protein synthesis (Chang et al., 2001; Nagaraju, 2007). Studies in insects suggest that MF may

also serve as a hormone in some insects. In the embryos of the cockroach N. cinerea, the

predominant product released by the embryonic CA is MF until the stage of breaking of the

chorion, and this substance circulates in embryonic haemolymph (Bürgin and Lanzrein, 1988;

Lanzrein et al., 1984). MF was also found to be biosynthesized by ring glands of larval D.

7

melanogaster (Richard et al., 1989a; Richard et al., 1989b), the embryonic CA of D. punctata

(Cusson et al., 1991b), the larval CA of Pseudaletia unipuncta (Cusson et al., 1991b) and the

adult CA of Phormia regina (Yin et al., 1995). In addition, MF was also identified from the

hemolymph of five orders of insects, including D. melanogaster (order Diptera), Schistocerca

americana (order Orthoptera), three species of true bugs (order Hemiptera), worker honeybees,

Apis mellifera (order Hymenoptera), and three species of beetle (order Coleoptera) (Teal et al.,

2014). Based on the study on the activity of MF, Goodman and Cusson (2012) reviewed the in

vivo biological role of MF in larvae and adults of D. melanogaster: During the larval stage, MF

is more active in blocking adult development than JH III or JHB3, whereas JH III or JHB3 is

more active once pupariation has been initiated.

2. JH biosynthetic pathway

The JH biosynthetic pathway, which comprises 13 discrete enzymatic steps, can be divided into

two distinct biosynthetic parts: the mevalonate pathway and the JH-specific pathway (Fig. 1.2).

The mevalonate pathway is an important cellular metabolic pathway present in all higher

eukaryotes and many bacteria. The isoprenoids produced by the mevalonate pathway are vital for

diverse cellular functions, including the synthesis of cholesterol, haem A, ubiquinone, dilochol,

and farnesylated proteins, growth control, and electron transport (Goldstein and Brown, 1990).

The most well-studied product of mevalonate pathway is cholesterol, because of its role in

maintaining cell membranes and its implications for human cardiovascular diseases (Goldstein

and Brown, 1990). The initial stages of biosynthesis of JH to the formation of farnesyl

diphosphate (FPP) proceeds through the mevalonate pathway, which is shared in vertebrates and

invertebrates (Belles et al., 2005; Goodman and Cusson, 2012).

8

S

O

CoA

Thiol

S

O

CoA

O

HMGS

Acetyl-CoA

HMG-CoA

Acetoaacetyl-CoA

S

O

CoA

OH

HO

O

Mevalonate

HMGR

OH

OOH

HO

O

MK

PMK

Mevalonate-PP (MPP)

Isopentenyl-PP (IPP)

PPMD

Dimethylallyl-PP (DMPP)

Farnesyl-PP

Farnesol

Farnesal

Farnesoic acid

FPPS

FPPP

FOLD

FALD

OPP

OOH

HO

O

OPPOPP

IPPI

OPP

OH

O

OH

O

OH

O OMe

O

OMe

O

O

O

JHAMT

JHAMT

CYP15A1

Mevolonate-POP

OOH

HO

O

Geranyl-PP (GPP)OPP

FPPS

Orthopteran and Dictyopteran

Lepidoptera

Methyl farnesoate (MF)

Juvenile hormone acid (JHA)

CYP15C1

Figure 1.2 The JH biosynthetic pathway. The mevalonate pathway was shown before the dash

line and the JH-specific pathway was after. Figure adapted from Goodman and Cusson (2012).

9

Insects and other arthropods, however, do not produce cholesterol as a final product of the

mevalonate pathway, because they lack the enzymes squalene synthetase (farnesyl-diphosphate

farnesyltransferase) and lanosterol synthase, which are required for the production of cholesterol

(Clark and Bloch, 1959). Thus, the second portion of the JH biosynthetic pathway comprises

enzymatic steps unique to JH-producing organisms. Earlier studies on the JH biosynthetic

pathway focused on the activity of enzymes in the mevalonate pathway (Casals et al., 1996;

Couillaud and Feyereisen, 1991; Feyereisen and Farnsworth, 1987a). Thanks to whole genome

sequencing and CA trancriptomics studies, the identification and characterization of JH

biosynthetic enzymes have greatly improved, especially for genes encoding enzymes in the JH-

specific pathway (Consortium, 2006; Group, 2004; Mita et al., 2004; Noriega et al., 2006). Since

JH III is the most common JH homologs in insects, our review focuses on the study of enzymes

directly involved in the biosynthesis of JH III.

2.1 Acetoacetyl-CoA thiolase (ACAT, Thiol)

The synthesis of JH begins with acetyl-CoA, which condenses with another acetyl-CoA through

the catalysis of Thiol to form Acetoacetyl-CoA (Fig. 1.2). The gene encoding Thiol has been

identified in the genome of D. melanogaster, B. mori, Anopheles gambiae and A. aegypti, A.

mellifera and in an EST of the scolytid beetle Ips pini (Bomtorin et al., 2014; Eigenheer et al.,

2003; Keeling et al., 2004; Kinjoh et al., 2007; Nouzova et al., 2011). In vertebrates, the

functional Thiol is comprised by a tetramer of identical subunits and has two cysteine residues at

the active sites (Gehring and Harris, 1970), which are conserved in insects. In B. mori, Thiol was

almost exclusively expressed in the CA-CC complex, wheras the transcription of Thiol in A.

aegypti and A. mellifera was expressed in many tissues, including the ovary and fat body

(Bomtorin et al., 2014; Kinjoh et al., 2007; Nouzova et al., 2011).

10

2.2 3-hydroxy-3-methylglutaryl-CoA synthase (HMGS)

The function of HMGS is to catalyze the condensation of acetyl-CoA and acetoacetyl-CoA to

yield HMG-CoA (Fig. 1.2). The enzymatic activity of HMGS in the CA was determined in the

adult female of D. punctata. The results show that enzymatic activity of HMGS has the same

pattern as JH III biosynthesis, which suggests that HMGS influences the rate of JH biosynthesis.

However, the fact that the enzyme activity declined after the dramatic decrease of JH on day 6

indicates that changes in the activity of HMGS enzymes are apparently not responsible for the

collapse in JH synthetic ability on day 6 (Couillaud and Feyereisen, 1991).

The gene encoding HMGS was first isolated in the cockroach B. germanica (Buesa et al., 1994;

Martinez-Gonzalez et al., 1993), and later in D. melanogaster (Spradling et al., 1999), the

scolytid beetle Dendroctonus jeffreyi (Tittiger et al., 2000), B. mori (Kinjoh et al., 2007), A.

aegypti (Nouzova et al., 2011), and A. mellifera (Bomtorin et al., 2014). HMGS predominantly

expressed in CA-CC complex of B. mori (Kinjoh et al., 2007), A. aegypti (Nouzova et al., 2011),

and A. mellifera (Bomtorin et al., 2014), and the transcript levels correspond to JH biosynthesis

in the CA.

In B. germanica, two HMGS enzymes (HMGS-1 and HMGS-2) with 69% amino acid identity

were demonstrated (Buesa et al., 1994), and the gene encoding HMGS-1 was considered to be a

functional retrogene derived from HMGS-2 by retrotransposition (Buesa et al., 1994; Cabano et

al., 1997; Casals et al., 2001). In mammals, two forms of HMGS (a mitochondrial form and a

cytoplasmic form) have been detected, which are encoded by two different genes (Ayte et al.,

1990). In B. germanica, none of the two HMGS enzymes show any recognizable N-terminal

leader peptide to target the protein to mitochondria, which suggests that the enzyme is cytosolic

11

in insects (Buesa et al., 1994). Both HMGS are highly expressed in the adult ovary, coordinately

regulated in the ovary during the gonadotrophic cycle, but expressed differently throughout

development (Ayte et al., 1990; Martinez-Gonzalez et al., 1993). The expression and enzymatic

activities of both HMGS were also determined in the fat body of B. germanica (Casals et al.,

1996). HMGS-1 did not show any significant mRNA level or detectable protein level in the fat

body, which indicates a limited role for HMGS-1 in the fat body. HMGS-2, on the other hand,

shows a clear pattern in the fat body, which was consistent with that of vitellogenin production.

In D. jeffreyi, HMGS transcript localizes mainly in the metathorax and abdomen (Tittiger et al.,

2000). Topical application of JH III induced a dose- and time-dependent increase in HMGS

transcripts in the male metathoracic-abdominal region, whereas no increase in the transcript

levels was observed in the JH III-treated female. The JH III-mediated regulation of HMGS

suggests that in addition to its function in the JH biosynthetic pathway, HMGS appears to control

the isoprenoid pathway (Tittiger et al., 2000).

2.3 3-hydroxy-3-methylglutaryl-CoA reductase (HMGR)

HMGR catalyzes the first committed step of the isoprenoid biosynthetic pathway, the conversion

of HMG-CoA to mevalonate (Fig. 1.2). It is believed to play an important role in the regulation

of sterol synthesis and is generally referred to as the rate-limiting enzyme in cholesterol synthesis

in vertebrates (Goldstein and Brown, 1990). Two distinct classes of HMGR have been identified

(Bochar et al., 1999): in eukaryotes, HMGR consist of a highly conserved C-terminal catalytic

domain and a poorly conserved N-terminal membrane anchor domain, which contains two to

eight inferred transmembrane helices. The second class of HMGR was discovered in the Archaea

and the true bacterium Pseudomonas mevalonii, which lack the N-terminal membrane anchor

domain. Both HMGR serve the same function in different species.

12

The presence of HMGR in the CA was first reported in M. sexta, which convert HMG-CoA to

mevalonate (Bergot et al., 1979). The activity of HMGR was further studied in D. punctata

(Feyereisen and Farnsworth, 1987a) and the grasshopper, Schistocerca nitens (Baker and

Schooley, 1981). HMGR in the insect undergoes phosphorylation (inactive form) and

dephosphorylation (active form), which could be altered by Mg-ATP or NaF (Monger and Law,

1982). In M. sexta, the activity of HMGR in the CA parallels, in most cases, the ability of the

gland to synthesize JH (Baker and Schooley, 1981). In D. punctata, the activity of HMGR

parallels JH biosynthesis until day 5 during the first gonadotrophic cycle. JH biosynthesis drops

dramatically on day 6, while the activity of HMGR remains high till day 8. In addition, the half-

life of HMGR was not related to the half-life of JH III biosynthesis (Feyereisen and Farnsworth,

1987a). The results suggest that HMGR is not the ‘the rate-limiting enzyme’ in JH biosynthesis

in D. punctata.

To date, the gene encoding HMGR was cloned in many insect species, including D.

melanogaster (Gertler et al., 1988), B. germanica (Martinezgonzalez et al., 1993), the I. pini

(Hall et al., 2002), Ips paraconfusus (Tittiger et al., 1999), D. jeffreyi (Tittiger et al., 2003), the

moth Agrotis ipsilon (Duportets et al., 2000), B. mori (Kinjoh et al., 2007), A. aegypti (Nouzova

et al., 2011), and A. mellifera (Bomtorin et al., 2014). The identification of HMGR genes has

permitted further study of the function and the regulation of HMGR. For instance, a HMGR gene

encoding 916 amino acid was identified in D. melanogaster. The 56% identity in C-terminal

region to the hamster HMGR reflects the essential function of the C-terminal region. On the

other hand, the high similarity in the membrane-spanning regions between the mammalian and

Drosophila HMGR suggests that the transmembrane domains may be essential for recognizing

specific mevalonate derivatives or their binding proteins. In addition, the identification of

13

HMGR has allowed the expression of Drosophila HMGR in Schneider cells. Addition of

mevalonate suppresses the transcript level and the enzymatic activity of Drosophila HMGR,

which indicates a feedback regulation involved in the mevalonate pathway (Gertler et al., 1988).

The expression and the function of HMGR were also determined in many other insects. In B.

germanica (Martinezgonzalez et al., 1993), A. ipsilon (Duportets et al., 2000) and A. mellifera

(Bomtorin et al., 2014), HMGR was found to express in many tissues, including fat body, ovary,

muscle, brain and CA, whereas its expression in B. mori (Kinjoh et al., 2007) and A. aegypti

(Nouzova et al., 2011) was predominantly in the CA. Study of HMGR in Ips paraconfusus and D.

jeffreyi suggests that HMGR is not only involved in the regulation of JH biosynthesis, but also

the production of monoterpenoid pheromones in the males. (Tittiger et al., 2003; Tittiger et al.,

1999).

2.4 Mevalonate kinase (MK)

MK is responsible for the phosphorylation of mevalonate to produce the 5-phosphomevalonate

(mevalonate-P), which is metabolized in fungal, plant, and vertebrate systems to isopentenyl

pyrophosphate (IPPI) (Cornforth et al., 1960) (Fig. 1.2). The isolation, purification, and

characterization of MK was first reported in larval S. bullata by Goodfellow and Barnes (1971).

MK is distributed in the muscle and brain complex cytosol fractions of S. bullata (Cornforth et

al., 1960). The gene encoding MK was identified in D. melanogaster and A. gambiae genomes,

the EST database of D. punctata (Noriega et al., 2006), B. mori (Kinjoh et al., 2007), A. aegypti

(Nouzova et al., 2011), and A. mellifera (Bomtorin et al., 2014). In B. mori and A. aegypti, the

transcript level of MK is highest in the CA, whereas in A. mellifera, the highest expression of

MK is in the brain (Bomtorin et al., 2014; Kinjoh et al., 2007; Nouzova et al., 2011). In addition,

14

the expression profile of MK in the CA of B. mori and A. aegypti corresponded to changes in JH

biosynthesis.

2.5 Phosphomevalonate kinase (PMK)

Phosphomevalonate kinase (PMK) catalyzes the phosphorylation of mevalonate-P into 5-

diphosphomevalonate (MPP), an essential step in isoprenoid biosynthesis (Fig. 1.2). Two non-

orthologous genes encoding PMK have been identified: the Saccharomyces cerevisiae ERG8

gene, which is found in eubacteria, fungi and plants, and the human PMK gene, which is present

only in animals (Houten and Waterham, 2001). In insects, the gene encoding PMK was

identified in D. melanogaster and A. gambiae genomes, B. mori (Kinjoh et al., 2007), A. aegypti

(Nouzova et al., 2011), and A. mellifera (Bomtorin et al., 2014). In B. mori, PMK is expressed in

multiple tissues, and the expression profile in the larval CA shows a similar pattern as that of the

JH biosynthesis (Kinjoh et al., 2007). In A. aegypti, the PMK was highly expressed in the CA,

followed by ovary. The expression of PMK in the CA also coordinated with JH biosynthesis

(Nouzova et al., 2011).

2.6 Diphosphomevalonate decarboxylase (PPMD)

Mevalonate diphosphate decarboxylase (PPMD) is an enzyme in the mevalonate pathway that

catalyzes the decarboxylation of the six-carbon MPP to the five-carbon isopentenyl diphosphate

(IPP). This reaction involves the dehydration of the substrate and the hydrolysis of one molecule

of ATP, and Mg2+ is required (Jabalquinto et al., 1988). In rats, PPMD is a key enzyme in the

MVA pathway that is essential for the biosynthesis of the isoprenoids. However, there are not

many studies on PPMD in insects. The gene encoding PPMD was identified from the genomes

of D. melanogaster and A. gambiae, the EST database of I. pini (Keeling et al., 2004) and later in

15

B. mori (Kinjoh et al., 2007), and A. aegypti (Nouzova et al., 2011). In B. mori, PPMD was

exclusively expressed in CC-CA complex of 4th instar larvae, and its expression paralleled the JH

titre in 4th and 5th instars. In A. aegypti, PPMD was also exclusively expressed in the CA and its

expression coordinated with JH biosynthesis by CA, which indicates that PPMD is involved in

the regulation of JH biosynthesis in insects (Kinjoh et al., 2007; Nouzova et al., 2011).

2.7 Isopentenyl diphosphate isomerase (IPPI)

In the mevalonate pathway, IPP is the sole product of the ATP-dependent decarboxylation of

MPP and must be isomerized to DMPP to form Geranyl pyrophosphate (GPP) (Fig. 1.2). The

enzyme isopentenyl diphosphate isomerase (IPPI), which catalyzes the isomerization of

isopentenyl pyrophosphate (IPP) to dimethylallyl pyrophosphate (DMPP) (Fig. 1.2), plays a

central role in isoprenoid biosynthesis (Ramos-Valdivia et al., 1997). Two isoforms of IPPI have

been identified. Type 1 IPPI (IPPI-1) is a metalloprotein that is found in eukaryotes, and the

optimal functioning of IPPI-1 requires a divalent metal cation (Mg2+ or Mn2+). Zinc has also

been identified as an essential cofactor for the catalysis activity of IPPI-1 from E. coli (Carrigan

and Poulter, 2003). The type 2 isoform (IPPI-2) is a flavoenzyme found in plant chloroplasts and

bacteria; the isomerase activity requires not only a divalent metal cation, but also a reduced

flavin coenzyme (de Ruyck et al., 2014).

In insect species, IPPI was first partially characterized from the extracts of B. mori, in which that

the isomerase activity is dependent on the metal ions. Mn2+ was a better activator than Mg2+,

especially at low concentrations (Koyama et al., 1985). Later study identified the sequence of

IPPI in the genome database of D. melanogaster and A. gambiae, the EST database of I. pini

(Keeling et al., 2004), B. mori (Kinjoh et al., 2007), A. aegypti (Diaz et al., 2012; Nouzova et al.,

16

2011), the spruce budworm, Choristoneura fumiferana (Sen et al., 2012), , M. sexta (Sen et al.,

2012) and A. mellifera (Bomtorin et al., 2014). The IPPI gene identified in A. aegypti (AaIPPI)

encodes a 244 amino acid (aa) protein with high similarity to IPPI-1 in other organisms (Diaz et

al., 2012). Two important motifs which are associated with the catalytic roles of IPPI-1 are well

conserved: a TNACCSHPL motif containing a conserved cysteine residue and a WGEHEIDY

motif that contains a conserved glutamate residue. The function of IPPI-1 in other organisms

requires the binding of a divalent metal cation (Mg2+, Mn2+ or Zn2+) (Carrigan and Poulter, 2003),

whereas the enzymatic assay shows that the full activity of AaIPPI requires Mg2+ or Mn2+ but not

Zn2+, and its activity can be completely inhibited by iodoacetamide (Diaz et al., 2012).

Insects in the order Lepidoptera produce five JH homologues (JH 0, JH I, 4-methyl JH I, JH II,

and JH III) (See section 1.1). For the biosynthesis of different JH homologues, homologs of IPP

and DMPP are involved into the mevalonate pathway, which requires the IPPI in Lepidoptera to

catalyze the isomerization of homoisopentenyl diphosphate (HIPP) to homodimethylallyl

diphosphate (HDMPP). Earlier studies in pig demonstrated that the isomerization of HIPP by

porcine IPPI produced very little HDMPP. However, studies in insects demonstrated that CA

homogenates of adult female M. sexta and purified IPPI from B. mori regiospecificly catalyzed

the isomerization of HIPP to HDMPP, which suggests that the lepidopteran IPPI enzyme is

structurally distinct from other isomerases (Baker et al., 1981; Koyama et al., 1985). Further

study on IPPI in C. fumiferana and M. sexta confirmed the function of IPPI in catalyzing the

isomerization of HIPP, and the homology models of the CfIPPI and HIPP isomerization study

revealed that the lepidopteran IPPI enzyme has a larger active site cavity, to allow binding of

larger substrates and to stabilize the high-energy intermediate formed during substrate

isomerization (Sen et al., 2012).

17

The expression of IPPI mRNA varies in different tissues, but in most insects, the highest

transcript level of IPPI was found in the CA. In B. mori, IPPI mRNA levels are expressed almost

exclusively in the CA of 4th instar larvae, with relative low levels in other tissues (Kinjoh et al.,

2007). In A. mellifera and C. fumiferana, mRNA of IPPI was expressed in multiple tissues, with

the highest transcript level in the CA (Bomtorin et al., 2014; Sen et al., 2012). IPPI mRNA of A.

aegypti expressed in various tissues, including CA-CC, ovary, hindgut, brain, midgut, fat body of

the female, and testis and accessory glands of the male (Diaz et al., 2012). The ubiquitous

expression of IPPI suggests that IPPI might be involved in many metabolic pathways. The

pattern of change of IPPI mRNA in the CA-CC during female pupal and adult development was

consistent with the changes in JH biosynthesis, which suggests that the transcription of IPPI is

partially responsible for JH biosynthesis (Diaz et al., 2012; Nouzova et al., 2011). Similar results

were found in B. mori (Kinjoh et al., 2007). The expression of BmIPPI in the CA of 4th, 5th

larvae, and pupae coordinate with the JH titre.

2.8 Farnesyl diphosphate synthase (FPPS)

The condensation of IPP and DMPP forms an intermediate compound GPP, which then

undergoes a second condensation step to generate farnesyl diphosphate (FPP) (Fig. 1.2). This

process is catalyzed by the enzyme Farnesyl diphosphate synthase (FPPS), a type of

prenyltransferase. FPPS is a homodimeric protein, which is formed by tightly coupled subunits

ranging from 32 to 44 kDa in size (Vandermoten et al., 2009a). The activity of FPPS requires

divalent metal cations (Mg2+ or Mn2+). Sequence analysis of FPPSs revealed seven conserved

regions, including two substrate binding regions, regions II and VI. Both regions contain an

aspartate-rich motif, DDx(xx)xD (x represents any amino acid) (Liang et al., 2002). Region II,

which includes the first aspartate-rich motif (FARM), is responsible for the determination of

18

chain-length, while region VI, which contains the second aspartate-rich motif (SARM), is

considered to be the IPP binding site (Liang et al., 2002). The first FPPS genes in insects were

cloned in A. ipsilon by Castillo-Gracia and Couillaud (1999) with high identity (about 40%) with

other FPPSs and high transcript level in the CA. Additional insect FPPSs were identified in many

insect orders, including Lepidoptera (Sen and Sperry, 2002), Diptera (Nouzova et al., 2011; Sen

et al., 2007), Coleoptera (Taban et al., 2009), Hemiptera (Lewis et al., 2008; Sun and Li, 2012;

Zhang and Li, 2008), Hymenoptera (Bomtorin et al., 2014) and Blattodea (Noriega et al., 2006).

In Lepidoptera, two distinct forms of FPPS were identified, designated type-1 and type-2 FPPS

(FPPS-I and FPPS-II) (Cusson et al., 2006), while two slightly different isoforms of type-2 FPPS

are present in B. mori (FPPS-2 and FPPS-3) (Kinjoh et al., 2007). Like the prenyltransferases in

other organisms, FPPS in M. sexta requires the divalent cation (Mg2+ or Mn2+) for its activity.

The presence of detergent, glycerol, and non-specific protein-protein interactions improves the

stability and catalytic activity of FPPS (Sen and Sperry, 2002). As described in section 1.1, the

Lepidoptera produce five JH homologues (Fig. 1.1), which requires the synthesis of

ethyl/methyl-substituted FPP by FPPS. The preference of prenyltransferase to ethyl/methyl-

substituted DMPP was determined using M. sexta CA homogenates, and the results suggest that

the selectivity of the enzyme incline to the ethyl-substituted substrate (Sen et al., 1996).

Additional studies on the selectivity of this prenyltransferase were performed using different

substrate analogs (Sen et al., 2006). Compared to pig liver FPPS, the lepidopteran enzyme

derived from CA homologues displays greater steric latitude around the C-3 and C-7 alkyl

positions of DMAPP and geranyl diphosphate (GPP). The enzymes generate more ethyl-

branched geranyl/farnesyl diphosphate and the substrate specificity related to the enzyme

localization (Sen et al., 2006). The structure analysis of FPPS-I and FPPS-II in C. fumiferana

19

revealed that FPPS-1 displays several unique active site substitutions, whereas FPPS-II has a

more conventional catalytic cavity which indicates FPPS-I are better suited than FPPS-II for

generating ethyl-substituted products. However, tissue distribution of FPPS mRNA showed that

FPPS-I is ubiquitous whereas FPPS-II is predominately expressed in the CA (Cusson et al.,

2006). The result is consistent with the distribution of FPPS 1-3 mRNA in 4th instar B. mori

(Kinjoh et al., 2007). These results suggest that FPPS-II may play a leading role in lepidopteran

JH biosynthesis despite its apparently more conventional catalytic cavity. In other species, FPPS

forms a homodimeric protein. The recombinant C. fumiferana FPPS-2 was active in producing

FPP; However, expression of FPPS-1 (CfFPPS1, Pseudaletia unipuncta FPPS1, and A. ipsilon

FPPS1) in E. coli failed to display any FPPS activity in vitro. Surprisingly, the combination of

CfFPPS1 and CfFPP2 enhanced the enzyme activity, and an association between CfFPPS1 and

CfFPPS2 was observed, which suggests that FPPS-I and FPPS-II may derive from a heteromer to

play a role in JH biosynthesis in moths. Whether these two enzymes form heterodimers in vivo

has not yet been verified.

Two FPPS genes that encode proteins with about 80% identity were identified in the green peach

aphid, Myzus persicae, and in the bird cherry-oat aphid Rhopalosiphum padi (Sun and Li, 2012;

Zhang and Li, 2008). Enzyme activity of FPPS in R. padi shows both enzymes could catalyze the

formation of FPP from IPP and DMAPP (Sun and Li, 2012). However, the function of FPPS in

M. persicae was not determined. Without the feature of the FPPS in M. persicae, it is premature

to conclude that these genes are involved JH biosynthesis. On the other hand, other

prenyltransferase genes displaying dual geranyl diphosphate (GPP)/farnesyl diphosphate (FPP)

synthase activity in vitro were identified in M. persicae. These two prenyltransferase genes

encode very similar proteins, apart from the presence of a mitochondrial leader sequence (Lewis

20

et al., 2008). It is interesting to note that the prenyltransferase enzyme in aphid is not unique to

the aphid species from which it was cloned. The molecular dynamics of the enzyme are

responsible to maintain the balance between the production of GPP and FPP (Vandermoten et al.,

2009b).

Only single copies of the FPPS gene were identified in insects from other orders, such as Diptera

(Nouzova et al., 2011; Sen et al., 2007) and Coleoptera (Taban et al., 2009). The scanning of the

A. mellifera genome showed the presence of seven copies (Consortium, 2006). The expression of

the first six copies of the FPPS gene revealed that FPPS3 is the bona fide gene involved in JH

biosynthesis in honey bees (Bomtorin et al., 2014).

2.9 Farnesyl diphosphate pyrophosphatase (FPPP)

Farnesyl diphosphate pyrophosphatase (FPPP) catalyzes the hydrolysis of farnesyl diphosphate

(FPP) to farnesol (FOL). Relatively little was known about FPPP in insects. Cao et al. (2009)

screened the D. melanogaster genome and identified the first FPPP genes in the insects. FPPP

belongs to the haloalkanoic acid dehalogenase (HAD) super family that catalyzes phosphoryl

transfer reactions (Allen and Dunaway-Mariano, 2004). Members of the HAD phosphatase

superfamily have four conserved amino acid signature motifs, which are also well conserved in

the DmFPPP. Nyati et al. (2013) identified 3 putative FPPP (AaFPPP-1, -2, and -3) paralogs

through a search for orthologs of the DmFPPP in the CA of A. aegypti. Recombinant AaFPPP-1

and AaFPPP-2 displayed the features of FPPP in their ability to hydrolyze FPP into FOL, and the

FPPP activity of the CA extracts was found to be Mg2+-dependent. The determination of the

function of FPPPs in JH biosynthesis using RNAi reveals that FPPP-1 plays the predominant

function in JH biosynthesis. Unlike mRNA of enzymes in the mevalonate pathway which are

21

predominately expressed in the CA, AaFPPPs mRNA are expressed in various tissues, with

FPPP-1 highly expressed in midgut and Malpighian tubules, FPPP-2 in Malpighian tubules, and

FPPP-3 in brain and ovary. The ubiquitous expression of FPPP may result from the pleiotropic

functions of farnesol and farnesal. In spite of the ubiquitous expression, the expression of FPPP-

1 and -2 in the CA correlated with JH biosynthesis in sugar-fed females, which suggests FPPP

may play a role in the regulation of JH biosynthesis.

2.10 Farnesol dehydrogenase (FOLD)

A Farnesol dehydrogenase (FOLD) is responsible for the catalysis of the conversion of farnesol

(FOL) to farnesal (FAL) (Fig. 1.2). In vertebrates, plants, and fungi, the oxidation of FOL to

FAL is mediated by nicotinamide-dependent dehydrogenases (Chayet et al., 1973; Inoue et al.,

1984; Keung, 1991). Study of FOL oxidation using CA homogenates of the adult female M sexta,

revealed that farnesol and/or farnesal dehydrogenase were NAD+-dependent enzymes (Baker et

al., 1983). However, the conversion of FOL to FAL in larval M. sexta was not affected by

nicotinamide. The enzyme, which oxidizes FOL to FAL in larval M. sexta, appears to be an

oxygen-dependent enzyme, perhaps a flavin and/or iron-dependent oxidase (Sperry and Sen,

2001). A FOLD enzyme was identified and functionally characterized in the CA of adult A.

aegypti (Mayoral et al., 2009a). In CA of adult female M sexta, the FOLD enzyme was

ineffective in the addition of NADP+ (Baker et al., 1983). However, the FOLD in A. aegypti was

characterized to be a NADP+-dependent farnesol-dehydrogenase. It is possible that moths utilize

a different mechanism for FOL oxidation. In A. aegypti, FOLD is expressed in various tissues,

and with a relatively low transcript level in the CA. On the other hand, the transcript levels of

FOLD in the CA coordinate with JH biosynthesis (Mayoral et al., 2009a).

22

2.11 Farnesal dehydrogenase (FALD)

Farnesal dehydrogenase (FALD), which catalyzes the oxidation of farnesal to FA, was one of the

less understood steps in JH synthesis (Fig. 1.2). An early study using the CA homogenates of the

adult female, M sexta predicted that FALD is an NAD+-dependent aldehyde dehydrogenase and

this aldehyde dehydrogenase showed some substrate specificity for the 2E isomer (Baker et al.,

1983). Rivera-Perez et al. (2013) identified and characterized a FALD enzyme in female A.

aegypti. Two FALD genes with 50% amino acid identity were identified, in which FALD-1

produces four different transcripts and FALD-2 produces one. All five FALD variants exhibit the

activity to convert FAL to farnesoic acid (FA), and the oxidation is stimulated in the presence of

NAD+. mRNA of FALD variants have unique tissue distribution profiles, with each

predominantly expressed in one unique tissue: FALD1-A in ovaries, FALD1-B in Malpighian

tubes, FALD1-C in hindgut, FALD1-D in nervous tissue, while FALD2 are relatively low in

transcript level comparing to FALD1. The reduction of FALD activity results in accumulation of

farnesol, which subsequently converts back into farnesol, resulting in farnesol leaking out of the

CA. Oxidation of farnesal may be a rate limiting step in JH synthesis in mosquito after blood

feeding.

2.12 Juvenile hormone acid O-methyltransferase (JHAMT)

In the final two steps of JH biosynthesis, FA is converted to JH III through a methyl transfer and

an epoxidation (Fig. 1.2). The order of the final two steps is insect order dependent. In

Orthoptera, Dictyoptera, Coleoptera and Diptera, FA undergoes the methylation of the carboxyl

group to produce methyl farnesoate (MF), which is epoxidized by a P450 monooxygenase at C10,

C11 position to generate JH III. In Lepidoptera, however, a reverse step order occurs:

23

epoxidation precedes methylation. Other ethyl-branched JH homologues are also synthesized

through the same order, but with different precursors derived from homomevalonate and

mevalonate (Shinoda and Itoyama, 2003). The modeling of A. gambiae, B. mori, D.

melanogaster and T. castaneum JHAMTs and docking simulation shows that all insect JHAMTs

are able to esterify both FA and JHA. The order of the methylation/epoxidation may be

controlled by the specificity of the epoxidase. The epoxidase in Lepidoptera might have higher

affinity than JHAMT for FA, which results in epoxidation precedes methylation. In other insects,

however, the epoxidase can only selectively catalyze MF. Thus, esterification of FA to MF by

JHAMT occurs in earlier step (Defelipe et al., 2011).

The first JHAMT was cloned and functionally characterized in, B. mori (Shinoda and Itoyama,

2003). The sequence analysis reveals that JHAMT belongs to the SAM-dependent

methyltransferases family, with a conserved S-adenosyl-L-methionine (SAM) binding motif. The

protein encoded by the JHAMT gene was able to not only convert JH III acid to JH III, but also

catalyze the conversion of JHA I, II, and FA to their cognate JH methyl esters in the presence of

S-adenosyl-L-methionine (SAM). Northern blot analysis shows that BmJHAMT mRNA is

exclusively expressed in the CA. Expression of JHAMT correlates well with the JH biosynthetic

activity of the CA in 4th and 5th instar larvae, pupae and adults, and especially in 5th instar larvae

in which the shutdown of the expression of JHAMT appears to be the primary reason for the

decline in JH biosynthesis. These results suggest that JHAMT is the rate limiting enzyme in the

JH biosynthesis in B. mori (Kinjoh et al., 2007; Shinoda and Itoyama, 2003). Orthologues of

JHAMT have also been cloned and characterized in other insect species, including T. castaneum

(Minakuchi et al., 2008a), D. melanogaster (Niwa et al., 2008), the Eri silkworm, Samia cynthia

ricini (Sheng et al., 2008), A. aegypti (Mayoral et al., 2009b), the desert locust Schistocerca

24

gregaria (Marchal et al., 2011) and A. mellifera (Bomtorin et al., 2014). In all species, JHAMT

is expressed predominately in the CA and the recombinant JHAMT protein from these insect can

catalyze the methylation of FA into MF, as well as JHA into JH III. In T. castaneum, silencing

the JHAMT gene using RNAi induced precocious metamorphosis, while in D. melanogaster

(Minakuchi et al., 2008a), JHAMT overexpression resulted in a pharate adult lethal phenotype

(Niwa et al., 2008). In S. gregaria, knockdown of JHAMT not only resulted in lower JH release,

but also a suppression in FA-stimulated JH release. A delay in sexual maturation was also

observed in JHAMT-silenced animals (Marchal et al., 2011). In many insects, such as T.

castaneum (Minakuchi et al., 2008a), D. melanogaster (Niwa et al., 2008), S. cynthia ricini

(Sheng et al., 2008), S. gregaria (Marchal et al., 2011) and A. mellifera (Bomtorin et al., 2014),

the expression of JHAMT in the CA correlates well with JH biosynthesis.

Farnesoic acid O-methyltransferase (FAMeT), which was first reported in a crustacean, was

initially considered to be the enzyme converting FA to MF in crustaceans (Gunawardene et al.,

2001). And a FAMeT was cloned in D. melanogaster (Burtenshaw et al., 2008).

Immunohistochemical analysis shows the presence of FAMeT in the CA portion of the ring

gland. However, recombinant FAMeT did not show any enzymatic activity in catalyzing the

conversion of FA or JHA. In S. gregaria, FAMeT mRNA was expressed in several tissues, and

its expression in the CA did not correlate with JH biosynthesis. In addition, silencing FAMeT

has no effect on either JH release or MF content of the CA (Marchal et al., 2011). These results

suggest that FAMeT does not encode a functional methyltransferase.

2.13 Juvenile hormone epoxidase (CYP15A1/CYP14C1)

25

Juvenile hormone epoxidase is involved in the epoxidation of FA in the Lepidotera or the

epoxidation of MF in the Orthoptera, Dictyoptera, Coleoptera, and Diptera (Fig. 1.2). Early

studies demonstrated that this epoxidase is a microsomal cytochrome P450 enzyme (Hammock,

1975). The first JH epoxidase, named CYP15A1, was cloned from D. punctata (Helvig et al.,

2004). This enzyme contains all the features of a typical microsomal P450, and its recombinant

protein catalyzed the epoxidation of MF to JH III in the presence of NADPH. The product of

epoxidation is mostly the (10R)-enantiomer. Studies on JHAMT reveal that the order of the final

two steps is primarily determined by the substrate specificity of epoxidase (Section 1.2.12).

CYP15A1 from D. punctata showed strong substrate specificity to the natural substrate MF.

Substitution of natural MF with geometrical isomers of MF and other terpenoids, such as

farnesol, farnesal, FA, and JH III showed little or no activity. DpCYP15A1 mRNA was

exclusively expressed in the CA and the expression level is higher in CA with high JH

biosynthetic activity. The expression and function of CYP15A1 has also been investigated in

many other insects, including A. aegypti (Nouzova et al., 2011), S. gregaria (Marchal et al., 2011)

and A. mellifera (Bomtorin et al., 2014). In all, CYP15A1 was selectively expressed in the CA.

In A. aegypti and A. mellifera, the expression of CYP15A1 does not correlate with JH

biosynthesis, whereas in S. gregaria, CYP15A1 shows high levels of transcription in active CA.

Silencing CYP15A1 in S. gregaria resulted in a reduction in JH release and an accumulation of

MF within the CA.

CYP15C1, an ortholog of CYP15A1, was recently identified and characterized in the

Lepidoptera, B. mori (Daimon et al., 2012). And CYP15C1 shares high homology with the

CYP15A1 in D. punctata. The dimolting (mod) mutation, which causes precocious larval-pupal

metamorphosis, results in a null mutation in the coding sequence of CYP15C1. CYP15C1 was

26

then expressed in Drosophila S2 cells and enzymological analysis revealed that CYP15C1

converts FA to JHA in a highly stereospecific manner. Further study showed that CYP15C1 is

responsible for the mod mutant of B. mori and its molecular defect results in the absence of JHs

(JH I and JH II) in B. mori, indicating CYP15C1 plays essential roles in JH biosynthesis. In

addition, CYP15C1 is specifically expressed in the CA. The expression of CYP15C1, on the

other hand, did not show any change during development, which suggests that CYP15C1 is not

the rate-limiting enzyme in the JH biosynthetic pathway. The CYP15 gene is not found in higher

dipterans such as D. melanogaster, probably because the production of JHB3 requires a different

epoxidase.

3. JH signaling pathway

3.1 Ultraspiracle (USP) as a JH receptor

Ultraspiracle (USP) is a homologue of the vertebrate retinoid X receptor (RXR), which can form

heteromers with other nuclear receptors to bind with genomic response elements (Henrich et al.,

1994). In Lepidoptera and Diptera, USP displays high identity with RXR only in the DNA-

binding domain, whereas the similarity in the ligand-binding domains is relatively low. In other

insect orders and other arthropods, USP shows higher similarity to vertebrate (Iwema et al.,

2007). USP can interact with several nuclear receptors, such as EcR and DHR38 (Sutherland et

al., 1995; Yao et al., 1992). The USP:EcR complex, which is formed by the heterodimerization

between EcR and USP, is required for the binding of the steroid hormone 20-hydroxyecdysone

(Yao et al., 1992).

Using a fluorescence assay, Jones and Sharp (1997) demonstrated a protein-ligand interaction

between the natural Drosophila JHs (JH III ester monoepoxide and bisepoxide, respectively) and

27

recombinant D. melanogaster USP. This interaction cannot be influenced by the addition of

farnesol or 20E, which indicates the specificity of the binding. On the other hand, JH acid, which

itself did not change the fluorescence of USP, affects the interaction between JH III ester and

USP. The action of JH acid indicates that JH acid may bind to USP in a different manner than JH

III ester. Further studies on D. melanogaster USP showed that it could also specifically bind to

JH III, which changes the conformation of USP and stabilizes the dimeric/oligomeric quaternary

structure of USP. The JH III agonist methoprene shows a competitive inhibition in the the JH III-

USP interaction (Jones et al., 2001). Furthermore, the binding affinities between USP and other

natural farnesoid products of the ring gland of D. melanogaster were determined (Jones et al.,

2006). MF exhibited a nanomolar affinity to USP, and the addition of an epoxide across a double

bond or any substitution on C1 (Fig. 1.1) other than methyl ester, resulted in a decrease in

affinity to USP. Mutational analysis showed that the binding of JH III to USP was strongly

reduced by the mutation C472A/H475L.

Although USP has been demonstrated to interact with JHs, further studies on USP did not show

its effect on JH action. The structure-based analysis of a Heliothis virescens USP protein shows

that JHs could fit into the ligand binding pocket (LBP) of USP. However, the percentage of

occupancy of LBP was relatively low, which raised concerns over the validity of USP as an JH

receptor (Sasorith et al., 2002). In vivo USP activation assays in third instar Drosophila larvae

demonstrated that neither natural JHs (JHI, JHII and JHIII) nor JH analogs (pyriproxifen and

methoprene) were able to activate USP, whereas fenoxycarb, a carbamate insecticide that mimics

the action of JH, induced a weak activation. Beck et al. (2009) determined the activation of USP

by JHs and their analogs in transgenic animals expressing either the GAL4-TcUSP (T.

castaneum) or the GAL4-DmUSP (D. melanogaster). Their results show that JHs and their

28

analogs were not able to activate USP. In particular, MF, which displayed a nanomolar affinity to

DmUSP (Jones et al., 2001), did not show any effect on the activation of USP. On the other hand,

pre-incubation of organs with JH III lead to the repression of GAL4-TcUSP and the GAL4-

DmUSP activated by 20E. Thus, USP may not act as an JH receptor, but may interact with JH in

the EcR/USP complex.

3.2 Methoprene-tolerant (Met) as a JH receptor

Methoprene-tolerant (Met) is a basic-helix-loop-helix (bHLH)/Per-Arnt-Sim (PAS) protein

containing a HLH structure and two PAS domains (A and B). The Met gene was first discovered

by an ethyl methane sulfonate mutagenesis screen (Wilson and Fabian, 1986). The Met mutation

conferred a 100-fold-increased resistance to methoprene and JH III, and was resistant to

methoprene-induced pseudotumor formation in larvae and to JH III- or methoprene-induced

vitellogenic oocyte development in adult females, suggesting Met might be a JH receptor. The

hydroxyapatite (HAP) binding assay of JH III to JH III-binding protein in fat body cells revealed

a 10-fold lower binding affinity in the Drosophila Met strain (Shemshedini and Wilson, 1990).

In vitro synthesized Drosophila Met bound to JH III with high affinity (Kd = 5.3 ± 1.5 nM, mean

± SD). And the effectiveness of JHs in the activation of Met expressed in Drosophila S2 cells is

JH III>JH II>JH I>methoprene (Miura et al., 2005). A similar affinity was also found in the

binding of Tribolium Met to JH III (Kd = 2.94 ± 0.68 nM, mean ± SD) (Charles et al., 2011). The

function of Met and the high affinity of Met to JH suggest that Met may act as a JH receptor.

As a potential JH receptor, the ligand-binding properties of Met were determined (Charles et al.,

2011). To examine which part of the Tribolium Met protein is responsible for binding JH III, the

truncated proteins with part of the conserved domain were synthesized for a ligand-binding assay.

29

The results show that Met specifically binds JH III through its C-terminal PAS domain (PAS-B

domain plus C-terminal region). As a bHLH protein, Met requires either a homo- or heterodimer

partner for its activity (Kewley et al., 2004). In A. aegypti, a Ftz-F1-interacting steroid receptor

coactivator (FISC) was identified as a functional partner of Met in mediating JH-induced gene

expression (Li et al., 2011). Microarray analysis and RNAi studies revealed an ortholog of FISC

in T. castaneum, named steroid receptor co-activator (SRC), which was responsible for the

formation of heterodimer with Met (Zhang et al., 2011). The closest relative of FISC/SRC in

Drosophila is Taiman (Tai) (Charles et al., 2011). The study on the interaction between Tai and

Met suggested a model of JH action on Met in Drosophila. Met forms a homodimer in the

absence of JH, while the binding of JH to the PAS-B domain of Met results in conformational

changes to release Met from the homophilic complex and allows it to bind Tai (Charles et al.,

2011). Another JH-dependent heterodimeric partner of Met, Cycle (CYC) was identified in A.

aegypti (Shin et al., 2012). The binding between Met and CYC only occurs in the presence of JH

III, and is induced by JH III in a dose dependent manner. Both Met and CYC specifically binds

to the E-box-like motif from the Kr-h1 gene promoter. Silencing CYC, Met or SRC/FISC using

RNAi, impaired the circadian activation of Kr-h1 and Hairy genes. Based on the previous studies,

it currently appears that the JH receptor is composed of two DNA-binding bHLH/PAS

transcription factors, in which Met is an obligatory component and the partner of Met varies in

different insects.

In Drosophila, the Met mutant showed a high resistance to the toxic and morphogenetic effects

of JHs and their analogs (Shemshedini et al., 1990; Wilson and Ashok, 1998; Wilson and Fabian,

1986). However, although Met-null mutants show reduced oogenesis, they are viable. The

phenotype of complete absence of Met is too subtle to conclude that Met is a genuine JH

30

receptor (Wilson and Ashok, 1998). The possible reasons for the lack of an expected phenotype

are: (1) JH has a weak effect on preadult Drosophila (Wilson and Ashok, 1998), (2) a paralogous

of Met, germ-cell expressed (gce), exists in Drosophila. gce function as a JH receptor in the

absence of Met (Abdou et al., 2011). To determine the clear role of Met, Met was studied in

another insect model, T. castaneum, which only possess one Drosophila Met/gce gene. In this

insect, silencing Met display a more clear effect. The loss of Met in early-instar larvae resulted in

the production of premature pupae or heterochronic larva-pupa intermediates (Konopova and

Jindra, 2007). In addition, the knockdown of Met in the final larval instars disrupted the larval-

pupal ecdysis and induced precocious development of adult structures (Parthasarathy et al.,

2008). In the true bug, Pyrrhocoris apterus, knockdown of Met also results in the similar

phenotype as JH depletion, which causes precocious development of adult color pattern, wings

and genitalia (Konopova et al., 2011). Met regulates premetamorphosis and metamorphosis in

both holometabolous and hemimetabolous insects as in the JH signaling pathway.

JHs not only play important roles in insect metamorphosis, but also regulate the reproduction of

insects. As a key factor in the JH signaling pathway, the function of Met in reproduction has also

been determined. In the migratory locust, Locusta migratoria, silencing Met or Tai results in an

arrest of ovarian development and a reduction in vitellogenin gene expression in the fat body

(Song et al., 2014). In D. punctata, the silencing of Met blocked basal oocyte development,

suppressed the transcription of vitellogenin in the fat body and the uptake of vitellogenin by

ovary. In addition, the typical profile of JH biosynthesis was disrupted in Met-knockdown

animals, which results in the failure of patency (Marchal et al., 2014).

3.3 Krüpel-homologues 1 (Kr-h1) in downstream of JH signaling pathway

31

Kr-h1, a transcription factor with a DNA-binding motif of eight C2H2 zinc fingers, was shown to

be the JH-inducible target of Met. In Drosophila, the expression of Kr-h1 is in JH-dependent

manner and ectopic expression of Kr-h1 caused a phenotype similar to application of JH

(Minakuchi et al., 2008b). In a B. mori cell line, subnanomolar levels of natural JHs were able to

induce BmKr-h1 rapidly (Kayukawa et al., 2012). This induction involves BmMet2 and BmSRC:

JH ligand to BmMet2 and interact with BmSRC to form a JH/BmMet2/BmSRC complex, which

activates BmKr-h1 by interacting with a JH response element (kJHRE). The transcription of Kr-

h1 demonstrates that the induction of BmKr-h1 by JH occurs only in the epidermis of

penultimate-instars, but not in the prepupal stage (Kayukawa et al., 2014).

The role of Kr-h1 in conveying the JH signal to regulate metamorphosis has been studied in both

hemimetabolous and holometabolous insects, including B. germanica (Lozano and Belles, 2011),

P. apterus (Konopova et al., 2011; Smykal et al., 2014b), R. prolixus (Konopova et al., 2011),

D. melanogaster (Minakuchi et al., 2008b), T. castaneum (Minakuchi et al., 2009)

and B. mori (Kayukawa et al., 2014; Smykal et al., 2014b). Kr-h1 represses the adult

morphogenesis. Knocking down Kr-h1 in the third instar or penultimate-instar larvae resulted in

precocious adult development. The repressing action of Kr-h1 on morphogenesis involves the

expression of the broad gene in holometabolous insects (Minakuchi et al., 2009), but not in

hemimetabolous insects (Konopova et al., 2011; Smykal et al., 2014b). Even though the absence

of Kr-h1 is necessary for adult morphogenesis, it is not required to maintain the larval program

during the first two larval instars. The early stages of insect development appear to be initially

independent of JH (Smykal et al., 2014b).

Aside from the function of Kr-h1 in metamorphosis, its roles in reproduction were also examined.

In the migratory locust, L. migratoria, Kr-h1 was demonstrated to convey the JH signal for the

32

induction of Vg. Depletion of Kr-h1 results in a drastic reduction in Vg expression in the fat

body, and subsequently resulted in unsuccessful egg production (Song et al., 2014). However, in

P. apterus, knockdown of Kr-h1 did not block ovarian development or suppress Vg expression in

the fat body(Smykal et al., 2014a). A similar result was also observed in

T. castaneum (Parthasarathy et al., 2010). Depletion of JHAMT or Met caused a significant

reduction in Vg mRNA level, whereas the knockdown of Kr-h1 only caused a 30% reduction.

The downstream signal of JH appears to vary among stages and species.

4. Diploptera punctata

The Pacific beetle cockroach, D. punctata, has proven to be a valuable model insect in the study

of the dynamics of regulation of juvenile hormone (JH) biosynthesis and metabolism,

particularly during late nymph development and reproduction, as a consequence of several other

unique physiological attributes: 1) strikingly high rates of JH biosynthesis compared to other

insects, 2) maintenance of constant in vitro rates of JH biosynthesis, 3) precise and predictable

reproductive events correlated with rates of JH production, 4) uniformity among colony

members of the same age. In addition, Diploptera are easy to rear and handle in the laboratory,

their CA are easily excised and the animals demonstrate high survival rates following surgical

manipulations (Roth and Stay, 1961; Stay, 1999; Tobe and Stay, 1977). Thus, I chose D.

punctata as my study model. In this section, the life cycle and the endocrine control of

development and reproduction of D. punctata will be reviewed.

4.1 Nymphal development

On average, female D. punctata have four nymphal stages, whereas males have three;, however,

the number of nymphal molts can vary depending on conditions (Holbrook and Schal, 1998,

33

2004). Based on observations of our laboratory colony of D. punctata, stadium durations of

females are 13, 14, 16, and 21 days for the first, second, penultimate and final stages,

respectively. Diploptera exhibit sexual dimorphism, and as adults, males are smaller and more

slender than females. In contrast to 64 days of female nymphal life, males generally undergo

imaginal ecdysis sooner, within approximately 50 days, with the first, penultimate and final

stadia lasting 13, 16, and 22 days, respectively (Yagi et al., 1991).

During the growth and development of Diploptera, JH and ecdysteroids play critical roles in

determining nymphal or imaginal developmental pathways. The circulating titer of JH and

ecdysteroids in each stage are shown in Fig. 1.3A (Kikukawa and Tobe, 1986b). In the early

nymphal stadia, higher rates of JH production are found in the latter half of the stadium. In final

instars, JH release declines to undetectable rates after day 10, as hemolymph ecdysteroid titer

starts to rise (Kikukawa and Tobe, 1986a, b). The decline in JH titer and the increase in

circulating ecdysteroids in the final instars allow imaginal ecdysis to occur. During the first 8

days of the penultimate 3rd stadium, JH is critical for maintaining nymphal characters at the next

molt. At this time, reduction in the JH titer results in precocious adult metamorphosis at the next

ecdysis (Kikukawa and Tobe, 1986a). Furthermore, JH and ecdysteroids are critical to regulate

the normal duration of the stadium (Kikukawa, 1989; Kikukawa and Tobe, 1986b). In final

instars, ecdysteroid titer becomes elevated only after JH release declines. How JH interacts with

ecdysteroids to regulate the development of D. punctata was investigated by the application of

JH analog prior to day 10 of last instar females. This resulted in the desynchronization of

ecdysteroid release (Kikukawa, 1989; Kikukawa and Tobe, 1986b; Tobe et al., 1985), possibly

because high JH titer blocks the release or synthesis of ecdysteroids. However, allatectomy on

the first day of the last stadium resulted in a prolongation of low levels of ecdysteroids and the

34

low titer of JH did not promote the release of ecdysteroids (Kikukawa and Tobe, 1986b). Further

investigation is needed to determine the interaction between JH and ecdysteroids in D. punctata.

4.2 Reproduction

After emergence, adult females mate immediately while still teneral, whereas adult males must

be about a week old to successfully court and mate (Stay and Roth, 1958; Woodhead, 1986).

Mating and spermatophore transfer stimulate CA activity. Subsequently, rates of JH biosynthesis

and oocyte growth increase rapidly (Roth and Stay, 1961; Ruegg et al., 1983; Stay and Tobe,

1977). By day 2 of the gonadotrophic cycle vitellogenesis commences (Fig. 1.3B) as Vg, the

precursor of vitellin (Vn), becomes detectable in the hemolymph and fat body. At this time,

oocytes reach about 0.8mm in length and begin to accumulate Vn (Mundall et al., 1981).

Elevated JH titer stimulates this synthesis and uptake, as the denervation of virgin female CA, or

the implantation of mated female CA with ovaries into males, similarly results in Vg synthesis

and uptake by the oocytes (Mundall et al., 1979; Stay and Tobe, 1977). As rates of JH

biosynthesis rise, CA volume and cell number also increase in parallel, reaching a peak around

day 4 to 5 (Szibbo and Tobe, 1981).

35

Figure 1.3. Dynamics of JH metabolism throughout the life cycle of D. punctata. (A) JH release during the development of female

nymphs (as measured in vitro). Dashed lines indicate ecdysteroid titer in the penultimate and final instars. JH critical periods in the

penultimate and final instar are indicated using bold lines. (B) JH release during the first gonadotrophic cycle of mated adult females.

The inset indicates growth of the oocytes during development. (C) JH release by embryonic CA during development in the brood sac.

MF release is indicated by a dashed line. Data based on (Kikukawa and Tobe, 1986b; Stay et al., 2002; Tobe et al., 1985)

36

As oocyte growth and Vn accumulation continues, rates of JH biosynthesis begin to decline

around day 5. When oocytes reach approximately 1.6mm in length, the accumulation of Vn

slows, the spaces between the follicle cells close and the cells begin to deposit the chorion (Stay

et al., 1984; Woodhead et al., 2003). Hemolymph Vg titer also declines at this time (Mundall et

al., 1981). As the oocytes approach their maximal length, 20-hydroxyecdysone becomes

detectable in the oocytes, and levels rise during chorion formation (Stay et al., 1984). Ecdysone

remains in the ovary where ecdysteroid conjugates are likely stored for use by the embryos

during development (Stay unpublished). By day 7, the rate of JH production has declined to day

1 levels (Fig. 1.3B), choriogenesis is complete, and ovulation occurs. As the basal oocytes leave

the common oviduct they are fertilized, pairs of fertilized eggs are then covered by a reduced

oothecal membrane, and finally retracted into the brood sac where embryogenesis proceeds

(Roth, 1970; Woodhead et al., 2003).

Unlike members of many other insect orders, cockroaches demonstrate a clear correlation

between cycles of JH biosynthetic activity by the CA, and patterns of oocyte growth and

vitellogenesis. Oocyte maturation, fat body vitellogenin (Vg) production and the uptake of Vg by

the ovaries are JH-dependent processes in this group (Engelmann, 1979; Rankin and Stay, 1984;

Roth and Stay, 1961; Stay and Tobe, 1978), whereas in higher orders, such as the Diptera,

endocrine regulation of oocyte development requires both JH and ecdysteroids and synthesis of

Vg occurs in both the fat body and ovary (see (Raikhel et al., 2005; Valle, 1993)).

4.3 Embryo development

Embryos, usually 12 to 14, develop within the brood sac for approximately 63 days, and grow

from 1.5mm at oviposition to 6.5mm by parturition. Dorsal closure of the body wall and

37

completion of the gut occurs relatively early, around 12 days after oviposition, when the embryos

are 1.6mm in length. At this time, the CA begin to migrate to their final dorsal position (Stay and

Coop, 1973). Early embryonic CA produce and release the JH precursor, MF, in the virtual

absence of JH, from day 16 until day 26 when embryos reach 2.6mm (Fig. 1.3C). During this

time, allatostatins (ASTs), a family of regulatory neuropeptides (see below) which inhibit JH

production in all post-embryonic developmental stages, stimulate MF and JH biosynthesis (Stay

et al., 2002). Surprisingly, this is similar to the function of these peptides in some crustaceans

(Kwok et al., 2005). The mechanism by which this occurs merits further study in Diploptera as

these finding may have important implications for the evolutionary history of these peptides

(Hult et al., 2008).

Unlike ovoviviparous blaberids, D. punctata is truly viviparous; females provide nourishment to

the embryos via nutritive milk, rich in carbohydrates and lipocalin-like proteins that is secreted

in large quantities by glandular cells in the brood sac epithelium. Secretion begins shortly after

dorsal closure. Once the gut is completely formed, the embryos start to drink the fluid and begin

to increase dramatically in dry weight (Stay and Coop, 1973; Stay and Coop, 1974; Williford et

al., 2004). Elevated levels of JH can disrupt milk production and trigger a decline in production,

but the humoral factor which promotes competence for milk production remains unidentified

(Evans and Stay, 1989, 1995; Stay and Lin, 1981). Currently, 25 Milk protein cDNAs have been

identified; these genes exist in multiple copies at several loci. Further study of this gene family is

critical, and may have implications for the evolution of viviparity in insects (Williford et al.,

2004).

Several days before parturition, a second cycle of oocyte growth begins, and re-mating is not

required as females store sperm after copulation (Mundall et al., 1981; Rankin and Stay, 1985).

38

As a consequence of viviparity Diploptera oviposit fewer oocytes, less frequently than do

oviparous species. However, nymphs are more developed at birth and reach reproductive

maturity more quickly, with fewer nymphal molts, than other blaberids (Roth, 1970; Stay and

Coop, 1973).

5 Regulation of JH titre

As a hormone which plays essential roles in insects, JH is tightly regulated to maintain the

development, metamorphosis and reproduction of insects. The circulating JH titer in the

hemolymph is determined not only by the activity of the CA, but also by other processes, such as

enzymatic degradation, binding to carrier proteins and uptake by target organs. In this section,

the factors involved in the regulation of JH titre in the CA will be summarized, mainly focusing

on the regulation in D. punctata.

CA are connected with the corpora cardiaca (CC) via nervi corporis allati (NCA) I, and in turn,

through the CC with the brain via the nervi corporis cardiaci (NCC) I and II to medial and lateral

neurosecretory cells of the brain respectively (Tobe and Stay, 1985a). JH, which is produced in

the CA, is regulated by both neural and humoral inputs, including neuropeptides,

neurotransmitters, ecdysteroids and even JH itself.

5.1 Allatostatins

ASTs are neuropeptides originally described to inhibit JH biosynthesis rapidly and reversibly in

insects. Their existence was originally hypothesized by Tobe and Stay (1980) based on the

relationship between JH biosynthesis and oocyte length. Aside from their ability to inhibit JH

biosynthesis, ASTs have also been demonstrated to suppress muscular activity, inhibit

vitellogenesis and modulate the activity of certain midgut digestive enzymes (Stay and Tobe,

39

2007). In insects, ASTs were initially grouped into three families, YXFGL-amide-allatostatins

(FLGa/ASTs), W2W9-allatostatins (MIP/ASTs) and PISCF-allatostatins (PISCF/ASTs) (Coast

and Schooley, 2011; Stay and Tobe, 2007). However, MIP/AST was originally characterized to

be a myoinhibitory peptide family, and therefore, the regulation of JH biosynthesis will be

discussed only in FLGa/ASTs and PISCF/ASTs (Coast and Schooley, 2011).

The FLGa/AST was first discovered from brain extracts of D. punctata (Pratt et al., 1989; Stay

and Tobe, 2007; Woodhead et al., 1989). In D. punctata, thirteen AST peptides are encoded by a

single neuropeptide precursor (Donly et al., 1993). Members of the FGLa family of ASTs share a

core C-terminal pentapeptide region (Tyr/Phe-Xaa-Phe-Gly-Leu/Ile-NH2) which is the main

functional region needed for inhibition of JH biosynthesis (Pratt et al., 1991; Stay et al., 1991).

Studies have demonstrated the pleiotropic nature of these peptides and their wide distribution.

ASTs have been found in many cells of the central nervous system, gut and ovary in D. punctata

(Garside et al., 2002; Stay et al., 1994b; Woodhead et al., 2003). A well-documented role for

FLGa/AST is the inhibitory effect on JH biosynthesis by the CA (Pratt et al., 1989; Stay et al.,

1994a; Woodhead et al., 1994; Woodhead et al., 1989), but they can also function as potent

inhibitors of muscle contraction in the gut (Lange et al., 1995; Stay et al., 1994b), stimulate

carbohydrate-metabolizing enzymes in the midgut and inhibit Vg production and cardiac activity

(Fuse et al., 1999; Garside et al., 2002; Martin et al., 1996; Vilaplana et al., 1999; Yu et al.,

1995b). Physiological and molecular work in D. punctata has confirmed the inhibitory function

of FGLa AST (for review see (Stay and Tobe, 2007)). This effect occurs through paracrine

release of FLGa/AST within the CA from cells in the pars lateralis of the brain sending axons in

the NCC II, to the CA (Stay et al., 1992). The expression of the FLGa/AST precursor in the

whole brain is negatively correlated with JH biosynthesis by the CA (Garside et al., 2003).

40

Furthermore, a negative correlation was also observed between sensitivity of CA to ASTs and JH

biosynthesis during the first gonadotrophic cycle of mated adult females. The sensitivity of the

CA to ASTs declines during vitellogenesis and increases again before choriogenesis occurs (Pratt

et al., 1990; Stay et al., 1991; Unnithan and Feyereisen, 1995).

Putative FLGa/AST receptors were characterised using affinity labeling and radio-ligand binding

assays in the CA and the brain (Cusson et al., 1991a, 1992a; Yu et al., 1995a). More recently, a

putative AST receptor was cloned belonging to the G-protein coupled receptor family of

mammalian galanin receptors. Its relative expression level shows a steep rise on day 6 in the CA

and the brain of the adult female, after which the level declines again. This downregulation is

consistent with the loss of sensitivity of CA to AST following ovulation (Pratt et al., 1990; Stay

et al., 1991; Unnithan and Feyereisen, 1995). The peak on day 6 suggests that AST and its

receptor may be responsible for the decline of JH biosynthesis at this time. Initial RNAi

experiments showed that silencing of the putative AST receptor results in an increase in JH

production (Lungchukiet et al., 2008a; Lungchukiet et al., 2008b). The signal transduction

pathway downstream of the FLGa/AST receptor likely involves the 1,4,5-inositol

triphosphate(IP3)/Diacylglycerol (DAG) pathway, as well as protein kinase C (PKC) (Rachinsky

et al., 1994). The IP3/DAG pathway is initiated upon binding of a neuropeptide to a G protein

bound to the membrane. Increased IP3 induces release of intracellular Ca2+, whereas DAG

activates PKC. Activators of PKC (phorbol esters) are potent inhibitors of JH III biosynthesis in

D. punctata (Feyereisen and Farnsworth, 1987b). Steps early in the JH pathway are likely the

target of AST, more specifically the export of citrate from the mitochondria or the cytosolic

conversion to acetyl-CoA, which will then enter the mevalonate pathway (Sutherland and

Feyereisen, 1996; Wang et al., 1994). However, more work is needed to further elucidate the

41

signal transduction pathway and to clearly identify the link between the AST receptor and the

exact target of this neuropeptide in JH biosynthesis.

The insect PISCF/AST was first identified in the tobacco hornworm, M. sexta (Kramer et al.,

1991). PISCF/AST is a 15 amino acid peptide with the sequence qIRYRQCYFNPISCF, with a

blocked N-terminus (pGlu), a free C-terminus, and a disulphide bridge linking two Cys residues

at positions 7 and 14. Using HPLC- MALDI-TOF, PISCF/AST was isolated and characterized in

the female mosquito A. aegypti (Li et al., 2006). AePISCF/AST inhibits JH biosynthesis in dose-

dependent manner, with maximal inhibition in the nanomolar range. This inhibitory effect can be

rescued by farnesoic acid, indicating that the target of AePISCF/AST is located before the

formation of farnesoic acid. PISCF/AST plays important roles in regulating JH biosynthesis in A.

aegypti. The levels of AePISCF/AST in the brain correlated with the sensitivity of CA to AST

and the activity of CA in JH biosynthesis (Li et al., 2006). However, in other insects, such as D.

melanogaster, PISCF/ASTs did not appear to regulate JH biosynthesis, but rather acts as a

myotropin (Price et al., 2002). In A. pisum and M. persicae, PISCF/AST shows a significant

dose-dependent feeding suppression effect, resulting in mortality, reduced growth and fecundity

(Down et al., 2010; Matthews et al., 2010).

Two PISCF/AST receptor (PISCF/AstR) genes were identified in D. melanogaster using a

reverse pharmacological approach. They are G-protein-coupled receptors, which are the insect

homologs of the mammalian opioid/somatostatin receptors. Site-directed mutagenesis studies

demonstrated that a residue in transmembrane region 3 and the loop between transmembrane

regions 6 and 7 affect ligand binding (Kreienkamp et al., 2002). Two PISCF/AstR paralogs were

also isolated in A. aegypti (Mayoral et al., 2010). The tissue distribution of the two PISCF/AstR

shows that both genes are expressed highest in the abdominal ganglia, whereas the expression

42

differs in other tissues, such as Malpighian tubules. In Malpighian tubules, PISCF/AstRB

displays high expression, whereas the transcript level of PISCF/AstRA is relatively low.

Developmental expression of PISCF/AstR mRNA also displays different patterns, which

suggests that the pleiotropic effects of PISCF/AST in mosquitoes might be mediated by the

different receptor paralogs (Mayoral et al., 2010). The PISCF/AstR identified in T. castaneum

could be activated by TcAST and T. castaneum allatostatin double C (Trica-ASTCC) as well as

M. sexta PISCF/AST in a dose-dependent manner (Audsley et al., 2013). TcPISCF/AstR is

widely distributed, with the highest transcript level in the head and the gut. Whole mount

immunocytochemistry localised TcPISCF/AstR in the median and lateral neurosecretory cells of

the brain, in the CC, throughout the ventral nerve cord and in midgut neurosecretory cells, but

not in Malpighian tubules, indicating that TcPISCF/AstR is close to PISCF/AstRA in A. aegypti.

5.2 Allatotropin

Allatotropin (AT) was first purified from the extracts of heads of pharate adult M. sexta (Kataoka

et al., 1989). This 13-residue peptide (GFKNVEMMTARGF-NH2) was shown to stimulate JH

biosynthesis in the CA of adult female M. sexta, but not in the CA of larval M. sexta, the

mealworm beetle Tenebrio molitor, the grasshopper S. nitens and the cockroach Periplaneta

americana (Kataoka et al., 1989). Further studies demonstrated that Manse-AT could also be

active on the CA in other adult insects, including the Lepidoptera Heliothis virescens (Teal,

2002), Spodoptera frugiperda (Oeh et al., 2000), and Lacanobia oleracea (Audsley et al., 1999).

In addition, Manse-AT was also found to stimulate JH biosynthesis in the larvae of L. oleracea

(Audsley et al., 2000), in Diptera Phormia regina (Tu et al., 2001) and in Orthoptera Romalea

microptera (Li et al., 2005).

43

Much work has attempted to identify other ATs. A putative AT has been extracted from the

subesophageal ganglion of male cricket G. bimaculatus. The extract was able to stimulate JH

biosynthesis in CA from G. bimaculatus and another cricket, Acheta domesticus (Lorenz and

Hoffmann, 1995). Extracts of the brain-subesophageal ganglion-CC-CA from P. apterus, also

display activity in the stimulation of JH biosynthesis (Hodkova et al., 1996). A methanolic

extract of the suboesophageal ganglia (SOG)-CC of the Mythimna loreyi virgin males stimulates

the synthesis of JH III acid and iso-JH II, whereas synthetic Manse-AT had no significant effect

(Kou and Chen, 2000). Unfortunately, ATs in G. bimaculatus or P. apterus have not been

isolated and identified. Another AT was isolated and identified from the abdominal ganglia of

the mosquito A. aegypti (Veenstra and Costes, 1999). The sequence of this peptide was

determined to be APFRNSEMMTARGF-NH2. The cDNA clone encoding this novel

neuropeptide was shown to encode a single copy of this peptide. AT in A. aegypti (refer as AaAT)

has a stimulatory effect on the JH biosynthesis of adult female CA, and it appears that the

stimulation results from the increased ability of CA to convert FA to JH III induced by AaAT (Li

et al., 2003). AaAT was also shown to induce JH biosynthesis in CA from newly emerged

females (Li et al., 2003).

In M. sexta, Manse-AT was shown to stimulate JH biosynthesis in the CA of adult females, but

not in larval or pupal CA of M. sexta. However, Manse-AT mRNA and immunoreactivity

showed the expression of Manse-AT not only in the nervous system of adults but also in that of

larvae, indicating that Manse-AT may play multiple roles in insects (Bhatt and Horodyski, 1999).

Later studies showed that in addition to regulating JH biosynthesis by the CA, AT are

multifunction in different insects, including the stimulation of myoactivity (Bhatt and Horodyski,

1999; Paemen et al., 1991; Rudwall et al., 2000), inhibition of midgut ion transport (Lee et al.,

44

1998), stimulation of foregut movement (Duve et al., 2000; Duve et al., 1999), and

cardioacceleratory effects (Koladich et al., 2002).

5.3 Neurotransmitters regulating JH biosynthesis

Several neurotransmitters are involved in the regulation of JH biosynthesis in insects.

Octopamine inhibits JH biosynthesis in the CA of D. punctata in vitro (Thompson et al., 1990).

In B. germanica, dopamine can have a stimulatory or inhibitory effect on the CA depending on

the stage of the ovarian cycle (Pastor et al., 1991). In M. sexta, dopamine stimulates JH

biosynthesis in CA from the first 2 days of the last larval stadium, but inhibits in CA from larvae

in the beginning of the prepupal period. The stimulatory or inhibitory effect of dopamine is

related to the the adenylyl cyclase system of CA (Granger et al., 1996). Ontogenetic differences

in the control of JH biosynthesis by dopamine have also been demonstrated in Drosophila

females in inhibiting or stimulating JH degradation (Gruntenko et al., 2005). The excitatory

neurotransmitter L-glutamate, which acts through ionotropic receptors to raise intracellular Ca2+

concentration, stimulates JH biosynthesis in the CA of D. punctata through the action of NMDA-,

kainate- and/or quisqualate-sensitive subtypes of ionotropic L-glutamate receptors (Chiang et al.,

2002b; Pszczolkowski et al., 1999).

5.4 Second messengers

The signal transduction pathways for neuropeptides and neurotransmitters usually involve

intracellular reaction cascades regulating levels of second messengers. The effect on JH

biosynthesis has been measured for several of these messengers.

Calcium is considered unique as a second messenger in the regulation of JH biosynthesis in D.

punctata, since intracellular Ca2+ can function as a second messenger on its own, and, at the same

45

time, Ca2+ is essential for the function of cyclic nucleotides and the IP3/DAG second messenger

system. Incubation of CA in medium lacking Ca2+, and blockage of non-specific Ca2+ channels,

inhibits JH release. Because no buildup of JH or MF occurs in the CA as a consequence of such

blockage, it seems likely that Ca2+ affects overall JH biosynthesis, instead of release (Kikukawa

et al., 1987). JH biosynthesis rises as Ca2+ concentrations are increased in vitro. Moreover,

elevated levels of extracellular Ca2+ can counteract the effect of brain extracts (containing AST)

on JH biosynthesis (Aucoin et al., 1987). Two types of Ca2+ channels have been proposed for this

action: 1) Ca2+ enters the CA cells through voltage-gated Ca2+ channels ((McQuiston et al., 1990;

Thompson and Tobe, 1986), as reviewed by (Rachinsky and Tobe, 1996)) or 2) there is a

glutamate-induced Ca2+ influx via an ionotropic L-glutamate receptor, later reported to be the

NMDA receptor (Chiang et al., 2002a; Chiang et al., 2002b; Pszczolkowski et al., 1999). Release

of Ca2+ from intracellular stores also appears to stimulate JH biosynthesis in vitro. The Ca2+-

ATPase inhibitor thapsigargin was shown to increase intracellular Ca2+ and induce JH

biosynthesis (Rachinsky and Tobe, 1996; Rachinsky et al., 1994).

Compounds that increase intracellular cAMP concentrations were found to inhibit JH

biosynthesis in an in vitro assay (Meller et al., 1985). The addition of brain extracts showed a

dose-dependent elevation of cAMP levels (Aucoin et al., 1987). Together, these results suggest

that cAMP could act as a second messenger for AST from the brain. A later study by Cusson et

al. (Cusson et al., 1992b) measured levels of cGMP and cAMP in virgin and mated females

following addition of AST. The mechanism of AST inhibition does not involve either cGMP or

cAMP, suggesting that cyclic nucleotides are likely second messengers of another inhibitory

signal affecting JH biosynthesis. A possible candidate is the neurotransmitter octopamine, which

46

in high concentrations can induce a rise in the level of cAMP in vitro (see above and (Thompson

et al., 1990)).

5.5 Ovarian factor

The normal pattern of JH biosynthesis in the CA of D. punctata during the first gonadotrophic

cycle of mated females requires the presence of the ovary (Rankin and Stay, 1984; Stay and

Tobe, 1978; Stay et al., 1983). Ovariectomy of females soon after the adult molt resulted in a low

level of JH biosynthesis throughout the first gonadotrophic cycle (Stay and Tobe, 1978). The

cyclic pattern in JH biosynthesis could be rescued following implantation into females of at least

one-half an ovary from a day 0 mated female (Stay et al., 1983). Similarly, implantation of

ovarioles with vitellogenic basal oocytes into male animals with denervated CA also resulted in

an increase in JH biosynthesis (Hass et al., 2003; Rankin and Stay, 1984).

The effect of the ovary on JH biosynthesis appears to be stage-dependent. In a study by

Sutherland et al. (2000), young ovaries were found to have a dual role: stimulation of JH

biosynthesis and repression of the mRNA levels of CYP4C7, a cytochrome P450 enzyme coding

a gene involved in JH catabolism. These roles were reversed in post-vitellogenic ovaries. Rankin

and Stay (1984) have shown that the size of the basal oocyte is a good indicator of the degree

with which the ovary can stimulate JH biosynthesis. The ovary acquires this ability at the start of

vitellogenesis but loses it post-vitellogenesis.

Currently, the exact nature of the ovarian factors involved in this modulation of JH production

remains unknown. A study by Unnithan et al. (1998) suggested that a factor produced by ovaries

in all stages can induce stable stimulation of JH synthesis. The effect of this ovarian factor is

antagonized by a factor from the brain, distinct from ASTs (Unnithan et al., 1998). In contrast, a

47

study by Elliott et al. (Elliott et al., 2006) provided evidence for a different stage-specific

peptidergic factor from the ovary that stimulates JH production. This ovarian factor is produced

when basal oocyte lengths range from 0.76 to 1.15 mm (day 2 and 3 adult females). Ovaries with

oocytes that are larger were no longer stimulatory in in vitro assays. This factor is considered the

same as that described earlier by Rankin and Stay (1984), but to date, the exact peptidergic

structure remains unknown.

An inhibitory peptidergic factor was reported from post-vitellogenic ovaries, which induced a

rise in the cGMP content of CA in vitro (Chang et al., 2005). Elevated cGMP signaling was

suggested to cause a long-term arrest on JH biosynthesis indirectly, through atrophy of the CA

cells, reduction in protein content and depletion of the cellular machinery. This factor is thought

to be the same post-vitellogenic ovary factor described by Sutherland et al. (2000) controlling the

expression of CYP4C7. However, this peptide is not likely to be AST since the quantities of AST

detected in the hemolymph are too low to exert a physiological effect on the CA (Stay et al.,

1994a). ASTs in the ovary probably act locally by facilitating ovulation or by preventing

glycosylation of Vg and its release from the fat body (Garside et al., 2002; Rankin and Stay,

1984; Woodhead et al., 2003).

5.6 JH feedback

Negative feedback regulation of an endocrine gland, induced by its own product is a well-known

endocrine control mechanism. A direct or indirect suppression of JH biosynthesis by the CA was

first shown in a study assessing the effect of a JH analog. Topical application of this analog

suppressed JH biosynthesis in a dose dependent fashion as did topical treatment with JH itself

(Tobe and Stay, 1979). A subsequent study suggested an indirect feedback of JH on the CA

48

through the brain, by comparing the effect of JH analog on males with CA connected or

disconnected with the brain (Stay et al., 1994a). In both cases, the rates of JH biosynthesis were

reduced compared to controls, intact CA more so than denervated ones. However, in animals in

which the CA were disconnected from the brain, AST levels in the hemolymph were greatly

elevated. These results indicate that the exogenous JH analog acts on the brain to inhibit JH

biosynthesis, by paracrine release of AST in the intact CA and through the hemolymph in the

disconnected CA (Ruegg et al., 1983; Stay et al., 1994a).

In vertebrates, the bulk end-product of mevalonate pathway, cholesterol, has a feedback

regulation on the transcription and post-transcription of enzymes in the mevalonate pathway,

which in turn mediates the production of cholesterol (Chang and Limanek, 1980; Clarke et al.,

1987; Goldstein and Brown, 1990). The addition of farnesol to cultured cells accelerated the

degradation of HMGR, while the addition of mevalonate raised the activity of FPPP expressed in

CHO cells (Meigs et al., 1996; Meigs and Simoni, 1997). In insects, the accumulation of JH

precursor or depletion of one enzyme in the pathway also results in a change in the activity of

other enzymes. In CA with low FALD activity, farnesal accumulated and was converted back to

farnesol that leaks from the CA, to balance the enzyme activity and the JH precursor pool

(Rivera-Perez et al., 2013). The expression and activity of enzymes in the JH biosynthetic

pathway appears to be involved in JH feedback regulation of JH titre, since the rate of JH

synthesis is regulated by the rate of flux of isoprenoids and the expression of genes in the

pathway (Details seen Section 1.2). However, further investigation is required to prove this

hypothesis.

5.7 Ecdysteroids

49

Injection of 20-hydroxyecdysone (20E) reduced oocyte growth and Vn content of basal oocytes

(Friedel et al., 1980) and inhibited JH biosynthesis in D. punctata (Stay et al., 1980). However,

no inhibitory effect was observed following treatment of CA with 20E in vitro (Paulson and

Stay, 1987). In M. sexta, incubation of brain-CC-CA complexes with 20E resulted in a reduction

in the production of JH acid (Granger and Janzen, 1987). The inhibitory effect of 20E in JH

biosynthesis may be related to dopamine: 20E down regulates dopamine, which is responsible

for the stimulation of JH biosynthesis (Granger et al., 1996). On the other hand, in D. punctata, a

significant rise in DNA synthesis in CA from male adults treated with 20E suggested that

ecdysteroids may affect JH biosynthesis by regulating proliferation of CA cells (Chiang et al.,

1995; Tsai et al., 1995). However, this conflicts with the observed 20E-induced decrease in CA

cell number of ovariectomized D. punctata females (Tobe et al., 1984). The pathway through

which ecdysteroids inhibit JH production by the CA is unclear.

6 Rationale and objectives of my study

The juvenile hormones are essential in regulating insect growth, development, metamorphosis,

aging, caste differentiation and reproduction. The important roles of JH in insect development

and reproduction and its value in pest control have led to multiple studies on the biosynthesis,

regulation and function of JH. Among cockroach families, endocrine regulatory mechanisms are

easier to discern in members of the Blaberidae. In these species, only the basal oocyte of each

ovariole undergoes vitellogenesis during a given gonadotrophic cycle, and a long period of

ovarian quiescence occurs while the oviposited fertilized eggs develop in the brood sac of the

female (Roth, 1970). Among these species, D. punctata (Dictyoptera; Blattaria; Blaberidae), the

only known truly viviparous species in that the brood sac provides nourishment to the embryos

50

(Roth, 1970), has become a model system for studies on reproduction and the regulation of JH

production.

Recently, the JH biosynthetic pathway has been established in insects such as B. mori (Kinjoh et

al., 2007), A. aegypti (Nouzova et al., 2011) and A. mellifera (Bomtorin et al., 2014). However,

there are fewer studies on the JH biosynthetic pathway in D. punctata. Thus, in Chapter 2 of this

thesis, the genes involved in the JH biosynthetic pathway were identified and characterized. This

is the first report of the characterization of the JH biosynthetic pathway in hemimetabolous

insects. The function of JH as a regulator of female cockroach reproduction, and the role of JH

biosynthetic enzymes in regulating JH biosynthesis were established. In particular, our research

revealed an interaction between the transcriptions of genes in the biosynthetic pathway i.e. a

feedback mechanism is involved in the regulation of the expression of enzymes in the JH

biosynthetic pathway.

In terms of the regulation of JH biosynthesis, ASTs have been demonstrated to be one of the

essential inhibitory factors, particularly for D. punctata, in which no AT or other allatoregulatory

neuropeptides have been discovered to date. Although the function of AST in the regulation of

JH has been well-studied, the mode of action of AST remains unclear. In Chapter 3, the mode of

action of AST was studied by determining the signaling pathway in AST action, and the target of

AST action. DpAstR was expressed in multiple vertebrate cell lines, and the activation activities

of the 13 ASTs were determined. Furthermore, my study showed that the activation of AstR

results in the elevation of intracellular calcium and cAMP, which indicates that Ca2+ and cAMP

can server as second messenger in the signaling pathway of AST action. In addition, the target of

AST action was studied using RNAi and the measurement of JH in the presence or absence of JH

51

precursors. Our result indicates that AST probably affects JH biosynthesis prior to the entry of

Acetyl-CoA into the JH biosynthetic pathway.

As a second messenger, Ca2+ has been demonstrated to modulate/regulate JH biosynthesis in D.

punctata. NMDAR is an ionotropic receptor with a high Ca2+ permeability, and its activators

glutamate and NMDA have been shown to stimulate JH biosynthesis, probably by activating the

NMDAR channel. Previous studies also showed that NMDA stimulates JH biosynthesis in vitro

in the D. punctata (Chiang et al., 2002a). Thus, it was of interest to determine the function of

NMDAR in JH biosynthesis and reproduction. In chapter 4, the genes encoding NMDAR were

identified in D. punctata and characterized in this thesis. The function of NMDAR in

reproduction was also examined by knocking down the genes encoding NMDAR subunit 2, and

by the application of NMDAR antagonist MK-801. However, neither JH biosynthesis nor oocyte

growth was affected, suggesting that NMDAR does not play important roles in the regulation of

JH biosynthesis or reproduction in female D. punctata.

52

References

Abdou, M.A., He, Q., Wen, D., Zyaan, O., Wang, J., Xu, J., Baumann, A.A., Joseph, J., Wilson, T.G., Li, S.,

2011. Drosophila Met and Gce are partially redundant in transducing juvenile hormone action. Insect Biochem

Mol Biol 41, 938-945.

Allen, K.N., Dunaway-Mariano, D., 2004. Phosphoryl group transfer: evolution of a catalytic scaffold. Trends

Biochem Sci 29, 495-503.

Aucoin, R.R., Rankin, S.M., Stay, B., Tobe, S.S., 1987. Calcium and cyclic AMP Involvement in the

regulation of juvenile hormone biosynthesis in Diploptera punctata. Insect Biochem 17, 965-969.

Audsley, N., Vandersmissen, H.P., Weaver, R., Dani, P., Matthews, J., Down, R., Vuerinckx, K., Kim, Y.J.,

Vanden Broeck, J., 2013. Characterisation and tissue distribution of the PISCF allatostatin receptor in the red

flour beetle, Tribolium castaneum. Insect Biochem Mol Biol 43, 65-74.

Audsley, N., Weaver, R.J., Edwards, J.P., 1999. Juvenile hormone synthesis by corpora allata of tomato moth

Lacanobia oleracea (Lepidoptera: Noctuidae), and the effects of allatostatins and allatotropin in vitro. . Eur J

Entomol 96, 287–293.

Audsley, N., Weaver, R.J., Edwards, J.P., 2000. Juvenile hormone biosynthesis by corpora allata of larval

tomato moth, Lacanobia oleracea, and regulation by Manduca sexta allatostatin and allatotropin. Insect

Biochem Mol Biol 30, 681-689.

Ayte, J., Gil-Gomez, G., Haro, D., Marrero, P.F., Hegardt, F.G., 1990. Rat mitochondrial and cytosolic 3-

hydroxy-3-methylglutaryl-CoA synthases are encoded by two different genes. Proc Natl Acad Sci U S A 87,

3874-3878.

Baker, F.C., Lanzrein, B., Miller, C.A., Tsai, L.W., Jamieson, G.C., Schooley, D.A., 1984. Detection of only

JH III in several life-stages of Nauphoeta cinerea and Thermobia domestica. Life Sciences 35, 1553-1560.

Baker, F.C., Lee, E., Bergot, B.J., Schooley, D.A., 1981. Isomerization of isopentenyl pyrophosphate and

homoisopentenyl pyrophosphate by Manduca sexta corpora cardiaca-corpora allata homogenates. Dev

Endocrinol 15, 67-80.

Baker, F.C., Mauchamp, B., Tsai, L.W., Schooley, D.A., 1983. Farnesol and farnesal dehydrogenase(s) in

corpora allata of the tobacco hornworm moth, Manduca sexta. J Lipid Res 24, 1586-1594.

Baker, F.C., Schooley, D.A., 1981. Biosynthesis of 3-hydroxy-3-methylglutaryl-CoA, 3-hydroxy-3-

ethylglutaryl-CoA, mevalonate and homomevalonate by insect corpus allatum and mammalian hepatic tissues.

Biochim Biophys Acta 664, 356-372.

Baker, F.C., Tsai, L.W., Reuter, C.C., Schooley, D.A., 1988. The absence of significant levels of the known

juvenile hormones and related compounds in the milkweed bug, Oncopeltus fasciatus. Insect Biochem 18,

453–462.

Beck, Y., Delaporte, C., Moras, D., Richards, G., Billas, I.M., 2009. The ligand-binding domains of the three

RXR-USP nuclear receptor types support distinct tissue and ligand specific hormonal responses in transgenic

Drosophila. Dev Biol 330, 1-11.

Belles, X., Martin, D., Piulachs, M.D., 2005. The mevalonate pathway and the synthesis of juvenile hormone

in insects. Annu Rev Entomol 50, 181-199.

Bergot, B.J., Baker, F.C., Cerf, D.C., Jamieson, G., Schooley, D.A., 1981. Qualitative and quantitative aspects

of juvenile hormone titers in developing embryos of several insect species: discovery of a new JH-like

substance extracted from eggs of Manduca sexta, in: Pratt, G.E., Brooks, G.T. (Eds.), Juvenile Hormone

Biochemistry, pp. 33-45.

Bergot, B.J., Baker, F.C., Lee, E., Schooley, D.A., 1979. Absolute configuration of homomevalonate and 3-

hydroxy-3-ethylglutaryl- and 3-hydroxy-3-methylglutaryl CoA, produced by cell-free extracts of insect

53

corpora allata; cautionary note on prediction of absolute stereochemistry based on liquid chromatographic

elution order of diastereomeric derivatives. J Am Chem Soc 101, 7432–7434.

Bhatt, T.R., Horodyski, F.M., 1999. Expression of the Manduca sexta allatotropin gene in cells of the central

and enteric nervous systems. J Comp Neurol 403, 407-420.

Bochar, D.A., Stauffacher, C.V., Rodwell, V.W., 1999. Sequence comparisons reveal two classes of 3-

hydroxy-3-methylglutaryl coenzyme A reductase. Mol Genet Metab 66, 122-127.

Bomtorin, A.D., Mackert, A., Rosa, G.C., Moda, L.M., Martins, J.R., Bitondi, M.M., Hartfelder, K., Simoes,

Z.L., 2014. Juvenile hormone biosynthesis gene expression in the corpora allata of honey bee (Apis mellifera

L.) female castes. PLoS One 9, e86923.

Borovsky, D., Carlson, D.A., Hancock, R.G., Rembold, H., van Handel, E., 1994. De novo biosynthesis of

juvenile hormone III and I by the accessory glands of the male mosquito. Insect Biochem Mol Biol 24, 437-

444.

Bowers, W.S., Marsella, P.A., Evans, P.H., 1983. Identification of an hemipteran juvenile hormone: in vitro

biosynthesis of JH III by Dysdercus fasciatus. J Exp Zool 228, 555–559.

Buesa, C., Martínez-Gonzalez, J., Casals, N., Haro, D., Piulachs, M.D., Bellés, X., Hegardt, F.G., 1994.

Blattella germanica has two HMG-CoA synthase genes. Both are regulated in the ovary during the

gonadotrophic cycle. J Biol Chem 269, 11707–11713.

Bürgin, C., Lanzrein, B., 1988. Stage dependent biosynthesis of methyl farnesoate and juvenile hormone III

and metabolism of juvenile hormone III in embryos of the cockroach, Nauphoeta cinera. . Insect Biochem 18,

3–9.

Burtenshaw, S.M., Su, P.P., Zhang, J.R., Tobe, S.S., Dayton, L., Bendena, W.G., 2008. A putative farnesoic

acid O-methyltransferase (FAMeT) orthologue in Drosophila melanogaster (CG10527): relationship to

juvenile hormone biosynthesis? Peptides 29, 242-251.

Bylemans, D., Borovsky, D., Ujvary, I., De Loof, A., 1998. Biosynthesis and regulation of Juvenile Hormone

III, Juvenile Hormone III Bisepoxide, and Methyl Farnesoate during the reproductive cycle of the grey fleshfly,

Neobellieria (Sarcophaga) bullata. Arch Insect Biochem Physiol 37, 248-256.

Cabano, J., Buesa, C., Hegardt, F.G., Marrero, P.F., 1997. Catalytic properties of recombinant 3-hydroxy-3-

methylglutaryl coenzyme A synthase-1 from Blattella germanica. Insect Biochem Mol Biol 27, 499-505.

Cao, L., Zhang, P., Grant, D.F., 2009. An insect farnesyl phosphatase homologous to the N-terminal domain of

soluble epoxide hydrolase. Biochem Biophys Res Commun 380, 188-192.

Carrigan, C.N., Poulter, C.D., 2003. Zinc is an essential cofactor for type I isopentenyl

diphosphate:dimethylallyl diphosphate isomerase. J Am Chem Soc 125, 9008-9009.

Casals, N., Buesa, C., Marrero, P.F., Belles, X., Hegardt, F.G., 2001. 3-Hydroxy-3-methylglutaryl coenzyme A

synthase-1 of Blattella germanica has structural and functional features of an active retrogene. Insect Biochem

Mol Biol 31, 425-433.

Casals, N., Buesa, C., Piulachs, M.D., Cabano, J., Marrero, P.F., Belles, X., Hegardt, F.G., 1996. Coordinated

expression and activity of 3-hydroxy-3-methylglutaryl coenzyme A synthase and reductase in the fat body of

Blattella germanica (L.) during vitellogenesis. Insect Biochem Mol Biol 26, 837-843.

Castillo-Gracia, M., Couillaud, F., 1999. Molecular cloning and tissue expression of an insect farnesyl

diphosphate synthase. Eur J Biochem 262, 365-370.

Chang, E.S., Chang, S.A., Mulder, E.P., 2001. Hormones in the lives of crustaceans: an overview. Am Zool 41,

1090–1097.

Chang, L.W., Tsai, C.M., Yang, D.M., Chiang, A.S., 2005. Cell size control by ovarian factors regulates

juvenile hormone synthesis in corpora allata of the cockroach, Diploptera punctata. Insect Biochem Mol Biol

35, 41-50.

54

Chang, T.Y., Limanek, J.S., 1980. Regulation of cytosolic acetoacetyl coenzyme A thiolase, 3-hydroxy-3-

methylglutaryl coenzyme A synthase, 3-hydroxy-3-methylglutaryl coenzyme A reductase, and mevalonate

kinase by low density lipoprotein and by 25-hydroxycholesterol in Chinese hamster ovary cells. J Biol Chem

255, 7787-7795.

Charles, J.P., Iwema, T., Epa, V.C., Takaki, K., Rynes, J., Jindra, M., 2011. Ligand-binding properties of a

juvenile hormone receptor, Methoprene-tolerant. Proc Natl Acad Sci U S A 108, 21128-21133.

Chayet, L., Pont-Lezica, R., George-Nascimento, C., Cori, O., 1973. Biosynthesis of sequiterpene alcohols and

aldehydes by cell free extracts from orange flavedo. . Phytochemistry 12, 95–101.

Chiang, A.S., Lin, W.Y., Liu, H.P., Pszczolkowski, M.A., Fu, T.F., Chiu, S.L., Holbrook, G.L., 2002a. Insect

NMDA receptors mediate juvenile hormone biosynthesis. Proc Natl Acad Sci U S A 99, 37-42.

Chiang, A.S., Pszczolkowski, M.A., Liu, H.P., Lin, S.C., 2002b. Ionotropic glutamate receptors mediate

juvenile hormone synthesis in the cockroach, Diploptera punctata. Insect Biochem Mol Biol 32, 669-678.

Chiang, A.S., Tsai, W.H., Schal, C., 1995. Neural and hormonal regulation of growth of corpora allata in the

cockroach, Diploptera punctata. Mol Cell Endocrinol 115, 51-57.

Clark, A.J., Bloch, K., 1959. The absence of sterol synthesis in insects. J Biol Chem 234, 2578–2582.

Clarke, C.F., Tanaka, R.D., Svenson, K., Wamsley, M., Fogelman, A.M., Edwards, P.A., 1987. Molecular

cloning and sequence of a cholesterol-repressible enzyme related to prenyltransferase in the isoprene

biosynthetic pathway. Mol Cell Biol 7, 3138-3146.

Coast, G.M., Schooley, D.A., 2011. Toward a consensus nomenclature for insect neuropeptides and peptide

hormones. Peptides 32, 620-631.

Consortium, T.H.B.G.S., 2006. Insights into social insects from the genome of the honeybee Apis mellifera. .

Nature 443, 931–949.

Cornforth, R.H., Fletcher, K., Hellig, H., Popjak, G., 1960. Stereospecificity of enzymic reactions involving

mevalonic acid. Nature 185, 923-924.

Couillaud, F., Feyereisen, R., 1991. Assay of HMG-CoA synthase in Diploptera punctata corpora allata. Insect

Biochem 21, 131-135.

Cusson, M., Beliveau, C., Sen, S.E., Vandermoten, S., Rutledge, R.G., Stewart, D., Francis, F., Haubruge, E.,

Rehse, P., Huggins, D.J., Dowling, A.P., Grant, G.H., 2006. Characterization and tissue-specific expression of

two lepidopteran farnesyl diphosphate synthase homologs: implications for the biosynthesis of ethyl-

substituted juvenile hormones. Proteins 65, 742-758.

Cusson, M., Prestwich, G.D., Stay, B., Tobe, S.S., 1991a. Photoaffinity labeling of allatostatin receptor

proteins in the corpora allata of the cockroach, Diploptera punctata. Biochem Biophys Res Commun 181, 736-

742.

Cusson, M., Prestwich, G.D., Stay, B., Tobe, S.S., 1992a. Evidence for allatostatin receptor Proteins in the

corpora allata of the cockroach, Diploptera punctata, as demonstrated by photoaffinity-labeling. Insect

Juvenile Hormone Research, 183-191.

Cusson, M., Yagi, K.J., Ding, Q., Duve, H., Thorpe, A., McNeil, J.N., Tobe, S.S., 1991b. Biosynthesis and

release of juvenile hormone and its precursors in insects and crustaceans: the search for a unifying arthropod

endocrinology. Insect Biochem 21, 1-6.

Cusson, M., Yagi, K.J., Guan, X.C., Tobe, S.S., 1992b. Assessment of the role of cyclic nucleotides in

allatostatin-induced inhibition of juvenile hormone biosynthesis in Diploptera punctata. Mol Cell Endocrinol

89, 121-125.

Dahm, K.H., Roller, H., Trost, B.M., 1968. The juvenile hormone. IV. Stereochemistry of juvenile hormone

and biological activity of some of its isomers and related compounds. Life Sci II 7, 129-137.

55

Daimon, T., Kozaki, T., Niwa, R., Kobayashi, I., Furuta, K., Namiki, T., Uchino, K., Banno, Y., Katsuma, S.,

Tamura, T., Mita, K., Sezutsu, H., Nakayama, M., Itoyama, K., Shimada, T., Shinoda, T., 2012. Precocious

metamorphosis in the juvenile hormone-deficient mutant of the silkworm, Bombyx mori. PLoS Genet 8,

e1002486.

Darrouzet, E., Mauchamp, B., Prestwich, G.D., Kerhoas, L., Ujvary, I., Couillaud, F., 1997. Hydroxy juvenile

hormones: new putative juvenile hormones biosynthesized by locust corpora allata in vitro. Biochem Biophys

Res Commun 240, 752-758.

de Ruyck, J., Janczak, M.W., Neti, S.S., Rothman, S.C., Schubert, H.L., Cornish, R.M., Matagne, A., Wouters,

J., Poulter, C.D., 2014. Determination of kinetics and the crystal structure of a novel type 2 isopentenyl

diphosphate: dimethylallyl diphosphate isomerase from Streptococcus pneumoniae. Chembiochem 15, 1452-

1458.

Defelipe, L.A., Dolghih, E., Roitberg, A.E., Nouzova, M., Mayoral, J.G., Noriega, F.G., Turjanski, A.G., 2011.

Juvenile hormone synthesis: "esterify then epoxidize" or "epoxidize then esterify"? Insights from the structural

characterization of juvenile hormone acid methyltransferase. Insect Biochem Mol Biol 41, 228-235.

Diaz, M.E., Mayoral, J.G., Priestap, H., Nouzova, M., Rivera-Perez, C., Noriega, F.G., 2012. Characterization

of an isopentenyl diphosphate isomerase involved in the juvenile hormone pathway in Aedes aegypti. Insect

Biochem Mol Biol 42, 751-757.

Donly, B.C., Ding, Q., Tobe, S.S., Bendena, W.G., 1993. Molecular cloning of the gene for the allatostatin

family of neuropeptides from the cockroach Diploptera punctata. Proc Natl Acad Sci U S A 90, 8807-8811.

Down, R.E., Matthews, H.J., Audsley, N., 2010. Effects of Manduca sexta allatostatin and an analog on the

pea aphid Acyrthosiphon pisum (Hemiptera: Aphididae) and degradation by enzymes from the aphid gut.

Peptides 31, 489-497.

Duportets, L., Belles, X., Rossignol, F., Couillaud, F., 2000. Molecular cloning and structural analysis of 3-

hydroxy-3-methylglutaryl coenzyme A reductase of the moth Agrotis ipsilon. Insect Mol Biol 9, 385-392.

Duve, H., Audsley, N., Weaver, R.J., Thorpe, A., 2000. Triple co-localisation of two types of allatostatin and

an allatotropin in the frontal ganglion of the lepidopteran Lacanobia oleracea (Noctuidae): innervation and

action on the foregut. Cell Tissue Res 300, 153-163.

Duve, H., East, P.D., Thorpe, A., 1999. Regulation of lepidopteran foregut movement by allatostatins and

allatotropin from the frontal ganglion. J Comp Neurol 413, 405-416.

Eigenheer, A.L., Keeling, C.I., Young, S., Tittiger, C., 2003. Comparison of gene representation in midguts

from two phytophagous insects, Bombyx mori and Ips pini, using expressed sequence tags. Gene 316, 127-136.

Eizaguirre, M., Schafellner, C., Lopez, C., Sehnal, F., 2005. Relationship between an increase of juvenile

hormone titer in early instars and the induction of diapause in fully grown larvae of Sesamia nonagrioides. J

Insect Physiol 51, 1127-1134.

Elliott, K.L., Woodhead, A.P., Stay, B., 2006. A stage-specific ovarian factor with stable stimulation of

juvenile hormone synthesis in corpora allata of the cockroach Diploptera punctata. J Insect Physiol 52, 929-

935.

Engelmann, F., 1979. Insect vitellogenin: identification, biosynthesis, and role in vitellogenesis. Adv in Insect

Phys 14, 59.

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

punctata. Invert Reprod Dev 15, 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. Invert Reprod Dev 28, 161-170.

Faulkner, D.J., Petersen, M.R., 1971. Synthesis of C-18 Cecropia juvenile hormone to obtain optically active

forms of known absolute configuration. J Am Chem Soc 93, 3766-3767.

56

Feldlaufer, M.F., Bowers, W.S., Soderlund, D.M., Evans, P.H., 1982. Biosynthesis of the sesquiterpenoid

skeleton of juvenile hormone 3 by Dysdercus fasciatus corpora allata in vitro. . J Exp Zool 223, 295–298.

Feyereisen, R., Farnsworth, D.E., 1987a. Characterization and regulation of HMG-CoA Reductase during a

cycle of Juvenile Hormone synthesis. Mol Cell Endocrinol 53, 227-238.

Feyereisen, R., Farnsworth, D.E., 1987b. Inhibition of Insect juvenile hormone synthesis by Phorbol 12-

myristate 13-acetate. FEBS Lett 222, 345-348.

Fisher, C.W., Mayer, R.T., 1982. Characterization of house fly microsomal mixed function oxidases: inhibition

by juvenile hormone i and piperonyl butoxide. Toxicology 24, 15-31.

Friedel, T., Feyereisen, R., Mundall, E.C., Tobe, S.S., 1980. The allatostatic effect of 20-hydroxyecdysone on

the adult viviparous cockroach, Diploptera punctata. J Insect Physiol 26, 665-670.

Fuse, M., Zhang, J.R., Partridge, E., Nachman, R.J., Orchard, I., Bendena, W.G., Tobe, S.S., 1999. Effects of

an allatostatin and a myosuppressin on midgut carbohydrate enzyme activity in the cockroach Diploptera punctata. Peptides 20, 1285-1293.

Garside, C.S., Bendena, W.G., Tobe, S.S., 2003. Quantification and visualization of Dippu-AST mRNA in the

brain of adult Diploptera punctata: mated females vs. virgin females vs. males. J Insect Physiol 49, 285-291.

Garside, C.S., Koladich, P.M., Bendena, W.G., Tobe, S.S., 2002. Expression of allatostatin in the oviducts of

the cockroach Diploptera punctata. Insect Biochem Mol Biol 32, 1089-1099.

Gehring, U., Harris, J.I., 1970. The active site cysteines of thiolase. Eur J Biochem 16, 492-498.

Gertler, F.B., Chiu, C.Y., Richter-Mann, L., Chin, D.J., 1988. Developmental and metabolic regulation of the

Drosophila melanogaster 3-hydroxy-3-methylglutaryl coenzyme A reductase. Mol Cell Biol 8, 2713-2721.

Goldstein, J.L., Brown, M.S., 1990. Regulation of the mevalonate pathway. Nature 343, 425-430.

Goodfellow, R.D., Barnes, F.J., 1971. Mevalonate kinase from the larva of the fleshfly, Sarcophaga bullata.

Insect Biochem 1, 271-275.

Goodman, W.G., Cusson, M., 2012. The juvenile hormones, in: Gilbert, L.I. (Ed.), Insect Endocrinology.

Academic Press, London, p. 55.

Granger, N.A., Bollenbacher, W.E., Vince, R., Gilbert, L.I., Baehr, J.C., Dray, F., 1979. In vitro biosynthesis

of juvenile hormone by the larval corpora allata of Manduca sexta: quantificationby radioimmunoassay. Mol

Cell Endocrinol 16, 1-17.

Granger, N.A., Janzen, W.P., 1987. Inhibition of Manduca sexta corpora allata in vitro by a cerebral

allatostatic neuropeptide. Mol Cell Endocrinol 49, 237-248.

Granger, N.A., Niemiec, S.M., Gilbert, L.I., Bollenbacher, W.E., 1982. Juvenile hormone synthesis in vitro by

larval and pupal corpora allata of Manduca sexta. Mol Cell Endocrinol 28, 587-604.

Granger, N.A., Sturgis, S.L., Ebersohl, R., Geng, C., Sparks, T.C., 1996. Dopaminergic control of corpora

allata activity in the larval tobacco hornworm, Manduca sexta. Arch Insect Biochem Physiol 32, 449-466.

Grossniklaus-Burgin, C., Lanzrein, B., 1990. Qualitative and quantitative analyses of juvenile hormone and

ecdysteroids from the egg to the pupal molt in Trichoplusia ni. Arch Insect Biochem Physiol 14, 13-30.

Group, B.B.A., 2004. A draft sequence for the genome of the domesticated silkworm (Bombyx mori). Science

306, 1937-1940.

Gruntenko, N.E., Karpova, E.K., Alekseev, A.A., Chentsova, N.A., Saprykina, Z.V., Bownes, M.,

Rauschenbach, I.Y., 2005. Effects of dopamine on juvenile hormone metabolism and fitness in Drosophila

virilis. J Insect Physiol 51, 959-968.

57

Gunawardene, Y.I.N.S., Chow, B.K.C., He, J.G., Chan, S.M., 2001. The shrimp FAMeT cDNA is encoded for

a putative enzyme involved in the methylfarnesoate (MF) biosynthetic pathway and is temporally expressed in

the eyestalk of different sexes. Insect Biochem Mol Biol 31, 1115-1124.

Hall, G.M., Tittiger, C., Andrews, G.L., Mastick, G.S., Kuenzli, M., Luo, X., Seybold, S.J., Blomquist, G.J.,

2002. Midgut tissue of male pine engraver, Ips pini, synthesizes monoterpenoid pheromone component

ipsdienol de novo. Naturwissenschaften 89, 79-83.

Hammock, B.D., 1975. NADPH dependent epoxidation of methyl farnesoate to juvenile hormone in the

cockroach Blaberus giganteus L. Life Sciences 17, 323-328.

Hass, J.K., Cassias, K.A., Woodhead, A.P., Stay, B., 2003. The effect of ovary implants on juvenile hormone

production by corpora allata of male Diploptera punctata. J Insect Sci 3, 30.

Helvig, C., Koener, J.F., Unnithan, G.C., Feyereisen, R., 2004. CYP15A1, the cytochrome P450 that catalyzes

epoxidation of methyl farnesoate to juvenile hormone III in cockroach corpora allata. Proc Natl Acad Sci U S

A 101, 4024-4029.

Henrich, V.C., Szekely, A.A., Kim, S.J., Brown, N.E., Antoniewski, C., Hayden, M.A., Lepesant, J.A., Gilbert,

L.I., 1994. Expression and function of the ultraspiracle (usp) gene during development of Drosophila

melanogaster. Dev Biol 165, 38-52.

Hiruma, K., Riddiford, L.M., 1985. Hormonal regulation of dopa decarboxylase during a larval molt. Dev Biol

110, 509-513.

Hodkova, M., Okuda, T., Wagner, R., 1996. Stimulation of corpora allata by extract from neuroendocrine

complex; Comparison of reproducing and diapausing Pyrrhocoris apterus. Eur J Entomol 93, 535–543.

Holbrook, G.L., Schal, C., 1998. Social influences on nymphal development in the cockroach, Diploptera

punctata. Physiological Entomology 23, 121-130.

Holbrook, G.L., Schal, C., 2004. Maternal investment affects offspring phenotypic plasticity in a viviparous

cockroach. Proc Natl Acad Sci U S A 101, 5595-5597.

Houten, S.M., Waterham, H.R., 2001. Nonorthologous gene displacement of phosphomevalonate kinase. Mol

Genet Metab 72, 273-276.

Hult, E.F., Weadick, C.J., Chang, B.S.W., Tobe, S.S., 2008. Reconstruction of ancestral FGLamide-type insect

allatostatins: A novel approach to the study of allatostatin function and evolution. J Insect Physiol 54, 959-968.

Inoue, H., Tsuji, H., Uritani, I., 1984. Characterization and activity change of farnesol dehydrogenase in black

rot fungus-infected sweet potato. . Agric Biol Chem 48, 733–738.

Iwabuchi, K., 1995. Effect of juvenile hormone on the embryogenesis of a polyembryonic wasp, Copidosoma

floridanum, in vitro. In Vitro Cell Dev Biol Anim 31, 803-805.

Iwema, T., Billas, I.M., Beck, Y., Bonneton, F., Nierengarten, H., Chaumot, A., Richards, G., Laudet, V.,

Moras, D., 2007. Structural and functional characterization of a novel type of ligand-independent RXR-USP

receptor. EMBO J 26, 3770-3782.

Jabalquinto, A.M., Alvear, M., Cardemil, E., 1988. Physiological aspects and mechanism of action of

mevalonate 5-diphosphate decarboxylase. Comp Biochem Physiol B 90, 671-677.

Jones, G., Jones, D., Teal, P., Sapa, A., Wozniak, M., 2006. The retinoid-X receptor ortholog, ultraspiracle,

binds with nanomolar affinity to an endogenous morphogenetic ligand. FEBS J 273, 4983-4996.

Jones, G., Sharp, P.A., 1997. Ultraspiracle: An invertebrate nuclear receptor for juvenile hormones. Proc Natl

Acad Sci U S A 94, 13499-13503.

Jones, G., Wozniak, M., Chu, Y., Dhar, S., Jones, D., 2001. Juvenile hormone III-dependent conformational

changes of the nuclear receptor ultraspiracle. Insect Biochem Mol Biol 32, 33-49.

58

Judy, K.J., Schooley, D.A., Dunham, L.L., Hall, M.S., Bergot, B.J., Siddall, J.B., 1973. Isolation, Structure,

and Absolute Configuration of a New Natural Insect Juvenile Hormone from Manduca sexta. Proc Natl Acad

Sci U S A 70, 1509-1513.

Kataoka, H., Toschi, A., Li, J.P., Carney, R.L., Schooley, D.A., Kramer, S.J., 1989. Identification of an

allatotropin from adult Manduca sexta. Science 243, 1481-1483.

Kayukawa, T., Minakuchi, C., Namiki, T., Togawa, T., Yoshiyama, M., Kamimura, M., Mita, K., Imanishi, S.,

Kiuchi, M., Ishikawa, Y., Shinoda, T., 2012. Transcriptional regulation of juvenile hormone-mediated

induction of Kruppel homolog 1, a repressor of insect metamorphosis. Proc Natl Acad Sci U S A 109, 11729-

11734.

Kayukawa, T., Murata, M., Kobayashi, I., Muramatsu, D., Okada, C., Uchino, K., Sezutsu, H., Kiuchi, M.,

Tamura, T., Hiruma, K., Ishikawa, Y., Shinoda, T., 2014. Hormonal regulation and developmental role of

Kruppel homolog 1, a repressor of metamorphosis, in the silkworm Bombyx mori. Dev Biol 388, 48-56.

Keeling, C.I., Blomquist, G.J., Tittiger, C., 2004. Coordinated gene expression for pheromone biosynthesis in

the pine engraver beetle, Ips pini (Coleoptera: Scolytidae). Naturwissenschaften 91, 324-328.

Keung, W.M., 1991. Human liver alcohol dehydrogenases catalyze the oxidation of the intermediary alcohols

of the shunt pathway of mevalonate metabolism. Biochem Biophys Res Commun 174, 701-707.

Kewley, R.J., Whitelaw, M.L., Chapman-Smith, A., 2004. The mammalian basic helix-loop-helix/PAS family

of transcriptional regulators. Int J Biochem Cell Biol 36, 189-204.

Kikukawa, S., Smith, C., Tobe, S.S., 1989. Morphogenetic and gonadotrophic effects of a juvenile hormone

analogue ((7S)-hydroprene) in last instar female larvae of Diploptera punctata. Physiological Entomology 14,

165-172.

Kikukawa, S., Tobe, S.S., 1986a. Critical periods for juvenile hormone sensitivity during larval life of female

Diploptera punctata. J Insect Physiol 32, 1035-1042.

Kikukawa, S., Tobe, S.S., 1986b. Juvenile hormone biosynthesis in female larvae of Diploptera punctata and

the effect of allatectomy on hemolymph ecdysteroid titer. J Insect Physiol 32, 981-986.

Kikukawa, S., Tobe, S.S., Solowiej, S., Rankin, S.M., Stay, B., 1987. Calcium as a regulator of juvenile

hormone biosynthesis and release in the cockroach Diploptera punctata. Insect Biochem 17, 179-187.

Kinjoh, T., Kaneko, Y., Itoyama, K., Mita, K., Hiruma, K., Shinoda, T., 2007. Control of juvenile hormone

biosynthesis in Bombyx mori: Cloning of the enzymes in the mevalonate pathway and assessment of their

developmental expression in the corpora allata. Insect Biochem Mol Biol 37, 808-818.

Koladich, P.M., Cusson, M., Bendena, W.G., Tobe, S.S., McNeil, J.N., 2002. Cardioacceleratory effects of

Manduca sexta allatotropin in the true armyworm moth, Pseudaletia unipuncta. Peptides 23, 645-651.

Konopova, B., Jindra, M., 2007. Juvenile hormone resistance gene Methoprene-tolerant controls entry into

metamorphosis in the beetle Tribolium castaneum. Proc Natl Acad Sci U S A 104, 10488-10493.

Konopova, B., Smykal, V., Jindra, M., 2011. Common and distinct roles of juvenile hormone signaling genes

in metamorphosis of holometabolous and hemimetabolous insects. PLoS One 6, e28728.

Kotaki, T., 1993. Biosynthetic products by heteropteran corpora allata in vitro. Invert Reprod Dev 31, 225–230.

Kotaki, T., 1996. Evidence for a new juvenile hormone in a stink bug, Plautia stali. J Insect Physiol 42, 279–

286.

Kotaki, T., Shinada, T., Kaihara, K., Ohfune, Y., Numata, H., 2009. Structure determination of a new juvenile

hormone from a heteropteran insect. Org Lett 11, 5234-5237.

Kotaki, T., Shinada, T., Kaihara, K., Ohfune, Y., Numata, H., 2011. Biological activities of juvenile hormone

III skipped bisepoxide in last instar nymphs and adults of a stink bug, Plautia stali. J Insect Physiol 57, 147-

152.

59

Kou, R., Chen, S.J., 2000. Allatotropic activity in the suboesophageal ganglia and corpora cardiaca of the adult

male loreyi leafworm, Mythimna loreyi. Arch Insect Biochem Physiol 43, 78-86.

Koyama, T., Matsubara, M., Ogura, K., 1985. Isoprenoid enzyme systems of silkworm. I. Partial purification

of isopentenyl pyrophosphate isomerase, farnesyl pyrophosphate synthetase, and geranylgeranyl

pyrophosphate synthetase. J Biochem 98, 449-456.

Kramer, S.J., Toschi, A., Miller, C.A., Kataoka, H., Quistad, G.B., Li, J.P., Carney, R.L., Schooley, D.A., 1991.

Identification of an allatostatin from the tobacco hornworm Manduca sexta. Proc Natl Acad Sci U S A 88,

9458-9462.

Kreienkamp, H.J., Larusson, H.J., Witte, I., Roeder, T., Birgul, N., Honck, H.H., Harder, S., Ellinghausen, G.,

Buck, F., Richter, D., 2002. Functional annotation of two orphan G-protein-coupled receptors, Drostar1 and -2,

from Drosophila melanogaster and their ligands by reverse pharmacology. J Biol Chem 277, 39937-39943.

Kwok, R., Zhang, J.R., Tobe, S.S., 2005. Regulation of methyl farnesoate production by mandibular organs in

the crayfish, Procambarus clarkii: A possible role for allatostatins. J Insect Physiol 51, 367-378.

Lakshmi, M., Dutta-Gupta, A., 1990. Juvenile hormone mediated DNA synthesis during larval development of

Corcyra cephalonica (Insecta). Biochem Int 22, 269-278.

Lange, A.B., Bendena, W.G., Tobe, S.S., 1995. The effect of the 13 Dip-allatostatins on myogenic and induced

contractions of the cockroach (Diploptera punctata) hindgut. J Insect Physiol 41, 581-588.

Lanzrein, B., Imboden, H., Biirgin, C., Briining, E., H., G., 1984. On titers, origin, and functions of juvenile

hormone III, methyl farnesoate and ecdysteroids in embryonic development of the ovoviviparous cockroach,

Nauphoeta cinerea. , in: Hoffmann, J., Porchet, M. (Eds.), Biosynthesis, Metabolism and Mode of Action of

lnvertebrate Hormones Springer, pp. 454-465.

Laufer, H., Borst, D., Baker, F.C., Reuter, C.C., Tsai, L.W., Schooley, D.A., Carrasco, C., Sinkus, M., 1987.

Identification of a juvenile hormone-like compound in a crustacean. Science 235, 202-205.

Lee, K.Y., Horodyski, F.M., Chamberlin, M.E., 1998. Inhibition of midgut ion transport by allatotropin (Mas-

AT) and Manduca FLRFamides in the tobacco hornworm Manduca sexta. J Exp Biol 201, 3067-3074.

Lefevere, K.S., Lacey, M.J., Smith, P.H., Roberts, B., 1993. Identification and quantification of juvenile

hormone biosynthesized by larval and adult Australian sheep blowfly Lucilia cuprina (Diptera: Calliphoridae).

Insect Biochem Mol Biol 23, 713-720.

Lewis, M.J., Prosser, I.M., Mohib, A., Field, L.M., 2008. Cloning and characterisation of a prenyltransferase

from the aphid Myzus persicae with potential involvement in alarm pheromone biosynthesis. Insect Mol Biol

17, 437-443.

Li, M., Mead, E.A., Zhu, J., 2011. Heterodimer of two bHLH-PAS proteins mediates juvenile hormone-

induced gene expression. Proc Natl Acad Sci U S A 108, 638-643.

Li, S., Ouyang, Y.C., Ostrowski, E., Borst, D.W., 2005. Allatotropin regulation of juvenile hormone synthesis

by the corpora allata from the lubber grasshopper, Romalea microptera. Peptides 26, 63-72.

Li, Y., Hernandez-Martinez, S., Fernandez, F., Mayoral, J.G., Topalis, P., Priestap, H., Perez, M., Navare, A.,

Noriega, F.G., 2006. Biochemical, molecular, and functional characterization of PISCF-allatostatin, a regulator

of juvenile hormone biosynthesis in the mosquito Aedes aegypti. J Biol Chem 281, 34048-34055.

Li, Y., Unnithan, G.C., Veenstra, J.A., Feyereisen, R., Noriega, F.G., 2003. Stimulation of JH biosynthesis by

the corpora allata of adult female Aedes aegypti in vitro: effect of farnesoic acid and Aedes allatotropin. J Exp

Biol 206, 1825-1832.

Liang, P.H., Ko, T.P., Wang, A.H., 2002. Structure, mechanism and function of prenyltransferases. Eur J

Biochem 269, 3339-3354.

Lorenz, M.W., Hoffmann, K.H., 1995. Allatotropic activity in the subesophageal ganglia of crickets, Gryllus bimaculatus and Acheta domesticus. . J Insect Physiol 41, 191–196.

60

Lozano, J., Belles, X., 2011. Conserved repressive function of Kruppel homolog 1 on insect metamorphosis in

hemimetabolous and holometabolous species. Sci Rep 1, 163.

Lungchukiet, P., Donly, B.C., Zhang, J.R., Tobe, S.S., Bendena, W.G., 2008a. Molecular cloning and

characterization of an allatostatin-like receptor in the cockroach Diploptera punctata. Peptides 29, 276-285.

Lungchukiet, P., Zhang, J.R., Tobe, S.S., Bendena, W.G., 2008b. Quantification of allatostatin receptor mRNA

levels in the cockroach, Diploptera punctata, using real-time PCR. J Insect Physiol 54, 981-987.

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, e106737.

Marchal, E., Zhang, J.R., Badisco, L., Verlinden, H., Hult, E.F., Van Wielendaele, P., Yagi, K.J., Tobe, S.S.,

Vanden Broeck, J., 2011. Final steps in juvenile hormone biosynthesis in the desert locust, Schistocerca

gregaria. Insect Biochem Mol Biol 41, 219-227.

Martin, D., Piulachs, M.D., Belles, X., 1996. Inhibition of vitellogenin production by allatostatin in the

German cockroach. Mol Cell Endocrinol 121, 191-196.

Martinez-Gonzalez, J., Buesa, C., Piulachs, M.D., Belles, X., Hegardt, F.G., 1993. 3-Hydroxy-3-

methylglutaryl-coenzyme-A synthase from Blattella germanica. Cloning, expression, developmental pattern

and tissue expression. Eur J Biochem 217, 691-699.

Martinezgonzalez, J., Buesa, C., Piulachs, M.D., Belles, X., Hegardt, F.G., 1993. Molecular cloning,

developmental pattern and tissue expression of 3-hydroxy-3-methylglutaryl coenzyme a reductase of the

cockroach Blattella germanica. Eur J Biochem 213, 233-241.

Matthews, H.J., Down, R.E., Audsley, N., 2010. Effects of Manduca sexta allatostatin and an analogue on the

peach-potato aphid Myzus persicae (hemiptera: aphididae) and degradation by enzymes in the aphid gut. Arch

Insect Biochem Physiol 75, 139-157.

Mauchamp, B., Couillaud, F., Malosse, C., 1985. Gas chromatography-mass spectroscopy analysis of juvenile

hormone released by insect corpora allata. Anal Biochem 145, 251-256.

Mauchamp, B., Darrouzet, E., Malosse, C., Couillaud, F., 1999. 4´-OH-JH-III: an additional hydroxylated

juvenile hormone produced by locust corpora allata in vitro. . Insect Biochem Mol Biol 29, 475–480.

Mayoral, J.G., Nouzova, M., Brockhoff, A., Goodwin, M., Hernandez-Martinez, S., Richter, D., Meyerhof, W.,

Noriega, F.G., 2010. Allatostatin-C receptors in mosquitoes. Peptides 31, 442-450.

Mayoral, J.G., Nouzova, M., Navare, A., Noriega, F.G., 2009a. NADP(+)-dependent farnesol dehydrogenase,

a corpora allata enzyme involved in juvenile hormone synthesis. Proc Natl Acad Sci U S A 106, 21091-21096.

Mayoral, J.G., Nouzova, M., Yoshiyama, M., Shinod, T., Hernandez-Martinez, S., Dolghih, E., Turjanski,

A.G., Roitberg, A.E., Priestap, H., Perez, M., Mackenzie, L., Li, Y.P., Noriega, F.G., 2009b. Molecular and

functional characterization of a juvenile hormone acid methyltransferase expressed in the corpora allata of

mosquitoes. Insect Biochem Mol Biol 39, 31-37.

McQuiston, A.R., Thompson, C.S., Tobe, S.S., 1990. Voltage-dependent sodium and calcium channels of the

corpora allata of the cockroach Diploptera punctata, in: Borkovec, A.B., Masler, E.P. (Eds.), Insect

Neurochemistry and Neurophysiology. Humana Press, Clifton, pp. 329-332.

Meigs, T.E., Roseman, D.S., Simoni, R.D., 1996. Regulation of 3-hydroxy-3-methylglutaryl-coenzyme A

reductase degradation by the nonsterol mevalonate metabolite farnesol in vivo. J Biol Chem 271, 7916-7922.

Meigs, T.E., Simoni, R.D., 1997. Farnesol as a regulator of HMG-CoA reductase degradation: characterization

and role of farnesyl pyrophosphatase. Arch Biochem Biophys 345, 1-9.

Meller, V.H., Aucoin, R.R., Tobe, S.S., Feyereisen, R., 1985. Evidence for an inhibitory role of Cyclic-AMP

in the control of juvenile hormone biosynthesis by cockroach corpora allata. Mol Cell Endocrinol 43, 155-163.

61

Meyer, A.S., Hanzmann, E., Murphy, R.C., 1971. Absolute configuration of Cecropia juvenile hormone. Proc

Natl Acad Sci U S A 68, 2312-2315.

Meyer, A.S., Hanzmann, E., Schneiderman, H.A., Gilbert, L.I., Boyette, M., 1970. The isolation and

identification of the two juvenile hormones from the Cecropia silk moth. Arch Biochem Biophys 137, 190-213.

Meyer, A.S., Schneiderman, H.A., Hanzmann, E., Ko, J.H., 1968. The two juvenile hormones from the

cecropia silk moth. Proc Natl Acad Sci U S A 60, 853-860.

Minakuchi, C., Namiki, T., Shinoda, T., 2009. Kruppel homolog 1, an early juvenile hormone-response gene

downstream of Methoprene-tolerant, mediates its anti-metamorphic action in the red flour beetle Tribolium

castaneum. Dev Biol 325, 341-350.

Minakuchi, C., Namiki, T., Yoshiyama, M., Shinoda, T., 2008a. RNAi-mediated knockdown of juvenile

hormone acid O-methyltransferase gene causes precocious metamorphosis in the red flour beetle Tribolium

castaneum. Febs Journal 275, 2919-2931.

Minakuchi, C., Zhou, X., Riddiford, L.M., 2008b. Kruppel homolog 1 (Kr-h1) mediates juvenile hormone

action during metamorphosis of Drosophila melanogaster. Mech Dev 125, 91-105.

Mita, K., Kasahara, M., Sasaki, S., Nagayasu, Y., Yamada, T., Kanamori, H., Namiki, N., Kitagawa, M.,

Yamashita, H., Yasukochi, Y., Kadono-Okuda, K., Yamamoto, K., Ajimura, M., Ravikumar, G., Shimomura,

M., Nagamura, Y., Shin, I.T., Abe, H., Shimada, T., Morishita, S., Sasaki, T., 2004. The genome sequence of

silkworm, Bombyx mori. DNA Res 11, 27-35.

Miura, K., Oda, M., Makita, S., Chinzei, Y., 2005. Characterization of the Drosophila Methoprene -tolerant

gene product. Juvenile hormone binding and ligand-dependent gene regulation. FEBS J 272, 1169-1178.

Miyawaki, R., Tanaka, S.I., Numata, H., 2006. Role of juvenile hormone in the control of summer diapause in

adult Poecilocoris lewisi (Heteroptera: Scutelleridae). Formosan Entomol 26, 1–10.

Monger, D.J., Law, J.H., 1982. Control of juvenile hormone biosynthesis. Evidence for phosphorylation of the

3-hydroxy-3-methylglutaryl coenzyme A reductase of insect corpus allatum. J Biol Chem 257, 1921-1923.

Moshitzky, P., Applebaum, S.W., 1995. Pathway and regulation of JHIII-Bisepoxide biosynthesis in adult

Drosophila melanogaster corpus allatum. Arch Insect Biochem Physiol 30, 225-237.

Moshitzky, P., Miloslavski, I., Aizenshtat, Z., Applebaum, S.W., 2003. Methyl palmitate: a novel product of

the Medfly (Ceratitis capitata) corpus allatum. Insect Biochem Mol Biol 33, 1299-1306.

Mundall, E.C., Tobe, S.S., Stay, B., 1979. Induction of vitellogenin and growth of implanted oocytes in male

cockroaches. Nature 282, 97-98.

Mundall, E.C., Tobe, S.S., Stay, B., 1981. Vitellogenin fluctuations in hemolymph and fat body and dynamics

of uptake into oocytes during the reproductive cycle of Diploptera punctata. J Insect Physiol 27, 821-827.

Nagaraju, G.P.C., 2007. Is methyl farnesoate a crustacean hormone? Aquaculture 272, 39-54.

Nakanishi, K., Schooley, D.A., Koreeda, M., Dillon, J., 1971. Absolute configuration of the C18-juvenile

hormone: application of a new circular dichroism method using tris(dipivaloylmethanato) praseodymium.

Chem Commun, 1235–1236.

Niwa, R., Niimi, T., Honda, N., Yoshiyama, M., Itoyama, K., Kataoka, H., Shinoda, T., 2008. Juvenile

hormone acid O-methyltransferase in Drosophila melanogaster. Insect Biochem Mol Biol 38, 714-720.

Noriega, F.G., Ribeiro, J.M.C., Koener, J.F., Valenzuela, J.G., Hernandez-Martinez, S., Pham, V.M.,

Feyereisen, R., 2006. Comparative genomics of insect juvenile hormone biosynthesis. Insect Biochem Mol

Biol 36, 366-374.

Nouzova, M., Edwards, M.J., Mayoral, J.G., Noriega, F.G., 2011. A coordinated expression of biosynthetic

enzymes controls the flux of juvenile hormone precursors in the corpora allata of mosquitoes. Insect Biochem

Mol Biol 41, 660-669.

62

Numata, H., Numata, A., Takahashi, C., Nakagawa, Y., Iwatani, K., Takahashi, S., Miura, K., Chinzei, Y.,

1992. Juvenile hormone I is the principal juvenile hormone in a hemipteran insect, Riptortus clavatus. .

Experientia 48, 606–610.

Nyati, P., Nouzova, M., Rivera-Perez, C., Clifton, M.E., Mayoral, J.G., Noriega, F.G., 2013. Farnesyl

phosphatase, a corpora allata enzyme involved in juvenile hormone biosynthesis in Aedes aegypti. PLoS One 8.

Oda, S., Tatarazako, N., Watanabe, H., Morita, M., Iguchi, T., 2005. Production of male neonates in Daphnia

magna (Cladocera, Crustacea) exposed to juvenile hormones and their analogs. Chemosphere 61, 1168-1174.

Oeh, U., Lorenz, M.W., Dyker, H., Losel, P., Hoffmann, K.H., 2000. Interaction between Manduca sexta

allatotropin and manduca sexta allatostatin in the fall armyworm Spodoptera frugiperda. Insect Biochem Mol

Biol 30, 719-727.

Paemen, L., Tips, A., Schoofs, L., Proost, P., Van Damme, J., De Loof, A., 1991. Lom-AG-myotropin: a novel

myotropic peptide from the male accessory glands of Locusta migratoria. Peptides 12, 7-10.

Parthasarathy, R., Sun, Z., Bai, H., Palli, S.R., 2010. Juvenile hormone regulation of vitellogenin synthesis in

the red flour beetle, Tribolium castaneum. Insect Biochem Mol Biol 40, 405-414.

Parthasarathy, R., Tan, A., Palli, S.R., 2008. bHLH-PAS family transcription factor methoprene-tolerant plays

a key role in JH action in preventing the premature development of adult structures during larval-pupal

metamorphosis. Mech Dev 125, 601-616.

Pastor, D., Piulachs, M.D., Cassier, P., Andre, M., Belles, X., 1991. Invivo and invitro studies of the action of

dopamine on oocyte growth and juvenile hormone production in Blattella germanica (L) (Dictyoptera,

Blattellidae). Comptes Rendus De L Academie Des Sciences Serie Iii-Sciences De La Vie-Life Sciences 313,

207-212.

Paulson, C.R., Stay, B., 1987. Humoral inhibition of the corpora allata in larvae of Diploptera punctata - Role

of the brain and ecdysteroids. J Insect Physiol 33, 613-622.

Pratt, G.E., Farnsworth, D.E., Feyereisen, R., 1990. Changes in the sensitivity of adult cockroach corpora

allata to a brain allatostatin. Mol Cell Endocrinol 70, 185-195.

Pratt, G.E., Farnsworth, D.E., Fok, K.F., Siegel, N.R., Mccormack, A.L., Shabanowitz, J., Hunt, D.F.,

Feyereisen, R., 1991. Identity of a 2nd type of allatostatin from cockroach brains - an octadecapeptide amide

with a tyrosine-rich address sequence. Proc Natl Acad Sci U S A 88, 2412-2416.

Pratt, G.E., Farnsworth, D.E., Siegel, N.R., Fok, K.F., Feyereisen, R., 1989. Identification of an allatostatin

from adult Diploptera punctata. Biochem Biophys Res Commun 163, 1243-1247.

Price, M.D., Merte, J., Nichols, R., Koladich, P.M., Tobe, S.S., Bendena, W.G., 2002. Drosophila

melanogaster flatline encodes a myotropin orthologue to Manduca sexta allatostatin. Peptides 23, 787-794.

Pszczolkowski, M.A., Lee, W.S., Liu, H.P., Chiang, A.S., 1999. Glutamate-induced rise in cytosolic calcium

concentration stimulates in vitro rates of juvenile hormone biosynthesis in corpus allatum of Diploptera

punctata. Mol Cell Endocrinol 158, 163-171.

Rachinsky, A., Tobe, S.S., 1996. Role of second messengers in the regulation of juvenile hormone production

in insects, with particular emphasis on calcium and phosphoinositide signaling. Arch Insect Biochem Physiol

33, 259-282.

Rachinsky, A., Zhang, J., Tobe, S.S., 1994. Signal transduction in the inhibition of juvenile hormone

biosynthesis by allatostatins - roles of diacylglycerol and calcium. Mol Cell Endocrinol 105, 89-96.

Raikhel, A.S., Brown, M.R., Belles, X., 2005. Hormonal control of reproductive processes, in: Gilbert, L.I.,

Iatrou, K., Gill, S.S. (Eds.), Comprehensive molecular insect science. Elsevier, Oxford, pp. 433-491.

Ramos-Valdivia, A.C., van der Heijden, R., Verpoorte, R., 1997. Isopentenyl diphosphate isomerase: a core

enzyme in isoprenoid biosynthesis. A review of its biochemistry and function. Nat Prod Rep 14, 591-603.

63

Rankin, S.M., Stay, B., 1984. The changing effect of the ovary on rates of juvenile hormone synthesis in

Diploptera punctata. Gen Comp Endocrinol 54, 382-388.

Rankin, S.M., Stay, B., 1985. Regulation of juvenile hormone synthesis during pregnancy in the cockroach,

Diploptera punctata. J Insect Physiol 31, 145-157.

Richard, D.S., Applebaum, S.W., Gilbert, L.I., 1989a. Developmental regulation of juvenile hormone

biosynthesis by the ring gland of Drosophila melanogaster. J Comp Physiol B 159, 383-387.

Richard, D.S., Applebaum, S.W., Sliter, T.J., Baker, F.C., Schooley, D.A., Reuter, C.C., Henrich, V.C., Gilbert,

L.I., 1989b. Juvenile hormone bisepoxide biosynthesis in vitro by the ring gland of Drosophila melanogaster: a

putative juvenile hormone in the higher Diptera. Proc Natl Acad Sci U S A 86, 1421-1425.

Rivera-Perez, C., Nouzova, M., Clifton, M.E., Garcia, E.M., LeBlanc, E., Noriega, F.G., 2013. Aldehyde

dehydrogenase 3 converts farnesal into farnesoic acid in the corpora allata of mosquitoes. Insect Biochem Mol

Biol 43, 675-682.

Röller, H., Dahm, D.H., Sweeley, C.C., Trost, B.M., 1967. The structure of the juvenile hormone. Angew

Chem Int Ed 6, 179–180.

Roth, L.M., 1970. Evolution and taxonomic significance of reproduction in Blattaria. Annual Review of

Entomology 15, 75-97.

Roth, L.M., Stay, B., 1961. Oocyte development in Diploptera punctata (Eschscholtz) (Blattaria). J Insect

Physiol 7, 186-202.

Rudwall, A.J., Sliwowska, J., Nassel, D.R., 2000. Allatotropin-like neuropeptide in the cockroach abdominal

nervous system: myotropic actions, sexually dimorphic distribution and colocalization with serotonin. J Comp

Neurol 428, 159-173.

Ruegg, R.P., Lococo, D.J., Tobe, S.S., 1983. Control of corpus allatum activity in Diploptera punctata - Roles

of the pars intercerebralis and pars lateralis. Experientia 39, 1329-1334.

Sasorith, S., Billas, I.M., Iwema, T., Moras, D., Wurtz, J.M., 2002. Structure-based analysis of the ultraspiracle

protein and docking studies of putative ligands. J Insect Sci 2, 25.

Sen, S.E., Ewing, G.J., Thurston, N., 1996. Characterization of lepidopteran prenyltransferase in Manduca

sexta corpora allata. . Arch Insect Biochem Physiol 32, 315–332.

Sen, S.E., Hitchcock, J.R., Jordan, J.L., Richard, T., 2006. Juvenile hormone biosynthesis in M. sexta:

substrate specificity of insect prenyltransferase utilizing homologous diphosphate analogs. Insect Biochem

Mol Biol 36, 827-834.

Sen, S.E., Sperry, A.E., 2002. Partial purification of a farnesyl diphosphate synthase from whole-body

Manduca sexta. Insect Biochem Mol Biol 32, 889-899.

Sen, S.E., Tomasello, A., Grasso, M., Denton, R., Macor, J., Beliveau, C., Cusson, M., Crowell, D.N., 2012.

Cloning, expression and characterization of lepidopteran isopentenyl diphosphate isomerase. Insect Biochem

Mol Biol 42, 739-750.

Sen, S.E., Trobaugh, C., Beliveau, C., Richard, T., Cusson, M., 2007. Cloning, expression and characterization

of a dipteran farnesyl diphosphate synthase. Insect Biochem Mol Biol 37, 1198-1206.

Shalaby, A.M., Abd el-Ghaffar, H., el Zayat, M.A., Youssef, N.S., 1990. Effect of juvenile hormone on Musca

domestica vicina Macq. II. The gonotrophic cycle. Folia Morphol (Praha) 38, 160-168.

Shemshedini, L., Lanoue, M., Wilson, T.G., 1990. Evidence for a juvenile hormone receptor involved in

protein synthesis in Drosophila melanogaster. J Biol Chem 265, 1913-1918.

Shemshedini, L., Wilson, T.G., 1990. Resistance to juvenile hormone and an insect growth regulator in

Drosophila is associated with an altered cytosolic juvenile hormone-binding protein. Proc Natl Acad Sci U S

A 87, 2072-2076.

64

Sheng, Z., Ma, L., Cao, M.X., Jiang, R.J., Li, S., 2008. Juvenile hormone acid methyl transferase is a key

regulatory enzyme for juvenile hormone synthesis in the Eri silkworm, Samia cynthica ricini. Arch Insect

Biochem Physiol 69, 143-154.

Shin, S.W., Zou, Z., Saha, T.T., Raikhel, A.S., 2012. bHLH-PAS heterodimer of methoprene-tolerant and

Cycle mediates circadian expression of juvenile hormone-induced mosquito genes. Proc Natl Acad Sci U S A

109, 16576-16581.

Shinoda, T., Itoyama, K., 2003. Juvenile hormone acid methyltransferase: a key regulatory enzyme for insect

metamorphosis. Proc Natl Acad Sci U S A 100, 11986-11991.

Smykal, V., Bajgar, A., Provaznik, J., Fexova, S., Buricova, M., Takaki, K., Hodkova, M., Jindra, M., Dolezel,

D., 2014a. Juvenile hormone signaling during reproduction and development of the linden bug, Pyrrhocoris apterus. Insect Biochem Mol Biol 45, 69-76.

Smykal, V., Daimon, T., Kayukawa, T., Takaki, K., Shinoda, T., Jindra, M., 2014b. Importance of juvenile

hormone signaling arises with competence of insect larvae to metamorphose. Dev Biol 390, 221-230.

Song, J., Wu, Z., Wang, Z., Deng, S., Zhou, S., 2014. Kruppel-homolog 1 mediates juvenile hormone action to

promote vitellogenesis and oocyte maturation in the migratory locust. Insect Biochem Mol Biol 52, 94-101.

Sperry, A.E., Sen, S.E., 2001. Farnesol oxidation in insects: evidence that the biosynthesis of insect juvenile

hormone is mediated by a specific alcohol oxidase. Insect Biochem Mol Biol 31, 171-178.

Spradling, A.C., Stern, D., Beaton, A., Rhem, E.J., Laverty, T., Mozden, N., Misra, S., Rubin, G., 1999. The

Berkely Drosophila Genome Project gene disruption project: single P-element insertions mutating 25% of vital

Drosophila genes. . Genetics 153, 135–177.

Stay, B., 1999. Diploptera punctata, in: Huber, I., Masler, E.P., Rao, B.R. (Eds.), Cockroaches as models for

neurobiology: Applications in biomedical research CRC press, Boca Raton, FL., pp. 880-889.

Stay, B., Bachmann, J.A.S., Stoltzman, C.A., Fairbairn, S.E., Yu, C.G., Tobe, S.S., 1994a. Factors affecting

allatostatin release in a cockroach (Diploptera punctata) - Nerve section, juvenile hormone analog and ovary. J

Insect Physiol 40, 365-372.

Stay, B., Chan, K.K., Woodhead, A.P., 1992. Allatostatin-immunoreactive neurons projecting to the corpora

allata of adult Diploptera punctata. Cell Tissue Res 270, 15-23.

Stay, B., Coop, A., 1973. Developmental stages and chemical composition in embryos of cockroach,

Diploptera punctata, with observations on effect of diet. J Insect Physiol 19, 147-171.

Stay, B., Coop, A.C., 1974. Milk Secretion for Embryogenesis in a Viviparous Cockroach. Tissue & Cell 6,

669-693.

Stay, B., Friedel, T., Tobe, S.S., Mundall, E.C., 1980. Feedback control of juvenile hormone synthesis in

cockroaches - Possible role for ecdysterone. Science 207, 898-900.

Stay, B., Joshi, S., Woodhead, A.P., 1991. Sensitivity to allatostatins of corpora allata from larval and adult

female Diploptera punctata. Journal of Insect Physiology 37, 63-70.

Stay, B., Lin, H.L., 1981. The inhibition of milk synthesis by juvenile hormone in the viviparous cockroach,

Diploptera punctata. J Insect Physiol 27, 551-&.

Stay, B., Ostedgaard, L.S., Tobe, S.S., Strambi, A., Spaziani, E., 1984. Ovarian and hemolymph titers of

ecdysteroid during the gonadotropic cycle in Diploptera punctata. J Insect Physiol 30, 643-651.

Stay, B., Roth, L.M., 1958. The reproductive behavior of Diploptera punctata (Blattaria: Diplopteridae).

Proceedings 10th international congress of entomology 2, 5.

Stay, B., Tobe, S.S., 1977. Control of juvenile hormone biosynthesis during the reproductive cycle of a

viviparous cockroach 1. Activation and inhibition of corpora allata. Gen Comp Endocrinol 33, 531-540.

65

Stay, B., Tobe, S.S., 1978. Control of juvenile hormone biosynthesis during reproductive cycle of a viviparous

cockroach 2. Effects of unilateral allatectomy, implantation of supernumerary corpora allata, and ovariectomy.

Gen Comp Endocrinol 34, 276-286.

Stay, B., Tobe, S.S., 2007. The role of allatostatins in juvenile hormone synthesis in insects and crustaceans.

Annual Review of Entomology 52, 277-299.

Stay, B., Tobe, S.S., Bendena, W.G., 1994b. Allatostatins - identification, primary structures, functions and

distribution. Advances in Insect Physiology, Vol 25 25, 267-337.

Stay, B., Tobe, S.S., Mundall, E.C., Rankin, S., 1983. Ovarian stimulation of juvenile hormone biosynthesis in

the viviparous cockroach, Diploptera punctata. Gen Comp Endocrinol 52, 341-349.

Stay, B., Zhang, J.R., Tobe, S.S., 2002. Methyl farnesoate and juvenile hormone production in embryos of

Diploptera punctata in relation to innervation of corpora allata and their sensitivity to allatostatin. Peptides 23,

1981-1990.

Steiner, B., Pfister-Wilhelm, R., Grossniklaus-Burgin, C., Rembold, H., Treiblmayr, K., Lanzrein, B., 1999.

Titres of juvenile hormone I, II and III in Spodoptera littoralis (Noctuidae) from the egg to the pupal moult and

their modification by the egg-larval parasitoid Chelonus inanitus (Braconidae). J Insect Physiol 45, 401-413.

Sun, X.F., Li, Z.X., 2012. In silico and in vitro analyses identified three amino acid residues critical to the

catalysis of two aphid farnesyl diphosphate synthase. Protein J 31, 417-424.

Sutherland, J.D., Kozlova, T., Tzertzinis, G., Kafatos, F.C., 1995. Drosophila hormone receptor 38: a second

partner for Drosophila USP suggests an unexpected role for nuclear receptors of the nerve growth factor-

induced protein B type. Proc Natl Acad Sci U S A 92, 7966-7970.

Sutherland, T.D., Feyereisen, R., 1996. Target of cockroach allatostatin in the pathway of juvenile hormone

biosynthesis. Mol Cell Endocrinol 120, 115-123.

Sutherland, T.D., Unnithan, G.C., Feyereisen, R., 2000. Terpenoid omega-hydroxylase (CYP4C7) messenger

RNA levels in the corpora allata: a marker for ovarian control of juvenile hormone synthesis in Diploptera

punctata. J Insect Physiol 46, 1219-1227.

Szibbo, C.M., Tobe, S.S., 1981. Cellular and volumetric changes in relation to the activity cycle in the corpora

allata of Diploptera punctata. J Insect Physiol 27, 655-665.

Taban, A.H., Tittiger, C., Blomquist, G.J., Welch, W.H., 2009. Isolation and characterization of farnesyl

diphosphate synthase from the cotton boll weevil, Anthonomus grandis. Arch Insect Biochem Physiol 71, 88-

104.

Teal, P.E., 2002. Effects of allatotropin and allatostatin on in vitro production of juvenile hormones by the

corpora allata of virgin females of the moths of Heliothis virescens and Manduca sexta. Peptides 23, 663-669.

Teal, P.E., Jones, D., Jones, G., Torto, B., Nyasembe, V., Borgemeister, C., Alborn, H.T., Kaplan, F., Boucias,

D., Lietze, V.U., 2014. Identification of methyl farnesoate from the hemolymph of insects. J Nat Prod 77, 402-

405.

Thompson, C.S., Tobe, S.S., 1986. Electrical properties of membranes and cells of the corpora allata of the

cockroach Diploptera punctata: evidence for the presence of voltage-sensitive calcium channels., in: Borkovec,

A.B., Gelman, D.B. (Eds.), Insect Neurochemistry and Neurophysiology. Humana Press, Totowa, pp. 375-378.

Thompson, C.S., Yagi, K.J., Chen, Z.F., Tobe, S.S., 1990. The effects of octopamine on juvenile hormone

biosynthesis, electrophysiology, and cAMP content of the corpora allata of the cockroach Diploptera punctata.

Journal of Comparative Physiology B-Biochemical Systemic and Environmental Physiology 160, 241-249.

Tittiger, C., Barkawi, L.S., Bengoa, C.S., Blomquist, G.J., Seybold, S.J., 2003. Structure and juvenile

hormone-mediated regulation of the HMG-CoA reductase gene from the Jeffrey pine beetle, Dendroctonus

jeffreyi. Mol Cell Endocrinol 199, 11-21.

66

Tittiger, C., Blomquist, G.J., Ivarsson, P., Borgeson, C.E., Seybold, S.J., 1999. Juvenile hormone regulation of

HMG-R gene expression in the bark beetle Ips paraconfusus (Coleoptera: Scolytidae): implications for male

aggregation pheromone biosynthesis. Cell Mol Life Sci 55, 121-127.

Tittiger, C., O'Keeffe, C., Bengoa, C.S., Barkawi, L.S., Seybold, S.J., Blomquist, G.J., 2000. Isolation and

endocrine regulation of an HMG-CoA synthase cDNA from the male Jeffrey pine beetle, Dendroctonus

jeffreyi (Coleoptera: Scolytidae). Insect Biochem Mol Biol 30, 1203-1211.

Tobe, S.S., Clarke, N., Stay, B., Ruegg, R.P., 1984. Changes in cell number and activity of the corpora allata

of the cockroach Diploptera punctata - a role for mating and the ovary. Canadian Journal of Zoology-Revue

Canadienne De Zoologie 62, 2178-2182.

Tobe, S.S., Ruegg, R.P., Stay, B.A., Baker, F.C., Miller, C.A., Schooley, D.A., 1985. Juvenile hormone titer

and regulation in the cockroach Diploptera punctata. Experientia 41, 1028-1034.

Tobe, S.S., Stay, B., 1977. Corpus allatum activity in vitro during reproductive cycle of viviparous cockroach,

Diploptera punctata (Eschscholtz). Gen Comp Endocrinol 31, 138-147.

Tobe, S.S., Stay, B., 1979. Modulation of juvenile hormone synthesis by an analog in the cockroach. Nature

281, 481-482.

Tobe, S.S., Stay, B., 1980. Control of juvenile hormone biosynthesis during the reproductive cycle of a

viviparous cockroach. III. Effects of denervation and age on compensation with unilateral allatectomy and

supernumerary corpora allata. Gen Comp Endocrinol 40, 89-98.

Tobe, S.S., Stay, B., 1985a. Structure and regulation of the corpus allatum. Advances in Insect Physiology 18,

305-432.

Tobe, S.S., Stay, B., 1985b. Structure and regulation of the corpus allatum. Advances in Insect Physiology 18,

305-432.

Tsai, W.H., Holbrook, G.L., Schal, C., Chiang, A.S., 1995. In vitro growth of corpora allata from Diploptera punctata. In Vitro Cellular & Developmental Biology-Animal 31, 542-546.

Tu, M., Kou, R., Wang, Z., Stoffolano, J.G., Jr., Yin, C., 2001. Immunolocalization and possible effect of a

moth allatotropin-like substance in a fly, Phormia regina (Diptera: Calliphoridae). J Insect Physiol 47, 233-

244.

Unnithan, G.C., Feyereisen, R., 1995. Experimental acquisition and loss of allatostatin sensitivity by corpora

allata of Diploptera punctata. J Insect Physiol 41, 975-980.

Unnithan, G.C., Sutherland, T.D., Cromey, D.W., Feyereisen, R., 1998. A factor causing stable stimulation of

juvenile hormone synthesis by Diploptera punctata corpora allata in vitro. J Insect Physiol 44, 1027-1037.

Valle, D., 1993. Vitellogenesis in insects and other groups--a review. Mem Inst Oswaldo Cruz 88, 1-26.

Vandermoten, S., Haubruge, E., Cusson, M., 2009a. New insights into short-chain prenyltransferases:

structural features, evolutionary history and potential for selective inhibition. Cell Mol Life Sci 66, 3685-3695.

Vandermoten, S., Santini, S., Haubruge, E., Heuze, F., Francis, F., Brasseur, R., Cusson, M., Charloteaux, B.,

2009b. Structural features conferring dual geranyl/farnesyl diphosphate synthase activity to an aphid

prenyltransferase. Insect Biochem Mol Biol 39, 707-716.

Veenstra, J.A., Costes, L., 1999. Isolation and identification of a peptide and its cDNA from the mosquito

Aedes aegypti related to Manduca sexta allatotropin. Peptides 20, 1145-1151.

Vilaplana, L., Maestro, J.L., Piulachs, M.D., Belles, X., 1999. Modulation of cardiac rhythm by allatostatins in

the cockroach Blattella germanica (L.) (Dictyoptera, Blattellidae). J Insect Physiol 45, 1057-1064.

Wang, Z.W., Ding, Q., Yagi, K.J., Tobe, S.S., 1994. Terminal stages in juvenile hormone biosynthesis in

corpora allata of Diploptera punctata - developmental changes in enzyme activity and regulation by

allatostatins. J Insect Physiol 40, 217-223.

67

Wigglesworth, V.B., 1934. The physiology of ecdysis in Rhodnius prolixus (Hemiptera). II. Factors controlling

molting and ‘metamorphosis’. Quart. J Microsc Sci s2–77, 191–222.

Williams, C.M., 1956. The juvenile hormone of insects. Nature 178, 212–213.

Williford, A., Stay, B., Bhattacharya, D., 2004. Evolution of a novel function: nutritive milk in the viviparous

cockroach, Diploptera punctata. Evo Dev 6, 67-77.

Wilson, T.G., Ashok, M., 1998. Insecticide resistance resulting from an absence of target-site gene product.

Proc Natl Acad Sci U S A 95, 14040-14044.

Wilson, T.G., Fabian, J., 1986. A Drosophila melanogaster mutant resistant to a chemical analog of juvenile

hormone. Dev Biol 118, 190-201.

Woodhead, A.P., 1986. Male age - Effect on mating behavior and success in the cockroach Diploptera punctata. Animal Behaviour 34, 1874-1879.

Woodhead, A.P., Khan, M.A., Stay, B., Tobe, S.S., 1994. 2 new allatostatins from the brains of Diploptera

punctata. Insect Biochem Mol Biol 24, 257-263.

Woodhead, A.P., Stay, B., Seidel, S.L., Khan, M.A., Tobe, S.S., 1989. Primary structure of 4 allatostatins -

neuropeptide inhibitors of juvenile hormone synthesis. Proc Natl Acad Sci U S A 86, 5997-6001.

Woodhead, A.P., Thompson, M.E., Chan, K.K., Stay, B., 2003. Allatostatin in ovaries, oviducts, and young

embryos in the cockroach Diploptera punctata. J Insect Physiol 49, 1103-1114.

Yagi, K.J., Konz, K.G., Stay, B., Tobe, S.S., 1991. Production and utilization of farnesoic acid in the juvenile

hormone biosynthetic pathway by corpora allata of larval Diploptera punctata. Gen Comp Endocrinol 81, 284-

294.

Yao, T.P., Segraves, W.A., Oro, A.E., McKeown, M., Evans, R.M., 1992. Drosophila ultraspiracle modulates

ecdysone receptor function via heterodimer formation. Cell 71, 63-72.

Yin, C.M., Zou, B.X., Jiang, M.G., Li, M.F., Qin, W.H., Potter, T.L., Stoffolano, J.G., 1995. Identification of

juvenile hormone III bisepoxide (JHB3), juvenile hormone III and methyl farnesoate secreted by the corpus

allatum of Phormia regina (Meigen), in vitro and function of JHB3. Either applied alone or as a part of a

juvenoid blend. J Insect Physiol 41, 473-479.

Yu, C.G., Hayes, T.K., Strey, A., Bendena, W.G., Tobe, S.S., 1995a. Identification and partial characterization

of receptors for allatostatins in brain and corpora allata of the cockroach Diploptera punctata using a binding

assay and photoaffinity labeling. Regul Pept 57, 347-358.

Yu, C.G., Stay, B., Ding, Q., Bendena, W.G., Tobe, S.S., 1995b. Immunochemical identification and

expression of allatostatins in the gut of Diploptera punctata. J Insect Physiol 41, 1035-1043.

Zhang, Y.L., Li, Z.X., 2008. Two different farnesyl diphosphate synthase genes exist in the genome of the

green peach aphid, Myzus persicae. Genome 51, 501-510.

Zhang, Z., Xu, J., Sheng, Z., Sui, Y., Palli, S.R., 2011. Steroid receptor co-activator is required for juvenile

hormone signal transduction through a bHLH-PAS transcription factor, methoprene tolerant. J Biol Chem 286,

8437-8447.

68

Chapter 2

Characterization of the Juvenile Hormone pathway in the

viviparous cockroach, Diploptera punctata

This chapter is an adapted reprint of my article:

Juan Huang, Elisabeth Marchal, Ekaterina F. Hult, and Stephen S. Tobe, Characterization of the

Juvenile Hormone pathway in the viviparous cockroach, Diploptera punctata. PloS one,

accepted.

Authors’ contribution:

E. Marchal performed part of the sequencing of the JH biosynthetic genes, and the tissue

distribution and expression profiles. I performed the RNAi and RCA. The paper was written by

E. Marchal and me with editing by E.F. Hult and S.S. Tobe

69

Summary

Juvenile hormones (JHs) are key regulators of insect development and reproduction. The JH

biosynthetic pathway is known to involve 13 discrete enzymatic steps. In the present study, we

have characterized the JH biosynthetic pathway in the cockroach Diploptera punctata. The effect

of exogenous JH precursors on JH biosynthesis was also determined. Based on sequence

similarity, orthologs for the genes directly involved in the pathway were cloned, and their spatial

and temporal transcript profiles were determined. The effect of shutting down the JH pathway in

adult female cockroaches was studied by knocking down genes encoding HMG-CoA reductase

(HMGR) and Juvenile hormone acid methyltransferase (JHAMT). As a result, oocyte

development slowed as a consequence of reduction in JH biosynthesis. Oocyte length, fat body

transcription of Vg and ovarian vitellin content significantly decreased. In addition, silencing

HMGR and JHAMT resulted in a decrease in the transcript levels of other genes in the pathway.

Introduction

Juvenile hormones (JHs) play key roles in regulating growth, development, metamorphosis,

aging, caste differentiation and reproduction in insects (as reviewed by Goodman and Cusson,

2012; Hartfelder, 2000). The multiple processes in which JH is involved and the critical role

which JH plays in metamorphosis and reproduction emphasize the importance of elucidating the

JH biosynthetic pathway and the factors that regulate its biosynthesis.

JHs are sesquiterpenoid compounds that are synthesized and secreted by specialized, paired

endocrine glands, the corpora allata (CA). The complete biosynthetic pathway of JH III (the most

widespread and predominant JH homologue in insects) comprises 13 discrete enzymatic steps.

This pathway can be divided into two metabolic parts (Fig. 2.1): the early portion comprises the

70

mevalonate pathway to the formation of farnesyl diphosphate (FPP) and is conserved in both

vertebrates and invertebrates; the later part of the pathway is specific to insects and other

arthropods. In this later part, FPP is cleaved to farnesol, which is then oxidized to the carboxylic

acid (farnesoic acid; FA), followed by methyl esterification, epoxidation and formation of JH

(Belles et al., 2005). The order in which these two final steps in JH biosynthesis occurs, appears

to be insect order dependent. In orthopteran, coleopteran, dipteran and dictyopteran insects, FA

is first methylated to MF, which in turn undergoes a C10, C11 epoxidation to the functional JH.

In Lepidoptera, however, the opposite situation prevails: epoxidation precedes methylation

(Defelipe et al., 2011; Marchal et al., 2011; Shinoda and Itoyama, 2003).

Recent studies have reported on the molecular elucidation of the JH pathway in several

holometabolous insects such as the silkworm Bombyx mori, the mosquito, Aedes aegypti and the

honeybee, Apis mellifera. In B. mori, all genes encoding enzymes involved in the mevalonate

pathway (Kinjoh et al., 2007) and the isoprenoid branch of JH biosynthesis (Cheng et al., 2014;

Shinoda and Itoyama, 2003) have been isolated. Each enzyme in the mevalonate pathway is

encoded by a single gene, except farnesyl diphosphate synthase (FPPS), which comprises three

homologs. The genes encoding enzymes in the isoprenoid branch of JH biosynthesis, however,

underwent gene duplication to create multiple copies. The transcripts for most JH enzymes are

highly enriched or exclusively expressed in the CA (Ueda et al., 2009). The expression pattern of

the genes encoding these enzymes in the CA correlates well with rates of JH biosynthesis

(Kinjoh et al., 2007). In A. aegypti, changes in the transcription of 11 of the enzymes are

responsible in part for the dynamic changes in JH biosynthesis (Nouzova et al., 2011). The

expression of genes in the JH biosynthetic pathway was also determined in female castes of A.

mellifera and was found to correlate with the JH hemolymph titre in adult worker bees, but, not

71

Fig. 2.1. Scheme of JH biosynthetic pathway (Adapted from Belles et al. (2005) and Nouzova

et al. (2011)). The Insect- and Arthropod-specific pathway is represented in the dashed box.

Precursors are in bold and connected by arrows. Enzymes are in italics. Abbreviations for the

enzymes are given in brackets.

Acetyl-CoA

Acetoacetyl-CoA

HMG-CoA

Mevalonic acid

Phosphomevalonate

Diphosphomevalonate

Isopentenyl diphosphate Dimethylallyl diphosphate

Geranyl diphosphate

Farnesyl diphosphate

Farnesol

Farnesal

Farnesoic acid

Methyl farnesoate

Juvenile hormone III

Acetoacetyl-CoA thiolase (1- Thiol)

HMG-CoA synthase (2- HMGS)

HMG-CoA reductase (3- HMGR)

Mevalonate kinase (4- MK)

Phosphomevalonate kinase (5- PMK)

Diphosphomevalonate decarboxylase (6- PPMD)

Isophentenyl diphosphate isomerase (7- IPPI)

Farnesyl diphosphate synthase (8- FPPS)

Farnesyl diphosphate synthase (8- FPPS)

Farnesyl diphosphate pyrophosphatase (9- FPPP)

Farnesol dehydrogenase (10 – FOLD)

Farnesal dehydrogenase (11 – FALD)

Juvenile hormone acid methytransferase (12- JHAMT)

Methyl farnesoate epoxidase (13- CYP15A1)

72

in larvae (Bomtorin et al., 2014). Until recently, no structural or molecular data were available

on FPP pyrophosphatase (FPPP) or farnesal dehydrogenase (FALD). Current studies in the

dipterans, the fly, Drosophila melanogaster and A. aegypti have characterized the gene encoding

an FPP phosphatase (FPPP) (Cao et al., 2009; Nyati et al., 2013). Moreover, genes encoding an

aldehyde dehydrogenase (FALD) have now been functionally characterized in A. aegypti

(Rivera-Perez et al., 2013).

Diploptera punctata, the only truly viviparous cockroach is a well-known model system in the

study of JH biosynthesis and its regulation. The physiology of this animal is characterized by

very stable and high rates of JH biosynthesis and precise and predictable reproductive events that

correlate well with rates of JH production (see review by Marchal et al. (2013a)). In adult

females, JH regulates oocyte maturation, fat body vitellogenin (Vg) production and the uptake of

Vg by the developing oocytes (Rankin and Stay, 1984; Stay and Tobe, 1978). Aside from the

original molecular identification of CYP15A1 (Helvig et al., 2004), the gene encoding the

epoxidase producing the functional JH, JH-related research in D. punctata has mainly focused on

examining JH titre and physiological aspects of JH function (see Marchal et al. (2013a)).

An earlier study predicted 4 genes encoding enzymes in JH biosynthetic pathway in D. punctata

based on sequence similarity with Drosophila and Anopheles gambiae genes (Noriega et al.,

2006). To further characterize the JH biosynthetic pathway in this important model system on a

molecular level, we have confirmed and identified 11 out of 13 genes encoding the JH

biosynthetic enzymes, and have also analyzed the tissue distribution and developmental

transcript profile of these genes during the first gonadotropic cycle of the female cockroach. The

predominant expression of these genes in the CA and the correlation between their transcript

levels and the rates of JH biosynthesis suggests that the 11 genes cloned in our study are

73

encoding functional enzymes. In addition, the transcript levels of the 11 genes indicate that the

expression of these genes partly regulate JH biosynthesis. Addition of exogenous JH precursors

was able to stimulate JH biosynthesis in CA with low activity suggesting that JH production was

also regulated by flux of substrates in the pathway.

We also used RNA interference (RNAi) to study the effect of silencing genes in the pathway on

rates of JH biosynthesis and ovarian development. HMGR and JHAMT were chosen as the target

genes for the RNAi, because both are well-studied genes whose expression patterns clearly

correlate with rates of JH biosynthesis (Goodman and Cusson, 2012). Our results show that JH

biosynthesis decreased and ovarian development slowed in HMGR-JHAMT dsRNA treated

animals. Of particular interest is our discovery that manipulation of individual genes encoding

the JH biosynthetic enzymes using dsRNA technology resulted in a significant decrease in the

transcript levels of other genes, which indicates a feedback mechanism is involved in the

regulation of the expression of enzymes in the JH biosynthetic pathway.

Materials and methods

Animals - The D. punctata colony was maintained at 27 °C in constant darkness and animals

were fed lab chow and water at libitum. To obtain pools of synchronised animals, newly molted

female adult cockroaches were picked from the colony, placed in separate containers and

provided with water and lab chow. Mated status was confirmed by the presence of a

spermatophore.

Tissue collection and RNA extraction - D. punctata were dissected in modified cockroach ringer

solution (150 mM NaCl, 12 mM KCl, 10 mM CaCl2.2H2O, 3 mM MgCl2.6H2O, 10 mM HEPES,

40 mM glucose, pH 7.2) using a dissecting microscope. Samples were flash-frozen in liquid N2

74

to prevent RNA degradation and were stored at -80 ºC until further processing. For each

dissected female, basal oocyte lengths were measured to determine the physiological age. CA

samples were taken from day 0 to 7 of adult females for the developmental profiles of genes

encoding enzymes in the JH biosynthetic pathway. For each time point, three biologically

independent pools of 10 animals each were collected. To determine the tissue distribution of the

genes of interest, the following tissues were dissected from 3 independent pools of 10 animals

each: brain (Br), nerve cord (NC), corpora allata (CA), fat body (FB), midgut (MG), Malpighian

tubules (MT), ovary (Ov) from females, and accessory gland (AG) and testes (Te) from males.

For the RNAi experiments (§2.6), 3 biologically independent pools comprising CA from 7

animals were collected. Pooled samples were homogenised with RNase-free pestles and total

RNA was extracted using the RNeasy Mini Kit (Qiagen) according to the manufacturer’s

instructions. An additional DNase treatment (RNase-free DNase set, Qiagen) was performed to

eliminate potential genomic DNA contamination. Because of the small size of the CA, RNA

from this tissue was extracted using the RNAqueous®-Micro Kit (Ambion), followed by the

recommended DNase step. Quality and concentration of the resulting RNA samples were

measured using a Nanodrop spectrophotometer (Thermo Scientific.).

Sequencing genes involved in the JH biosynthetic pathway - Since no genome or full

transcriptome sequence data are currently available for D. punctata, a first set of degenerate

primers was developed based on a multiple sequence alignment of known orthologous sequences

from different insects. Such an alignment was made for Acetoacetyl-CoA thiolase (Thiol),

HMG-CoA reductase (HMGR), Phosphomevalonate kinase (PK), Diphosphomevalonate

decarboxylase (PPMD), Isopentenyl diphosphate isomerase (IPPI), Juvenile hormone acid

methyltransferase (JHAMT) and the JH target vitellogenin (Vg). Degenerate primers are listed in

75

Supplementary table S1. For Methyl farnesoate epoxidase (CYP15A1) the full sequence was

already characterised (GenBank accession number AY509244) (Helvig et al., 2004). For HMG-

CoA synthase (HMGS), Mevalonate kinase (MK), Farnesyl diphosphate synthase (FPPS) and

Farnesol dehydrogenase (FOLD), a partial sequence was found in the D. punctata CA EST

database (Noriega et al., 2006). Their sequence was partially confirmed using gene-specific

primers (Table S1). A temperature-gradient polymerase chain reaction (PCR) was run using Taq

DNA polymerase (Sigma-Aldrich Co.) with D. punctata day 4 CA cDNA as a template. Bands

of the expected size were cut out and further purified using the GenEluteTM

Gel extraction Kit

(Sigma-Aldrich Co.). The resulting DNA fragments were subcloned into a CloneJET™ cloning

vector using the CloneJET™ PCR Cloning Kit (Fermentas) and sequenced. Sequences for PMK,

PPMD and JHAMT identified using degenerate primers, were too short to submit to NCBI’s

GenBank; therefore, RACE (Rapid Amplification of cDNA Ends) was performed. Their

sequences were confirmed using primers listed in Table S1.

Radiochemical assay (RCA) - The in vitro radiochemical assay (RCA) for JH biosynthesis was

performed (Feyereisen and Tobe, 1981; Tobe and Clarke, 1985). CA were incubated in TC199

medium for 3h, then transferred to new medium containing JH precursors for another 3h

incubation. JH biosynthesis in both incubations was measured, and first incubation

measurements were used as a control. JH precursors acetyl CoA, mevalonic acid (MA),

diphosphomevalonate (DPPM) and farnesol (Sigma-Aldrich, Canada) were dissolved in water

before use.

cDNA synthesis and quantitative real-time PCR (q-RT-PCR) - cDNA was transcribed from an

equal amount of RNA using the SuperScript™ III First-Strand Synthesis SuperMix for q-RT-

PCR in a final volume of 20 µl following the manufacturer’s instructions (Invitrogen Life

76

Technologies). All samples were reverse transcribed together in a single run. The resulting

cDNA samples were diluted 10-fold with PCR grade water. A calibrator sample was prepared by

pooling 5 µl of each cDNA sample. In the same run, negative control reactions were set up

without reverse transcriptase enzyme to test for genomic DNA contamination.

q-RT-PCR primers were designed using IDT’s (Integrated DNA Technologies) PrimeQuest

design tool (http://eu.idtdna.com/PrimerQuest/Home/Index). Primer sets were subsequently

validated by determining relative standard curves for each gene transcript using a five-fold serial

dilution of the calibrator cDNA sample. Efficiency and correlation coefficients (R²) were

determined for each primer pair. Primers used for q-RT-PCR profiling are listed in Table 2.1.

q-RT-PCR reactions were carried out in triplicate in a total volume of 10 µl containing 5 µl of

IQ™ SYBR® Green Supermix (Bio-Rad), 1 µl forward and reverse primer (5 µM), 2 µl of MQ-

water and 1 µl of cDNA. All reactions were performed using Bio-Rad’s CFX384 Touch ™ Real-

Time PCR Detection System using a two-step thermal cycling profile: 95ºC for 3 min, followed

by 40 cycles of 95ºC for 10 s and 59ºC for 30 s. Upon completion of every run, a dissociation

protocol (melt curve analysis) was performed to check for formation of primer dimers. A few

representative PCR products were run on a 1.2% agarose gel containing GelRedTM (Biotium)

and visualised under UV to confirm target specificity. Prior to target gene profiling, previously

described housekeeping genes were tested for their stability in the designed tissue distribution,

temporal profiling and RNAi experiments. The optimal housekeeping genes were selected using

the geNorm and Normfinder software as described previously (Marchal et al., 2013b). For each

tested cDNA sample, the normalization factor for the reference genes relative to the calibrator

samples was calculated and used to determine the normalized expression levels of the target

genes relative to the calibrator (Vandesompele et al., 2002).

77

Table 2.1. q-RT-PCR primer sequences and reaction efficiencies and correlation coefficients in the q-RT-PCR assay.

Gene name F primers R primers Efficiency (%) Correlation

coefficient (R2)

Thiol 5’-TGCCTTCCAAAAGGAGAATG-3’ 5’- ACATCACCTGCCATCAACAC-3’ 90 0.97

HMGS 5’- TGCTGGGAAGTACACAGCAGG-3’ 5’- CTCCACGAGCTTGCTGACTG-3’ 83 0.994

HMGR 5’-TGGGAGCATGTTGTGAAAAT-3’ 5’-ACCAAGCAGCCCTCAGTAGT-3’ 95 0.988

MK 5’- TACGGCAAAACTGCCCTTGC-3’ 5’- AATGGAGGAGGTTCGGCGA-3’ 93 0.996

PMK 5’- TACGAAAACAACGAGGATGG-3’ 5’- TTCTGCATCATCTACACCTTCA-3’ 100 0.985

PPMD 5’- TGGAAGGTGACATAACAGCAA-3’ 5’- ATCCTTGATGCCAGTGAACA-3’ 90 0.967

IPPI 5’- CCTTCCCCAACCATGTAACT-3’ 5’- ACCAACGCCATTTGTCTCTT-3’ 100 0.992

FPPS 5’-TGCTTTGGAGATCCTGAGGT-3’ 5’- TGTTCAGGAGTGGTTCGTTG-3’ 96 0.987

FOLD 5’- TGGCGCGTAGGGTAGACAA-3’ 5’- GACCCATTTGAAAGCCTCCTTGA-3’ 93 0.993

JHAMT 5’- ATCCAGGTGCTGGAAGGAGAG-3’ 5’- CTGCCCAGAGTCGAACAGG-3’ 99 0.984

CYP15A1 5’- GTTGGGATCTCGGAGCATGG-3’ 5’- CGAACACGTCATGCATCGGT-3’ 100 0.992

Vg AAAGGTGTCCTCAGCCAGC TCCTCCATCTCGGATTGGGA 95 0.998

78

RNA interference (RNAi) - CA cDNA was used as a template to amplify fragments of genes to be

used in dsRNA construction. These fragments were subcloned and sequenced as described above.

Two separate constructs were designed in different regions of the genes to eliminate off-target

effects. Primers used are given in Table 2.2. dsRNA was synthesized using the MEGAscript

RNAi kit (Ambion). The procedure is based on the high-yield transcription reaction of a user-

provided linear transcript with a T7 promoter sequence. Transcription was carried out at 37 °C

overnight. The reaction mixture was treated with DNase I and RNase, and then purified by

phase-solid phase adsorption purification, according to the manufacturer’s instructions (Ambion).

The dsRNA concentration was determined using a Nanodrop spectrophotometer (Thermo

Scientific.). Diluted dsRNA was run on a 1.2% agarose gel to examine integrity of the construct

and efficiency of duplex formation. The negative control construct (-pJET) was designed in a

non-coding region of the pJET 1.2/blunt cloning vector (CloneJet PCR Cloning Kit, Thermo

Scientific).

Several RNAi trials with different injection timings and dsRNA construct concentrations were

investigated to obtain an efficient gene silencing (data not shown). Newly molted mated females

were injected with 3 μg HMGR dsRNA or control dsRNA on day 0 and day 2, and with 3 μg

JHAMT dsRNA or control dsRNA on day 1 and day 3. CA, ovary and fat body samples were

taken on day 4 as described above. Fat body was immediately stored in liquid N2 prior to RNA

extraction. CA were dissected and cleaned in TC199 medium (GIBCO; 1.3 mM Ca2+, 2% Ficoll,

methionine-free) for use in the RCA or flash frozen in liquid N2 to prevent RNA degradation.

Basal oocyte length was measured and the ovaries were collected for vitellogenin measurements

(§ 2.7) and histology (§ 2.8).

79

Table 2.2. Primers for dsRNA construction

Name F Primer(5'-3') R Primer(5'-3')

HMGR dsRNA ACATGGACAGTTCTGTGCCT CCCAACTTTTGCAGATGACAG

CACTTCTCGCATTGTGGCT CAGTACCCTTGGAGAGC

JHAMT dsRNA AAAAGAGACGCAGCCCACGCA CGATCCTCGTGGGAACAGATG

GTACAGCACGCCACCTCCA AACTACGGCACTCTGGAGC

Control dsRNA TTGCGCTCACTGCCAATTGC CTGGCCTTTTGCTCACATGTT

*The T7 promoter sequence was added at the 5' end of each dsRNA primer.

80

Vitellogenin ELISA - Single dissected ovaries were homogenized and extracted twice in 50 µl

PBS according to Mundall et al. (1981). Total ovary protein content was determined using a

Bradford assay (Bradford, 1976) performed according to the manufacturer’s (Sigma-Aldrich) 96

well plate protocol.

Vitellin was quantified using an indirect enzyme-linked immunosorbent assay (ELISA)

following the protocol described in Marchal et al. (2014). A rabbit polyclonal antibody made

against D. punctata mature egg homogenate was used as the first antibody (Stoltzman and Stay,

1997). Goat anti-rabbit IgG, HRP-linked antibody (1:3000 in 1% BSA) was then added as the

second antibody. Wells were treated with 100ul of TMB for 10 min. Absorbance was measured

at 650 nm with a Molecular Devices SpectraMax Pus 384 microplate reader.

Histology and microscopy - To view overall structure, a subset of oocytes was fixed for 3 days in

3% glutaraldehyde in 0.1 M phosphate buffer, post-fixed for 1 hr with 1% OsO4 in 0.1 M

phosphate buffer, dehydrated in ascending ethanol series, and embedded in Spurr’s resin.

Sections (1 µm) were cut with a Leica EM UC6 ultramicrotome, then mounted and stained with

methylene blue and toluidine blue. All light microscopy was conducted using a Leica DMI3000

inverted microscope.

Results

Identification of genes encoding JH biosynthetic enzymes and Vg in D. punctata.

CYP15A1 was previously identified in D. punctata and the genes encoding HMGS, MK, FPPS

and FOLD were present in a CA EST database of D. punctata (Helvig et al., 2004; Noriega et al.,

2006). We confirmed the sequence of HMGS, MK, FPPS and FOLD. For the remaining enzymes

in the JH biosynthetic pathway, we aligned (predicted) orthologous sequences from different

81

insect orders and designed degenerate primers based on conserved domains. We succeeded in

cloning an additional five (partial) sequences for enzymes in the conserved mevalonate pathway:

Thiol, HMGR, PMK, PPMD and IPPI and one extra sequence encoding an enzyme in the JH-

specific part of the pathway, JHAMT. The complete sequences for HMGR and JHAMT were

obtained. The ORFs of HMGR and JHAMT encode proteins of 825 and 274 amino acids in

length, respectively. Amino acid alignments of these proteins with known insect orthologs are

given in Supplementary Fig. S2.1. Moreover, a partial sequence of Vg was cloned from D.

punctata adult female fat body. The sequences for the genes were deposited in NCBI’s GenBank

with accession numbers shown in Table S2.1.

Tissue specificity and developmental transcript profiles during the first gonadotropic cycle are

consistent with roles of these genes in JH biosynthesis.

Using q-RT-PCR, the relative transcript levels of 11 genes in the JH biosynthetic pathway were

determined in several tissues of day 4 mated females and males. Relative transcript levels for the

Diploptera orthologs of genes in the JH pathway were normalized to transcript levels of the

reference genes EF1a and Tubulin (Marchal et al., 2013b). To represent the tissue distribution

data for the 11 enzyme-encoding genes in one figure, the transcript levels were normalized to the

transcript level of PMK measured in the calibrator sample. The orthologous genes encoding

enzymes in the JH biosynthetic pathway were most highly expressed in the CA, which is

consistent with the functions of the enzymes (Fig. 2.2). Most genes appear to be exclusively

expressed in the day 4 female CA: Thiol, HMGS, HMGR, MK, PPMD, FPPS, JHAMT and

CYP15A1; whereas a few show a broader tissue distribution: PMK, IPPI and FOLD. These latter

3 appear to be expressed not only in nervous tissues but also in peripheral tissues such as the

ovary, fat body and Malpighian tubules. Moreover, there are 1000-fold differences in relative

82

Figure 2.2. Tissue specific expression of genes encoding JH biosynthetic enzymes. All tissues were dissected from day 4 mated

adult females, except accessory glands (AG) and testes (Test) were from day 4 adult male. Abbreviations on the X-axis: Brain (Br),

corpora allata (CA), nerve cord (NC), midgut (Mg), Malpighian tubules (MT), fat body (Fb), and ovary (Ov). Bars represent the mean

of three biologically independent pools of ten animals run in triplicate and normalized to Tubulin and EF1α. Vertical error bars

indicate SEM.

83

mRNA levels of the enzyme-encoding genes, with PMK, PPMD, IPPI and FOLD being present

in low abundance, Thiol, HMGS and HMGR being of intermediate abundance and MK, FPPS,

JHAMT and CYP15A1 being represented in high abundance (Fig. S2.2).

Our next step was to follow the relative transcript levels in the CA of mated females throughout

the first gonadotropic cycle. Target gene expression was normalized to transcript levels of EF1a

and Armadillo according to a previous study (Marchal et al., 2013b). In general, the expression

of the 11 genes is correlated with the in vitro rates of JH biosynthesis in the CA and relative

mRNA levels for DippuVg measured in the female fat body (Fig. 2.3). The relative mRNA levels

in the CA for most of the genes of interest were low at the beginning of the adult female stage

when JH biosynthesis is low; rose during the vitellogenic portion of the first gonadotropic cycle

reaching a peak on day 3-4 and thereafter began to decline on day 5, attaining a low level during

oviposition on day 7. However, there appear to be three exceptions to this pattern: PMK, IPPI

and FPPS. Relative mRNA quantities for PMK and IPPI were very low compared to other genes

and did not display dramatic changes during the first gonadotropic cycle. FPPS transcripts were

high on day 0, remained high until day 4 and then declined on day 5 when vitellogenesis slows.

Addition of JH precursors stimulates JH biosynthesis in CA with low JH biosynthetic activity in

vitro.

Previous studies suggested that JH synthesis is controlled by the rate of flux of isoprenoids in A.

aegypti (Nouzova et al., 2011). To determine the role of other JH precursors in regulating JH

biosynthesis, we tested the rate of JH biosynthesis with the addition of JH precursors in the early

steps of mevalonate pathway or the addition of farnesol. The addition of acetyl CoA, DPPM or

farnesol to the incubation medium had a significant stimulatory effect on JH biosynthesis,

84

Figure 2.3. Developmental expression of genes encoding JH biosynthetic enzymes during the first gonadotrophic cycle of D.

punctata. Measurements were taken every day during the cycle (day 0 to day 7 after the final molt). Bars represent the mean of three

biologically independent pools of ten animals run in triplicate and normalized to Armadillo and EF1α. Vertical error bars indicate

SEM. Inset at the right bottom shows JH biosynthesis per individual CA (n=12) and the transcript level of DpVg in fat body during the

first gonadotrophic cycle (Marchal et al., 2014). Vertical error bars indicate SEM.

85

whereas MA did not (Fig. 2.4A). For the first time, we demonstrated that acetyl CoA, as the first

precursor in the JH biosynthetic pathway, was able to stimulate JH biosynthesis. The rank order

of the stimulatory effects of the different JH precursors on JH biosynthesis is as follows:

farnesol > acetyl CoA > DMMP > MA.

We also evaluated the sensitivity of CA to farnesol during the first gonadotrophic cycle. CA

were dissected from day 0, 3, 4, 5, and 7, and incubated with medium containing 40 µM farnesol.

On days 3 and 4, when the CA show high JH biosynthetic activity, the addition of farnesol had

no effect on in vitro JH biosynthesis (Fig. 2.4B). On the other hand, JH biosynthesis in CA of

day 0, 5 and 7 mated females was significantly increased.

Injection of HMGR-JHAMT dsRNA resulted in a significant downregulation of the target genes

but also of other genes in the JH biosynthetic pathway.

A systemic RNAi response was observed following injection of HMGR- and JHAMT dsRNA

into animals every other day during the first gonadotropic cycle. Relative transcript levels were

measured in the CA of day 4 animals using q-RT-PCR. A significant knockdown of 64% and 94%

was measured for HMGR and JHAMT, respectively. Off-target effects were investigated by

checking the Ct value of housekeeping genes . Moreover, two different dsRNA constructs were

used to eliminate off-target effects (primers listed in Table 2). Both constructs yielded a similar

phenotype.

q-RT-PCR was used to determine the relative mRNA levels of the other genes involved in the JH

biosynthetic pathway using the control and treated CA described above. A significant reduction

in relative expression levels was observed for all genes encoding the enzymes directly involved

in the pathway, with the exception of the genes encoding FOLD and the epoxidase, CYP15A1

(Fig. 2.5).

86

Figure 2.4. The effect of JH precursors on JH biosynthesis by CA from mated female D.

punctata. JH biosynthesis was determined in CA that were first incubated in medium TC199

(control), and then in medium with JH precursor (treatment). (A) JH precursors stimulate JH

biosynthesis by CA from day 7 mated female cockroach, D. punctata. 100µM of JH precursor

was added to the medium during the second incubation. (B) The sensitivity of CA to JH

precursors during the first gonadotrophic cycle. 40µM farnesol was added during the second

incubation. Values represent mean ± SEM (n≥10). Significant differences are indicated ***

p <

0.001.

Figure 2.5. Efficiency of HMGR-JHAMT RNAi-mediated knockdown and the effect of

silencing on the transcription of the other genes encoding enzymes in the JH biosynthetic

pathway in day 4 mated female D. punctata. The data represent averages of 3 pools (7 pairs of

CA per pool), run in triplicate using q-RT-PCR and normalised to Tubulin and EF1a transcript

levels. Values represent mean ± SEM. Significant differences are indicated by asterisks (*p <

0.05, **

P < 0.005, ***

p < 0.001, ****

p < 0.0001).

87

Silencing HMGR and JHAMT resulted in a significant reduction in rates of JH biosynthesis and

slows ovarian development.

The downregulation of HMGR and JHAMT resulted in a 73% reduction in the rates of JH

biosynthesis in the CA of day 4 adult females as measured using the in vitro RCA (Fig. 2.6A).

To confirm the role of JH in inducing Vg transcription in the fat body and uptake in the

developing basal oocytes, fat body and ovaries were dissected from four day old control and

treated adult cockroaches. The basal oocyte length was significantly decreased following

silencing of HMGR and JHAMT. The average oocyte length in the HMGR-JHAMT RNAi

animals measured 0.8 mm compared to the control of 1.21 mm (Fig. 2.6B). The transcript level

of the Vg in the fat body of the HMGR-JHAMT dsRNA-treated animals was reduced 80% (Fig.

2.6C). In addition, the knockdown of HMGR-JHAMT mRNA resulted in a significant reduction

in vitellin content compared to the controls (Fig. 2.6D).

We also studied the histology of the developing basal oocytes in control and treated females.

Control oocytes were fully patent on day 4, showing large intercellular spaces between the

follicle cells and the clear presence of yolk spheres. In the HMGR-JHAMT RNAi animals, on the

other hand, patency was not fully induced and as a result, yolk was not deposited in these

oocytes (Fig. 2.7).

88

Figure 2.6. JH regulates ovarian development: (A) The application of HMGR-JHAMT dsRNA

results in a dramatic decrease of JH biosynthesis (n=16). (B) Oocyte length in HMGR, JHAMT

dsRNA-treated animals compared to control animals (n≥27). (C) Relative Vg mRNA levels in

HMGR-JHAMT dsRNA-treated animals compared to control animals. The transcription data

represent averages of 3 pools (5 ovaries per pool), run in triplicate using q-RT-PCR and

normalised to Tubulin and EF1a transcript levels. (D) Vitellin content measured by ELISA in

HMGR-JHAMT dsRNA-treated animals compared to controls (n=9). Animals were dissected on

day 4 after the final molt in control and treated groups. Values represent mean ± SEM.

Significant differences are indicated by asterisks (**

P < 0.005, ***

p < 0.001, ****

p < 0.0001).

89

Figure 2.7. Transverse sections of the basal oocytes from day 4 control and HMGR-JHAMT dsRNA-treated animals. Follicle

cells (Fc), intercellular gaps (arrows), yolk spheres (arrowheads) and lipid spheres (L) are indicated. Scale bars represent 200 µm.

90

Discussion

Sequence of genes encoding enzymes in the JH biosynthetic pathway - Through in silico data

mining and degenerate primer PCR and RACE, we have successfully characterized 11 of the 13

enzyme-encoding genes in the JH biosynthetic pathway of D. punctata. Cheng et al (2014)

showed that enzymes in the early steps of the mevalonate pathway were generated by single

copy genes in many insects, whereas genes encoding FPPS and enzymes in the JH-specific

pathway probably underwent gene duplication in Lepidoptera. In D. punctata, only a single copy

of these genes was identified. However, at this point, the existence of multiple copies of several

of these genes cannot be ruled out and this issue will have to await the assembly of the

Diploptera genome.

The full-length HMGR and JHAMT were cloned from D. punctata CA. HMGR encodes a protein

of 825 amino acids in length. Following a conserved domain search, HMGR was observed to

contain the typical conserved motifs found in members of the HMG-CoA reductase superfamily

of proteins. JHAMT encodes a protein of 274 amino acids in length and contains the motifs

typically found in AdoMet-dependent methyltransferases. Still missing are sequences for

farnesyl diphosphate pyrophosphatase (FPPP) and farnesal dehydrogenase (FALD). Molecular

analysis of these genes was long hampered as a consequence of the minute size of the CA but

using a newly developed sensitive assay for measuring JH precursor pools employing fluorescent

tags, the genes encoding FPPP and FALD were recently characterized in A. aegypti (Nyati et al.,

2013; Rivera-Perez et al., 2013; Rivera-Perez et al., 2012). These were, however, not included in

our study because the availability of only dipteran sequences made their characterization in the

phylogenetic basal D. punctata problematic and therefore remain uncharacterized in this species

to date.

91

Transcription of enzymes in the JH biosynthetic pathway partly regulates JH biosynthesis - The

11 orthologs of JH biosynthesis genes cloned in our model D. punctata are predominately

expressed in the CA, and their relative transcript levels correlate well with the rates of JH

biosynthesis (Fig. 2.2 and 2.3). Our results suggest that CA are highly specialized tissues for the

biosynthesis of JH, and genes identified in our study are encoding functional enzymes in the JH

pathway. Enzymes in the mevalonate pathway are responsible for not only JH production but

also for the production of other terpenoids such as defensive secretions and pheromones (as

reviewed by (Belles et al., 2005; Goldstein and Brown, 1990)). It is therefore of interest that all

mevalonate enzyme-encoding genes except PMK and IPPI are exclusively expressed in the CA

with trace amounts in other tissues. This can be explained by the strikingly high rates of JH

biosynthesis in the CA of day 4 adult female D. punctata. For the later enzymes involved in the

JH specific steps of the pathway, the transcription of JHAMT and CYP15A1 is CA-specific,

whereas FOLD is expressed in many tissues. Similar results were observed in A. aegypti and A.

mellifera (Bomtorin et al., 2014; Mayoral et al., 2009a; Nouzova et al., 2011). Because of the

multiple functions of farnesol, the oxidation of farnesol to farnesal does not appear to be a JH

biosynthesis-specific reaction (Mayoral et al., 2009a).

Changes in rates of JH biosynthesis during the first gonadotropic cycle of D. punctata are

dynamic, corresponding to specific reproductive events (see below) and are therefore tightly

regulated. The transcript levels of enzyme-encoding genes in JH biosynthesis are highly

coordinated with the rates of JH biosynthesis (Fig. 2.3), suggesting that at least part of the

regulation of JH biosynthesis involves coordinated changes in the transcription of the genes in

the biosynthetic pathway, in agreement with studies performed in A. aegypti (Nouzova et al.,

2011) and B. mori (Kinjoh et al., 2007), although the ability of each gene to regulate JH

92

biosynthesis appears to differ. In addition, the expression patterns of HMGS and HMGR mRNA

were similar to the patterns of their enzyme activities (Couillaud and Feyereisen, 1991;

Feyereisen and Farnsworth, 1987). Although the transcription of most genes in the biosynthetic

pathway showed a correlation with the JH biosynthesis, there are a few exceptions, including

PMK, IPPI and FPPS. PMK and IPPI, which are expressed in multiple tissues (Fig. 2.2), and

may not be specifically regulated in the CA. For FPPS, 3 FPPS homologs were found in B. mori

(Ueda et al., 2009), and 7 in A. mellifera (Bomtorin et al., 2014). Although we currently have

identified only 1 FPPS gene in D. punctata, additional FPPS genes may exist in the CA that

regulate JH biosynthesis.

Effect of exogenous JH precursors on rates of JH biosynthesis - We have found that JH

precursors are able to stimulate JH biosynthesis (Fig. 2.4A), which suggests that the rate of JH

biosynthesis is not only controlled by transcription of the genes in the pathway, but also by the

flux of substrates in the pathway. Farnesol stimulates JH biosynthesis by CA from day 0, 5 and 7

mated females. However, application of farnesol to CA showing high JH biosynthetic activity

(day 3 and 4) does not result in a significant change in JH biosynthesis (see Fig. 2.4B). These

results suggest that during the first gonadotrophic cycle of D. punctata, the supply of this

precursor is rate-limiting in CA with showing low JH biosynthetic activity, whereas there are

other factors, including nutrients and neurotransmitters that control JH production at high rates of

JH biosynthesis. Nevertheless, 14 JH precursors are involved in the JH biosynthetic pathway,

and each JH precursor connects to the ‘upstream’ and ‘downstream’ enzymes. Many factors

could affect the stimulatory effect of JH precursor in JH production, including cell permeability

of the added compounds, the activity of upstream/downstream enzymes, and the size of other JH

precursor pools.

93

A feedback mechanism involved in the regulation of the transcription of JH biosynthetic genes -

HMGR and JHAMT have been studied in the JH pathway since they are representative enzymes

involved in the well-conserved mevalonate pathway and in the JH-specific portion of the

pathway respectively (Debernard et al., 1994; Li et al., 2013; Marchal et al., 2011;

Martinezgonzalez et al., 1993; Mayoral et al., 2009c; Monger et al., 1982; Wang et al., 2013).

We therefore chose to silence HMGR and JHAMT on a posttranscriptional level using RNAi

during the first gonadotropic cycle of D. punctata. RNAi trials have shown that fewer injections

of HMGR or JHAMT dsRNA and with lower concentrations of construct, did not result in a

significant silencing of our target gene mRNA levels (data not shown). We therefore conclude

that the RNAi response in D. punctata appears to be systemic, but is not as effective as in

another hemimetabolous insect model, the desert locust, Schistocerca gregaria in which smaller

quantities of the JHAMT dsRNA construct induced a longer downregulation of the same target

gene during the first gonadotropic cycle (Marchal et al., 2011). A further in-depth study on the

RNAi machinery D. punctata should be performed to explain the differences in sensitivity

compared to other hemimetabolous insect species.

JH biosynthesis is controlled by many factors including neuropeptides (allatostatins, allatotropins,

short neuropeptide F) (Kaneko and Hiruma, 2014; Weaver and Audsley, 2009),

neurotransmitters (octopamine, dopamine and glutamate) (Granger et al., 1996; Pszczolkowski et

al., 1999; Thompson et al., 1990) and JH itself (Goodman and Cusson, 2012; Marchal et al.,

2013a). Topical application of JH or a JH analog resulted in a suppression of JH biosynthesis,

which suggests a negative feedback regulation of CA in D. punctata (Tobe and Stay, 1979). In

higher animals, cholesterol, the bulk end-product of mevalonate pathway, shows a feedback

regulation on the transcription and post-transcription of enzymes in the mevalonate pathway

94

(Chang and Limanek, 1980; Clarke et al., 1987; Goldstein and Brown, 1990). Later studies

showed that farnesol, an intermediate product in the JH biosynthetic pathway, is able to

accelerate the degradation of HMGR, and the increase of mevalonate flux raises the activity of

FPPP expressed in CHO cells (Meigs et al., 1996; Meigs and Simoni, 1997). In our study,

regulating transcription of the JH biosynthetic enzymes seems to affect the entire biosynthetic

pathway, rather than individual steps. The silencing of HMGR and JHAMT resulted in a

significant decrease in the transcript level of several other genes in the pathway (Fig. 2.5). A

possible explanation is that the accumulation of JH precursors as a result of RNAi treatment

resulted in a feedback on the expression of other genes in the JH biosynthetic pathway to balance

the size of JH precursor pools and the enzyme activity.

The essential role of JH in reproduction - Our results clearly show that select silencing of genes

encoding enzymes in the JH biosynthetic pathway effectively reduces JH biosynthesis in vitro

(Fig. 2.6). As a well-known model for the study of the regulation of JH biosynthesis, D. punctata

displays a consistent characteristic profile of JH production during the first gonadotropic

cyclethat is closely correlated with specific reproductive events such as vitellogenesis and

chorionation. Once JH production rises on day 2, vitellogenin synthesis in the fat body

commences, and vitellogenin is taken up from the hemolymph and is incorporated into the

developing basal oocytes. The elevated JH titer results in patency in the follicular epithelium of

the basal oocytes, permitting the uptake of Vg and subsequent yolk formation (Marchal et al.,

2014; Pratt and Davey, 1972). On day 5, the spaces in the follicular epithelium close as rates of

JH biosynthesis decline and Vg transcription in the fat body and circulating Vg levels decrease.

At this point, choriogenesis begins and JH biosynthesis remains low until oviposition on day 7

(as reviewed by Marchal et al., 2013a). Our RNAi study has focused on day 4 of the first

95

gonadotropic cycle, a time when JH titre is high in controls and animals are vitellogenic (Tobe et

al., 1985) as characterized by high DippuVg mRNA transcript levels in the fat body and JH-

dependent induction of patency in the maturing basal oocytes. Following the reduction in JH

production by the silencing of DippuHMGR and DippuJHAMT, mRNA levels of DippuVg in the

fat body are significantly reduced, incorporation of vitellin into the oocytes is impaired and

patency does not occur (Fig. 2.6 and 2.7). Previous reports have shown that in S. gregaria,

oocyte length is significantly affected following treatment of females with JHAMT dsRNA

(Marchal et al., 2011). RNAi was also used in the cotton bollworm Helicoverpa armigera to

effectively silence HMGR, resulting in reduced Vg expression (Wang et al., 2013). Each of these

studies highlights specific aspects of JH-regulated reproductive events. The results described in

the current paper confirm the central role played by JH during the female reproductive cycle of D.

punctata.

JH biosynthesis is a fundamental process in regulating insect development, metamorphosis and

reproduction (Goodman and Cusson, 2012). Although the genes directly involved in JH

biosynthesis have been characterized using molecular techniques in several insect species, there was

no similar study in the hemimetabolous insects. Thus, our study is the first to characterize the

majority of the genes directly involved in JH biosynthesis in a hemimetabolous insect and a well-

known model for studying the physiology of JH. This paper now provides the molecular tools to

study the regulatory mechanisms of JH production in D. punctata (for a first study, see (Huang et

al., 2014)). In addition, the role of JH biosynthetic enzymes and the JH precursor supplies in

regulating JH biosynthesis have been demonstrated, as has the critical function of JH as the

master regulator of cockroach female reproduction.

96

Supplementary data

A.

BlageHMGR 1 M-VGRLFRAHGQFCASHPWEVIVATLTLTVCMLTVDQ-------------RP-LGLPPGWGHNC---------------------------------ITLEEYNAADMIVMTLIRCVAVLYSYYQFCHLQKLGSKYILGI

DippuHMGR 1 M-VGWLFLAHGQFCASHPWEVIVATFTLTACMLTVDP-------------HP-LGLPPGWRKNC---------------------------------ISLEEYNAADVIVMTLIRCVAVLYSYYQFCHLQKLGSKYILGI

ApimeHMGR 1 M-LTRLFEIHGRFCAGHPLEVIVTTFTLTACILNMETG----------NGHPSVPLTAHCGPGR---------------------------------CNSDDLNAADVIVMTIIRCLAILFTYHQFRNLQKLGSKYILGI

TricaHMGR 1 M-TTRLFRAHGEFCATHPWEVIVATLTLAACMLTVDQQ----------HPAPPPKPTLRYCAEC---------------------------------LQEAEYNAADLIVMTLIRCLAVLYCYYQFCNLEKLGSKYILGI

AedaeHMGR 1 MKVGRLFRAHGEFCASHPWEVIVALVTLTACIITFDKG----------SADPFQQSSSRSGSGRPC----------------PSWRFSSVSPDEAFRCEEVEQNGIDVILMTIVRCSAILYCYYHFCNLQKLGSKYILGI

DromeHMGR 1 M-IGRLFRAHGEFCASHPWEVIVALLTITACMLTVDKNNTLDASSGLGTATASAAAAGGSGSGAGSGASGTIPPSSMGGSATSSRHRPCHGWSQSCDGLEAEYNAADVILMTIVRCTAVLYCYYQFCSLHRLGSKYVLGI

BommoHMGR 1 M---KVWGAHGEFCARHQWEVIVATLALLACAASVERHG----------PGNRSEHCAGWARACP--------------------------------GLEAEYQAADAVIMTFVRCAALLYAYYQVLNLHKIASKYLLII

BlageHMGR 93 AGLFTVFSSFVFSSSVINFLGSDVSDLKDALFFFLLLIDLSKATVLAQFALSSRSQDEVKHNIARGIAMLGPTITLDTVVETLVIGVGMLSGVRRLEVLCCFACMSVIVNYVVFMTFYPACLSLILELSRSGESGRPAWH

DippuHMGR 93 AGLFTVFSSFVFSSSVINFLGSDISDLKDALFFFLLLIDLSKATLLAQFALSSCSQDEVKHNIARGIAMLGPTITLDTVVETLVIGVGTLSGVRRLEVLCCFACMSVIVNYIVFMTFYPAFLSLILELARSGEGGRPAWH

ApimeHMGR 97 AGLFTVFSSVVFTSSVVNFGRSDISDLKDALFFFLLIIDLSKAAVLAQLALSSRNKEEVRANIARGMSLLGPTITLDTLVETLLISIGSLSGVRRLEILCSFACLGVVVNYIVFMTFYPACLSLILELSRETNTMKPLSA

TricaHMGR 97 AGLFTVFSSFVFTSTVLNLLWIDVSDLKDALFFFLLLIDLSKAAMLAQSALSASNQEEVKSNIARGMAVLGPTITLDTIVETLVIGVGTLSGVHRLEMLSYFACLSVLVNYIVFMTFYPACLSLILELSRTTN---IYGN

AedaeHMGR 115 AGLFTVFSSFIFTSTVINFLGSEVSDLKDALFFFLLLIDLSKAAILAQLALCGSSQSEVTMNIARGMEILGPAISLDTLVETLVIGVGTLSGVQRLELLSGFAVLSAIVNYIVFMTFYPACLSLILDLSRNAGNLIQKNK

DromeHMGR 140 AGLFTVFSSFIFTTAIIKFLGSDISELKDALFFLLLVIDLSNSGRLAQLALSGSNQAEVTQNIARGLELLGPAISLDTIVEVLLVGVGTLSGVQRLEVLCMFAVLSVLVNYVVFMTFYPACLSLIFDLSRSGVDMSVVRE

BommoHMGR 96 AGVFSTFASFIFTSALASLFWSELASIKDAPFLFLLVADVARGARMAKAGWSAG--EDQGKGVGRALSLLGPTATLDTLLAILLVGVGALSGVPRLEHMCTFACLALLVDYIVFITFYPACLSLVADFATNRK---EIAH

BlageHMGR 233 DKS--LIIKALHEEDQKPNPVVQRVKVIMSAGLMLVHAHRWVRCLSIALWPDLTSLRYFCTHCDTGVSYSRWSFASEGEELPTVKLVTG--DSVVNSNST----DDAQLHYYIMRWLTVSADHIVILILLLALAVKFVFF

DippuHMGR 233 DRS--LIMKALHEEDQKPNPVVQRVKVIMSAGLMLVHAH------------------------------SRWSFASEGEELPIVNLVTG--DSVVITNST----QDAQIHDYIMRWLTVSADHIVILILLLALAVKFVFF

ApimeHMGR 237 DKI--FMMHPLDEEDQKPNPVVERVKLIMIAGLFVVHAN------------------------------SRFKSE-ESEDTVEGKVSST--NSHVIVNSYNETEDSSEVKGYLMNWLSVSADNIVILILLLALAIKFIFF

TricaHMGR 234 KQS--LIARALKEEDHKSNPVVQRVKLIMSAGLMLVHAR------------------------------SRWPFK--EDDVENIRPLVV--EQHMTLNRT----EDTTLHEYIMKWVTVSADHIVILILLLALVVKFVFF

AedaeHMGR 255 KEN--LLARVLTEEDQKPNPVVQRVKLIMSSGLMIVHVL------------------------------SRLAISEKDSDTAEHIIASHSHEHLAAMNKT----EPNEISEFIMRWLSISTEQIVTYILIIALGVKFVFF

DromeHMGR 280 KAKGSLLLKSLTEEEQKANPVLQRVKLIMTTGLMAVHIY------------------------------SRVAFSGSDYDAVDKTLTPT-LSLNVSNNRT----ESGEIADIIIKWLTMSADHIVISIVLIALVVKFICF

BommoHMGR 231 DSP-------FSEEDLKPNPVVQRVKMIMAAGLLCVHLT------------------------------SRWPWS------ANHGIIEG--PIDASIPVP----HDNILLHSYVKWFSVSADYIVIATLLCALIIKFIFF

BlageHMGR 365 ETRDELTTTRGMDGWVEVSSP------VEHKYVQTE----------------------------------QPSCSAP------EQPLEEPPASN------------RSIDECLSVCK-SDVGAQ--ALSDCEVMALVTS-

DippuHMGR 335 EVKDELNTTRGMDGWVEVSSP------VEHKFVQTD----------------------------------QVCWELPSSS---QEPDEQPLPCD------------RPVQECLAICK-SDAGAN--SLSDSEVMSLVSS-

ApimeHMGR 342 EDKEDIAKQLQFKVEDDTEEK------IENEYMKEKKFEIEYVKDKENENEDMNFSFLTTKMPFKLSFAKMQTISIPWIDGKEEQIDEKQCTVNTSNDKNLFSQIPRSVEECLKIYK-SELGAN--GLTDEEVIQLVKN-

TricaHMGR 334 ENKEELAEQLRAHISTESVDPGKKDNRFKMPLIKTQSFFLTNNTKED---------------------SACEDKEVQTDIGRLESEFEKLPAGN-----DVSVKTCRSLEECLKIYNDSNLGGA--ALSDDEVILLVKN-

AedaeHMGR 359 D-RHNLSDQILLSVANEAAAAAAAAAAAAQKQKQLELELERPQPVATFT--------LAETVPVEEKATQSELCLLPGRGQRAKSMGADDELNLDDLEDELVEREPRPLEMCLKILNETDDGAV--GLTDEEIKMIVRAG

DromeHMGR 385 DNRDPLPDQLRQSGPVAIAAKASQTTPIDEEHVEQEKDTENS--------------------------AAVRTLLFTIEDQSSANASTQTDLLPLRHRLVGPIKPPRPVQECLDILNSTEEGSGPAALSDEEIVSIVHAG

BommoHMGR 322 EEQRNWVYDMDDMTVKEVINDTDLS--RKPKFSVGD---------------------------------DSNSEVSTQTDEAGNVEDMEWPTLSPSSSASKLNAKKRPMVECLELYR---SGACT-SLSDEEVIMLVEQ-

BlageHMGR 443 -GHIAGYQLEKVVRNPERGVGIRRQILTKTADL-KDALDNLPYKNYDYLKVMGACCENVIGYMPVPVGVAGPLNLDGRLVHVPLATTEGCLVASTNRGMRALMRCGVTSRIVADGMTRGPVVRFPNIDRASEAMLWMQVP

DippuHMGR 416 -GHVAAYQLEKMVGDPERGVGIRRKILTQKADL-KDALDNLPYKNYDYTKVMGACCENVIGYMPVPVGVAGPLKLDGCLVYVPLATTEGCLVASTNRGSRALMRCGVTSRIVADGMTRGPVVRFPNIERASEAMLWMQAA

ApimeHMGR 472 -NHIAAYQLEKAVGDMERGVEIRRFIIGEAGNF-LDYLSNLPYKDYDYSKVLGACCENVMGYVPVPLGIAGPLLLDGELYYVPMATTEGCLVASTNRGSRALLKCGVTSRVVADGMTRGPVVRFPNIVRASEAMAWMQDP

TricaHMGR 445 -KHIPAYQIEKAVDDPERGVGIRRKILAREGNF-SEALTDLPFRNYDYAKVMGACCENVIGYMPVPVGYAGPLNLDGRHVYVPMATTEGCLVASTNRGCRALLDCGVTSRVVSDGMTRGPVVRFPSITKASEAMSWMKCS

AedaeHMGR 488 NGYCPLYKIETVIGDAERGVKIRRDMIQKEANLPANAFKHLSYKNYDYSKVMNACCENVLGYVQIPVGYAGPLILDGVRYYVPMATTEGALVASTNRGCKAISTRGVTSFVEDIGMTRAPCIKFPNVLRAAQAKRWMETP

DromeHMGR 499 GTHCPLHKIESVLDDPERGVRIRRQIIGSRAKMPVGRLDVLPYEHFDYRKVLNACCENVLGYVPIPVGYAGPLLLDGETYYVPMATTEGALVASTNRGCKALSVRGVRSVVEDVGMTRAPCVRFPSVARAAEAKSWIEND

BommoHMGR 422 -SHIPMHRLEAVLEDPLRGVRLRRRVIASRFNN-ETAIKQLPYLNYDYSKVLNACCENVIGYIGVPVGYAGPLVVDGKPYMIPMATTEGALVASTNRGAKAIGSRGVTSVVEDVGMTRAPAVKLPNVVRAHECRQWIDNK

BlageHMGR 581 YNFEQIKKNFDSTSRFARLSKIHIRVAGRHLFIRFIATTGDAMGMNMLSKGTEVALAYVQQVYPDMEILSLSGNFCTDKKPAAVNWIEGRGKSVVCEAIVPADIIKSVLKTSVQALMDVNITKNLIGSAVAGSIGGFNAH

DippuHMGR 554 QNFEAMKKHFDSTSRYARLSKIHIRVAGRHLFIRFVATTGDAMGMNMLSKGTEVALSFVQQMFPDMEILSLSGNFCTDKKPAAVNWIEGRGKSVVCEAVVPADVVSSVLKTSVQALVDLNITKNFYGSAIAGSVGGNNAH

ApimeHMGR 610 DNFKEMKNSFNLTSRFARLTKINIRIAGRHLFIRFVATTGDAMGMNMLSKGTEKSLNTVKEHFPDMEILSLSGNFCTDKKPAAVNWVCGRGKSVVCEAVVPADIVTNVLKTSVHALVDVNISKNMIGSAIAGSIGGFNAH

TricaHMGR 583 QNFEAMKQQFDSSSRFARLSKLLIKIAGRHLFVRFEAKTGDAMGMNMVSKGTEMSLKYVQKQFPEMEILSLSGNFCTDKKPAAVNWIEGRGKSVVCEAIIPSEIVKKVLKTSTPALVDVNNSKNMIGSAVAGSIGGFNAH

AedaeHMGR 628 ENFAVIKKAFDSTSRFARLQELHIAMDGPILYARFRALTGDAMGMNMVSKGSEMALREVHRSFPDMQIISLSGNFCSDKKPAAINWIKGRGKRVICEAIVPADKLRTILKTNARTLVQCNKLKNMTGSAVAGSIGGNNAH

DromeHMGR 639 ENYRVVKTEFDSTSRFGRLKDCHIAMDGPQLYIRFVAITGDAMGMNMVSKGAEMALRRIQLQFPDMQIISLSGNFCCDKKPAAINWIKGRGKRVVTECTISAATLRSVLKTDAKTLVECNKLKNMGGSAMAGSIGGNNAH

BommoHMGR 560 ENYALLKEAFDSTSRFARLQEIHVGVDGATLYLRFRATTGDAMGMNMVSKGAENALKLLKTFFRDMEVISLSGNYCSDKKAAAINWIKGRGKRVVCETVISSENLRTIFKTDAKTLSRCNKIKNLSGSALAGSIGGNNAH

BlageHMGR 721 AANIVTAIFIATGQDPAQNVGSSNCMTLMEPWGEDGKDLYVSCTMPSIEIGTIGGGTVLPPQAACLDMLGVRGANEMCPGENANTLARIVCGTVLAGELSLMSALAAGHLVKSHMRHNRSSVSTS----------GSEPS

DippuHMGR 694 AANIVTAIFIATGQDPAQNVGSSNCITIMEPWGDDKKDLYVSCSMPSIEIGTVGGGTILPPQAACLDMLGVKGANAVCPGENANMLARIVCGTVLAGELSLMSALAAGHLVKSHLRHNRSSVTTS----------GSEPS

ApimeHMGR 750 AANIVTAIFIATGQDPAQNVGSSNCMTLMEPWGTDGSDLYVSCTMPSIEIGTVGGGTILPAQGACLSILGVKGAHSDEPGENASRLARIVCATVLAGELSLMAALTAGHLVKSHLRHNRSSTTVTNAMSVPQKYTGTKLS

TricaHMGR 723 AANIVTAIYIATGQDPAQNIGSSNCMTLMEPWGETGEDLYVSCTMPSIEIGTVGGGTVLPAQSSCLEMLRVKGSHPDCPGENASQLARIVCGTVLAGELSLMAALTAGHLVRSHLRHNRSTTTIP-----------EEFS

AedaeHMGR 768 AANMVTAIYIATGQDPAQNVTSSNCSTNMEPYGDNGEDLYMTCTMPSLEVGTVGGGTGLPGQGACLDMLGVRGAHPTHPAENSKQLARVICATVMAGELSLMAALVNSDLVKSHMRHNRSSVAVNPG----LPATAQLQS

DromeHMGR 779 AANMVTAVFLATGQDPAQNVTSSNCSTAMECWAENSEDLYMTCTMPSLEVGTVGGGTGLPGQSACLEMLGVRGAHATRPGDNAKKLAQIVCATVMAGELSLMAALVNSDLVKSHMRHNRSSIAVN--------------S

BommoHMGR 700 AANMVTAIFIATGQDPAQNVTSSNCSTNMEAYGENGEDLYVTCTMPSLEVGTVGGGTVLTGQGACLEILGVKGAG-TRPAENSARLASLICATVLAGELSLMAALVNSDLVKSHMRHNRSTLNVQ------------TAN

BlageHMGR 851 TP--------ACKS-------------

DippuHMGR 824 KS-------------------------

97

ApimeHMGR 890 VPNLLQPVQNVCKGLLEKS--------

TricaHMGR 852 QNRYHIP---PCKDI------------

AedaeHMGR 904 APSLLTACNSSSSGSSSSIGTSLSAKQ

DromeHMGR 905 ANNPLNVTVSSCSTIS-----------

BommoHMGR 827 VEPYTVALKVPPS--------------

B.

SchgrJHAMT 1 MDKAELYSRSNGLQRWEASAALEAAWPALRWPAPP-LRVLDVGCGAGDVTVDLLLPRLPP-HTQLVGTDVSAAMVEHAAELYGAAHPGLSFQLLDIADPDIDASPVYQLAPFDKIFSFFCLHWVPEQRQAAENLHRLLKP

DippuJHAMT 1 MHKAELYSSSHGLQKRDAAHALTEYLDHMTWR-PG-DRVLDVGCGPGFVTAQELMPRLPQDFAILVGTDVSHAMVQHATSTY--VQPKLKFAHLDISSTHIDK-ELWEPG-FDKIFSFYCLHWIPDQRTAVNNIYHLLRP

TricaJHAMT 1 MNKASLYSKYSGLQKNDASFVIDNYLRLIKWK-PN-ANILDIGSGDGNVIFELLLPKIPKHFAKFVGTDISEEMVLFAKNQCN--DPKIDFLQMDIS----ATIPPEFHEYFDHIFSFYCLHWVVEQRQAMKNIFDMLKP

DromeJHAMT 1 MNQASLYQHANQVQRHDAKLILDEFASTMQWRSDGEDALLDVGSGSGNVLMDFVKPLLP-IRGQLVGTDISSQMVHYASKHYQR-EERTRFQVLDIGCER-LPEELS--GRFDHVTSFYCLHWVQNLKGALGNIYNLLKP

AedaeJHAMT 1 MNKPNLYHRANGVQRRDAKEILDEHGHLLRWKEENEDSLLDIGCGSGDVLIDFVIPMVPPKRARVLGTDVSEQMVRFARKVHSD-VENLFFETLDIEGD--ISSFLNKWGCFDHITSFYCLHWVRSQRSAFSNIYNLMAP

BommoJHAMT 1 MNNADLYRKSNSLQKRDALRCLEEHANKIKWKKIG-DRVIDLGCADG-SVTDILKVYMPKNYGRLVGCDISEEMVKYANKHHG--FGRTSFRVLDIEGD--LTADLK--QGFDHVFSFYTLHWIRDQERAFRNIFNLLGD

ApimeJHAMT 1 MFLVEEYVKASTIQYRDAADIIGEFAEEMSEMKGK---CLDIGCGPGIVTKELILPNLSP-EAKLVGMDISRPMIEYAKNMYHD-EERLSFQLLDIET---MDLPKDTFDQFNNVLSFYCLHWCQNFRKAFDNIYKLLRP

SchgrJHAMT 139 GG-EVVLSLLAHCPIFSVYEGLAHKPQWKEYMEDARRFISPYHHSEDPAREMNELLCRAGFRVTLCTRQQRSFTFPGHSALIEAVTAVNPFVERLPETLQQEFLEDCMKEVLRQKLVSIEDDADSNNNSNGSNRSGNNAV

DippuJHAMT 135 AG-EALVLLMAKCPVFNVYTAQSNKPKWQQYMKDASRYISPYHQLKDPKSEFINIVEDVGFHVVDCDCRQNKFNYGTLERLKDAIKAVNPFMDRIPEELQEEYLNDCLSEARRIKCT------ESNNN-----------V

TricaJHAMT 133 GG-EMLLTFLASNPIYDIYERMAKSNKWGPYMNNLKKYISPYHHSEDPETELENLLKKEGFITHLCRVENRSYTFPSFSVLSKSVSAVNPFIKKLPENEIDTYIEDYLKEVRKLKTIT----IETCNN------------

DromeJHAMT 136 EGGDCLLAFLASNPVYEVYKILKTNDKWSTFMQDVENFISPLHYSLSPGEEFSQLLNDVGFVQHNVEIRNEVFVYEGVRTLKDNVKAICPFLERMPADLHEQFLDDFIDIVISMNLQQ-------GEN------------

AedaeJHAMT 138 NG-DCLLGFLARNPIFDIYDQLSNSAKWSMYMTDVDKYISPYQYCENPVGEIEEILSSVGFTKYKIHIADKIYVYEGIDSLKKAVQAVNPFSERMPLDLQEDFLNDYIAVVRRMSLSEN-----CCGN------------

BommoJHAMT 133 EG-DCLLLFLGHTPIFDVYRTLSHTEKWHSWLEHVDRFISPYHDNEDPEKEVKKIMERVGFSNIEVQCKTLFYVYDDLDVLKKSVAAINPFN--IPKDILEDFLEDYIDVVREMRLLDR-----CNNN------------

ApimeJHAMT 133 GG-KGLFMLLSWNDGFDVYKKLYANPRYRPYMQEPERFIPIFHECKDRRVNLRKILETTGFEILHCSEREKSYIYKNSEIMKKHIMAINPFISRIPNSLKKEFEDEITREIVNMKIQLLN----KDEN------------

SchgrJHAMT 278 DANSKKLTTRYSVLTVVAAKAAAQNGVRTVR-----------

DippuJHAMT 257 EEVT---TVSYDIIVAHIKKP---------------------

TricaJHAMT 256 NDNEEKIHVPYKLFVTFASKPV--------------------

DromeJHAMT 257 NEDQKFLSP-YKLVVAYARKTPEFVNNVFLEPTHQNLVKGIN

AedaeJHAMT 260 ENDYKFITP-YKLVVVYAVK----------------------

BommoJHAMT 253 VGESVSIKFNYKVISVYARKLCLSLM----------------

ApimeJHAMT 256 GEQEYNILDRYQIFVTYIRKPVC-------------------

Figure S2.1. Amino acid sequence alignment of Diploptera proteins with several functionally characterized ortholog insect

proteins. (A) DippuHMGR with HMGR from the German cockroach Blattella germanica (GenBank accession number: CAA49628.1

(Martinezgonzalez et al., 1993), the honeybee Apis mellifera (GenBank accession number: XP_623118.1), the red flour beetle

Tribolium castaneum (GenBank accession number: XP_973850.1), the mosquito Aedes aegypti (GenBank accession number:

XP_001659923.1), the fly Drosophila melanogaster (GenBank accession number: NP_732900.1) and the silkworm Bombyx mori

(GenBank accession number: BAF62108.1 (Kinjoh et al., 2007)) (B) DippuJHAMT with JHAMT from the desert locust, S. gregaria

(GenBank accession number: ADV17350.1 (Marchal et al., 2011)), the red flour beetle T. castaneum (GenBank accession number:

NP_001120783.1 (Minakuchi et al., 2008)), the fly D. melanogaster (GenBank accession number: AB113579.1 (Niwa et al., 2008)),

the mosquito A. aegypti (GenBank accesson number: ABD65474.1 (Mayoral et al., 2009b)), the silkworm B. mori (GenBank

accession number: NP_001036901.1 (Shinoda and Itoyama, 2003)) and the honeybee A. mellifiera (GenBank accession number:

AGG79412.1 (Li et al., 2013)).

98

PM

K

PP

MD

IPP

I

FO

LD

Th

iol

HM

GS

HM

GR

MK

FP

PS

JH

AM

T

CY

P15A

1

0

1 0

2 0

5 0

1 0 0

1 5 0

2 0 0

5 0 0

1 0 0 0

1 5 0 0

2 0 0 0

2 5 0 0

3 0 0 0

3 5 0 0

Re

lati

ve

mR

NA

qu

an

tity

Figure S2.2. Comparison of transcript levels of genes encoding JH biosynthetic enzymes in

CA of day 4 mated female. Bars represent the mean of three biologically independent pools of

ten animals run in triplicate and normalized to Tubulin and EF1α. Vertical error bars indicate

SEM. Enzyme abbreviations are as in Fig. 2.1.

99

Table S2.1. (Degenerate) primer sequences for cloning of (partial) sequences for orthologous genes encoding JH biosynthesis

enzymes in D. punctata. Gene name abbreviations were shown in Fig. 2.1.

Gene name F Primer(5'-3') R Primer(5'-3') Accession number (NCBI)

Thiol GGDCARAATCCWGCDAGRCARGC TTGHGCWGCRAANGCTTCRTT KJ188021

HMGS AGAGTGAACTGGAAGTGCA TATGAAGCACGCACTCCTC KJ188022

HMGR AATGGTAGGGTGGCTGTTTCTA GTATTTCATGATTTGCTGGGTTC KJ188023

MK CTGCCCCCGGTAAAGTTATC GATGCTTGACTTCCTGCACC KJ188024

PMK ATGGAAGTAAACAAAATTTCAG CACTCTGTTTCTGCATCATCTAC KJ188025

PPMD TGTTCAGAAAAYAATTTYCCNAC TGATTNSWRTCYTTCATTGT KJ188026

IPPI CAYMGDGCVTTYAGTKTDTTY TCDAWYTCRTGTTCKCCCCA KJ188027

FPPS TGGCGCGTAGGGTAGACAA GATCCATAACATTCTGCAAACATTG KJ188028

FOLD ATGCAGCGTTGGACTGGA TCCAGCATTGTTGATGAGAA KJ188029

JHAMT ATGCACAAAGCAGAACTGTATTC TCAAGGCTTTTTAATATGAGCGACAAT KJ188030

CYP15A1 ATGGTCATCGCTCTTATTGTCATC ACCATTCATTTCCTTGGAATCAACT AY509244

Vg TGGAACRCDCTICTCTGYTGYCT TTKAKSAYGTTRAYTTCCCAG KJ188031

100

References

Belles, X., Martin, D., Piulachs, M.D., 2005. The mevalonate pathway and the synthesis of juvenile hormone

in insects. Annu Rev Entomol 50, 181-199.

Bomtorin, A.D., Mackert, A., Rosa, G.C., Moda, L.M., Martins, J.R., Bitondi, M.M., Hartfelder, K., Simoes,

Z.L., 2014. Juvenile hormone biosynthesis gene expression in the corpora allata of honey bee (Apis mellifera

L.) female castes. PLoS One 9, e86923.

Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of microgram quantities of protein

utilizing the principle of protein-dye binding. Anal Biochem 72, 248-254.

Cao, L., Zhang, P., Grant, D.F., 2009. An insect farnesyl phosphatase homologous to the N-terminal domain of

soluble epoxide hydrolase. Biochem Biophys Res Commun 380, 188-192.

Chang, T.Y., Limanek, J.S., 1980. Regulation of cytosolic acetoacetyl coenzyme A thiolase, 3-hydroxy-3-

methylglutaryl coenzyme A synthase, 3-hydroxy-3-methylglutaryl coenzyme A reductase, and mevalonate

kinase by low density lipoprotein and by 25-hydroxycholesterol in Chinese hamster ovary cells. J Biol Chem

255, 7787-7795.

Cheng, D., Meng, M., Peng, J., Qian, W., Kang, L., Xia, Q., 2014. Genome-wide comparison of genes

involved in the biosynthesis, metabolism, and signaling of juvenile hormone between silkworm and other

insects. Genet Mol Biol 37, 444-459.

Clarke, C.F., Tanaka, R.D., Svenson, K., Wamsley, M., Fogelman, A.M., Edwards, P.A., 1987. Molecular

cloning and sequence of a cholesterol-repressible enzyme related to prenyltransferase in the isoprene

biosynthetic pathway. Mol Cell Biol 7, 3138-3146.

Couillaud, F., Feyereisen, R., 1991. Assay of HMG-CoA synthase in Diploptera punctata corpora allata. Insect

Biochem 21, 131-135.

Debernard, S., Rossignol, F., Couillaud, F., 1994. The HMG-CoA reductase inhibitor fluvastatin inhibits insect

juvenile hormone biosynthesis. Gen Comp Endocrinol 95, 92-98.

Defelipe, L.A., Dolghih, E., Roitberg, A.E., Nouzova, M., Mayoral, J.G., Noriega, F.G., Turjanski, A.G., 2011.

Juvenile hormone synthesis: "esterify then epoxidize" or "epoxidize then esterify"? Insights from the structural

characterization of juvenile hormone acid methyltransferase. Insect Biochem Mol Biol 41, 228-235.

Feyereisen, R., Farnsworth, D.E., 1987. Characterization and regulation of HMG-CoA reductase during a cycle

of juvenile hormone synthesis. Mol Cell Endocrinol 53, 227-238.

Feyereisen, R., Tobe, S.S., 1981. A rapid partition assay for routine analysis of juvenile hormone release by

insect corpora allata. Anal Biochem 111, 372-375.

Goldstein, J.L., Brown, M.S., 1990. Regulation of the mevalonate pathway. Nature 343, 425-430.

Goodman, W.G., Cusson, M., 2012. The Juvenile Hormones, in: Gilbert, L.I. (Ed.), Insect Endocrinology, pp.

310-365.

Granger, N.A., Sturgis, S.L., Ebersohl, R., Geng, C., Sparks, T.C., 1996. Dopaminergic control of corpora

allata activity in the larval tobacco hornworm, Manduca sexta. Arch Insect Biochem Physiol 32, 449-466.

Hartfelder, K., 2000. Insect juvenile hormone: from "status quo" to high society. Braz J Med Biol Res 33, 157-

177.

Helvig, C., Koener, J.F., Unnithan, G.C., Feyereisen, R., 2004. CYP15A1, the cytochrome P450 that catalyzes

epoxidation of methyl farnesoate to juvenile hormone III in cockroach corpora allata. Proc Natl Acad Sci USA

101, 4024-4029.

101

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

Biochem Mol Biol 54C, 61-68.

Kaneko, Y., Hiruma, K., 2014. Short neuropeptide F (sNPF) is a stage-specific suppressor for juvenile

hormone biosynthesis by corpora allata, and a critical factor for the initiation of insect metamorphosis. Dev

Biol 393, 312-319.

Kinjoh, T., Kaneko, Y., Itoyama, K., Mita, K., Hiruma, K., Shinoda, T., 2007. Control of juvenile hormone

biosynthesis in Bombyx mori: cloning of the enzymes in the mevalonate pathway and assessment of their

developmental expression in the corpora allata. Insect Biochem Mol Biol 37, 808-818.

Li, W.F., Huang, Z.Y., Liu, F., Li, Z.G., Yan, L.M., Zhang, S.W., Chen, S.L., Zhong, B.X., Su, S.K., 2013.

Molecular cloning and characterization of juvenile hormone acid methyltransferase in the honey bee, Apis

mellifera, and its differential expression during caste differentiation. PLoS One 8.

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,

e106737.

Marchal, E., Hult, E.F., Huang, J., Stay, B., Tobe, S.S., 2013a. Diploptera punctata as a model for studying the

endocrinology of arthropod reproduction and development. Gen Comp Endocr 188, 85-93.

Marchal, E., Hult, E.F., Huang, J., Tobe, S.S., 2013b. Sequencing and validation of housekeeping genes for

quantitative real-time PCR during the gonadotrophic cycle of Diploptera punctata. BMC Res Notes 6, 237.

Marchal, E., Zhang, J., Badisco, L., Verlinden, H., Hult, E.F., Van Wielendaele, P., Yagi, K.J., Tobe, S.S.,

Vanden Broeck, J., 2011. Final steps in juvenile hormone biosynthesis in the desert locust, Schistocerca

gregaria. Insect Biochem Mol Biol 41, 219-227.

Martinezgonzalez, J., Buesa, C., Piulachs, M.D., Belles, X., Hegardt, F.G., 1993. Molecular cloning,

developmental pattern and tissue expression of 3-hydroxy-3-methylglutaryl coenzyme a reductase of the

cockroach Blattella germanica. European Journal of Biochemistry 213, 233-241.

Mayoral, J.G., Nouzova, M., Navare, A., Noriega, F.G., 2009a. NADP(+)-dependent farnesol dehydrogenase,

a corpora allata enzyme involved in juvenile hormone synthesis. Proc Natl Acad Sci U S A 106, 21091-21096.

Mayoral, J.G., Nouzova, M., Yoshiyama, M., Shinod, T., Hernandez-Martinez, S., Dolghih, E., Turjanski,

A.G., Roitberg, A.E., Priestap, H., Perez, M., Mackenzie, L., Li, Y.P., Noriega, F.G., 2009b. Molecular and

functional characterization of a juvenile hormone acid methyltransferase expressed in the corpora allata of

mosquitoes. Insect Biochem Mol Biol 39, 31-37.

Mayoral, J.G., Nouzova, M., Yoshiyama, M., Shinoda, T., Hernandez-Martinez, S., Dolghih, E., Turjanski,

A.G., Roitberg, A.E., Priestap, H., Perez, M., Mackenzie, L., Li, Y., Noriega, F.G., 2009c. Molecular and

functional characterization of a juvenile hormone acid methyltransferase expressed in the corpora allata of

mosquitoes. Insect Biochem Mol Biol 39, 31-37.

Meigs, T.E., Roseman, D.S., Simoni, R.D., 1996. Regulation of 3-hydroxy-3-methylglutaryl-coenzyme A

reductase degradation by the nonsterol mevalonate metabolite farnesol in vivo. J Biol Chem 271, 7916-7922.

Meigs, T.E., Simoni, R.D., 1997. Farnesol as a regulator of HMG-CoA reductase degradation: characterization

and role of farnesyl pyrophosphatase. Arch Biochem Biophys 345, 1-9.

Minakuchi, C., Namiki, T., Yoshiyama, M., Shinoda, T., 2008. RNAi-mediated knockdown of juvenile

hormone acid O-methyltransferase gene causes precocious metamorphosis in the red flour beetle Tribolium castaneum. Febs Journal 275, 2919-2931.

Monger, D.J., Lim, W.A., Kezdy, F.J., Law, J.H., 1982. Compactin inhibits insect HMG-CoA reductase and

juvenile hormone biosynthesis. Biochem Biophys Res Commun 105, 1374-1380.

102

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. J Insect Physiol 27, 821-827.

Niwa, R., Niimi, T., Honda, N., Yoshiyama, M., Itoyama, K., Kataoka, H., Shinoda, T., 2008. Juvenile

hormone acid O-methyltransferase in Drosophila melanogaster. Insect Biochem Mol Biol 38, 714-720.

Noriega, F.G., Ribeiro, J.M., Koener, J.F., Valenzuela, J.G., Hernandez-Martinez, S., Pham, V.M., Feyereisen,

R., 2006. Comparative genomics of insect juvenile hormone biosynthesis. Insect Biochem Mol Biol 36, 366-

374.

Nouzova, M., Edwards, M.J., Mayoral, J.G., Noriega, F.G., 2011. A coordinated expression of biosynthetic

enzymes controls the flux of juvenile hormone precursors in the corpora allata of mosquitoes. Insect Biochem

Mol Biol 41, 660-669.

Nyati, P., Nouzova, M., Rivera-Perez, C., Clifton, M.E., Mayoral, J.G., Noriega, F.G., 2013. Farnesyl

phosphatase, a corpora allata enzyme involved in juvenile hormone biosynthesis in Aedes aegypti. PLoS One 8.

Pratt, G.E., Davey, K.G., 1972. Corpus Allatum and Oogenesis in Rhodnius-Prolixus (Stal) .1. Effects of

Allatectomy. Journal of Experimental Biology 56, 201-214.

Pszczolkowski, M.A., Lee, W.S., Liu, H.P., Chiang, A.S., 1999. Glutamate-induced rise in cytosolic calcium

concentration stimulates in vitro rates of juvenile hormone biosynthesis in corpus allatum of Diploptera

punctata. Mol Cell Endocrinol 158, 163-171.

Rankin, S.M., Stay, B., 1984. The changing effect of the ovary on rates of juvenile hormone synthesis in

Diploptera punctata. Gen Comp Endocr 54, 382-388.

Rivera-Perez, C., Nouzova, M., Clifton, M.E., Garcia, E.M., LeBlanc, E., Noriega, F.G., 2013. Aldehyde

dehydrogenase 3 converts farnesal into farnesoic acid in the corpora allata of mosquitoes. Insect Biochem Mol

Biol 43, 675-682.

Rivera-Perez, C., Nouzova, M., Noriega, F.G., 2012. A quantitative assay for the juvenile hormones and their

precursors using fluorescent tags. PLoS One 7, e43784.

Shinoda, T., Itoyama, K., 2003. Juvenile hormone acid methyltransferase: a key regulatory enzyme for insect

metamorphosis. Proc Natl Acad Sci U S A 100, 11986-11991.

Stay, B., Tobe, S.S., 1978. Control of juvenile hormone biosynthesis during reproductive cycle of a viviparous

cockroach 2. Effects of unilateral allatectomy, implantation of supernumerary corpora allata, and ovariectomy.

Gen Comp Endocr 34, 276-286.

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). Eur J

Entomol 94, 335-348.

Thompson, C.S., Yagi, K.J., Chen, Z.F., Tobe, S.S., 1990. The effects of octopamine on juvenile hormone

biosynthesis, electrophysiology, and cAMP content of the corpora allata of the cockroach Diploptera punctata.

J Comp Physiol B 160, 241-249.

Tobe, S.S., Clarke, N., 1985. The effect of L-methionine concentration on juvenile hormone biosynthesis by

corpora allata of the cockroach Diploptera punctata. Insect Biochem 15, 175-179.

Tobe, S.S., Ruegg, R.P., Stay, B.A., Baker, F.C., Miller, C.A., Schooley, D.A., 1985. Juvenile hormone titer

and regulation in the cockroach Diploptera punctata. Experientia 41, 1028-1034.

Tobe, S.S., Stay, B., 1979. Modulation of juvenile hormone synthesis by an analog in the cockroach. Nature

281, 481-482.

Ueda, H., Shinoda, T., Hiruma, K., 2009. Spatial expression of the mevalonate enzymes involved in juvenile

hormone biosynthesis in the corpora allata in Bombyx mori. J Insect Physiol 55, 798-804.

103

Vandesompele, J., De Preter, K., Pattyn, F., Poppe, B., Van Roy, N., De Paepe, A., Speleman, F., 2002.

Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal

control genes. Genome Biol 3.

Wang, Z., Dong, Y., Desneux, N., Niu, C., 2013. RNAi silencing of the HaHMG-CoA reductase gene inhibits

oviposition in the Helicoverpa armigera cotton bollworm. PLoS One 8, e67732.

Weaver, R.J., Audsley, N., 2009. Neuropeptide regulators of juvenile hormone synthesis: structures, functions,

distribution, and unanswered questions. Ann N Y Acad Sci 1163, 316-329.

104

Chapter 3

Mode of action of allatostatins in the regulation of juvenile hormone

biosynthesis in the cockroach, Diploptera punctata

This chapter is an adapted reprint of my article:

Juan Huang, Elisabeth Marchal, Ekaterina F. Hult, Sven Zels, Jozef Vanden Broeck, Stephen S.

Tobe, Mode of action of allatostatins in the regulation of juvenile hormone biosynthesis in the

cockroach, Diploptera punctata, Insect Biochem Mol Biol, 2014, 54, 61-68.

Authors’ contribution:

I designed and performed all experiments in this study and wrote the paper with editing by E.

Marchal, E.F. Hult, and S.S. Tobe.

105

Summary

The FGLamide allatostatins (FGL/ASTs) are a family of neuropeptides with pleiotropic

functions, including the inhibition of juvenile hormone (JH) biosynthesis, vitellogenesis and

muscle contraction. In the cockroach, Diploptera punctata, thirteen FGLa/ASTs and one

allatostatin receptor (AstR) have been identified. However, the mode of action of ASTs in

regulation of JH biosynthesis remains unclear. Here, we determined the tissue distribution of

Dippu-AstR and Dippu-AST. The transcript level of Dippu-AstR in the CA corresponds to the JH

biosynthesis, while the expression of Dippu-AST mRNA in the brain did not show any

significant change during the first gonadotrophic cycle. In addition, silencing Dippu-AstR results

in a significantly increase of JH biosynthesis, while the knockdown of Dippu-AST has no effect.

These results suggest that it is the change of AST receptor in the CA, instead of AST levels in

the brain that control the AST-mediated inhibition of JH production.

To determine the signal pathway of Dippu-AstR, we expressed Dippu-AstR in vertebrate cell

lines, and activated the receptor with the Dippu-ASTs. Our results show that all thirteen ASTs

activated Dippu-AstR in a dose dependent manner, albeit with different potencies. Functional

analysis of AstR in multiple cell lines demonstrated that activation of the AstR receptor resulted

in elevated levels of Ca2+ and cAMP, which suggests that Dippu-AstR can act through the Gαq

and Gαs protein pathways. The study on the target of AST action reveals that FGL/AST affects

JH biosynthesis prior to the entry of acetyl-CoA into the JH biosynthetic pathway.

Introduction

Allatostatins (ASTs), a family of pleiotropic neuropeptides, were originally named for their

ability to inhibit juvenile hormone (JH) biosynthesis by corpora allata (CA) rapidly and

106

reversibly (Bendena et al., 1999). Three families of ASTs have been identified in insects and

named as: FGLa/ASTs (A-type), MIP/ASTs (B-type) and the PISCF/ASTs (C-type) (Coast and

Schooley, 2011). The distribution and function of the three families of ASTs has been reviewed

by Stay and Tobe (Stay and Tobe, 2007). The best documented role of FGLa/AST is their ability

to inhibit juvenile hormone (JH) biosynthesis by corpora allata (CA) (Stay and Tobe, 2007).

Later studies demonstrated other functions of FGLa/ASTs, including regulation of myotropic

activity in gut tissues (Duve et al., 1995; Lange et al., 1995) and cardiac rhythm (Vilaplana et al.,

1999), inhibition of vitellogenin synthesis in the fat body (Martin et al., 1996), and stimulation of

enzyme activity in the lumen of the midgut (Fuse et al., 1999).

As neuropeptides, ASTs exert their effects by binding to a G protein-coupled receptor, AstR.

FGLa/AST receptors, which are structurally related to the mammalian galanin receptor, were

first identified in the fly Drosophila melanogaster DAR-1 (Birgul et al., 1999; Lenz et al., 2000)

and DAR-2 (Birgul et al., 1999; Lenz et al., 2000), and later in the silkworm, Bombyx mori

(Secher et al., 2001) and the stick insect, Carausius morosus (Auerswald et al., 2001). In D.

punctata, putative AstRs have previously been partially characterized using photoaffinity

labeling and a radioligand-binding assay in the CA and brain (Cusson et al., 1991, 1992; Yu et

al., 1995). Lungchukiet et al (Lungchukiet et al., 2008b) identified a putative FGLa/AST

receptor gene in D. punctata. Silencing this gene resulted in a significant increase in JH

biosynthesis (Lungchukiet et al., 2008a; Lungchukiet et al., 2008b).

The responses of FGLa/AST receptor to FGLa/ASTs have been assayed in D. melanogaster

(DAR-1 and DAR-2) (Birgul et al., 1999; Larsen et al., 2001), B. mori (Secher et al., 2001) and

the cockroach, Periplaneta americana (Gade et al., 2008). The studies in B. mori and P.

americana focused on the response of AstR to exogenous ASTs. The signaling pathway of AstR

107

was, however, not completely elucidated. Larsen et al (Larsen et al., 2001) expressed Drosophila

FGLa/AST receptors, DAR-1 and DAR-2 in CHO cells and activated them with four putative

FGLa/ASTs from D. melanogaster AST and four FGLa/ASTs from D. punctata in the presence

or absence of pertussis toxin (PTX). They found that PTX caused a complete loss of Ca2+ signal

in cells expressing DAR-1, and a decreased Ca2+ signal in cells expressing DAR-2. Their results

suggested that the activation of DAR-1 and DAR-2 by FGLa/ASTs coupled to multiple signaling

pathways, including Gi/o protein and other, PTX-insensitive G-proteins.

Although the function of AST has been well-studied, little is known about the precise target of

AST action. Previous studies focused on select enzymes in the JH biosynthetic pathway. It is

now known that thirteen enzymes are involved in the JH biosynthetic pathway (Fig. 2.1) (Belles

et al., 2005; Nouzova et al., 2011). The potential targets of action of ASTs were originally

studied by employing different known JH precursors (Pratt et al., 1991; Pratt et al., 1989). The

results suggested that the inhibitory action of AST on JH biosynthesis resides in step(s) prior to

mevalonate. However, neither HMG-CoA synthase nor HMG-CoA reductase activity was

affected by ASTs (Sutherland and Feyereisen, 1996). These authors proposed that the target of

AST action on JH biosynthesis is in step(s) prior to the JH biosynthetic pathway, and may be

related to the transport of citrate from mitochondria to cytosol and/or to the cleavage of citrate to

yield acetyl-CoA.

Our study focuses on the viviparous cockroach, Diploptera punctata, in which ASTs were first

characterized. The first gonadotropic cycle is characterized by a precise regulation of JH

biosynthesis necessary to coordinate a specific series of reproductive events closely correlated

with oocyte growth. This makes D. punctata an ideal model for a more in depth study of the

mode of action of AST and its signal transduction pathway. We have analyzed the spatial

108

expression pattern of the AST precursor and AstR which is consistent with their roles in

regulating JH biosynthesis. Moreover, by using an aequorin-based assay and expressing the

receptor in a mammalian system, we have unambiguously identified that AST is a ligand for the

candidate AstR (Lungchukiet et al., 2008a; Lungchukiet et al., 2008b) and that this receptor can

couple to Ca2+ and cAMP.

Thirteen FGLa/ASTs have been identified in D. punctata. These peptides share a conserved C-

terminal Tyr (Phe)- Xaa-Phe-Gly-Leu-NH2, which is believed to be the main functional region

for the inhibition of JH biosynthesis (Donly et al., 1993; Marchal et al., 2013a). The FGLa/ASTs

inhibit JH biosynthesis by CA at low concentrations in vitro but with different potencies (Tobe et

al., 2000). Our study has also examined the relationship between binding affinity and potency

with a view to understanding the sites of action of the peptides.

The precise target of AST action remains unclear so far. Our study did not find any significant

changes in the transcript level of genes encoding enzymes in JH biosynthetic pathway. The

rescue of AST-induced JH inhibition by JH precursors suggest that the target of AST action is

prior to the entry of Acetyl-CoA into the JH biosynthetic pathway.

Materials and Methods

Insects - D. punctata were reared in cages and fed with lab chow and water at libitum at 27 ºC in

a dark room. Newly molted female adult cockroaches were picked from the colony and raised in

separate containers. Mated status was confirmed by the presence of a spermatophore.

Tissue collection - Cockroach tissues were dissected under a dissecting microscope. Basal oocyte

length was measured to determine the physiological age of cockroaches. Selected tissues were

dissected and cleaned in sterile cockroach ringer solution (150 mM NaCl, 12 mM KCl, 10 mM

109

CaCl2.2H2O, 3 mM MgCl2.6H2O, 10 mM HEPES, 40 mM Glucose, pH 7.2), and stored at -80 ºC

to prevent degradation.

RNA extraction and cDNA synthesis - Pooled samples were homogenized using a plastic pestle

and RNA was extracted using the RNeasy Mini Kit (Qiagen) according to the manufacturer’s

instructions. An additional DNase treatment (RNase-free DNase set, Qiagen) was performed to

eliminate potential genomic DNA contamination. RNA of CA was extracted using the

RNAqueous®-Micro Kit (Ambion). DNase treatment was performed to eliminate genomic DNA

contamination. The quantity and quality of RNA was determined using a Nanodrop

spectrophotometer (Thermo Scientific). An equal amount of RNA was transcribed with

Superscript III reverse transcriptase (Invitrogen Life Technologies) utilizing random hexamers as

described in the protocol. The resulting cDNA was diluted tenfold.

Quantitative Real Time-PCR (q-RT-PCR) - Prior to target gene profiling, Tubulin and EF1a were

chosen as the optimal housekeeping genes according to a previous study (Marchal et al., 2013b).

The q-RT-PCR reactions were performed in triplicate on a CFX384 Touch ™ Real-Time PCR

Detection System (Bio-Rad) in a final volume of 10 µl, containing 1 µl of cDNA, 5 µl IQ™

SYBR® Green Supermix (Bio-Rad), 1 µl forward and reverse primer (5 µM) and 2 µl of MQ-

water. The reaction was incubated for 3 min in 95ºC, followed by 40 cycles with following

thermal profile: 95ºC, 10 s; 59ºC, 30 s. Target specificity was confirmed by performing a

dissociation protocol (melt curve analysis) and running a few representative q-RT-PCR products

on an agarose gel containing GelRed™ (Biotium). Realtime primers used for q-RT-PCR are

listed in Table S3.1. Primer sets were validated by determining relative standard curves for each

gene transcript using a five-fold serial dilution of a calibrator cDNA sample. Efficiency and

correlation coefficient (R²) are shown in Table S1. The primer sets for genes in the JH

110

biosynthetic pathway were chosen as in Chapter 2. The quantity of mRNA for each tested gene

relative to reference genes was determined as described by Vandesompele et al. (Vandesompele

et al., 2002).

Peptides and substrates - Thirteen Dippu-ASTs were custom synthesized by GL Biochemical

Ltd., Shanghai (China). Peptides were purified by high performance liquid chromatography

(HPLC) (purity ≥95%). Peptides were dissolved in water to obtain a concentration of 1mM.

Peptide solutions were stored at -80 ºC prior to further processing and dilution.

Acetyl-CoA, mevalonic acid (MA), diphosphomevalonate (DPPM) and farnesol were purchased

from Sigma-Aldrich Canada. Substrates were dissolved in water before use.

Cell culture and transfection - Chinese hamster ovary (CHO) WTA11 and PAM28 cells stably

expressing apoaequorin (Euroscreen, Belgium) and human embryonic kidney (HEK) 293 cells

were cultured in monolayers in Dulbecco’s Modified Eagles Medium nutrient mixture F12-Ham

(DMEM/F12) (Sigma) supplemented with 10% heat-inactivated fetal calf serum (Invitrogen),

100 IU/ml penicillin and 100 µg/ml streptomycin (Invitrogen). An additional 250 µg/ml Zeocin

(Invitrogen) was added to the medium for CHO-WTA11 cells, and an additional 5 µg/ml

Puromycin (Sigma) was added to the medium for PAM cells. The cells were cultured at 37ºC

with a constant supply of 5% CO2.

Transfections with pcDNA3.1D-Dippu-AstR or empty pcDNA3.1D vector were carried out in

T75 flasks at 60 to 80% confluency. Transfection medium for CHO cells was prepared using the

Lipofectamine LTX kit (Invitrogen) with 3.75ml Opti-MEM, 7.5 µg vector construct and 18.75

µl Plus™ Reagent in a 5 ml polystyrene round-bottom tube. After 5 min incubation at room

temperature, 45 µl LTX (Invitrogen) was added. Transfection medium was added dropwise to

the cells after 30 min incubation at room temperature. The transfection medium of the HEK293

111

cells was similar to that of the CHO cells, except that 9 µg DNA construct (6 µg of vector

construct and 3 µg of reporter gene plasmid) was added to the medium. The luciferase ORF,

downstream of a cAMP responsive element (CRE) served as the reporter gene. After transfection,

cells were incubated overnight and an additional 10 ml of culture medium was added. The cells

were allowed to grow for another night prior to intracellular Ca2+ or cAMP measurements.

Aequorin assay - CHO cells were detached using 1× Phosphate Buffered Saline (PBS)

containing 0.2% EDTA (pH 8.0), collected and pelleted by centrifugation in DMEM/F12

medium. The number of viable and nonviable cells was determined using a NucleoCounter NC-

100™ (Chemometic). The cells were then resuspended to a density of 5×106 cells/ml in sterile

filtered DMEM/bovine serum albumin (BSA) medium (DMEM/F12 with L-glutamine and 15

mM HEPES, without Phenol red, supplemented with 0.1% BSA). A concentration of 5 µM

coelenterazine h (Invitrogen) was added and the cells were incubated for 4 h in the dark at room

temperature with gentle shaking to reconstitute the holo-enzyme aequorin. The cells were then

diluted 10-fold in BSA medium and incubated for another 30 min in the dark with gentle shaking.

Peptides were dissolved in BSA medium and dispensed in 50 µl aliquots into the wells of a white

96-well plate. Fifty µl of the cell suspension was injected into each well and light emission was

recorded using a Mithras LB 940 multimode microplate reader (Berthold Technologies) over 30

sec. The cells were lysed by injection of 50 µl 0.3% Triton X-100 and light emission was

monitored for an additional 8 sec. The total response (ligand + Triton-X100) is the representative

for the quantity of viable cells present in the well. BSA medium was used as a negative control

in each row of the plate, and 1 µM ATP served as a positive control. The negative response was

subtracted from the luminescence measured in wells of the same row. Calculations were made

112

using the output file from Microwin software (Berthold Technologies) in Excel (Microsoft).

Further analysis was done in Excel and GraphPad Prism 6.

cAMP reporter assay - HEK cells were detached using 1× PBS containing 0.2% EDTA (pH 8.0),

collected and pelleted by centrifugation in DMEM/F12 medium. The number of viable and

nonviable cells was determined using a NucleoCounter NC-100™ (Chemometic). The cells were

then resuspended to a concentration of 1×106 cell/ml in DMEM/F12 containing 200 µM 3-

isobutyl-1-methylxanthine (IBMX, Sigma) to prevent cAMP breakdown. Peptides were

dissolved in IBMX medium in the presence or absence of 20 µM forskolin, and were dispensed

in 50 µl aliquots into the wells of a white 96-well plate. Fifty µl of cell suspension was added

into each well and the plate was incubated in a CO2 incubator (5% CO2) at 37ºC for 3.5 h. Fifty

µl of Steadylite Plus substrate (PerkinElmer) was then added to each well and the plate was

incubated in the dark for 15 min while gently shaking. Light emission resulting from the

luciferase enzymatic activity was recorded for 5 s/well using a Mithras LB940. Results were

analyzed using the output of MicroWin and further processed by Excel and GraphPad Prism 6.

Radiochemical assay (RCA) - The in vitro RCA for JH biosynthesis was performed as described

by Feyereisen and Tobe (Feyereisen and Tobe, 1981) and modified by Tobe and Clarke (Tobe

and Clarke, 1985). Two incubations were conducted in the JH biosynthesis rescue experiments.

CA were incubated in TC199 medium with 10-6 M AST7 for 3h, and then transferred to fresh

medium with 10-6 M AST7 and JH precursors for a second 3h incubation. JH biosynthesis was

determined after each incubation. The rate of JH biosynthesis during the first incubation was

used as the control value.

RNA interference (RNAi) - Dippu-AstR and Dippu-AST dsRNA constructs were prepared using

the MEGAscript® RNAi Kit (Ambion). Primers used are given in Table S3.2. The Fragments

113

were amplified by PCR using forward and reverse primers with T7 promoters

(TAATACGACTCACTATAGGGAGA) attached to the 5’ end. Fragments were then subcloned

and sequenced to verify the presence of the T7 promoter. The amplified fragment was used in an

RNA transcription reaction and incubated overnight to yield annealed dsRNA transcripts. A

nuclease digestion was subsequently performed to remove ssRNA and DNA remaining in the

product. The dsRNA was further purified according to the manufacturer’s instructions (Ambion).

Concentration of the dsRNA construct was determined using a nanodrop instrument (Thermo

Fisher Scientific Inc.). Five-fold diluted dsRNA was run on a 1.2% agarose gel to examine the

quality and integrity of the construct. The control dsRNA was prepared using a PCR fragment

amplified from a non-coding region of pJET1.2 cloning vector (Thermo Scientific).

Dippu-AstR dsRNA was diluted in cockroach saline to a concentration of 250ng/µl. Each adult

female was injected with 4 µl of dsRNA solution on day 0, 2 and 4 after the final moult. CA

were dissected on day 6 and stored in liquid nitrogen prior to RNA extraction. A second set of

cockroaches was injected with Dippu-AstR dsRNA in the same scheme as described above. CA

were dissected on day 6 and cleaned in TC199 medium (GIBCO; 1.3 mM Ca2+, 2% Ficoll,

methionine-free) for their use in the radiochemical assay (RCA) determining the effect of AST

on JH biosynthesis. Basal oocyte length was measured during dissection.

Dippu-AST dsRNA was diluted in cockroach saline to a concentration of 250ng/µl. Each adult

female was injected with 4 µl of dsRNA solution on day 0, 2, 4 and 5 after the final moult. Brain

were dissected on day 6 and stored in liquid nitrogen prior to RNA extraction. CA were dissected

on day 6 and cleaned in TC199 medium (GIBCO; 1.3 mM Ca2+, 2% Ficoll, methionine-free) for

their use in the radiochemical assay (RCA) determining the effect of AST on JH biosynthesis.

Basal oocyte length was measured during dissection.

114

Data analysis – Data set were analyzed by T-test to determine the significant difference. All

results were expressed as mean ± SEM, and considered significantly different at P ≤ 0.05.

Results

Tissue distribution of Dippu-AST and Dippu-AstR

The tissue distribution of Dippu-AST and Dippu-AstR was determined in adult male and female

cockroaches using q-RT-PCR (Fig. 3.1). Nine tissues were used to examine the tissue specificity

of Dippu-AST and Dippu-AstR: brain (Br), nerve cord (NC), corpora allata (CA), fat body (Fb),

ovary (Ov), midgut (MG) and Malpighian tubules (MT) from females and accessory gland (AG)

and testes (Te) from males. The Dippu-AST gene is expressed in brain, nerve cord and midgut,

which is consistent with the pleiotropic functions of ASTs (Fig. 3.1). Dippu-AstR shows the

highest transcription levels in the CA, followed by nerve cord, brain and fat body. All the other

tissues tested showed either negligible or undetectable levels of Dippu-AstR mRNA.

Developmental expression during the first gonadotrophic cycle of females

We determined the developmental profile of Dippu-AstR in the CA, brain and fat body of mated

females during their first gonadotrophic cycle (Fig. 3.2). In CA, the Dippu-AstR mRNA declined

on days 2-4 and then increased significantly on day 6. The transcript level of Dippu-AstR in the

CA was inversely correlated with the rate of JH biosynthesis (Fig. 3.2A). In brain, on the other

hand, transcript levels of Dippu-AstR and Dippu-AST showed only minor variations (Fig. 3.2B).

The transcript level of Dippu-AstR in fat body was relatively low, with the highest level of

mRNA expression on day 4, and negligible on all other days (data not shown).

115

B r N C C A F b O v T e AG M G M T

0

2

4

6

Re

lati

ve

mR

NA

qu

an

tity

D ip p u -A s tR

D ip p u -A S T

Figure 3.1. Relative expression levels of Dippu-AstR (black) and Dippu-AST (gray) mRNA

in tissues of day 4 males and mated females. The data represent averages of 3 pools (10

animals per pool), run in triplicate. Tissues tested are brain (Br), nerve cord (NC), corpora allata

(CA), fat body (FB), ovary (Ov), midgut (MG) and Malpighian tubules (MT) from females and

testes (Te) and accessory gland (AG) from males. Values represent mean ± SEM.

116

D 0 D 1 D 2 D 3 D 4 D 5 D 6 D 7

0

5

1 0

1 5

2 0

0

5

1 0

1 5

2 0

2 5

Re

lati

ve

mR

NA

qu

an

tity

of

As

tR JH

bio

sy

nth

es

is(p

mo

l/h/C

A)

D ip p u -A s tR

J H b io s y n th e s is

D 0 D 1 D 2 D 3 D 4 D 5 D 6 D 7

0 .0

0 .5

1 .0

1 .5

2 .0

0

1

2

3

4

Re

lati

ve

mR

NA

qu

an

tity

of

As

tRR

ela

tive

mR

NA

qu

an

tity o

f AS

T

D ip p u -A s tR

D ip p u -A S T

A

B

Figure 3.2. Relative expression levels of Dippu-AstR and Dippu-AST mRNA during the first

gonadotrophic cycle. (A) Dippu-AstR mRNA expression and JH biosynthesis in CA. Dippu-

AstR mRNA levels were quantified by q-RT-PCR and normalized against levels of Armadillo

and EF1a mRNA (Marchal et al., 2013b). The data represent the average of 3 biologically

independent pools (10 animals per pool), run in triplicate. Rate of JH biosynthesis was measured

using the RCA, n≥10. (B) Dippu-AstR and Dippu-AST mRNA expression in brain. Dippu-AstR

and Dippu-AST mRNA levels were normalized against levels of Tubulin and EF1a mRNA

(Marchal et al., 2013b). Values represent mean ± SEM.

117

Effect of Dippu-AST dsRNA on JH biosynthesis

To determine the controlling role of AST in JH biosynthesis, we knocked down Dippu-AST

using RNAi. Injecting Dippu-AST dsRNA results in a 91% decrease of Dippu-AST mRNA in the

brain. However, JH biosynthesis of CA did not show any significant change in the dsRNA

treated animals (Fig. 3.3).

Functional activation of Dippu-AstR with ASTs

We initially expressed Dippu-AstR in CHO-WTA11 cells. This cell line expresses the Ca2+

reporter apoaequorin and Gα16, a promiscuous G protein that couples to most GPCRs with

subsequent mobilization of intracellular Ca2+ (Stables et al., 1997). As shown in Fig. 3.4, all 13

tested ASTs induced clear dose-dependent bioluminescence responses in AstR expressing cells.

Most of the peptides showed similar degrees of biological efficacy (ability to activate AstR) but

differed considerably in potency, with EC50 values ranging from 0.2 nM for AST6 to 30 nM, in

the case of AST13 (Table 3.1). In general, the abilities of ASTs to activate AstR corresponded to

their potencies as inhibitors of JH production by the CA, with the exception of AST1, AST5 and

AST6 (Tobe et al., 2000) (Table 3.1).

To determine the second messenger pathways involved in Dippu-AstR activation, we first

expressed Dippu-AstR in CHO-PAM28 cells (lacking the promiscuous Gα16). The expression of

apoaequorin in the cell allows testing whether the receptor can couple naturally through

intracellular Ca2+. Two ASTs (AST5 and 6) were chosen because of their high potency in

inhibiting JH biosynthesis or in activating Dippu-AstR. As shown in Fig. 3.5A, AST5 and 6

induced dose-dependent intracellular Ca2+ responses in AstR-transfected CHO-PAM28 cells,

with EC50 values of 21.4 nM and 1.1 nM, respectively.

118

We subsequently tested whether the receptor coupled with cAMP in the signal transduction

pathway. We expressed Dippu-AstR in HEK293 cells, which contain the luciferase gene under

the control of a cAMP response element. We assayed different concentrations of AST5 and

AST6 in the presence or absence of forskolin. Forskolin activates adenylate cyclase (AC), which

increases intracellular levels of cAMP. If AstR would couple negatively to AC, we would expect

lower levels of cAMP following application of ASTs. This was not the case. Application of

AST5 or AST6 to transfected HEK cells did not cause any significant change in cAMP level.

However, in the absence of forskolin, AST5 and AST6 treatments resulted in a dose-dependent

increase in cAMP concentration, as measured by assay of luciferase activity (Fig. 3.5B). The

EC50 values were 287.8 nM for AST5 and 24.3 nM for AST6. The EC50 values for the cAMP-

based luciferase assay in HEK293 cells were higher than for the Ca2+ responses detected in either

of the two CHO cell lines, but the relative order of ligand potency was maintained. CHO-

WTA11, CHO-PAM28 and HEK293 cells transfected with pcDNA3.1D (empty vector) did not

show any response to ASTs.

AST does not affect the transcript level of enzymes in the JH biosynthetic pathway

The injection of AstR dsRNA on day 0, 2 and 4 resulted in a 67% knockdown of AstR mRNA

levels in CA of day 6 females (Fig. 3.6A). The JH biosynthetic activity of CA was measured

using the RCA. In AstR knockdown animals, JH production increased by 60% and the response

of CA to AST decreased by 28% (Fig. 3.6B and 3.6C).

Eleven of 13 genes encoding enzymes in the JH biosynthetic pathway have been identified

(Chapter 2). To determine the target of AST action in the JH biosynthetic pathway, we measured

the mRNA levels of the 11 genes encoding enzymes catalyzing different steps in the JH

biosynthetic pathway. Although AstR dsRNA treatment resulted in an increase in the rate of JH

119

biosynthesis, none of the relative transcript levels showed significant changes (Fig. 3.6D). The

high variation of the transcript level of JHAMT on day 6 can be explained by previous results

that show a great degree of fluctuation in JHAMT mRNA levels at times when rates of JH

biosynthesis are highly dynamic (see Chapter 2).

To confirm our results, we incubated CA in medium containing 10-7 M AST7 for 3 hours, which

results in a 60 to 70% decrease in JH biosynthesis (Tobe et al., 2000), and determined the

transcript level of the enzymes in the CA. As in the AstR RNAi experiments, the mRNA levels of

the 11 genes tested show no significant changes (Fig. 3.7).

JH precursors reverse AST-induced inhibition of JH biosynthesis

Earlier studies showed that mevalonate and farnesol were able to partially reverse AST-induced

inhibition of JH biosynthesis (Pratt et al., 1991; Pratt et al., 1989). To determine which enzyme(s)

is/are affected by ASTs in the JH biosynthetic pathway, we treated CA with select JH precursors

in the presence of 10-6 M AST7. If the activity of enzymes was inhibited by AST, the JH

precursors prior to the enzyme will not be able to reverse the inhibition of AST. As shown in Fig.

3.8, all exogenous JH precursors acetyl-CoA, mevalonic acid, diphosphomevalonate and farnesol

significantly stimulated the rate of JH biosynthesis in the presence of AST7.

120

c o n tro l D ip p u -A S T d s R N A

0 .0

0 .5

1 .0

1 .5

Re

lati

ve

mR

NA

qu

an

tity

of

AS

T

***

C o n tro l D ip p u -A S T d s R N A

0

5

1 0

1 5

2 0

2 5

JH

bio

sy

nth

es

is(p

mo

l/h

/CA

) C o n tro l

D ip p u -A S T dsR N A

A B

Figure 3.3. The effect of Dippu-AST dsRNA on JH biosynthesis by the CA. (A) Efficiency of

Dippu-AstR RNAi-mediated knockdown in the brain of day 6 mated females. Relative quantity

of Dippu-AstR mRNA levels in CA was compared between control and AST dsRNA treated

animals. (B) JH biosynthesis by CA from Dippu-AST dsRNA-treated animals. Glands were taken

from day 6 mated females, n≥10. The quantity of mRNA data represent averages of 3 pools (8

pairs of CA per pool) run in triplicates using q-RT-PCR and normalised to Tubulin and EF1a

mRNA (Marchal et al., 2013b) . Vertical bars indicate S.E.M. . Significant differences are

indicated by asterisks (*P < 0.05)

121

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 1

lo g [A S T 1 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 1 .2 9 5 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 2

lo g [A S T 2 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .0 3 3 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 3

lo g [A S T 3 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .8 2 9 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 4

lo g [A S T 4 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .4 8 9 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 5

lo g [A S T 5 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 4 .2 1 4 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 6

lo g [A S T 6 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 1 .9 2 9 e -0 1 0

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 7

lo g [A S T 7 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .1 3 5 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 8

lo g [A S T 8 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 4 .0 8 7 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 9

lo g [A S T 9 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 5 .6 5 4 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 1 0

lo g [A S T 1 0 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 7 .7 1 7 e -0 1 0

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 1 1

lo g [A S T 1 1 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .7 4 9 e -0 0 9

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 1 2

% b

iolu

min

es

ce

nc

e

E C 5 0 = 1 .0 4 4 e -0 0 8

lo g [A S T 1 2 ] (M )

-1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

A S T 1 3

lo g [A S T 1 3 ] (M )

% b

iolu

min

es

ce

nc

e

E C 5 0 = 2 .9 9 9 e -0 0 8

Figure 3.4. Dose-response curves for ASTs in CHO-WTA11 cells expressing Dippu-AstR. Data points represent the average ±

SEM of three independent measurements performed in duplicate and are expressed as percentage of the maximal response. The zero

response level corresponds to treatment with BSA buffer only.

122

-1 4 -1 3 -1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

D ip p u -A s tR in C H O -P A M 2 8

C o n c e n t r a t io n lo g [ M ]

% b

iolu

min

es

ce

nc

e A S T 5

A S T 6

-1 2 -1 1 -1 0 -9 -8 -7 -6 -5 -4

-2 0

0

2 0

4 0

6 0

8 0

1 0 0

1 2 0

D ip p u -A s tR in H E K 2 9 3

C o n c e n t r a t io n lo g [ M ]%

bio

lum

ine

sc

en

ce A S T 5

A S T 6

A B

Figure 3.5. Dose-response curves for the bioluminescence response induced in (A) CHO-

PAM28 and (B) HEK293 cells expressing Dippu-AstR. Data points represent the average ±

SEM of three independent measurements performed in duplicate and are expressed as a

percentage of the maximal response. The zero response level corresponds to treatment with BSA

buffer only. Peptides tested are AST5 (squares) and AST6 (triangles).

123

c o n tro l A s tR d s R N A

0 .0

0 .5

1 .0

1 .5

2 .0

Re

lati

ve

mR

NA

qu

an

tity

of

As

tR

*

c o n tro l A s tR d s R N A

0

2 0

4 0

6 0

8 0

% J

H b

ios

yn

the

sis

in

hib

itio

n

*

c o n tro l A s tR d s R N A

0

5

1 0

1 5

JH

bio

sy

nth

es

is(p

mo

l/h

/CA

)

*

T h io l H M G S H M G R M K P M K P P -M e v D IP P I F P P S F O L D J H A M T C Y P 1 5

0

1

2

3

4

5

Re

lati

ve

mR

NA

qu

an

tity

c o n tro l

A s tR d s R N A

A B C

D

Figure 3.6. The effect of Dippu-AstR dsRNA on JH biosynthesis by the CA and on the

expression of genes encoding enzymes in the JH biosynthetic pathway of D. punctata. (A)

Efficiency of Dippu-AstR RNAi-mediated knockdown in mated females. Relative quantity of

Dippu-AstR mRNA levels in CA was compared between control and AstR dsRNA treated

animals. (B) JH biosynthesis by CA from Dippu-AstR dsRNA-treated animals. Glands were

taken from day 6 mated females, n≥10. (C) Effect of 10-7 M Dippu-AST7 on JH biosynthesis by

CA from control and AstR knockdown animals. CA were incubated first in normal medium for

3h and subsequently incubated in medium with 10-7 M AST7 for another 3h. The percentage of

inhibition was calculated using the rate of JH biosynthesis during the first and second incubation.

(D) The effect of silencing Dippu-AstR on the transcript levels of genes encoding enzymes in the

JH biosynthetic pathway. Enzyme abbreviations are as described in Figure S1. The mRNA

quantity data represent averages of 3 pools (8 pairs of CA per pool) run in triplicates using q-RT-

PCR. Vertical bars indicate S.E.M.. Significant differences are indicated by asterisks (*P < 0.05).

124

AC AT H M G S H M G R M K P M K P P M e v D IP P I F P P S F AL D J H AM T E P O X

0

5

1 0

1 5

2 0R

ela

tiv

e m

RN

A q

ua

nti

ty c o n tro l

tre a tm e n t

Figure 3.7. The effect of AST on the expression levels of genes encoding enzymes in the JH

biosynthetic pathway in CA of day 6 mated female D. punctata. CA were incubated for 3h in

normal (control) medium or medium supplemented with 10-7 M AST7 (treatment) and then

washed with cockroach saline prior to the RNA extraction. RNA was extracted from CA in

control and treatment groups to determine the transcript level of genes encoding enzymes in JH

biosynthetic pathway. Enzyme abbreviations are as described in Figure S1. The data represent

averages of 3 pools (8 pairs of CA per pool) run in triplicate using q-RT-PCR. Vertical bars

indicate S.E.M.

Ac e tyl C o A M A M AP T F a r n e s o l

0

5

1 0

1 5

JH

re

lea

se

(p

mo

l/C

A/h

)

c o n tro l

J H p re c u rs o r

***

******

***

Figure 3.8. JH precursors rescue the AST-induced JH inhibition. JH biosynthesis was

evaluated in CA that were first incubated in medium TC199 with 10-6 M AST7 (control), and

then in medium with 10-6 M AST7 together with the individual precursors: Acetyl-CoA (100

µM), MA (100 µM), DPPM (100 µM) or farnesol (40 µM). CA were dissected from day 7 mated

females. Each data point represents mean ± SEM (n ≥ 10) Significant differences are indicated

by asterisks (***P < 0.001).

125

Table 3.1. Potency of Dippu-ASTs a: activation of AstR in CHO-WTA11 cells (EC50) or

inhibitory effect on JH release (IC50)

Allatostatin EC50

1

/nM

Rank1

order

(EC50)

IC502

/nM

Rank2

order

(IC50)

Sequence

AST1 1.295 3 9.566 11 Leu-Tyr-Asp-Phe-Gly-Leu-NH2

AST2 2.033 4 0.4254 3 Ala-Tyr-Ser-Tyr-Val-Ser-Glu-Tyr-Lys-Arg-Leu-Pro-Val- Tyr-Asn-Phe-Gly-Leu-NH2

AST3 2.829 8 12.54 12 Ser-Lys-Met-Tyr-Gly-Phe-Gly-Leu-NH2

AST4 2.489 6 1.139 5 Asp-Gly-Arg-Met-Tyr-Ser-Phe-Gly-Leu-NH2

AST5 4.214 10 0.1063 1 Asp-Arg-Leu-Tyr-Ser-Phe-Gly-Leu-NH2

AST6 0.1929 1 2.725 8 Ala-Arg-Pro-Tyr-Ser-Phe-Gly-Leu-NH2

AST7 2.135 5 0.4118 2 Ala-Pro-Ser-Gly-Ala-Gln-Arg-Leu-Tyr-Gly-Phe-Gly-Leu-NH2

AST8 4.087 9 3.275 10 Gly-Gly-Ser-Leu-Tyr-Ser-Phe-Gly-Leu-NH2

AST9 5.654 11 2 6 Gly-Asp-Gly-Arg-Leu-Tyr-Ala-Phe-Gly-Leu-NH2

AST10 0.7717 2 0.9225 4 Pro-Val-Asn-Ser-Gly-Arg-Ser-Ser-Gly-Ser-Arg-Phe-Asn-Phe-Gly-Leu-NH2

AST11 2.749 7 2.795 9 Tyr-Pro-Gln-Glu-His-Arg-Phe-Ser-Phe-Gly-Leu-NH2

AST12 10.44 12 2.011 7 Pro-Phe-Asn-Phe-Gly-Leu-NH2

AST13 29.99 13 12.84 13 Ile-Pro-Met-Tyr-Asp-Phe-Gly-Ile-NH2

a Potency is defined as the dose required to achieve a given level of activation of AstR or

inhibition of JH biosynthesis, and are listed in rank order. 1 EC50 was determined by AstR

activation assay in CHO-WTA11 cells (details seen Fig. 1). 2 IC50 values are as reported by

(Tobe et al., 2000) (CA from day 2 virgin female).

126

Discussion

Our observations on the expression of the Dippu-AstR and Dippu-AST provide insights into the

tissue-specific interaction between the ligands and the receptor. In brain and nerve cord, both

Dippu-AstR and Dippu-AST are expressed, whereas in CA, only the receptor is expressed. This is

consistent with observations that AST is delivered to the CA by nerves from neurosecretory cells

in the brain (Stay et al., 1992). In the midgut, however, Dippu-AST is expressed, but its receptor

is not (Fig. 3.1); tissue distribution of Dippu-AstR and Dippu-AST in day 7 animals provided

similar results (Fig. S3.1). Previous studies suggest that ASTs can induce myotropic activity in

gut tissues (Duve et al., 1995; Lange et al., 1995). Furthermore, putative receptors for ASTs in

midgut were partially characterized using a radioligand-binding assay (Bowser and Tobe, 2000).

Therefore, the undetectable transcript level of Dippu-AstR in midgut suggests that one or more

additional AstR in D. punctata may exist, just as two receptors for allatostatins have been

identified in Drosophila melanogaster (Birgul et al., 1999; Lenz et al., 2000).

The sensitivity of CA to ASTs differs between animals of different ages (Pratt et al., 1990; Stay

et al., 1991). CA appear more sensitive to AST treatment when JH biosynthetic activity is low

than at stages of high JH production. In the present study, we have shown that the transcript

levels of Dippu-AstR in the CA correlate with changes in the sensitivity of the CA to ASTs (Fig.

3.2A). It is therefore reasonable to suggest that the changing transcript levels of Dippu-AstR are

partly responsible for the changes in JH biosynthesis during the first gonadotropic cycle. The

ability of ASTs to inhibit JH biosynthesis depends not only on the sensitivity of the CA, but also

on the concentration of the ASTs (Stay et al., 1996). q-RT-PCR of the AST precursor gene

showed that there was no significant change in the transcript level of ASTs (Fig. 3.2B). Lloyd et

al (2000) determined the quantity of AST in CA and brain by ELISA using antibody against

127

Dippu-AST7, and found no significant change in the content of AST7 during the first

gonadotropic cycle. Meanwhile, silencing the AST precursor gene did not cause any significant

increase of JH production (Fig. 3.3). Same result was observed in the German cockroach, B.

germanica (Maestro and Belles, 2006). These results suggest that the AST-inhibition of JH

biosynthesis by AST is controlled by the change of expression of the AST receptor.

To functionally analyze AstR, we expressed the Dippu-AstR in mammalian cells and activated it

with Dippu-AST. We first expressed Dippu-AstR in CHO cells, demonstrating that it is indeed a

Diploptera allatostatin receptor. All thirteen ASTs are very potent in activating AstR expressed

in CHO-WTA11 cells with EC50 values in the low nanomolar range (Fig. 3.4). Despite the

structural similarities, there were some differences in the response to individual peptides. AST13,

which contains isoleucine at the C-terminus, exhibited the highest EC50. This result reinforces

the importance of the C-terminal pentapeptide motif Y/FXFGL-NH2. Moreover, the amino acids

outside of the C-terminal pentapeptide motif appear to determine the affinity of AST for AstR as

well. AST5 and AST6 have the same pentapeptide motif (Table 3.1), whereas their ability to

activate the receptor is significantly different. The change of Pro to Leu, and Ala to Asp

decreases AST affinity of AstR by 20-fold.

A P. americana allatostatin receptor was expressed in Xenopus oocytes, and the potency of

different Periplaneta ASTs in activating Peram-AstR was determined (Gade et al., 2008). Five

Dippu-ASTs (AST 1, 2, 3, 6 and 13) have the same structure as in P. americana (AST 1, 2, 3, 6

and 14). And the rank order of potency of these ASTs in activating Dippu-AstR expressed in

CHO-WTA11 cells is similar to that of P. americana receptor (Gade et al., 2008). However, the

potencies of AST 1, 5 and 6 in activating AstR differ from their capacities to inhibit JH

production in vitro (Table 1). AST 5, which showed the highest inhibitory activity, demonstrated

128

only a moderate potency in activating the expressed Dippu-AstR, ranking 10 out of 13. In

contrast, AST1, which was one of the weakest inhibitors of JH biosynthesis in vitro (ranking 11

out of 13), appears to be one of the most potent peptides in receptor activation. AST6, the

structure of which is conserved in several insect orders, showed the highest potency in activating

expressed Dippu-AstR (Bendena et al., 1999). Its ability to inhibit JH biosynthesis, on the other

hand, was moderate. Different reasons could contribute to the differences between potency in

activating the receptor and inhibiting JH biosynthesis in vitro. The susceptibility of ASTs to

degradation by enzymes in the CA differs between ASTs, which could result in different

potencies in inhibiting JH biosynthesis. AST5, which showed high resistance to the degradation

by hemolymph and membrane preparations, presented the highest activity in inhibiting JH

biosynthesis (Garside et al., 1997a, b). Furthermore, tests with cloned receptors in mammalian

cells might not fully reflect the actual situation in insects. The posttranslational modifications of

AstR or the interacting proteins present in receptor-expressing cells could affect the

conformational and functional properties of this receptor. Nevertheless, we here chose

mammalian cell lines because there may be fewer interacting proteins in this heterologous

system than in insect cell lines.

The inhibition of JH biosynthesis by Dippu-ASTs is a complex process, probably involving more

than one second messenger. Previous studies showed that ASTs regulate JH biosynthesis by

acting through the inositol trisphosphate (IP3)/diglyceride (DAG) second messenger systems, in

which protein kinase C (PKC) and Ca2+ are involved (Feyereisen and Farnsworth, 1987b;

Rachinsky and Tobe, 1996; Rachinsky et al., 1994). However, the role of Ca2+ in regulation of

JH biosynthesis is confounding. JH production shows dose-dependency on extracellular Ca2+ in

the medium, whereas the Ca2+ ionophore A23187 caused a rapid decline in JH release

129

(Kikukawa et al., 1987). Inhibition of JH biosynthesis by brain extracts was antagonized by

inorganic Ca2+channel blockers (Feyereisen and Farnsworth, 1987a). However, elevation of

intracellular Ca2+ levels by treating CA with the Ca2+-mobilizing drug thapsigargin diminished

the inhibitory effect of ASTs (Rachinsky et al., 1994). The activation of Dippu-AstR induced a

dose-dependent intracellular Ca2+ response in CHO-PAM28 cells (Fig. 3.5A), which supports the

involvement of Ca2+ as a second messenger of AST. Aucoin et al (1987) suggested that there

may be both a Ca2+-dependent and a Ca2+-independent pathway (cAMP) involved in the

inhibition of JH biosynthesis. Incubating CA with brain extracts resulted in an increase in the

level of cAMP and a decrease in JH biosynthesis. Forskolin, which causes a dose-dependent

accumulation of cAMP, led to a rapid and dose-dependent inhibition of JH biosynthesis.

Moreover, the sensitivity of CA to forskolin shows the same pattern as its sensitivity to AST

during the first gonadotrophic cycle (Feyereisen and Farnsworth, 1987a; Meller et al., 1985). In

contrast, another study suggest that levels of cAMP do not increase following treatment of CA

with ASTs. Our findings demonstrated that the activation of Dippu-AstR induced a dose-

dependent cAMP response in HEK293 cells, which suggests that cAMP may be involved in the

signaling pathway of AST cells (Fig. 3.5B). Dippu-AstR appears to dually couple through the

Gαs and Gαq pathways and triggers the release of intracellular Ca2+ and the increase in cAMP

levels. Studies on Drosophila AstRs suggested that DAR-1 and DAR-2 couple to Gi/o mediated

signaling and other G-proteins (Birgul et al., 1999; Larsen et al., 2001). However, activation of

AstR in the presence of forskolin did not significantly influence the cAMP signal. This result

suggests that there is no involvement of Gαi in the Dippu-AstR pathway.

AST regulates JH biosynthesis by affecting either enzymes directly involved in the JH

biosynthetic pathway or steps prior to the production of acetyl-CoA. We have studied the effect

130

of AST on the transcript levels of genes in the JH biosynthetic pathway. The inhibitory effect of

AST on JH biosynthesis was rapid and reversible, which suggests that AST may not regulate JH

biosynthesis through transcript levels ((Pratt et al., 1991; Pratt et al., 1989). Our study confirmed

this earlier assumption by revealing that the transcript levels of genes in the JH biosynthetic

pathway were not affected by AstR silencing or AST treatment in vitro (Fig. 3.6 and 3.7).

Previous studies indicate that ASTs regulate JH biosynthesis through steps(s) prior to the

production of mevalonate (Pratt et al., 1991; Pratt et al., 1989). However, the enzyme activities

of HMG-CoA synthase or HMG-CoA reductase were not affected by AST (Sutherland and

Feyereisen, 1996). To determine which enzyme in the JH biosynthetic pathway was influenced

by AST, we used several JH precursors to reverse the AST-induced inhibition of JH biosynthesis.

All the JH precursors (including acetyl-CoA) were able to partially restore JH biosynthesis,

thereby suggesting that the enzyme activities were not affected by AST. The target of AST

action probably lies prior to the start of the JH biosynthetic pathway in regulating the production

of precursors of acetyl-CoA such as the transport of citrate from mitochondria to cytosol and/or

the cleavage of citrate to yield acetyl-CoA (Sutherland and Feyereisen, 1996). Further

investigation is needed to identify the exact target of AST.

In D. punctata, ASTs play an important role in the regulation of JH biosynthesis. The potency of

the 13 ASTs in activating AstR reveals the structure-activity relationship of AST action. The

activation of AstR in CHO and HEK cells suggests that both Ca2+ and cAMP can be involved in

the signal transduction of AST. Our results show that AST does not affect the transcript levels

and that the activities of enzymes in the JH biosynthetic pathway remain normal following

treatment with AST. AST probably affects JH biosynthesis prior to the entry of precursors into

the JH biosynthetic pathway. The exact target of AST action remains an open question.

131

Supplementary data

Table S3.1 Oligonucleotide sequences for primers used in q-RT-PCR for reference and target genes. Efficiencies and R² values

are indicated.

Reference genes F-primer R-primer Efficiency (%) R2

Tubulin 5'-AAATTACCAACGCTTGCTTTGAA-3' 5'-TGGCGAGGATCGCATTTT-3' 95.1 0.993

EF1α 5'-TCGTCTTCCTCTGCAGGATGTCT-3' 5'-GGGTGCAAATGTCACAACCATACC-3' 99.2 0.994

Armadillo 5'-GCTACTGCACCACTCACAGAATTATT-3' 5'-CTGCAGCATACGTTGCAACA-3' 94.5 0.980

Target genes F-primer R-primer Efficiency (%) R2

AstR 5’-GCCATTTGGAGAAATCTGGT-3’ 5’-GAACCGACATGGAAGTGATG-3’ 95.0 0.974

AST 5’-GGCAAAAGAGCAAGACCTTACAG-3’ 5’-CCTCCTCGCTTGCCAAGTC-3’ 95.6 0.997

Thiol 5’-TGCCTTCCAAAAGGAGAATG-3’ 5’- ACATCACCTGCCATCAACAC-3’ 90 0.970

HMGS 5’- TGCTGGGAAGTACACAGCAGG-3’ 5’- CTCCACGAGCTTGCTGACTG-3’ 83 0.994

HMGR 5’-TGGGAGCATGTTGTGAAAAT-3’ 5’-ACCAAGCAGCCCTCAGTAGT-3’ 95 0.988

MK 5’- TACGGCAAAACTGCCCTTGC-3’ 5’- AATGGAGGAGGTTCGGCGA-3’ 93 0.996

PMK 5’- TACGAAAACAACGAGGATGG-3’ 5’- TTCTGCATCATCTACACCTTCA-3’ 100 0.985

PP-MevD 5’- TGGAAGGTGACATAACAGCAA-3’ 5’- ATCCTTGATGCCAGTGAACA-3’ 90 0.967

132

Table S3.1 (Continue) Oligonucleotide sequences for primers used in q-RT-PCR for reference and target genes. Efficiencies

and R² values are indicated.

Target genes F-primer R-primer Efficiency (%) R2

IPPI 5’- CCTTCCCCAACCATGTAACT-3’ 5’- ACCAACGCCATTTGTCTCTT-3’ 100 0.992

FPPS 5’-TGCTTTGGAGATCCTGAGGT-3’ 5’- TGTTCAGGAGTGGTTCGTTG-3’ 96 0.987

FOLD 5’- TGGCGCGTAGGGTAGACAA-3’ 5’- GACCCATTTGAAAGCCTCCTTGA-3’ 93 0.993

JHAMT 5’- ATCCAGGTGCTGGAAGGAGAG-3’ 5’- CTGCCCAGAGTCGAACAGG-3’ 99 0.984

CYP15 5’- GTTGGGATCTCGGAGCATGG-3’ 5’- CGAACACGTCATGCATCGGT-3’ 100 0.992

Table S3.2. Primers for dsRNA construction

Name F Primer(5'-3') R Primer(5'-3')

AstR dsRNA TGTAATCATACGGCTAACGGATC

AATGGTAGAACGTAGTCTGTTGC

AST dsRNA GTCTACCGTCGGCTCTTGTA

TGCTTCTCTTGCCGAGACC

Control dsRNA TTGCGCTCACTGCCAATTGC CTGGCCTTTTGCTCACATGTT

*The T7 promoter sequence was added at the 5' end of each dsRNA primer.

133

B r N C C A F b O v T e AG M G M T

0

5

1 0

1 5

Re

lati

ve

mR

NA

qu

an

tity

D ip p u -A s tR

D ip p u -A S T

Figure S3.1. Relative transcript levels of Dippu-AstR (black) and Dippu-AST (gray) mRNA

in tissues of day 7 male and mated female D. punctata. Dippu-AstR/AST mRNA were

quantified by q-RT-PCR. The data represent the average of 3 biologically independent pools (10

animals per pool), run in triplicate. Tissue abbreviations are as described in Figure 2.1. Values

represent mean ± SEM.

134

References

Aucoin, R.R., Rankin, S.M., Stay, B., Tobe, S.S., 1987. Calcium and cyclic-AMP involvement in the

regulation of juvenile hormone biosynthesis in Diploptera Punctata. Insect Biochem 17, 965-969.

Auerswald, L., Birgul, N., Gade, G., Kreienkamp, H.J., Richter, D., 2001. Structural, functional, and

evolutionary characterization of novel members of the allatostatin receptor family from insects. Biochem

Biophys Res Commun 282, 904-909.

Belles, X., Martin, D., Piulachs, M.D., 2005. The mevalonate pathway and the synthesis of juvenile hormone

in insects. Annu Rev Entomol 50, 181-199.

Bendena, W.G., Donly, B.C., Tobe, S.S., 1999. Allatostatins: a growing family of neuropeptides with

structural and functional diversity. Annals of the New York Academy of Sciences 897, 311-329.

Birgul, N., Weise, C., Kreienkamp, H.J., Richter, D., 1999. Reverse physiology in Drosophila: identification

of a novel allatostatin-like neuropeptide and its cognate receptor structurally related to the mammalian

somatostatin/galanin/opioid receptor family. Embo J 18, 5892-5900.

Bowser, P.R.F., Tobe, S.S., 2000. Partial characterization of a putative allatostatin receptor in the midgut of the

cockroach Diploptera punctata. Gen Comp Endocrinol 119, 1-10.

Coast, G.M., Schooley, D.A., 2011. Toward a consensus nomenclature for insect neuropeptides and peptide

hormones. Peptides 32, 620-631.

Cusson, M., Prestwich, G.D., Stay, B., Tobe, S.S., 1991. Photoaffinity-Labeling of Allatostatin Receptor

Proteins in the Corpora Allata of the Cockroach, Diploptera-Punctata. Biochem Biophys Res Commun 181,

736-742.

Cusson, M., Prestwich, G.D., Stay, B., Tobe, S.S., 1992. Evidence for allatostatin receptor proteins in the

corpora allata of the cockroach, Diploptera punctata, as demonstrated by photoaffinity labeling. Insect

Juvenile Hormone Res, 183-191.

Donly, B.C., Ding, Q., Tobe, S.S., Bendena, W.G., 1993. Molecular cloning of the gene for the allatostatin

family of neuropeptides from the cockroach Diploptera punctata. Proc Natl Acad Sci U S A 90, 12055-12055.

Duve, H., Wren, P., Thorpe, A., 1995. Innervation of the foregut of the cockroach Leucophaea maderae and

inhibition of spontaneous contractile activity by callatostatin neuropeptides. Physiol Entomol 20, 33-44.

Feyereisen, R., Farnsworth, D.E., 1987a. Comparison of the Inhibitory Effects of Brain Extract, High K+ and

Forskolin on Juvenile-Hormone Synthesis by Diploptera-Punctata Corpora Allata. Insect Biochem 17, 939-942.

Feyereisen, R., Farnsworth, D.E., 1987b. Inhibition of Insect Juvenile-Hormone Synthesis by Phorbol 12-

Myristate 13-Acetate. Febs Lett 222, 345-348.

Feyereisen, R., Tobe, S.S., 1981. A Rapid Partition Assay for Routine Analysis of Juvenile-Hormone Release

by Insect Corpora Allata. Anal Biochem 111, 372-375.

Fuse, M., Zhang, J.R., Partridge, E., Nachman, R.J., Orchard, I., Bendena, W.G., Tobe, S.S., 1999. Effects of

an allatostatin and a myosuppressin on midgut carbohydrate enzyme activity in the cockroach Diploptera

punctata. Peptides 20, 1285-1293.

Gade, G., Marco, H.G., Richter, D., Weaver, R.J., 2008. Structure-activity studies with endogenous

allatostatins from Periplaneta americana: Expressed receptor compared with functional bioassay. J Insect

Physiol 54, 1557-1557.

Garside, C.S., Hayes, T.K., Tobe, S.S., 1997a. Degradation of Dip-allatostatins by hemolymph from the

cockroach, Diploptera punctata. Peptides 18, 17-25.

Garside, C.S., Hayes, T.K., Tobe, S.S., 1997b. Inactivation of Dip-allatostatin 5 by membrane preparations

from the cockroach Diploptera punctata. Gen Comp Endocrinol 108, 258-270.

135

Kikukawa, S., Tobe, S.S., Solowiej, S., Rankin, S.M., Stay, B., 1987. Calcium as a regulator of juvenile

hormone biosynthesis and release in the cockroach Diploptera punctata. Insect Biochem 17, 179-187.

Lange, A.B., Bendena, W.G., Tobe, S.S., 1995. The effect of the 13 Dip-allatostatins on myogenic and induced

contractions of the cockroach (Diploptera punctata) hindgut. J Insect Physiol 41, 581-588.

Larsen, M.J., Burton, K.J., Zantello, M.R., Smith, V.G., Lowery, D.L., Kubiak, T.M., 2001. Type A

allatostatins from Drosophila melanogaster and Diplotera puncata activate two Drosophila allatostatin

receptors, DAR-1 and DAR-2, expressed in CHO cells. Biochem Biophys Res Commun 286, 895-901.

Lenz, C., Williamson, M., Grimmelikhuijzen, C.J.P., 2000. Molecular cloning and genomic organization of a

second probable allatostatin receptor from Drosophila melanogaster. Biochem Biophys Res Commun 273,

571-577.

Lloyd, G.T., Woodhead, A.P., Stay, B., 2000. Release of neurosecretory granules within the corpus allatum in

relation to the regulation of juvenile hormone synthesis in Diploptera punctata. Insect Biochem Mol Biol 30,

739-746.

Lungchukiet, P., Donly, B.C., Zhang, J.R., Tobe, S.S., Bendena, W.G., 2008a. Molecular cloning and

characterization of an allatostatin-like receptor in the cockroach Diploptera punctata. Peptides 29, 276-285.

Lungchukiet, P., Zhang, J.R., Tobe, S.S., Bendena, W.G., 2008b. Quantification of allatostatin receptor mRNA

levels in the cockroach, Diploptera punctata, using real-time PCR. J Insect Physiol 54, 981-987.

Maestro, J.L., Belles, X., 2006. Silencing allatostatin expression using double-stranded RNA targeted to

preproallatostatin mRNA in the German cockroach. Archives of insect biochemistry and physiology 62, 73-79.

Marchal, E., Hult, E.F., Huang, J., Stay, B., Tobe, S.S., 2013a. Diploptera punctata as a model for studying the

endocrinology of arthropod reproduction and development. Gen Comp Endocrinol 188, 85-93.

Marchal, E., Hult, E.F., Huang, J., Tobe, S.S., 2013b. Sequencing and validation of housekeeping genes for

quantitative real-time PCR during the gonadotrophic cycle of Diploptera punctata. BMC Res Notes 6, 237.

Martin, D., Piulachs, M.D., Belles, X., 1996. Inhibition of vitellogenin production by allatostatin in the

German cockroach. Mol Cell Endocrinol 121, 191-196.

Meller, V.H., Aucoin, R.R., Tobe, S.S., Feyereisen, R., 1985. Evidence for an inhibitory role of cyclic AMP in

the control of juvenile hormone biosynthesis by cockroach corpora allata. Mol Cell Endocrinol 43, 155-163.

Nouzova, M., Edwards, M.J., Mayoral, J.G., Noriega, F.G., 2011. A coordinated expression of biosynthetic

enzymes controls the flux of juvenile hormone precursors in the corpora allata of mosquitoes. Insect Biochem

Mol Biol 41, 660-669.

Pratt, G.E., Farnsworth, D.E., Feyereisen, R., 1990. Changes in the sensitivity of adult cockroach corpora

allata to a brain allatostatin. Mol Cell Endocrinol 70, 185-195.

Pratt, G.E., Farnsworth, D.E., Fok, K.F., Siegel, N.R., McCormack, A.L., Shabanowitz, J., Hunt, D.F.,

Feyereisen, R., 1991. Identity of a second type of allatostatin from cockroach brains: an octadecapeptide amide

with a tyrosine-rich address sequence. Proc Natl Acad Sci U S A 88, 2412-2416.

Pratt, G.E., Farnsworth, D.E., Siegel, N.R., Fok, K.F., Feyereisen, R., 1989. Identification of an Allatostatin

from Adult Diploptera-Punctata. Biochem Biophys Res Commun 163, 1243-1247.

Rachinsky, A., Tobe, S.S., 1996. Role of second messengers in the regulation of juvenile hormone production

in insects, with particular emphasis on calcium and phosphoinositide signaling. Archives of insect

biochemistry and physiology 33, 259-282.

Rachinsky, A., Zhang, J., Tobe, S.S., 1994. Signal transduction in the inhibition of juvenile hormone

biosynthesis by allatostatins: roles of diacylglycerol and calcium. Mol Cell Endocrinol 105, 89-96.

136

Secher, T., Lenz, C., Cazzamali, G., Sorensen, G., Williamson, M., Hansen, G.N., Svane, P.,

Grimmelikhuijzen, C.J.P., 2001. Molecular cloning of a functional allatostatin gut/brain receptor and an

allatostatin preprohormone from the silkworm Bombyx mori. J Biol Chem 276, 47052-47060.

Stables, J., Green, A., Marshall, F., Fraser, N., Knight, E., Sautel, M., Milligan, G., Lee, M., Rees, S., 1997. A

bioluminescent assay for agonist activity at potentially any G-protein-coupled receptor. Anal Biochem 252,

115-126.

Stay, B., Chan, K.K., Woodhead, A.P., 1992. Allatostatin-immunoreactive neurons projecting to the corpora

allata of adult Diploptera punctata. Cell Tissue Res 270, 15-23.

Stay, B., Fairbairn, S., Yu, C.G., 1996. Role of allatostatins in the regulation of juvenile hormone synthesis.

Archives of insect biochemistry and physiology 32, 287-297.

Stay, B., Joshi, S., Woodhead, A.P., 1991. Sensitivity to allatostatins of corpora allata from larval and adult

female Diploptera Punctata. J Insect Physiol 37, 63-70.

Stay, B., Tobe, S.S., 2007. The role of allatostatins in juvenile hormone synthesis in insects and crustaceans.

Annual review of entomology 52, 277-299.

Sutherland, T.D., Feyereisen, R., 1996. Target of cockroach allatostatin in the pathway of juvenile hormone

biosynthesis. Mol Cell Endocrinol 120, 115-123.

Tobe, S.S., Clarke, N., 1985. The Effect of L-Methionine Concentration on Juvenile-Hormone Biosynthesis by

Corpora-Allata of the Cockroach Diploptera-Punctata. Insect Biochem 15, 175-179.

Tobe, S.S., Zhang, J.R., Bowser, P.R., Donly, B.C., Bendena, W.G., 2000. Biological activities of the

allatostatin family of peptides in the cockroach, Diploptera punctata, and potential interactions with receptors.

J Insect Physiol 46, 231-242.

Vandesompele, J., De Preter, K., Pattyn, F., Poppe, B., Van Roy, N., De Paepe, A., Speleman, F., 2002.

Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal

control genes. Genome biology 3, RESEARCH0034.

Vilaplana, L., Maestro, J.L., Piulachs, M.D., Belles, X., 1999. Modulation of cardiac rhythm by allatostatins in

the cockroach Blattella germanica (L.) (Dictyoptera, Blattellidae). J Insect Physiol 45, 1057-1064.

Yu, C.G., Hayes, T.K., Strey, A., Bendena, W.G., Tobe, S.S., 1995. Identification and partial characterization

of receptors for allatostatins in brain and corpora allata of the cockroach Diploptera punctata using a binding

assay and photoaffinity labeling. Regul Pept 57, 347-358.

137

Chapter 4

Identification and characterization of the NMDA receptor and its

role in regulating reproduction in the cockroach, Diploptera

punctata

This chapter is an adapted reprint of my article:

Juan Huang, Ekaterina F. Hult, Elisabeth Marchal, Stephen S. Tobe, Identification and

characterization of the NMDA receptor and its role in regulating reproduction in the cockroach,

Diploptera punctata, 2014, paper submitted to ‘Journal of experimental biology’

Authors’ contribution:

I designed and performed all experiments in this study and wrote the paper with editing by E.

Marchal, E.F. Hult, and S.S. Tobe.

138

Summary

The NMDA receptor (NMDAR) plays important roles in excitatory neurotransmission and in the

regulation of reproduction in mammals. NMDAR in insects comprises two subunits, NR1 and

NR2. In this study, we identified two NR1 paralogs and eleven NR2 alternative splicing variants

in D. punctata. This is the first report of NR1 paralogs in insects. The tissue distribution and

expression profile of DpNR1A, DpNR1B and DpNR2 in different tissues were also investigated.

Previous studies have demonstrated NMDA-stimulated JH biosynthesis in the corpora allata (CA)

through the influx of extracellular Ca2+ in Diploptera punctata. However, our data show that the

transcript levels of DpNR1A, DpNR1B and DpNR2 were low in the CA. In addition, neither

partial knockdown of DpNR2 nor in vivo treatment with a physiologically relevant dose of MK-

801 resulted in any significant change in JH biosynthesis by CA or basal oocyte growth.

Injection of animals with a high dose of MK-801 (30 µg/animal/injection), which paralyzed the

animals for 4-5 h, resulted in a significant decrease in JH biosynthesis on days 4 and 5. However,

the reproductive events during the first gonadotrophic cycle in female D. punctata were

unaffected. Thus, NMDAR does not appear to play important roles in the regulation of JH

biosynthesis or mediate reproduction of female D. punctata.

Introduction

L-glutamate (Glu), a major excitatory amino acid transmitter, mediates diverse physiological

functions in the vertebrate nervous system (Mahesh and Brann, 2005). The Glu receptors (GluR)

have been classified into three major subtypes: the α-amino-3-hydroxy-5-methyl-4-

isoxazolepropionic acid (AMPA) receptor, N-methyl-D-aspartate (NMDA) receptor and kainate

receptor (Madden, 2002). The NMDA receptors (NMDAR) are distinguished from other

139

ionotropic receptors by their unique properties, including selective agonists and antagonists, high

Ca2+ permeability, and voltage-dependent Mg2+ blockade (McBain and Mayer, 1994). The

unique properties of NMDAR allow it to play key roles in excitatory neurotransmission and

important neurological processes, including learning, memory and behavior (Mussig et al., 2010;

Newcomer and Krystal, 2001; Xia et al., 2005).

NMDAR in vertebrates is composed of two subunits, NR1 and NR2, and in some cases NR3

subunits (Madden, 2002). The NR1 subunit is essential for the basic channel activity of NMDAR,

whereas the NR2 subunit contributes to enhance and modulate the receptor function (Sydow et

al., 1996). Although much is known about the function of NMDAR in vertebrates, little

information is available on NMDARs in insects. Thus far, two subunits, NR1 and NR2, have

been identified in insects. The NR1 and NR2 subunits were previously reported to be distributed

throughout the brain of Drosophila melanogaster and Apis mellifera (Wu et al., 2007; Xia et al.,

2005; Zannat et al., 2006).

The importance of NMDAR in the regulation of reproduction of mammals is well-known.

NMDAR mediates reproduction through the regulation of pulse and surge gonadotropin-

releasing hormone (GnRH)/luteinizing hormone (LH) secretion (Maffucci et al., 2009; Mahesh

and Brann, 2005). In insects, juvenile hormones (JHs) are key regulators of growth, development,

metamorphosis, aging, caste differentiation and reproduction (Goodman and Granger, 2005;

Hartfelder, 2000). As a result of the importance of JH in physiological processes, its biosynthesis

is tightly regulated by many factors, including neuropeptides (allatostatins, allatotropins) (Stay

and Tobe, 2007) and neurotransmitters (octopamine, dopamine and glutamate) (Granger et al.,

1996; Pszczolkowski et al., 1999; Thompson et al., 1990). Chiang et al (2002) demonstrated that

JH biosynthesis by corpora allata (CA) of the cockroach, Diploptera punctata, is stimulated by

140

an NMDA-induced influx of Ca2+ ions and this elevation was significantly reduced by NMDAR

antagonists including Mg2+, MK-801 or conantokin T. Furthermore, a Drosophila larval mutant

for NMDAR1 showed reduced mRNA levels of the gene encoding JH acid methyltransferase

(JHAMT), a key regulatory enzyme of JH biosynthesis in this species (Huang et al., 2011).

These studies suggest that NMDAR plays a role in the regulation of JH biosynthesis.

As a high-affinity antagonist of NMDARs, MK-801 has been used to study the function of

NMDAR in both vertebrates and invertebrates (Rawls et al., 2009; Sircar et al., 1987; Troncoso

and Maldonado, 2002). In addition to the inhibitory effect of MK-801 in NMDA-stimulated JH

biosynthesis in D. punctata, MK-801 was found to influence ovarian development and

vitellogenesis in the flesh fly Neobellieria bullata and the locust Schistocerca gregaria (Begum

et al., 2004; Chiang et al., 2002). In S. gregaria, the inhibition of vitellogenesis was overcome by

treatment with JH. A later study on the butterfly Bicyclus anynana and the cricket Gryllus

bimaculatus showed that MK-801 affects JH biosynthesis in vitro and JH titres in both species,

and subsequently regulates insect reproduction (Geister et al., 2008).

D. punctata is a well-known model in studying the physiology of JH biosynthesis and regulation;

in this animal, JH biosynthesis is high and stable, and the reproductive events correlate very well

with rates of JH production (see review by Marchal et al. (2013a)). In this study, we chose D.

punctata as our model to determine the role of NMDAR in the regulation of JH biosynthesis and

reproduction. We identified the genes encoding the subunits of NMDAR in D. punctata, and

examined the expression of NMDAR in several tissues. In addition, we investigated the roles of

NMDAR in the regulation of JH biosynthesis, vitellogenesis and oocyte growth in vivo using

RNA interference (RNAi) and MK-801 treatments.

141

Materials and Methods

Insects - D. punctata were reared in cages and fed with lab chow and water at libitum at 27 -28

ºC in a dark room. Newly molted male and female adult cockroaches were picked from the

colony and raised in separate containers. Mated status in the females was confirmed by the

presence of a spermatophore.

Tissue collection - Cockroach tissues were dissected under a dissecting microscope. Basal oocyte

length was measured to determine the physiological age of female cockroaches. Selected tissues

were dissected and cleaned in sterile cockroach ringer solution (150 mM NaCl, 12 mM KCl, 10

mM CaCl2.2H2O, 3 mM MgCl2.6H2O, 10 mM HEPES, 40 mM Glucose, pH 7.2), flash-frozen in

liquid nitrogen to prevent RNA degradation and stored at -80 ºC until further processing.

RNA extraction and cDNA synthesis - Selected tissues were collected from adult females and

male for gene sequence, tissue distribution and developmental profiling. RNA extraction and

cDNA synthesis were performed as described by Marchal et al. (2013b). For the tissue

distribution and developmental profiling, three biologically independent pools of 10 animals

each were collected. For RNAi and MK-801 treatment experiments, 3 biologically independent

pools of brain and fat body were collected, each pool containing tissue from 5 animals.

Sequencing of DpNR1A, DpNR1B and DpNR2 - Degenerate primer sequences were designed for

DpNR1A, DpNR1B and DpNR2 based on conserved amino acid sequences of several insect

orthologs. Primers used for degenerate PCR are listed in Table S4.1. Partial sequences were

obtained using these primers in a standard T-gradient PCR using Taq DNA polymerase (Sigma-

Aldrich) and a D. punctata brain cDNA sample. After purification, the resulting DNA fragments

were subcloned into a pJET 1.2/blunt cloning vector (CloneJet PCR Cloning Kit, Thermo

142

Scientific) and sequenced following the protocol outlined in the ABI PRISM BigDye Terminator

Ready Reaction Cycle Sequencing Kit (Applied Biosystems). The complete sequence of

DpNR1A, DpNR1B and DpNR2 was obtained using 5’-RACE (Rapid Amplification of cDNA

Ends) and 3’-RACE strategies, following the protocol outlined in the Roche 5'/3' RACE Kit.

Primers used for RACE are listed in Table S4.1.

Phylogenetic analysis - For the phylogenetic analysis of DpNR1 genes, the NR1 sequences of 15

insect species (identified or predicted) were used. These sequences were aligned using ClustalW

as implemented within MEGA 6.06 (Tamura et al., 2013). Poorly aligned positions and gaps

were removed, which resulted in 854 amino acid residues. The obtained alignment was used to

construct a phylogenetic tree in PhyML 3.0 (Guindon and Gascuel, 2003) based on the

maximum-likelihood principle, using the WAG substitution model (Whelan and Goldman, 2001).

Four substitution rate categories were used to estimate the gamma parameter shape with 100

bootstrap replicates to assess branch support (Felsenstein, 1985; Yang, 1994). The resulting tree

was then rooted using the sea slug, Aplysia californica as an outgroup.

Quantitative Real Time-PCR (q-RT-PCR) - Primers used for q-RT-PCR are shown in Table S4.2.

Primer sets were validated by determining relative standard curves for each gene transcript using

a five-fold serial dilution of a calibrator cDNA sample. Efficiency and correlation coefficient (R²)

can be found in Table S2. Reactions were performed in triplicate on a CFX384 Touch ™ Real-

Time PCR Detection System (Bio-Rad) as described previously by Marchal et al. (2013b).

Target specificity was confirmed by running a few representative q-RT-PCR products on an

agarose gel containing GelRed™ (Biotium). The optimal housekeeping genes for target gene

profiling and RNAi experiments were chosen according to a previous study (Marchal et al.,

143

2013b). The quantity of mRNA for each tested gene relative to reference genes was determined

as described by Vandesompele et al. (2002).

RNA interference (RNAi) - Primers for DpNR2 and control pJET dsRNA constructs with T7

promoters are shown in Table S4.3. Double-stranded RNA (dsRNA) constructs were prepared

using the MEGAscript® RNAi Kit (Ambion). A PCR using forward and reverse primers with

attached T7 promoters was performed to amplify the fragment, which was subcloned and

sequenced to verify the presence of the T7 promoter. The amplified fragment was used in an

RNA transcription reaction which was incubated overnight to obtain a high yield of annealed

dsRNA construct. A nuclease digestion was subsequently performed to remove ssRNA and DNA

remaining in the product. The dsRNA was further purified according to the manufacturer’s

instructions (Ambion). Concentration of the dsRNA construct was determined using a Nanodrop

instrument (Thermo Fisher Scientific Inc., Canada). Five-fold diluted dsRNA was run on a 1.2%

agarose gel to examine the quality and integrity of the construct.

Newly molted adult female cockroaches (day 0) were injected with 2 µg of either DpNR2

dsRNA or pJET dsRNA diluted in 5 µl of cockroach saline. This treatment was repeated on days

1, 2 and 3, and the effect of DpNR2 dsRNA was determined on day 4. Brains were dissected and

stored in liquid nitrogen prior to RNA extraction. CA were dissected and cleaned in TC199

medium (GIBCO; 1.3 mM Ca2+, 2% Ficoll, methionine-free) for use in the radiochemical assay

(RCA) (see below). Basal oocyte length was measured during dissection.

MK-801 in vivo assay - MK-801 was dissolved in ddH2O to a concentration of 6 µg/µl and

injected into animals using a Hamilton syringe. Newly molted adult females were injected with

30 µg of MK-801 on days 0, 1 and 3, and the effect of MK-801 was determined from day 4 to

day 8. Fat body was dissected from day 4 animals and stored in liquid nitrogen prior to RNA

144

extraction. CA were dissected and cleaned in TC199 medium (GIBCO; 1.3 mM Ca2+, 2% Ficoll,

methionine-free) for use in the RCA (see below). Basal oocyte length was also measured.

Results

Identification of DpNR1 subunits

Two DpNR1 were identified with open reading frames of 2871bp (DpNR1A) and 2703bp

(DpNR1B). This was accomplished employing a degenerate PCR approach with adult brain

cDNA. 5’-RACE and 3’-RACE experiments were performed to complete the sequences. The two

DpNR1 sequences were deposited in the NCBI GenBank and received accession numbers:

KJ747198 and KJ747199. Unlike the alternative splicing variants of NR1 genes in other insects,

the differences between the two variants in D. punctata are distributed throughout the whole

gene (Fig. 4.1) and may be the result of gene duplication. A phylogenetic tree of NR1 was

constructed by maximum likelihood methods (Fig. 4.2). Generally, the sequences of NR1 were

grouped based on insect orders, and the two DpNR1s (Blattodea) cluster together with

confidence. On the other hand, the relationship between orders is not well-resolved as indicated

by low bootstrap values at deeper nodes. The lack of power to resolve these nodes could be

explained by insufficient sampling among hemimetabolous insects.

145

Fig. 4.1. Amino acid sequence alignment of the two Diploptera NR1 subunit (DpNR1A, DpNR1B), and homologous receptors

from D. melanogaster (DNR1, GenBank acc. no. NP_730940.1) and T. castaneum (TNR1, GenBank acc. no. XP_969654.1,

predicted sequence). Conservatively substituted residues are highlighted in yellow, and the different residues between DpNR1A and

DpNR1B in green. Three putative hydrophobic transmembrane regions (TM1, TM3, and TM4) and one hydrophobic pore-forming

segment (TM2) are highlighted in the boxes. Agonist-binding domain (S1 and S2 domains) are underlined.

146

Fig. 4.2. Phylogram depicting the relationship between the NR1 subunits from Diploptera

and orthologues of this receptor from other insects. Phylogenetic analysis was conducted in

PhyML 3.0 using WAG substitution model with 100 bootstrap replicates. Poorly aligned amino

acid were eliminated by eye resulting in 854 positions. The bar represents 0.2 substitutions per

site. The sea slug, Aplysia californica was used as outgroup to root the tree.

147

Identification of DpNR2 subunits

DpNR2 undergoes alternative splicing, mostly at the 3’ end, generating eleven different

transcripts, which may encode nine different proteins. Full-length cDNAs for all eleven variants

have been isolated and their sequences were deposited in the NCBI GenBank with accession

numbers: KJ747200 to KJ747210. The deduced amino acid sequence and nucleotide sequence of

DpNR2 subunits was shown in Supplementary Fig. S4.1. In the eleven transcripts, there are two

5’ untranslated region, two insertions, and four different 3’ ends (Fig. 4.3A). The sequence of

insertions, deletions and different 3’-ends are shown in Fig. 4.3B. All eleven DpNR2 variants

contain the domain structure, four hydrophobic regions (TM1-TM4) and two ligand binding

domains. Amino acid sequence comparisons between DpNR2A-1 and NR2 subunits from D.

melanogaster, A. mellifera, and A. aegypti were determined (Table S4.4). DpNR2A-1 shares the

highest similarity with NR2 subunits from A. mellifera.

Expression of DpNR1A, DpNR1B and DpNR2

Previous studies of NMDAR in insects have focused on its function in brain (Wu et al., 2007;

Xia et al., 2005; Zannat et al., 2006). In the present study, we were interested in assessing the

role of NMDAR in other tissues. We therefore determined the localization and relative

abundance of DpNR1A, DpNR1B and DpNR2 mRNA in day 4 adult male and female

cockroaches using q-RT-PCR (Fig. 4.4). Specific q-RT-PCR primers for each DpNR1 variant

were designed to determine whether transcripts for the different DpNR1 paralogs are present in

the same tissue. Q-RT-PCR primers for DpNR2 are located in the conserved region of DpNR2

common to all putative splice variants. In mated female cockroaches, DpNR1A, DpNR1B and

DpNR2 were highly expressed in brain, followed by nerve cord. Previous study has shown that

NMDA stimulated JH biosynthesis in the CA of Diploptera (Chiang et al., 2002). However, the

148

transcript level of DpNR1A, DpNR1B and DpNR2 in the CA was relatively low compared to

other tissues. In male cockroaches, the highest transcript levels of DpNR1A, DpNR1B and

DpNR2 were measured in the brain (Fig. 4.4). It is interesting to see that the transcript level of

DpNR2 in testes was very low whereas both genes encoding NR1 were high.

To further study the function of NMDAR in D. punctata, we determined the developmental

profile of DpNR1A, DpNR1B and DpNR2 in the brain, CA and testes. DpNR1A and DpNR1B

were stably transcribed in brain of day 0 to day 7 post-emergence mated female D. punctata.

DpNR2 mRNA in the brain showed a slight increase on days 2, 3 and 7, but none of the changes

were significant (Fig. 4.5). In the CA, the transcript level of DpNR1B and DpNR2 remained very

low and did not show any significant change during the first gonadotrophic cycle (Fig. 6).

DpNR1A mRNA levels exhibited a significant increase on day 6. In male cockroaches, we

studied the expression of NMDAR in testes of males of differing ages. The transcript level of

DpNR2 was very low throughout all the ages tested (Fig. 4.7).

Effect of DpNR2 dsRNA on JH biosynthesis and oocyte growth

To investigate the role of NMDAR in reproduction, DpNR2 was silenced using RNAi. The

injection of 2 µg of DpNR2 dsRNA on days 0, 1, 2 and 3 resulted in a 48.7% knockdown of

DpNR2 mRNA levels in brains of day 4 females (Fig. 4.8a). Knockdown of DpNR2 also resulted

in a decrease in the transcript levels of DpNR1A and DpNR1B. However, the JH biosynthetic

activity of CA and basal oocyte lengths in dsRNA-treated animals did not show any significant

difference from control animals (Fig. 4.8b, 4.8c). A similar result was also observed in DpNR1B

dsRNA- treated animals (Fig. S4.2).

149

Fig. 4.3. Molecular characterization of DpNR2. (A) The schematic structures of the 11

variants of the NR2. NR2 variants are generated by alternative splicing, including two 5’

untranslated region, two insertions, and four different 3’ ends. The 5’ end untranslated region is

shown with black or dash line. The position of insertion 1 is shown by the black arrow and

insertion 2 by a gray arrow. The alternated carboxyl-terminal sequences are indicated by colored

boxes: A1 (Green), A2 (Red), B (Gray), and C (Purple). Four putative transmembrane segments

(TM I – IV) are shown by bold black lines. The agonist-binding domains S1 and S2 are

indicated by the hatched boxes. The accession number is displayed under the gene number. (B)

The sequence of insertion, deletion and the alternative created carboxyl-terminal.

150

B r N C C A F b M G M T O v B r T e A G

0

1

2

3

Re

lati

ve

mR

NA

qu

an

tity

of

NR

1A

B r N C C A F b M G M T O v B r T e A G

0 .0

0 .5

1 .0

1 .5

2 .0

2 .5

Re

lati

ve

mR

NA

qu

an

tity

of

NR

1B

B r N C C A F b M G M T O v B r T e A G

0

1

2

3

4

5

Re

lati

ve

mR

NA

qu

an

tity

of

NR

2

A

B

C

F em a le M ale

F em a le M ale

F em a le M ale

Fig. 4.4. Graphic representation of the relative tissue distribution of (A) DpNR1A transcript

levels, (B) DpNR1B transcript levels and (C) DpNR2 transcript levels in tissues of day 4

adult male and mated female D. punctata. Relative mRNA quantity was normalized against

levels of Tubulin and EF1α mRNA (Marchal et al., 2013b). The data represents an average of 3

pools (10 animals per pool), run in triplicate. Abbreviations used: Br brain, NC nerve cord, CA

corpora allata, Fb fat body, Ov ovary, MG midgut, MT Malpighian tubules, AG accessory gland

and Te testes. Values represent mean ± SEM.

151

D 0 D 1 D 2 D 3 D 4 D 5 D 6 D 7

0

5

1 0

1 5

Re

lati

ve

mR

NA

qu

an

tity N R 1A

N R 1B

N R 2

Fig. 4.5. Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in brains of mated

female D. punctata from day 0-day 7 after ecdysis. mRNA levels were normalized against

levels of Tubulin and EF1a mRNA (Marchal et al., 2013b). The data represent the average of 3

biologically independent pools (10 animals per pool), run in triplicate. Values represent mean ±

SEM.

D 0 D 1 D 2 D 3 D 4 D 5 D 6 D 7

0 .0

0 .1

0 .2

0 .3

0 .4

0 .5

Re

lati

ve

mR

NA

qu

an

tity N R 1A

N R 1B

N R 2

Fig. 4.6. Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in CA of of mated

female D. punctata from day 0-day 7 after ecdysis. mRNA levels were normalized against

levels of Armadillo and EF1a mRNA (Marchal et al., 2013b). The data represent the average of

3 biologically independent pools (10 animals per pool), run in triplicate. Values represent mean ±

SEM.

152

D 0 D 2 D 4 D 6 D 8 D 1 0 D 1 5 D 2 5

0

1

2

3

4

Re

lati

ve

mR

NA

qu

an

tity

N R 1A

N R 1B

N R 2

Fig. 4.7. Relative transcript levels of DpNR1A, DpNR1B and DpNR2 in testes of different

ages of male D. punctata. mRNA levels were normalized against levels of Tubulin and EF1a

mRNA (Marchal et al., 2013b). The data represent the average of 3 biologically independent

pools (10 animals per pool), run in triplicate. Values represent mean ± SEM.

N R 1 A N R 1 B N R 2

0 .0

0 .2

0 .4

0 .6

0 .8

1 .0

Re

lati

ve

mR

NA

qu

an

tity

*

c o n tro l

N R 2 dsR N A

c o n tro l N R 2 d s R N A

0

1 0

2 0

3 0

4 0

JH

bio

sy

nth

es

is(p

mo

l/h

/CA

)

c o n tro l N R 2 d s R N A

0 .0

0 .5

1 .0

1 .5

Oo

cy

te l

en

gth

(m

m)

AB C

Fig. 4.8. The effect of DpNR2 dsRNA treatment on JH biosynthesis and basal oocyte growth,

and the interactions among these genes in mated female D. punctata. (A) Relative quantity of

DippuNR1A, DippuNR1B and DippuNR2 mRNA levels in brain between control and dsRNA

treated animals. mRNA levels were normalized against levels of Tubulin and EF1a mRNA

(Marchal et al., 2013b). The data represent the average of 3 biologically independent pools (5

animals each pool), run in triplicate. (B) JH biosynthesis by CA from control and dsRNA-treated

animals. (C) Basal oocyte length in control and dsRNA-treated animals. Values represent mean ±

SEM. Levels of significance to the control are indicated with the asterisk symbol: *P < 0.05

153

In vivo effect of NMDAR antagonist MK-801 on JH biosynthesis and ovarian development

The effect of NMDAR on JH biosynthesis and ovarian development was also determined by

injecting the NMDAR non-competitive antagonist MK-801 into the animals. To determine the

optimal dose of injection, newly molted adult females were injected with 0 (control), 3, 12 and

30 µg MK-801 on days 0, 1 and 3. JH biosynthesis and basal oocyte length were determined on

day 4. As shown in Fig. S4.3, there was no significant effect on JH biosynthesis in animals

treated with 3 µg or 12 µg MK-801.

To further study the effect of MK-801 on reproduction, mated female cockroaches were injected

with 30 µg MK-801 on days 0, 1, and 3 following ecdysis. Rates of JH biosynthesis and basal

oocyte lengths in control and treated animals were determined from day 4 to day 8. As shown in

Fig. 4.9A, application of MK-801 resulted in a 34% and 26% decrease of JH biosynthesis in day

4, and 5 animals, respectively. On day 6, however, MK-801 treated animals showed higher rates

of JH biosynthesis than control animals. On days 7 and 8, there was no significant difference in

JH biosynthesis between control and MK-801 treated animals. Furthermore, MK-801 did not

have any effect on basal oocyte length (Fig. 4.9B). All animals oviposited on day 8. To study the

effect of MK-801 on the Vg synthesis, we determined the transcript level of DpVg in the fat body

of day 4 animal. No significant change was found in the transcription of DpVg in MK-801-

treated animals (Fig. 4.9C).

154

Figure 4.9. In vivo effect of MK-801 on JH biosynthesis (A), basal oocyte growth (B) and relative Vg mRNA levels (C). Females

were injected with MK-801 on days 0, 1 and 3 following ecdysis (30µg/animal). Control was injected with ddH2O. JH biosynthesis (A)

and oocyte length (B) were determined on day 4 to day 8. All animals oviposited on day 8. The mRNA level of Vg was determined in

the fat body of day 4 animals. mRNA levels were normalized to levels of Tubulin and EF1a mRNA (Marchal et al., 2013b). The data

represent the average of 3 biologically independent pools (5 animals each pool), run in triplicate. Values represent mean ± SEM. (A)

N≥15; (B) N≥10. Levels of significance to the control are indicated with an asterisk: *P < 0.05, **P < 0.01, ***P < 0.001

d a y 4 d a y 5 d a y 6 d a y 7 d a y 8

0

1 0

2 0

3 0

4 0

5 0

6 0

7 0

A G E

JH

bio

sy

nth

es

is(p

mo

l/h

/CA

)

c o n tro l

M K -8 0 1

***

**

*

d a y 4 d a y 5 d a y 6 d a y 7

1 .2

1 .4

1 .6

1 .8

2 .0

2 .2

A G E

Ba

sa

l o

oc

yte

le

ng

th (

mm

)

c o n tro l

M K -8 0 1

c o n tro l M K -8 0 1

0 .0

0 .5

1 .0

1 .5

Re

lati

ve

mR

NA

qu

an

tity

of

VG

A B C

155

Discussion

We have identified two distinct NR1 genes and eleven NR2 variants in D. punctata. The major

functional domains appear to be well conserved in both DpNR1 and DpNR2 amino acid

sequences (Fig. 4.1 and 4.3). The protein contains three hydrophobic transmembrane regions

(TM1, 3-4), a hydrophobic pore-forming segment (TM2), and two ligand binding domains (S1

and S2) (Fig. 4.1 and 4.3) (Dingledine et al., 1999; Kuryatov et al., 1994; Stern-Bach et al.,

1994). The asparagine residue, which was predicted to control Ca2+ permeability and voltage-

dependent Mg2+ blockade, is present in the TM2 domain of DpNR1 (N630 in DpNR1A and

N608 in DpNR1B) (Burnashev et al., 1992). The overall amino acid sequence identity between

DpNR1, DpNR2 and NMDAR subunits in other species is shown in Table S4.4. DpNR1A has

higher sequence identity to NR1 in other species than DpNR1B. Both DpNR1 and DpNR2

subunits show highest homology to NMDAR of A. mellifera.

The two NR1 subunits are 65.6% identical at the amino acid level and the differences are

distributed throughout the entire gene, which suggests that the two NR1 paralogs result from

gene duplication (Fig. 4.1). The other NR1 paralogs were isolated from the Zebrafish (Cox et al.,

2005). They encode the same length of protein with only a few amino acid differences. No NR1

paralogs other than in Zebrafish have been identified in vertebrate or invertebrate species.

Phylogenetic analysis of DpNR1 with NR1s from other insect species shows that the two

DpNR1s are grouped together, which suggests that these are paralogs resulting from a gene

duplication event in D. punctata that did not occur in other insects. However, given that our

dataset was composed primarily of higher insects, and the small number of insect species in

which NR1 has been identified, it is difficult to make definitive conclusions about the origin of

the paralogs. DpNR2 undergoes alternative splicing, generating eleven different transcripts that

156

may encode nine protein isoforms (Fig. 4.3). Unlike NR2 in Drosophila in which alternative

splicing occurs mainly at the 5’ untranslated region, DpNR2 undergoes alternative splicing

principally at the 3’-end (Xia et al., 2005).

Tissue distribution shows that both DpNR1 and DpNR2 mRNA accumulated in brain and nerve

cord, which is consistent with the role of NMDAR in learning and memory (Xia et al., 2005).

The NR1 subunit is essential for the basic channel activity of NMDAR, whereas the NR2 subunit

is regarded as the rate-limiting molecule in controlling the optimal channel properties of

NMDAR (Monyer et al., 1992; Sprengel et al., 1998; Tang et al., 1999). In D. punctata, we

observed that both DpNR1 paralogs are stably expressed in brain, whereas the relative DpNR2

mRNA level underwent a slight change, probably to regulate the activity of NMDAR (Fig. 4.5).

In rats, D-aspartic acid (D-Asp) can induce testosterone synthesis through the activation of

NMDAR in testis (Santillo et al., 2014). A high transcript level of DpNR1 was observed in adult

D. punctata testes, indicating that NMDAR may also play a role in regulation of reproduction in

our model insect (Fig. 4.6). However, the transcript level of DpNR2 in testes was negligible

compared to that of DpNR1, which suggests that NR1 may form a homomeric functional channel,

as was observed in an earlier study in which expression of Drosophila NR1 alone produced a

weak but significant NMDA response (Ultsch et al., 1993; Xia et al., 2005). An alternative

explanation for the low expression of DpNR2 in the testes could be the existence of one or more

additional DpNR2 subunits in D. punctata. Four distinct NR2 subunits (A-D) were identified in

mammalian species. All four NR2 subunits are expressed in brain (Cull-Candy et al., 2001). In

testis, however, the expression of NR2 subunits varies (Hu et al., 2004; Santillo et al., 2014). In

the rat testis, only NR2A and NR2D are strongly expressed, whereas NR2B and NR2C are

undetectable (Santillo et al., 2014). In mouse testis, on the other hand, only NR2B subunit is

157

highly expressed (Hu et al., 2004). Thus, it is possible that an additional NR2 gene is expressed

in the testis of D. punctata.

The function of NMDAR in reproduction is well-studied in mammals (Mahesh and Brann, 2005).

However, clear evidence linking NMDAR to reproduction in insects remains limited. The

possible role of NMDAR in reproduction of insects was described by Chiang et al. (2002). In

that study, NMDA stimulates JH biosynthesis by inducing a Ca2+-influx from the extracellular

environment in the CA of D. punctata (Chiang et al., 2002). However, our study demonstrates

that the transcript levels of DpNR1 and DpNR2 in the CA are very low throughout the first

gonadotrophic cycle.

Treatment with MK-801 in vivo blocked vitellogenesis in S. gregaria and resulted in a reduction

in JH titre in G. bimaculatus (Begum et al., 2004; Geister et al., 2008). For this reason, the role

of NMDAR in JH biosynthesis and reproduction was further examined by the administration of

MK-801 in vivo in D. punctata. Administration of a high dose of MK-801 results in a significant

decrease in JH biosynthesis by day 4 and 5 CA, indicating that MK-801 does have an effect on

JH biosynthesis. However, the effect of MK-801 on JH biosynthesis is not apparent at doses less

than 12 µg/animal/injection (about 60 µg/g body mass) (Fig. S4.3). The result of administration

of MK-801 in S. gregaria suggests that the synthesis of Vg in the fat body was affected by MK-

801 injection (Begum et al., 2004). In D. punctata, the production of Vg in fat body and the

uptake of Vg by the basal oocytes are JH-dependent events (Marchal et al., 2013a; Rankin and

Stay, 1984; Stay and Tobe, 1978). Although rates of JH biosynthesis on days 4 and 5 in MK-

801-treated animals are lower than in controls, the change in transcript level of DpVg was not

significant relative to controls (Fig. 4.9C), and the pattern of JH biosynthesis during the first

gonadotrophic cycle is similar to that in control animals (Fig. 4.9A). Oocyte growth in treated

158

animals is also similar to the controls (Fig. 4.9B). All females oviposited on day 8. Overall,

administration of MK-801 does not have a significant effect on reproduction in female D.

punctata.

Treatment of animals with an extremely high dose of MK-801 in vivo significantly reduced JH

biosynthesis on days 4 and 5, even though MK-801 did not show any significant effect on JH

biosynthesis in vitro. Partial knockdown of DpNR2 did not show any effect on JH biosynthesis or

on basal oocyte growth (Fig. 4.8). To date, there is no clear evidence showing that the effect of

MK-801 on JH biosynthesis is mediated through NMDAR. In rats, injection of 0.2 µg/g MK-801

resulted in a failure to elevate LH, FSH, or progesterone (Luderer et al., 1993). In rats and mice,

a dose of 0.1 µg/g MK-801 was the maximum dose that could be used without causing

sensorimotor impairments and/or signs of intoxication (Van der Staay et al., 2011). Injection of

30 µg/animal of MK-801 paralyzed the cockroaches for 4-5 h. Therefore, it is possible that the

significant decrease in JH biosynthesis on days 4 and 5 in D. punctata was the result of

physiological stress induced by MK-801 treatment, rather than the action of NMDAR.

Accordingly, NMDAR does not appear to play important roles in the regulation of JH

biosynthesis or reproduction in female D. punctata.

In conclusion, two NR1 paralogs and eleven NR2 alternative splicing variants have been

identified in D. punctata. The expression of NMDAR subunits suggests that NMDAR may play

a role in the reproduction of male cockroaches. However, in the female cockroach, although a

previous study suggested that NMDAR mediates JH biosynthesis in D. punctata, our data reveal

a different story. Neither in vivo treatment of MK-801 nor partial knockdown of DpNR2 have

any effect on JH biosynthesis. The decrease in JH biosynthesis at a high dose in MK-801-treated

animals appears to result from physiological stress, rather than the direct action of NMDAR. In

159

addition, no reproductive events were affected following the blocking of NMDAR activity

through RNAi or MK-801 treatment. A reexamination of the function of NMDA and receptors in

the reproduction of insects now appears to be appropriate and timely.

160

Supplementary data

Table S4.1 Degenerate primer and RACE primer sequences for cloning DpNR1A, DpNR1B

and DpNR2 cDNAs.

Name Symbol Degenerate/RACE primer sequences

NMDA receptor

subunit 1 variant

A

NR1A F 5'- CTRTCGCCCGATGGTCARTTYGG -3'

R 5'- CAGTATGAAYACYCCTGCCATRT -3'

Fn 5'- ACGTATTTCAACATYGGYGGHGT-3'

Rn 5'- CCAAATTGGCCATCCGGCGACAA-3'

3’Race 5'- GGATTCGCCATGATCATTGTGGCA -3'

5’Race 5'- GGATCCATTGAAATCGCTGTGGA -3'

NMDA receptor

subunit 1 variant

B

NR1B F 5'- CTRTCGCCCGATGGTCARTTYGG -3'

R 5'- CAGTATGAAYACYCCTGCCATRT -3'

Fn 5'- ATCCACAGCTCCGACACDGAYGG -3'

Rn 5'- GTACCTTCTCCTATGCCACTATT-3'

3’Race 5'- TAAATGATGCGAGACTGCGTA-3'

5’Race 5'- CCACACATCAGCTTGATGGGA -3'

NMDA receptor

subunit 2

NR2 F 5'- ATACCGGTCATCKCVTGGAAYGC -3'

R 5'- AACATCCAAGARGCYGTRTCRAA -3'

Fn 5' - CTCGGAGAGGGAAGCAGTTGTGG -3’

Rn 5' -TCGAGCAGTCGYTTRTTRAACAT - 3'

3’Race 5'- CCAGAACCGATCCACTGTAGCA -3'

5’Race 5'- CTCTCCGCTCAAGGCCGGAATT -3'

161

Table S4.2 Oligonucleotide sequences for primers used in q-RT-PCR for reference and target genes. Efficiencies and R² values are

indicated.

Reference

genes F-primer R-primer

Amplicon

size (bp)

Efficiency

(%) R2

Tubulin 5'-AAATTACCAACGCTTGCTTTGAA-3' 5'-TGGCGAGGATCGCATTTT-3' 58 95.1 0.993

EF1α 5'-TCGTCTTCCTCTGCAGGATGTCT-3' 5'-GGGTGCAAATGTCACAACCATACC-3' 109 99.2 0.994

Armadillo 5'-GCTACTGCACCACTCACAGAATTATT-

3' 5'-CTGCAGCATACGTTGCAACA-3' 64 94.5 0.980

Target genes F-primer R-primer Amplicon

size (bp)

Efficiency

(%) R2

DpNR1A 5’- ATCGAGAAGCGGAAAACACT -3’ 5’- GTTGCTGGATCATTGACACC -3’ 80 90.2 0.984

DpNR1B 5’- GCACACTTTGGGACTCAAGA -3’ 5’- CCTCCAGCAACTAGCATGAA -3’ 54 100 0.983

DpNR2 5’- AAGAACCAGAACCGATCCAC -3’ 5’- GGCCACTAGGAAATCCAAAA -3’ 105 91.7 0.974

DpVg 5’-AAAGGTGTCCTCAGCCAGC-3’ 5’-TCCTCCATCTCGGATTGGGA-3’ 105 95.1 0.998

162

Table S4.3 Nucleotide sequences of primers used in making the dsRNA constructs.

Gene F-primer (5'-3') R-primer (5'-3')

DpNR2 ACAGGATATGGTATCGCCTTTAGC TCTTGAGCTTCGAAAACTGCAC

DpNR12 TAACTGGGGACAGTCCACAC TCAGGTGAGAGTCCAACATGG

pJET TTGCGCTCACTGCCAATTGC CTGGCCTTTTGCTCACATGTT

Table S4.4 Amino acid identities (in %) between NMDAR subunits in D. punctata and

other insects. DpNR1 was compared with NR1 in other species, and DpNR2A-1 was

compared with NR2 in other species.

DpNR1B D. melanogaster A. mellifera A. aegypti

DpNR1A 65.6 65.4 75.4 68.2

DpNR1B --- 55.6 62.7 58.8

DpNR2A-1 --- 67.3% 75.6% 71.0%

163

AGCTCGACCTTTGGGACTATGATTGTTGGGCAAAGTTGCATCGTAATTCTCTTGTTTACGGTTACCATGGTGACTACTTCGAGGTCTCCTTGGTTGGGTGTGACTCCAGCGGTATCCAACCTTTCGTCGAGTTCGCCCAGATTGTCAGCTTGGAGAGAAAATAACTCTGGGCCGCACCATCACCATCATCAAGGGGAGAAGAAA

GGAAGAGAAGGAAACTTGTCCATTGGACTTATTGTGCCGTACACTAACTTTGGCGTAAGGGACTATATTCGGGCAGTGAAGAGTGCTGTGGAGAAATTGGCGAAACCAAGAGGCAGGAGACTCAACTTTTTCAAGAAGTACAATTTCTCTCCCAACGAAGTTCACAGCGTCATGATGTCACTAACTCCAAGCCCCACTGCCATTCTTAACTCACTCTGCAAGGAGTTTCTCTCTGTGAACGTCTCCGCCATATTGTACCTG

1 - ATGAACTATGAGAAATATGGAAGGAGTACGGCGTCTGCGCAATATTTCTTGCAGCTTGCGGGCTACTTGGGAATCCCGGTTATCGCCTGG - 90 1 - M N Y E K Y G R S T A S A Q Y F L Q L A G Y L G I P V I A W - 30

91 - AACGCGGATAATTCCGGCCTTGAGCGGAGAGCGTCTCAATCAAGCCTTCAGCTGCAGTTAGCACCATCTCTGGAGCACCAGACGGCCGCC - 180 31 - N A D N S G L E R R A S Q S S L Q L Q L A P S L E H Q T A A - 60

181 - ATGTTGAGTATATTGGAGAGATACAAGTGGCATCAGTTCTCTGTGGTCACCAGTCAGATAGCGGGTCACGATGACTTCCTACAGGCGGTC - 270 61 - M L S I L E R Y K W H Q F S V V T S Q I A G H D D F L Q A V - 90

271 - AGAGAGAGGATCACCGAAGTGCAAGATAGATTCAAGTTCACGATCCTGAATCAAGTGTTGGTCACAAAGCCGGTAGATTTACTAGACCTG - 360 91 - R E R I T E V Q D R F K F T I L N Q V L V T K P V D L L D L - 120

361 - GTCAACTCTGAGTCTCGAGTGATGTTACTCTACGCCACCAGAGAAGAGGCCATACACATCTTGAAAGCCGCCAGAGATTACCAGATCACT - 450 121 - V N S E S R V M L L Y A T R E E A I H I L K A A R D Y Q I T - 150

451 - GGAGAGAACTACGTATGGGTCGTCACCCAGAGCGTCATGGAGAACCTCCAGACACCTTTCGGTTTCCCTGTCGGCATGCTCGGTGTCCAT - 540 151 - G E N Y V W V V T Q S V M E N L Q T P F G F P V G M L G V H - 180

541 - TTCGATACAAGCAGCACATCCCTTGTGAACGAAATCACCACCGCCATTAGGGTGTACGCATACGGGGTAGAGGATTTTGTGAATGATCCC - 630 181 - F D T S S T S L V N E I T T A I R V Y A Y G V E D F V N D P - 210

631 - AGAAATGTCAATCTCTCTCTCAGCACACAATTATCTTGTGAAGGGATGGGAGACTCCAGATGGAAAACAGGAGACAGATTCTTCAGGTAC - 720 211 - R N V N L S L S T Q L S C E G M G D S R W K T G D R F F R Y - 240

721 - CTCCGGAACGTGAGTGTTGAGGGGGATACAGGAAAGCCACATGTAGAATTCACACCGGAAGGAGTTCTGAAGGCAGCAGAGTTGAAAATA - 810 241 - L R N V S V E G D T G K P H V E F T P E G V L K A A E L K I - 270

811 - ATGAATTTGAGGCCTGGTGTTAGCAAGCAGCTTGTGTGGGAAGAGATCGGAGTGTGGAAATCTTGGGAGAAGGAAGGTCTGGACATCAAA - 900 271 - M N L R P G V S K Q L V W E E I G V W K S W E K E G L D I K - 300

901 - GACATCGTGTGGCCTGGGAACAGTCACACTCCGCCTCAGGGAGTTCCTGAAAAGTTCCACCTGAAGATAACTTTCCTGGAAGAACCTCCA - 990 301 - D I V W P G N S H T P P Q G V P E K F H L K I T F L E E P P - 330

991 - TATATCAACCTTGCGCCCCCGGATCCCGTCACCGGAAAATGCAGCATGAACAGGGGCGTTCTCTGCCGGGTGGCCAAGGAAGAAGAAATG - 1080 331 - Y I N L A P P D P V T G K C S M N R G V L C R V A K E E E M - 360

1081 - GAAAAGGTGGATGTACCGATGGCACATAAAAACGGCAGTTTCTACCAATGCTGCTCTGGGTTCTGCATAGACCTGCTAGAAAAATTCGCA - 1170 361 - E K V D V P M A H K N G S F Y Q C C S G F C I D L L E K F A - 390

S1 1171 - GAAGAACTTGGGTTCACATACGAACTCGTGAGGGTGGAAGATGGAAAATGGGGGACTCTGGAGAACGGGAAGTGGAACGGATTGATAGCG - 1260 391 - E E L G F T Y E L V R V E D G K W G T L E N G K W N G L I A - 420

1261 - GATCTCGTCAACCGGAAGACAGACATGGTGATGACGTCACTGATGATAAACTCGGAGAGGGAAGCAGTTGTGGACTTCACTGTGCCCTTT - 1350 421 - D L V N R K T D M V M T S L M I N S E R E A V V D F T V P F - 450

1351 - ATGGAGACTGGCATTGCCATTTTGGTGGCCAAGAGAACTGGTATCATATCTCCCACAGCTTTTCTTGAACCGTTTGACACGGCCTCCTGG - 1440

451 - M E T G I A I L V A K R T G I I S P T A F L E P F D T A S W - 480

1441 - ATGCTGGTAGGTGTAGTAGCGATCCAGGCAGCCACTTTTACTATTTTCCTTTTCGAATGGCTGAGTCCTTCAGGATTCGACATGAAGAAT - 1530 481 - M L V G V V A I Q A A T F T I F L F E W L S P S G F D M K N - 510

TM 1 1531 - GTGGATGGTTTCCAGCAAGTGTCTCCGTCCCCAAATCACCGGTTTTCATTGTTCAGGACCTATTGGTTAGTGTGGGCTGTCCTCTTCCAA - 1620 511 - V D G F Q Q V S P S P N H R F S L F R T Y W L V W A V L F Q - 540

TM 2 1621 - GCAGCCGTACATGTAGATTCCCCAAGAGGCTTCACTGCCAGGTTCATGACCAATGTGTGGGCCATGTTTGCAGTGGTGTTCCTTGCTATT - 1710 541 - A A V H V D S P R G F T A R F M T N V W A M F A V V F L A I - 570

TM 3 1711 - TACACTGCCAACCTGGCTGCCTTCATGATCACAAGAGAAGAATTTCATGAATTTTCTGGCCTCGACGATTCTAGGCTGTCAAAACCATTC - 1800 571 - Y T A N L A A F M I T R E E F H E F S G L D D S R L S K P F - 600

1801 - AGTCACAAACCCATGTACAGGTTTGGTACTATCCCATGGAGCCACACCGACTCAACTCTTAGTAAATATTTTGCCCCCATGCATGCCTAC - 1890 601 - S H K P M Y R F G T I P W S H T D S T L S K Y F A P M H A Y - 630

S2 1891 - ATGAAGAACCAGAACCGATCCACTGTAGCAGAGGGAATAGAAGCTGTTCTTAGTGGTGAACTGGATGCTTTCATTTATGATGGCACAGTT - 1980 631 - M K N Q N R S T V A E G I E A V L S G E L D A F I Y D G T V - 660

1981 - TTGGATTTCCTAGTGGCCCAGGATGAGGACTGCCGTCTGCTCACTGTGGGATCATGGTATGCCATGACAGGATATGGTATCGCCTTTAGC - 2070 661 - L D F L V A Q D E D C R L L T V G S W Y A M T G Y G I A F S - 690

2071 - CGCAACTCTAAATATGTTCAAATGTTCAACAAGCAAATGCTTGATTTTCGGGAAAATGGGGACTTGGAGAGACTACGGAGGTATTGGATG - 2160 691 - R N S K Y V Q M F N K Q M L D F R E N G D L E R L R R Y W M - 720

2161 - ACCGGGACATGTAAGCCTGGTAAACAAGAACATAAATCCAGTGACCCTCTAGCCCTGGAGCAGTTTTTATCTGCATTTTTGCTGCTCATG - 2250 721 - T G T C K P G K Q E H K S S D P L A L E Q F L S A F L L L M - 750

TM 4 2251 - TCTGGTATCCTACTAGCTGCAGTACTCTTAGCACTAGAACATGTGTATTTCAAATATGTTAGAAAACATTTGGCCAAAACAGACAGAGGA - 2340 751 - S G I L L A A V L L A L E H V Y F K Y V R K H L A K T D R G - 780

2341 - GGTTGCTGTGCTCTAATAAGTTTGAGCATGGGCAAATCATTAACGTTCCGCGGTGCAGTTTTCGAAGCTCAAGATATATTAAGACATCAT - 2430 781 - G C C A L I S L S M G K S L T F R G A V F E A Q D I L R H H - 810

2431 - CGTTGTCGGGATCCTATATGTGATACACATTTATGGAAAGTGAAACACGAGTTGGATCTCGCCCGTATGAAGATTCGACAATTGCAGAAA - 2520 811 - R C R D P I C D T H L W K V K H E L D L A R M K I R Q L Q K - 840

2521 - GAACTTGAAGCTCACGGCATCAAGCCAAGTAGAAGGAAGAAGAAGAAACGCGTCCCGTCCTGCTGGGTCTCCTGCTTCACACGAAACCAG - 2610 841 - E L E A H G I K P S R R K K K K R V P S C W V S C F T R N Q - 870

164

2611 - CCCGCGGGAGAGCAGGCGTCCCAATTGTTGAATGCAGATGACGTCCTCAAACCGCGTGATATGACAAGAAATCACGTGACATCTACAGAA - 2700 871 - P A G E Q A S Q L L N A D D V L K P R D M T R N H V T S T E - 900

2701 - TTCTCCGGCCGTTTTCACAGTGCGGGCCAGTTATACAGGTAA - 2742 901 - F S G R F H S A G Q L Y R * - 930

Fig S4.1. The nucleotide sequence and deduced amino acid sequence of DpNR2B-2. Agonist-

binding domains (S1 and S2 domains) are underlined; three hydrophobic transmembrane regions

(TM1, TM3, and TM4) and one hydrophobic pore-forming segment (TM2) are highlighted in the

boxes. Two insertions, and alternative 3’-end are highlighted in yellow, green and purple,

respectively.

Fig. S4.2. The effect of DpNR1B dsRNA treatment on JH biosynthesis and basal oocyte

growth, and the interactions among these genes in mated female D. punctata. (A) Relative

quantity of DippuNR1A, DippuNR1B and DippuNR2 mRNA levels in brain between control and

dsRNA treated animals. mRNA levels were normalized against levels of Tubulin and EF1a

mRNA. The data represent the average of 3 biologically independent pools (5 animals each pool),

run in triplicate. (B) JH biosynthesis by CA from control and dsRNA-treated animals. (C) Basal

oocyte length in control and dsRNA-treated animals. Values represent mean ± SEM. Levels of

significance to the control are indicated with the asterisk symbol: *P < 0.05, **P < 0.01

N R 1 1 N R 1 2 N R 2

0 .0

0 .5

1 .0

1 .5

2 .0

Re

lati

ve

mR

NA

qu

an

tity

***

N R 12 d sR N A

c o n tro l

C o n tro l N R 1 2 d s R N A

0

2 0

4 0

6 0

8 0

JH

bio

sy

nth

es

is(p

mo

l/h

/CA

)

C o n tro l N R 1 2 d s R N A

0 .0

0 .5

1 .0

1 .5

Ba

sa

l o

oc

yte

le

ng

th (

mm

)

A B C

165

Fig. S4.3. JH biosynthesis of CA in day 4 mated female D. punctata injected with different

doses of MK-801 on days 0, 1 and 3. Control was injected with ddH2O. Values represent mean

± SEM (n ≥17). Levels of significance to the control are indicated with the asterisk symbol: ***P

< 0.001.

3 µ g 1 2 µ g 3 0 µ g

0

2 0

4 0

6 0

JH

bio

sy

nth

es

is (

pm

ol/

h/C

A) c o n tro l

M K -8 0 1

***

166

References

Begum, M., Breuer, M., Kodrik, D., M, M.R., De Loof, A., 2004. The NMDA receptor antagonist MK-801

inhibits vitellogenesis in the flesh fly Neobellieria bullata and in the desert locust Schistocerca gregaria. J.

Insect Physiol. 50, 927-934.

Burnashev, N., Schoepfer, R., Monyer, H., Ruppersberg, J.P., Gunther, W., Seeburg, P.H., Sakmann, B., 1992.

Control by asparagine residues of calcium permeability and magnesium blockade in the NMDA receptor.

Science 257, 1415-1419.

Chiang, A.S., Lin, W.Y., Liu, H.P., Pszczolkowski, M.A., Fu, T.F., Chiu, S.L., Holbrook, G.L., 2002. Insect

NMDA receptors mediate juvenile hormone biosynthesis. P. Natl. Acad. Sci. U.S.A. 99, 37-42.

Cox, J.A., Kucenas, S., Voigt, M.M., 2005. Molecular characterization and embryonic expression of the family

of N-methyl-D-aspartate receptor subunit genes in the zebrafish. Dev. Dynam. 234, 756-766.

Cull-Candy, S., Brickley, S., Farrant, M., 2001. NMDA receptor subunits: diversity, development and disease.

Curr. Opin. Neurobiol. 11, 327-335.

Dingledine, R., Borges, K., Bowie, D., Traynelis, S.F., 1999. The glutamate receptor ion channels. Pharmacol.

Rev. 51, 7-61.

Felsenstein, J., 1985. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 39, 783-

791.

Geister, T.L., Lorenz, M.W., Hoffmann, K.H., Fischer, K., 2008. Effects of the NMDA receptor antagonist

MK-801 on female reproduction and juvenile hormone biosynthesis in the cricket Gryllus bimaculatus and the

butterfly Bicyclus anynana. J. Exp. Biol. 211, 1587-1593.

Goodman, W.G., Granger, N.A., 2005. The juvenile hormones, in: Gilbert, L.I., Iatrou, K., Gill, S.S. (Eds.),

Comprehensive Molecular Insect Science. Elsevier, Oxford, pp. 319-408.

Granger, N.A., Sturgis, S.L., Ebersohl, R., Geng, C., Sparks, T.C., 1996. Dopaminergic control of corpora

allata activity in the larval tobacco hornworm, Manduca sexta. Arch. Insect Biochem. Physiol. 32, 449-466.

Guindon, S., Gascuel, O., 2003. A simple, fast, and accurate algorithm to estimate large phylogenies by

maximum likelihood. Syst. Biol. 52, 696-704.

Hartfelder, K., 2000. Insect juvenile hormone: from "status quo" to high society. Braz. J. Med. Biol. Res. 33,

157-177.

Hu, J.H., Yang, N., Ma, Y.H., Jiang, J., Zhang, J.F., Fei, J., Guo, L.H., 2004. Identification of glutamate

transporters and receptors in mouse testis. Acta Pharmacol. Sin. 25, 366-371.

Huang, J., Tian, L., Peng, C., Abdou, M., Wen, D., Wang, Y., Li, S., Wang, J., 2011. DPP-mediated TGFbeta

signaling regulates juvenile hormone biosynthesis by activating the expression of juvenile hormone acid

methyltransferase. Development 138, 2283-2291.

Kuryatov, A., Laube, B., Betz, H., Kuhse, J., 1994. Mutational analysis of the glycine-binding site of the

NMDA receptor: structural similarity with bacterial amino acid-binding proteins. Neuron 12, 1291-1300.

Luderer, U., Strobl, F.J., Levine, J.E., Schwartz, N.B., 1993. Differential gonadotropin responses to N-methyl-

D,L-aspartate in metestrous, proestrous, and ovariectomized rats. Biol. Reprod. 48, 857-866.

Madden, D.R., 2002. The structure and function of glutamate receptor ion channels. Nat. Rev. Neurosci. 3, 91-

101.

Maffucci, J.A., Noel, M.L., Gillette, R., Wu, D., Gore, A.C., 2009. Age- and hormone-regulation of N-methyl-D-aspartate receptor subunit NR2b in the anteroventral periventricular nucleus of the female rat: implications

for reproductive senescence. J. Neuroendocrinol. 21, 506-517.

167

Mahesh, V.B., Brann, D.W., 2005. Regulatory role of excitatory amino acids in reproduction. Endocrine 28,

271-280.

Marchal, E., Hult, E.F., Huang, J., Stay, B., Tobe, S.S., 2013a. Diploptera punctata as a model for studying the

endocrinology of arthropod reproduction and development. Gen. Comp. Endocr. 188, 85-93.

Marchal, E., Hult, E.F., Huang, J., Tobe, S.S., 2013b. Sequencing and validation of housekeeping genes for

quantitative real-time PCR during the gonadotrophic cycle of Diploptera punctata. BMC Res. Notes. 6, 237.

McBain, C.J., Mayer, M.L., 1994. N-methyl-D-aspartic acid receptor structure and function. Physiol. Rev. 74,

723-760.

Monyer, H., Sprengel, R., Schoepfer, R., Herb, A., Higuchi, M., Lomeli, H., Burnashev, N., Sakmann, B.,

Seeburg, P.H., 1992. Heteromeric NMDA receptors: molecular and functional distinction of subtypes. Science

256, 1217-1221.

Mussig, L., Richlitzki, A., Rossler, R., Eisenhardt, D., Menzel, R., Leboulle, G., 2010. Acute disruption of the

NMDA receptor subunit NR1 in the honeybee brain selectively impairs memory formation. J. Neurosci. 30,

7817-7825.

Newcomer, J.W., Krystal, J.H., 2001. NMDA receptor regulation of memory and behavior in humans.

Hippocampus 11, 529-542.

Pszczolkowski, M.A., Lee, W.S., Liu, H.P., Chiang, A.S., 1999. Glutamate-induced rise in cytosolic calcium

concentration stimulates in vitro rates of juvenile hormone biosynthesis in corpus allatum of Diploptera

punctata. Mol. Cell. Endocrinol. 158, 163-171.

Rankin, S.M., Stay, B., 1984. The changing effect of the ovary on rates of juvenile hormone synthesis in

Diploptera punctata. Gen. Comp. Endocrinol. 54, 382-388.

Rawls, S.M., Thomas, T., Adeola, M., Patil, T., Raymondi, N., Poles, A., Loo, M., Raffa, R.B., 2009.

Topiramate antagonizes NMDA- and AMPA-induced seizure-like activity in planarians. Pharmacol. Biochem.

Be. 93, 363-367.

Santillo, A., Falvo, S., Chieffi, P., Burrone, L., Chieffi Baccari, G., Longobardi, S., Di Fiore, M.M., 2014. D-

aspartate affects NMDA receptor-extracellular signal-regulated kinase pathway and upregulates androgen

receptor expression in the rat testis. Theriogenology 81, 744-751.

Sircar, R., Rappaport, M., Nichtenhauser, R., Zukin, S.R., 1987. The novel anticonvulsant Mk-801 - a potent

and specific ligand of the brain phencyclidine sigma-receptor. Brain Res. 435, 235-240.

Sprengel, R., Suchanek, B., Amico, C., Brusa, R., Burnashev, N., Rozov, A., Hvalby, O., Jensen, V., Paulsen,

O., Andersen, P., Kim, J.J., Thompson, R.F., Sun, W., Webster, L.C., Grant, S.G., Eilers, J., Konnerth, A., Li,

J., McNamara, J.O., Seeburg, P.H., 1998. Importance of the intracellular domain of NR2 subunits for NMDA

receptor function in vivo. Cell 92, 279-289.

Stay, B., Tobe, S.S., 1978. Control of juvenile hormone biosynthesis during the reproductive cycle of a

viviparous cockroach. II. Effects of unilateral allatectomy, implantation of supernumerary corpora allata, and

ovariectomy. Gen. Comp. Endocrinol. 34, 276-286.

Stay, B., Tobe, S.S., 2007. The role of allatostatins in juvenile hormone synthesis in insects and crustaceans.

Annu. Rev. Entomol. 52, 277-299.

Stern-Bach, Y., Bettler, B., Hartley, M., Sheppard, P.O., O'Hara, P.J., Heinemann, S.F., 1994. Agonist

selectivity of glutamate receptors is specified by two domains structurally related to bacterial amino acid-

binding proteins. Neuron 13, 1345-1357.

Sydow, S., Kopke, A.K.E., Blank, T., Spiess, J., 1996. Overexpression of a functional NMDA receptor subunit

(NMDAR1) in baculovirus-infected Trichoplusia ni insect cells. Mol. Brain Res. 41, 228-240.

Tamura, K., Stecher, G., Peterson, D., Filipski, A., Kumar, S., 2013. MEGA6: Molecular Evolutionary

Genetics Analysis version 6.0. Mol. Biol. Evol. 30, 2725-2729.

168

Tang, Y.P., Shimizu, E., Dube, G.R., Rampon, C., Kerchner, G.A., Zhuo, M., Liu, G., Tsien, J.Z., 1999.

Genetic enhancement of learning and memory in mice. Nature 401, 63-69.

Thompson, C.S., Yagi, K.J., Chen, Z.F., Tobe, S.S., 1990. The effects of octopamine on juvenile hormone

biosynthesis, electrophysiology, and cAMP content of the corpora allata of the cockroach Diploptera punctata.

J. Comp. Physiol. B 160, 241-249.

Troncoso, J., Maldonado, H., 2002. Two related forms of memory in the crab Chasmagnathus are differentially

affected by NMDA receptor antagonists. Pharmacol. Biochem. Be. 72, 251-265.

Ultsch, A., Schuster, C.M., Laube, B., Betz, H., Schmitt, B., 1993. Glutamate receptors of Drosophila

melanogaster. Primary structure of a putative NMDA receptor protein expressed in the head of the adult fly.

Febs Lett 324, 171-177.

Van der Staay, F.J., Rutten, K., Erb, C., Blokland, A., 2011. Effects of the cognition impairer MK-801 on

learning and memory in mice and rats. Behav. Brain Res. 220, 215-229.

Vandesompele, J., De Preter, K., Pattyn, F., Poppe, B., Van Roy, N., De Paepe, A., Speleman, F., 2002.

Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal

control genes. Genome Biol. 3, 1-12.

Whelan, S., Goldman, N., 2001. A general empirical model of protein evolution derived from multiple protein

families using a maximum-likelihood approach. Mol. Biol. Evol. 18, 691-699.

Wu, C.L., Xia, S., Fu, T.F., Wang, H., Chen, Y.H., Leong, D., Chiang, A.S., Tully, T., 2007. Specific

requirement of NMDA receptors for long-term memory consolidation in Drosophila ellipsoid body. Nat.

Neurosci. 10, 1578-1586.

Xia, S., Miyashita, T., Fu, T.F., Lin, W.Y., Wu, C.L., Pyzocha, L., Lin, I.R., Saitoe, M., Tully, T., Chiang,

A.S., 2005. NMDA receptors mediate olfactory learning and memory in Drosophila. Curr. Biol. 15, 603-615.

Yang, Z., 1994. Maximum likelihood phylogenetic estimation from DNA sequences with variable rates over

sites: approximate methods. J. Mol. Evol. 39, 306-314.

Zannat, M.T., Locatelli, F., Rybak, J., Menzel, R., Leboulle, G., 2006. Identification and localisation of the

NR1 sub-unit homologue of the NMDA glutamate receptor in the honeybee brain. Neurosci. Lett. 398, 274-

279.

169

Chapter 5

General Discussion

1. Function of JH in reproduction

The original function of JH in insects was considered to be a regulator of reproduction (Tobe and

Bendena, 1999). Several different types of hormones are involved in the endocrine control of

female reproduction, of which JHs and ecdysteroids are most studied. In most hemimetabolous

insects, the major events of reproduction are controlled by JH, while in higher insects, both JH

and ecdysteroids were involved (Raikhel et al., 2005). In D. punctata, allatectomy or

implantation of CA with ovaries into male animals showed that JH stimulates Vg synthesis and

Vg uptake (Mundall et al., 1979; Stay and Tobe, 1977). Moreover, in several other cockroaches

species, treatment with JH or JH analogs achieved the same result (Comas et al., 1999; Don-

Wheeler and Engelmann, 1997; Weaver and Edwards, 1990). In this thesis, the function of JH in

regulating female reproduction was further investigated by repressing JH biosynthesis using

RNAi. As expected, the lowered JH biosynthesis on day 4 in the CA resulted in a significant

reduction of Vg mRNA levels, uptake of Vn into the oocytes, and in blocking patency (Chapter

2). The increase of JH biosynthesis on day 3 and 4 is essential for the development of oocytes.

On the other hand, the decrease in JH production starting on day 5 is required to complete the

regular growth of oocytes. An abnormal high titre of JH after day 5 resulted in a failure in

oviposition (Hult et al, data unpublished). It is obvious that JH plays critical functions in

regulating cockroach female reproduction and therefore must be tightly regulated.

170

2. Evolution of the JH biosynthetic pathway

2.1 The loss of sterols biosynthesis in arthropods

In most eukaryotic animal and plant systems, cholesterol and related sterols, which are essential

component of cell membranes, are biosynthesized through the mevalonate pathway. Arthropods,

however, do not appear to synthesize cholesterol, but rather produce sesquiterpenoids. The final

product of the mevalonate pathway is JH, rather than cholesterol, as a consequence of the

absence of genes encoding squalene synthase and other subsequent enzymes to produce sterols

(Belles et al., 2005). The biosynthesis of JH and cholesterol share the mevalonate pathway until

the production of FPP. In Arthropoda, FPP is converted to FOL and thereafter to FA (Figure

1.1), while it is catalyzed to squalene in animals producing cholesterol. In some plants, on the

other hand, cholesterol and JH III and their related compounds are synthesized, suggesting that

both the JH and cholesterol biosynthetic pathway are present in plant system (Yang et al., 2013).

Tobe and Bendena (1999) suggest that the ability to biosynthesize cholesterol was maintained in

the ancestor of the arthropod/annelid lineage, but lost in the arthropod line, probably because of

the abundance of cholesterol in the diet. The biosynthesis of JH in some plants is the result of

convergence, which is unrelated to its occurrence in arthropods. Nevertheless, the reason for

losing the ability to biosynthesize cholesterol de novo in arthropods remains unknown.

2.2 Enzymes in the JH biosynthetic pathway controlling the end product of the JH

biosynthetic pathway

As shown in Figure 1.1, various JHs are produced in insects. These JHs shares the main carbon

skeleton, but with different methyl/ethyl substitutions and carbon-sites for epoxidation. Most

ethyl substituted JH III products occur in the order of the Lepidoptera (Details seen Chapter 1

171

section 1). The reason for the various end products of the JH biosynthetic pathway in

Lepidoptera can be found in the catalytic selectivity of enzymes in the pathway. Unlike in other

insects, both acetyl-CoA and propionyl-CoA are involved in the early steps of JH biosynthesis to

produce ethyl-branched JHs in M. sexta (Brindle et al., 1987; Brindle et al., 1988). The

branched-chain amino acid transaminase, which generates propionyl-CoA by the metabolism of

leucine and valine, is found only in the CA of Lepidoptera (Brindle et al., 1988). In addition,

enzymes in the JH biosynthetic pathway have evolved to utilize different substrates for JH

biosynthesis. For instance, IPPI in M. sexta and C. fumiferana contains an active site cavity,

which allows it to bind to larger substrates (ethyl-substituted IPP) and to stabilize the high-

energy intermediates to form ethyl- substituted DMPP (Sen et al., 2012). FPPS in M. sexta is also

adapted to catalyze both ethyl- and methyl-substituted DMPP and GPP, and the selectivity of the

enzyme inclines to the ethyl-substituted substrate (Sen et al., 1996; Sen et al., 2006).

As noted in Chapter 1, section 1.2.13, CYP15 genes have been characterized to encode

cytochrome P450 (CYP) enzymes that epoxidize MF to JH in Orthoptera, Dictyoptera and

Coleoptera, or FA to JHA in Lepidotera. However, in higher Diptera that produce bisepoxy

forms of JH III, no CYP15 gene has been found (Daimon and Shinoda, 2013). Different CYP

enzymes that could catalyze the bisepoxidization of MF may be present in the higher Diptera to

compensate the loss of CYP15, resulting in the production of JHB3. However, in P. stali that

produces JHSB3, a CYP15 ortholog was identified. The function of this CYP15 ortholog has not

yet been determined. Whether there are other epoxidases involved in the synthesis of JHSB3

remains unclear. In addition, in the crustacean D. pulex, no CYP15 ortholog was found in the

genome. The lack of CYP15 genes may be the reason for crustaceans using MF as their JH-like

bioactive compound.

172

The catalytic selectivity of the enzymes in the JH biosynthetic pathway determines not only the

structure of the JH products, but also the order of steps in the JH biosynthetic pathway. As

shown in Figure 1.2, the last two steps (methylation and epoxidation) of the JH biosynthetic

pathways differ between Lepidoptera and other insects. Two different epoxidase enzymes,

CYP15A1 and CYP15C1 are involved in the last steps of JH biosynthesis in different insect

orders. The order of the methylation and epoxidation steps appears to be controlled by the

selectivity of the epoxidase. CYP15C1, the functional epoxidase in Lepidoptera, specifically

converts FA to JHA (Daimon et al., 2012). Therefore, epoxidation occurs before methylation in

Lepidoptera. However, in other insect orders, CYP15A1 has been characterized as the active

epoxidase. The substrate selectivity of CYP15A1 to MF determines that methylation precedes

epoxidation (Helvig et al., 2004).

D. punctata, as a more basal species in the Insecta, did not appear to undergo any significant

modification of the enzymes in the JH biosynthetic pathway. Thus, only JH III has been

identified as the JH biosynthetic product of CA.

3. Regulation of JH biosynthesis

3.1 Enzymes in the biosynthetic pathway

Regulation of JH biosynthesis is a complicated process involving multiple regulators. The best-

characterized regulators are the neuropeptides, including ATs and ASTs. Enzymes in the JH

biosynthetic pathway were also shown to regulate JH biosynthesis (Couillaud and Feyereisen,

1991; Feyereisen and Farnsworth, 1987; Kinjoh et al., 2007; Nouzova et al., 2011). Some

enzymes have been proposed to catalyze rate-limiting steps in the pathway (e.g. FALD in A.

aegypti (Rivera-Perez et al., 2013); JHAMT in B. mori (Shinoda and Itoyama, 2003)). In this

173

thesis, the transcript levels of genes in the JH biosynthetic pathway were determined in relation

to the JH biosynthesis in D. punctata. Similar to other species, the expression of most genes is

well-correlated with the JH biosynthetic activity, suggesting that JH biosynthesis is in part

controlled by the transcription of enzymes. However, no rate-limiting enzyme could be identified

in D. punctata. My study has also shown that precursor supply mediates JH biosynthesis. In

addition, silencing individual genes encoding the JH biosynthetic enzymes resulted in a

significant decrease in the transcript levels of other genes. Therefore, a feedback among genes in

JH biosynthetic pathway was proposed: the expression of enzymes in the pathway is correlated

with other enzymes to obtain the balance between enzyme activity and JH precursor supply.

Further experiments to determine the amount of JH precursors in the CA after silencing

individual genes would provide evidence to support the hypothesis.

3.2 Neuropeptides

In cockroaches, ASTs compose a family of 13 to 14 neuropeptides with the conserved C-

terminal sequence (Y/FXFGL/I-NH2) and variable N-terminal sequences. All the peptides are

encoded by a single gene. Evolutionary analysis of this gene indicates that the AST sequences

were generated by a gene duplication event before the species diverged (Belles et al., 1999).

Structure-activity studies show that all ASTs are able to inhibit JH biosynthesis effectively, albeit

with different potencies (Tobe et al., 2000). This result suggests that AstRs in cockroaches are

relatively unselective, binding ASTs with different structures. In my thesis, the selectivity of

DpAstR to its ligand was determined by the activation of DpAstR expressed in CHO cells using

13 DpASTs and AST analogs. All ASTs and AST analogs were able to activate DpAstR, and the

selectivity of DpAstR to DpASTs was shown to be structure-dependent (Table 3.1 and Table 6.1).

174

The core sequence (Y/FXFGL/I-NH2) is essential for binding, while the N-terminal region

regulates binding affinity.

In general, the ability of AST to bind AstR correlates well with its potency to inhibit JH

production by the CA (Table 3.1). AST is a pleiotropic neuropeptide, mediating different

physiological effects, such as midgut enzyme activity and hindgut muscle contraction. However,

in these functions, binding of AST to its receptor does not seem to correlate with its activity

(Lange et al., 1995). In addition, the AstR identified in the CA of D. punctata was not found to

be expressed in the midgut, where the precursor of its ligand AST was highly expressed. My

study supports the hypothesis proposed by Tobe and Bendena (1999) that there are different AST

receptor subtypes in different tissues. Unfortunately, no second AstR could be identified in D.

punctata.

As mentioned before (Chapter 1, section 3.1), there are different families of allatostatic

neuropeptides. The FGLa/ASTs which inhibits JH biosynthesis in cockroaches, do not show the

allato-inhibitory effect on the CA of Lepidoptera or Diptera. Instead, the PISCF/AST family is

involved to regulate JH biosynthesis in these insect orders. Unlike in cockroaches where 14

FGLa/ASTs with very high sequence homology are encoded by a single gene, the PISCF/AST

precursor gene only encodes one AST peptide in Lepidoptera and Diptera (Kramer et al., 1991;

Li et al., 2006). In addition to PISCF/AST, another type of neuropeptide, termed allatotropin

(AT), which stimulates JH biosynthesis in the CA, was also identified in Lepidoptera and Diptera.

3.3 Second messenger (Ca2+)

Calcium, an intracellular second messenger, is a critical factor in controlling JH biosynthesis.

However, the role of calcium in the JH biosynthesis seems to vary depending on species and

175

stage (Allen et al., 1992; Dale and Tobe, 1988; Kikukawa et al., 1987). To make it more

complicated, calcium was determined to act as a second messenger in the signal pathway of AST

action to inhibit JH biosynthesis (Chapter 3). Calcium can also act as the second messenger of

Manse-AT to enhance CA activity (Horodyski et al., 2011). In this thesis, the function of

NMDAR as an ionotropic calcium channel was investigated. A previous study demonstrated that

NMDA stimulates JH biosynthesis in the CA through the influx of Ca2+ into the CA cells

(Chiang et al., 2002). However, our study, using RNAi and in vivo treatment with NMDAR

antagonists, revealed that NMDAR does not appear to regulate JH biosynthesis (Chapter 4).

Clearly, a complex mechanism is involved in the regulation of JH biosynthesis by calcium.

4. The value of my study in insect control

As an essential hormone controlling many physiological processes in insects, JH production is

tightly regulated to allow normal development. This characteristic makes JH a potential target for

insect control. Thousands of JH analogs have been synthesized and bioassayed to discover

potential insecticides (Goodman and Cusson, 2012). In addition to JH analogs, ASTs which

inhibit JH biosynthesis at low doses, have been employed in the design of potential insect growth

regulators (IGRs). Previous studies have determined the distribution, activity and structure-

activity relationship of the thirteen ASTs in D. punctata, which has permitted the development of

analogs of the naturally occurring ASTs (Donly et al., 1993; Rankin and Stay, 1987; Tobe et al.,

2000). The development of AST analogs as potential insecticides need to overcome two major

problems: (1) maintaining the high allato-inhibitory activity being a shorter peptide; (2) having a

high resistance to the peptidases in the CA and hemolymph (Marchal et al., 2013). In this thesis,

two series of AST analogs were designed and synthesized based on the previous rules. Some

AST analogs not only exhibited similar potencies in inhibiting JH biosynthesis in vitro to the

176

naturally occurring AST (Table 6.2), but were also able to regulate JH biosynthesis and oocyte

growth in an in vivo bioassay (Fig. 6.1). Preliminary degradation results show that the AST

analogs have higher resistance to peptidases in hemolymph than natural AST. These AST

analogs can be considered as potential IGRs. Further modification of these AST analgos is now

required to obtain higher allato-inhibitory activity and higher resistance to peptidases in the

cockroach.

5. Future perspectives

This thesis aimed to characterize the JH biosynthetic pathway and determine factors regulating

JH biosynthesis in D. punctata. Genes in the JH biosynthetic pathways were identified,

functionally characterized, and regulatory factors including AST and NMDAR were studied.

However, many questions still remain unanswered. The exact target of AST action remains

unknown. The role of AST in regulating JH biosynthesis is well-established, while the roles of

other regulatory factors including second messengers, JH feedback and ecdysteroids remain

unclear. In addition to the regulation of JH biosynthesis, the JH action and cross-talking with

other hormone also require further investigate.

D. punctata is a great model in studying the endocrinology of arthropods. In this animal the JH

titre is relatively high and stable, only JH III is is synthesized, and reproductive events are well-

correlated with JH biosynthesis. The genome of D. punctata is currently being assembled which

will allow high-throughput and more in depth studies aimed at identifying the regulatory

mechanisms targeting JH biosynthesis. A first study is currently ongoing involving differential

transcriptomics of CA with low versus high JH biosynthetic activity.

177

References

Allen, C.U., Janzen, W.P., Granger, N.A., 1992. Manipulation of intracellular calcium affects in vitro juvenile

hormone synthesis by larval corpora allata of Manduca sexta. Mol Cell Endocrinol 84, 227-241.

Belles, X., Graham, L.A., Bendena, W.G., Ding, Q., Edwards, J.P., Weaver, R.J., Tobe, S.S., 1999. The

molecular evolution of the allatostatin precursor in cockroaches. Peptides 20, 11-22.

Belles, X., Martin, D., Piulachs, M.D., 2005. The mevalonate pathway and the synthesis of juvenile hormone

in insects. Annu Rev Entomol 50, 181-199.

Brindle, P.A., Baker, F.C., Tsai, L.W., Reuter, C.C., Schooley, D.A., 1987. Sources of propionate for the

biogenesis of ethyl-branched insect juvenile hormones: Role of isoleucine and valine. Proc Natl Acad Sci U S

A 84, 7906-7910.

Brindle, P.A., Schooley, D.A., Tsai, L.W., Baker, F.C., 1988. Comparative metabolism of branched-chain

amino acids to precursors of juvenile hormone biogenesis in corpora allata of lepidopterous versus

nonlepidopterous insects. J Biol Chem 263, 10653-10657.

Chiang, A.S., Lin, W.Y., Liu, H.P., Pszczolkowski, M.A., Fu, T.F., Chiu, S.L., Holbrook, G.L., 2002. Insect

NMDA receptors mediate juvenile hormone biosynthesis. Proc Natl Acad Sci U S A 99, 37-42.

Comas, D., Piulachs, M.D., Belles, X., 1999. Fast induction of vitellogenin gene expression by juvenile

hormone III in the cockroach Blattella germanica (L.) (Dictyoptera, Blattellidae). Insect Biochem Mol Biol 29,

821-827.

Couillaud, F., Feyereisen, R., 1991. Assay of HMG-CoA synthase in Diploptera punctata corpora allata. Insect

Biochem 21, 131-135.

Daimon, T., Kozaki, T., Niwa, R., Kobayashi, I., Furuta, K., Namiki, T., Uchino, K., Banno, Y., Katsuma, S.,

Tamura, T., Mita, K., Sezutsu, H., Nakayama, M., Itoyama, K., Shimada, T., Shinoda, T., 2012. Precocious

metamorphosis in the juvenile hormone-deficient mutant of the silkworm, Bombyx mori. PLoS Genet 8,

e1002486.

Daimon, T., Shinoda, T., 2013. Function, diversity, and application of insect juvenile hormone epoxidases

(CYP15). Biotechnol Appl Biochem 60, 82-91.

Dale, J.F., Tobe, S.S., 1988. Differences in the stimulation by calcium ionophore of juvenile hormone III

release from corpora allata of solitarious and gregarious Locusta migratoria. Experientia 44, 240-242.

Don-Wheeler, G., Engelmann, F., 1997. The biosynthesis and processing of vitellogenin in the fat bodies of

females and males of the cockroach Leucophaea maderae. Insect Biochem Mol Biol 27, 901-918.

Donly, B.C., Ding, Q., Tobe, S.S., Bendena, W.G., 1993. Molecular cloning of the gene for the allatostatin

family of neuropeptides from the cockroach Diploptera punctata. Proc Natl Acad Sci U S A 90, 8807-8811.

Feyereisen, R., Farnsworth, D.E., 1987. Characterization and regulation of HMG-CoA Reductase during a

cycle of Juvenile Hormone synthesis. Mol Cell Endocrinol 53, 227-238.

Goodman, W.G., Cusson, M., 2012. The juvenile hormones, in: Gilbert, L.I. (Ed.), Insect Endocrinology.

Academic Press, London, p. 55.

Helvig, C., Koener, J.F., Unnithan, G.C., Feyereisen, R., 2004. CYP15A1, the cytochrome P450 that catalyzes

epoxidation of methyl farnesoate to juvenile hormone III in cockroach corpora allata. Proc Natl Acad Sci U S

A 101, 4024-4029.

Horodyski, F.M., Verlinden, H., Filkin, N., Vandersmissen, H.P., Fleury, C., Reynolds, S.E., Kai, Z.P., Broeck,

J.V., 2011. Isolation and functional characterization of an allatotropin receptor from Manduca sexta. Insect Biochem Mol Biol 41, 804-814.

178

Kikukawa, S., Tobe, S.S., Solowiej, S., Rankin, S.M., Stay, B., 1987. Calcium as a regulator of juvenile

hormone biosynthesis and release in the cockroach Diploptera punctata. Insect Biochem 17, 179-187.

Kinjoh, T., Kaneko, Y., Itoyama, K., Mita, K., Hiruma, K., Shinoda, T., 2007. Control of juvenile hormone

biosynthesis in Bombyx mori: Cloning of the enzymes in the mevalonate pathway and assessment of their

developmental expression in the corpora allata. Insect Biochem Mol Biol 37, 808-818.

Kramer, S.J., Toschi, A., Miller, C.A., Kataoka, H., Quistad, G.B., Li, J.P., Carney, R.L., Schooley, D.A., 1991.

Identification of an allatostatin from the tobacco hornworm Manduca sexta. Proc Natl Acad Sci U S A 88,

9458-9462.

Lange, A.B., Bendena, W.G., Tobe, S.S., 1995. The effect of the thirteen Dip-Allatostatins on myogenic and

induced contractions of the cockroach (Diploptera Punctanta) hindgut. J Insect Physiol 41, 581-588.

Li, Y., Hernandez-Martinez, S., Fernandez, F., Mayoral, J.G., Topalis, P., Priestap, H., Perez, M., Navare, A.,

Noriega, F.G., 2006. Biochemical, molecular, and functional characterization of PISCF-allatostatin, a regulator

of juvenile hormone biosynthesis in the mosquito Aedes aegypti. J Biol Chem 281, 34048-34055.

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. Gen Comp Endocr 188, 85-93.

Mundall, E.C., Tobe, S.S., Stay, B., 1979. Induction of vitellogenin and growth of implanted oocytes in male

cockroaches. Nature 282, 97-98.

Nouzova, M., Edwards, M.J., Mayoral, J.G., Noriega, F.G., 2011. A coordinated expression of biosynthetic

enzymes controls the flux of juvenile hormone precursors in the corpora allata of mosquitoes. Insect Biochem

Mol Biol 41, 660-669.

Raikhel, A.S., Brown, M.R., Belles, X., 2005. Hormonal control of reproductive processes, in: Gilbert, L.I.,

Iatrou, K., Gill, S.S. (Eds.), Comprehensive molecular insect science. Elsevier, Oxford, pp. 433-491.

Rankin, S.M., Stay, B., 1987. Distribution of allatostatin in the adult cockroach, Diploptera punctata and

effects on corpora allata in vitro. Journal of Insect Physiology 33, 551-558.

Rivera-Perez, C., Nouzova, M., Clifton, M.E., Garcia, E.M., LeBlanc, E., Noriega, F.G., 2013. Aldehyde

dehydrogenase 3 converts farnesal into farnesoic acid in the corpora allata of mosquitoes. Insect Biochem Mol

Biol 43, 675-682.

Sen, S.E., Ewing, G.J., Thurston, N., 1996. Characterization of lepidopteran prenyltransferase in Manduca sexta corpora allata. . Arch Insect Biochem Physiol 32, 315–332.

Sen, S.E., Hitchcock, J.R., Jordan, J.L., Richard, T., 2006. Juvenile hormone biosynthesis in M. sexta:

substrate specificity of insect prenyltransferase utilizing homologous diphosphate analogs. Insect Biochem

Mol Biol 36, 827-834.

Sen, S.E., Tomasello, A., Grasso, M., Denton, R., Macor, J., Beliveau, C., Cusson, M., Crowell, D.N., 2012.

Cloning, expression and characterization of lepidopteran isopentenyl diphosphate isomerase. Insect Biochem

Mol Biol 42, 739-750.

Shinoda, T., Itoyama, K., 2003. Juvenile hormone acid methyltransferase: a key regulatory enzyme for insect

metamorphosis. Proc Natl Acad Sci U S A 100, 11986-11991.

Stay, B., Tobe, S.S., 1977. Control of juvenile hormone biosynthesis during the reproductive cycle of a

viviparous cockroach 1. Activation and inhibition of corpora allata. Gen Comp Endocrinol 33, 531-540.

Tobe, S.S., Bendena, W.G., 1999. The regulation of juvenile hormone production in arthropods. Functional

and evolutionary perspectives. Ann N Y Acad Sci 897, 300-310.

Tobe, S.S., Zhang, J.R., Bowser, P.R.F., Donly, B.C., Bendena, W.G., 2000. Biological activities of the

allatostatin family of peptides in the cockroach, Diploptera punctata, and potential interactions with receptors.

Journal of Insect Physiology 46, 231-242.

179

Weaver, R.J., Edwards, J.P., 1990. The role of the corpora allata and associated nerves in the regulation of

ovarian cycles in the oviparous cockroach Periplaneta americana. J Insect Physiol 36, 51-59.

Yang, H., Kim, H.S., Jeong, E.J., Khiev, P., Chin, Y.W., Sung, S.H., 2013. Plant-derived juvenile hormone III

analogues and other sesquiterpenes from the stem bark of Cananga latifolia. Phytochemistry 94, 277-283.

180

Chapter 6

Appendices

ASTs which inhibit JH biosynthesis at low doses, have been employed in the design of potential

insecticides. Furthermore, molecular cloning, structure determination, distribution and activity of

the thirteen ASTs in D. punctata, have permitted the development of analogs of the naturally

occurring ASTs (Donly et al., 1993; Rankin and Stay, 1987; Tobe et al., 2000). Although ASTs

inhibit JH biosynthesis effectively in vitro, they have some shortcomings as potential pesticides.

Firstly, the longer the peptide, the more unwanted side-products that are produced during

chemical synthesis. For ASTs, even the shortest AST peptide has 6 amino acids (Tobe et al.,

2000). Secondly, ASTs are susceptible to metabolic inactivation by peptidases in hemolymph

and midgut (Garside et al., 1997a, b). For example the half-life of Dippu-AST 5 (AST5: Asp-

Arg-Leu-Tyr-Ser-Phe-Gly-Leu-amide) is less than 1 hour when incubated with membrane

preparations of brain and midgut (Garside et al., 1997b). Therefore, much effort has been

expended to overcome this problem and create better AST pseudopeptides. In this thesis, two

series of AST anaologs were designed and their activities in regulating JH biosynthesis in vitro

and in vivo in D. punctata were shown in this chapter.

181

Table 6.1. Structure of AST analogs

Analogs Structure

W201

Gly

O

Phe Gly Leu NH2

W202 Gly

O

Phe Gly Leu NH2

W203

Gly

O

Phe Gly Leu NH2

W204 Gly

O

Phe Gly Leu NH2

W205

Gly

O

Phe Gly Leu NH2

W206

W70

Gly

O

Phe Gly Leu NH2

Br

W71

Gly

O

Phe Gly Leu NH2

Br

K15

Gly

O

Phe Gly Leu NH2

Br

182

Table 6.2. Potency of Dippu-AST analogs a: inhibitory effect on JH release (IC50) and activation

of AstR in CHO-WTA11 cells (EC50)

Sample Group (R1) IC50 /nM Rank (IC50) EC50 /nM Rank (EC50)

AST6

1.45 2 0.2 1

W201 CH2 957 10 2952 10

W202 (CH2)2 93.5 8 472.7 9

W203 (CH2)3 77.0 7 153.3 7

W204 (CH2)4 69.5 5 443.3 8

W205 (CH2)5 37.8 4 62.6 2

W206 (CH2)6 27.2 3 65.5 4

Sample Group (R2) IC50 /nM Rank (IC50) EC50 /nM Rank (EC50)

W70 1-Br 128 9 147.4 6

W71 2-Br 78.4 6 95.6 5

K15 3-Br 6.98 1 62.8 3

a Potency is defined as the dose required to achieve a given level of inhibition of JH biosynthesis or activation of

AstR. 1 EC50 was determined by AstR activation assay in CHO-WTA11 cells (details seen Fig. 2). 2 IC50 values are

measured using RCA (CA from day 7 mated female).

183

Fig. 6.1. The effect of topical application of K15 and W206 on JH biosynthesis (A) and

oocyte growth (B). AST analogs were diluted in 20% DMSO and 80% acetone to a

concentration of 10-3 M, and applied 5 µl to the dorsal abdomen of each D. punctata female on

day 0. The effect of AST analogs on JH biosynthesis and basal oocyte growth was determined

from day 4 to day 8.

184

References

Donly, B.C., Ding, Q., Tobe, S.S., Bendena, W.G., 1993. Molecular cloning of the gene for the allatostatin

family of neuropeptides from the cockroach Diploptera punctata. Proceedings of the National Academy of

Sciences of the United States of America 90, 8807-8811.

Garside, C.S., Hayes, T.K., Tobe, S.S., 1997a. Degradation of dip-allatostatins by hemolymph from the

cockroach, Diploptera punctata. Peptides 18, 17-25.

Garside, C.S., Hayes, T.K., Tobe, S.S., 1997b. Inactivation of dip-allatostatin 5 by membrane preparations

from the cockroach Diploptera punctata. General and comparative endocrinology 108, 258-270.

Rankin, S.M., Stay, B., 1987. Distribution of allatostatin in the adult cockroach, Diploptera punctata and

effects on corpora allata in vitro. Journal of Insect Physiology 33, 551-558.

Tobe, S.S., Zhang, J.R., Bowser, P.R.F., Donly, B.C., Bendena, W.G., 2000. Biological activities of the

allatostatin family of peptides in the cockroach, Diploptera punctata, and potential interactions with receptors.

Journal of Insect Physiology 46, 231-242.

185

Copyright Acknowledgements

The author would like to thank:

Elisabeth Marchal, Ekaterina F. Hult, and Stephen S. Tobe for permitting to use the following

material: (1) Diploptera punctata as a model for studying the endocrinology of arthropod

reproduction and development. Gen Comp Endocr , 2013, 188, 85-93. (2) Characterization of the

Juvenile Hormone pathway in the viviparous cockroach, Diploptera punctata. PloS one,

accepted. (3) Identification and characterization of the NMDA receptor and its role in regulating

reproduction in the cockroach, Diploptera punctata, 2014, paper submitted to ‘Journal of

experimental biology’

Elisabeth Marchal, Ekaterina F. Hult, Sven Zels, Jozef Vanden Broeck, and Stephen S. Tobe for

permit to use the following material: Mode of action of allatostatins in the regulation of juvenile

hormone biosynthesis in the cockroach, Diploptera punctata, Insect Biochem Mol Biol, 2014, 54,

61-68.

And Journal: ‘Insect biochemistry and molecular biology’ for allowing the author to reprint the

article in the institutions open scholarly website (institutional repository).