bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1 Insight into mosquito GnRH-related neuropeptide receptor 2 specificity revealed through analysis of naturally occurring and 3 synthetic analogs of this neuropeptide family 4 Azizia Wahedi1, Gerd Gäde2* and Jean-Paul Paluzzi1* 5 1 Department of Biology, York University, 4700 Keele Street, Toronto, Ontario, M3J1P3, 6 Canada 7 2 Department of Biological Sciences, University of Cape Town, Private Bag, Rondebosch 8 7701, South Africa 9

10 *corresponding authors 11 Prof. Jean-Paul Paluzzi, Department of Biology, York University, 4700 Keele Street, 12 Toronto, Ontario, M3J1P3, Canada, Email: [email protected] 13 Prof. Gerd Gäde, Department of Biological Sciences, University of Cape Town, Private Bag, 14 Rondebosch 7701, South Africa, Email: [email protected] 15 16 17 18 19 20 21 Keywords: 22 GnRH-related neuropeptides, G protein-coupled receptor, structure activity relationships, 23 adipokinetic hormone (AKH), corzazonin (CRZ), AKH/CRZ-related peptide (ACP) 24

1 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

25 Abstract 26 Adipokinetic hormone (AKH), corzazonin (CRZ) and the AKH/CRZ-related peptide (ACP)

27 are peptides considered homologous to the vertebrate gonadotropin-releasing hormone

28 (GnRH). All three Aedes aegypti GnRH-related neuropeptide receptors have been

29 characterized and functionally deorphanized, which individually exhibit high specificity for

30 their native ligands, which prompted us to investigate the contribution of ligand structures in

31 conferring receptor specificity. In the current study, we designed a series of analogs based on

32 the native ACP sequence in A. aegypti and screened them against the ACP receptor using a

33 heterologous system to identify critical residues required for receptor activation. Specifically,

34 analogs lacking the carboxy-terminal amidation, replacing aromatic residues, as well as

35 truncated analogs were either completely inactive or had very low activities even at high

36 concentration. The next most critical residues were the polar threonine in position 3 and the

37 blocked amino-terminal pyroglutamate, with activity of the latter partially recovered using an

38 alternatively blocked analog. ACP analogs with alanine substitutions at position 2 (valine), 5

39 (serine), 6 (arginine) and 7 (aspartic acid) positions were less detrimental as were

40 replacements of charged residues. Interestingly, replacing asparagine with an alanine at

41 position 9, creating a C-terminal WAA-amide, resulted in a 5-fold more active analog which

42 may be useful as a lead superagonist compound. Similarly, we utilized this high-throughput

43 approach against an A. aegypti AKH receptor (AKHR-IA) testing a number of mostly

44 naturally-occurring AKH analogs from other to determine how substitutions of

45 specific amino acids in the AKH ligand influences receptor activation. AKH analogs having

46 single substitutions compared to the endogenous A. aegypti AKH revealed position 7 (serine)

47 was well tolerated whereas changes to position 6 (proline) had pronounced effects, with

48 receptor activity compromised nearly ten-fold. Substitution of position 3 (threonine) or

49 analogs with combinations of substitutions were quite detrimental with a significant decrease

2 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

50 in AKHR-IA activation. Interestingly, analogs with an asparagine residue at position seven

51 displayed improved receptor activation compared to the native mosquito AKH. Collectively,

52 these results advance our understanding of how two GnRH-related systems in A. aegypti

53 sharing the most recent evolutionary origin sustain independence of function and signalling

54 despite their relatively high degree of ligand and receptor homology.

55

56

3 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

57 Introduction 58 In invertebrates there exist three neuropeptide systems of which the mature peptides as well

59 as their cognate G protein-coupled receptors (GPCRs) are structurally similar, and

60 collectively, they are related to the vertebrate gonadotropin releasing hormone (GnRH)

61 system and suggested to form a large peptide superfamily (Gäde et al., 2011; Hansen et al.,

62 2010; Li et al., 2016; Roch et al., 2011). These neuropeptide systems are called the

63 adipokinetic hormone (AKH)/red pigment-concentrating hormone (RPCH) family, the

64 corazonin (CRZ) family and the structurally intermediate adipokinetic hormone/corazonin-

65 related peptide (ACP) family. AKHs are primarily or exclusively produced in neurosecretory

66 cells of the corpus cardiacum, and a major function is mobilization of energy reserves stored

67 in the fat body, thus providing an increase of the concentration of diacylglycerols, trehalose

68 or proline for locomotory active phases in the haemolymph. In accordance with this function,

69 AKH has been shown to activate the enzymes glycogen phosphorylase and triacylglycerol

70 lipase and transcripts of the AKHR are most prominently found in fat body tissue (see

71 reviews by (Gäde, 2004; Gäde and Auerswald, 2003; Marco and Gäde, 2019a). As with most

72 endocrine regulatory peptides, additional functions such as inhibition of anabolic processes

73 (protein and lipid syntheses), involvement in oxidative stress reactions and egg production

74 inter alia are known making AKH a truly pleiotropic hormone (see reviews by (Kodrík, 2008;

75 Kodrík et al., 2015). This is also true for CRZ which is mainly synthesized in neuroendocrine

76 cells of the pars lateralis of the protocerebrum and released via the corpora cardiaca (Predel et

77 al., 2007). Although originally described as a potent cardiostimulatory peptide (Veenstra,

78 1989), it does not generally fulfil this role but is known for many additional functions such as

79 involvement in the (1) release of pre-ecdysis and ecdysis triggering hormones, (2) reduction

80 of silk spinning rates in the silk , (3) pigmentation events (darkening) of the epidermis in

81 locusts during gregarization and (4) regulation of caste identity in an ant species (Gospocic et

4 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

82 al., 2017; Kim et al., 2004; Tanaka et al., 2002; Tawfik et al., 1999). The functional role of

83 ACP is less clear. All previous studies had not found a clear-cut function for this peptide until

84 work by Zhou and colleagues claimed that ACP in the cricket Gryllus bimaculatus regulates

85 the concentration of carbohydrates and lipids in the haemolymph (Zhou et al., 2018).

86 Experiments in the current study have been conducted with the mosquito species Aedes

87 aegypti for a number of reasons. Primarily because of being such an infamous disease vector

88 for pathogens such as Yellow and Dengue fever, Chikungunya and, as latest addition, Zika

89 arboviruses, all of which summarily are responsible for affecting approximately 50-100

90 million people per year, with the majority of these attributed to Dengue virus infection (Baud

91 et al., 2017; Price et al., 2015; Saiz et al., 2016). Secondly, knowledge on the interaction of

92 the ligands with their respective receptor are thought to be very helpful for drug research

93 using specific GPCRs as targets for development of selective biorational insecticides

94 (Audsley and Down, 2015; Hill et al., 2018; Verlinden et al., 2014). Another reason was that

95 each of the three neuropeptide systems has already been partially investigated, thus, this

96 study could build on previous research data. With respect to the peptides, these were first

97 predicted from the genomic work on A. aegypti (Nene et al., 2007). The presence of mature

98 AKH and corazonin, but not of ACP, was shown by direct mass profiling (Predel et al.,

99 2010). The AKH and ACP precursors have been cloned (Kaufmann et al., 2009) as have one

100 CRZ receptor (CRZR) and various variants of the AKHR and ACPR (Kaufmann et al., 2009;

101 Oryan et al., 2018; Wahedi and Paluzzi, 2018). Each receptor is very selective and responds

102 only to its cognate peptide, thus there are indeed three completely separate and independently

103 working neuroendocrine systems active in this mosquito species.

104 With respect to the ACP system in A. aegypti, expression studies revealed transcripts of the

105 major ACPR form as well as the ACP precursor were enriched during development in adults,

106 specifically during day one and four after eclosion (Kaufmann et al., 2009; Wahedi and

5 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

107 Paluzzi, 2018). On the tissue level, ACP precursor transcripts are mainly present in the

108 head/thorax region (Kaufmann et al., 2009) and, more specifically, within the brain and

109 thoracic ganglia (Wahedi and Paluzzi, 2018) and the ACPR transcripts are mainly detected in

110 the central nervous system with significant enrichment in the abdominal ganglia, particularly

111 in males (Wahedi and Paluzzi, 2018). The AKH precursor transcripts are present in

112 head/thorax region of pupae and adult females and in the abdomen of males (Kaufmann et al.,

113 2009). Expression of AKHR was shown in all life stages (Kaufmann et al., 2009), but similar

114 to the ACPR and ACP transcripts, AKHR transcript variants were found to be significantly

115 enriched in early adult stages (Oryan et al., 2018). On the tissue level, highest expression in

116 adults was found in thorax and abdomen with low levels detected in ovaries (Kaufmann et al.,

117 2009). Similar expression profiles were determined recently by Oryan et al. (2018) using RT-

118 qPCR: expression in head, thoracic ganglia, accessory reproductive tissues and carcass of

119 adult females as well as in abdominal ganglia and specific enrichment in the carcass and

120 reproductive organs of adult males. Similarly, CRZR was mostly expressed during later

121 development, including pupal and adult stages. In terms of tissue expression profiles, this

122 was examined in adult stages where CRZR was found in the heads and thoracic ganglia of

123 both sexes as well as in ovary, abdominal ganglia and Malpighian tubules in females and in

124 the testes and carcass of males (Oryan et al., 2018).

125 Thus, although quite a substantial body of information on the three signalling systems in A.

126 aegypti is known, we lack knowledge of how the various ligands interact with their cognate

127 receptor. For instance, which amino acid residue in the ligand is specifically important for

128 this interaction? By replacing each residue successively with simple amino acids such as

129 glycine or alanine one can probe into the importance of amino acid side chains and their

130 characteristics in relation to facilitating receptor specificity. Such structure-activity studies

131 have been conducted on the AKH system in a number of insects and a crustacean either by

6 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

132 measuring physiological actions in vivo (Gäde and Hayes, 1995; Ziegler et al., 1998) or in a

133 cellular mammalian expression system in vitro (Caers et al., 2012; Marchal et al., 2018;

134 Marco et al., 2017). In conjunction with nuclear magnetic resonance (NMR) data on the

135 secondary structure of AKHs (Nair et al., 2001; Zubrzycki and Gäde, 1994) and the

136 knowledge of the receptor sequence, molecular dynamic methods can devise models how the

137 ligand is interacting with its receptor (Jackson et al., 2018; Mugumbate et al., 2013).

138 Since no such data is available for any ACP system, the objective of the current study was to

139 fill this knowledge gap by determining critical amino acids of the ACP neuropeptide

140 necessary for activation of its cognate receptor, ACPR. We therefore designed a series of

141 analogs based on the endogenous A. aegypti ACP sequence and screened them using an in

142 vitro receptor assay system to determine the crucial residues for ACPR activation. Given the

143 closer evolutionary relationship between the ACP and AKH systems in (Hansen et

144 al., 2010; Hauser and Grimmelikhuijzen, 2014; Tian et al., 2016; Zandawala et al., 2018), a

145 second aim of the current study was to advance our knowledge and understanding of critical

146 residues and properties of uniquely positioned amino acids necessary for the specificity of the

147 AKH ligand to its receptor, AKHR-IA. Here, we took advantage of naturally occurring AKHs

148 from other insects to examine amino acid substitutions and determine the consequences onto

149 AKHR-IA activation using the same in vitro heterologous assay. Collectively, determining

150 indispensable amino acid residues of these two GnRH-related mosquito neuropeptides that

151 are necessary for activation of their prospective receptors will help clarify how these

152 evolutionarily-related systems uphold specific signalling networks avoiding cross activation

153 onto the structurally related, but functionally distinct signalling system.

154 155

7 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

156 Materials and Methods 157 158 ACP and AKH Receptor Expression Construct Preparation

159 A mammalian expression construct containing the ACPR-I ORF (hereafter denoted as

160 ACPR) with Kozak translation initiation sequence had been previously prepared in

161 pcDNA3.1+ (Wahedi and Paluzzi, 2018); however, this earlier study revealed that coupling of

162 the ACPR with calcium mobilization in the heterologous expression system was not very

163 strong, even following application of high concentrations of the native ACP ligand.

164 Therefore, to improve the bioluminescent signal linked to calcium mobilization following

165 receptor activation, which was required in order to provide greater signal to noise ratio when

166 testing ACP analogs with substitutions leading to intermediate levels of activation of the

167 ACPR, we created a new mammalian expression construct using the dual promoter vector,

168 pBudCE4.1. Specifically, we utilized a cloning strategy as reported previously (Paluzzi et al.,

169 2015) whereby the ACPR was inserted into the multiple cloning site (MCS) downstream of

170 the CMV promoter in pBudCE4.1 using the sense primer SalI-KozakACPR, 5’-

171 GGTCGACGCCACCATGTATCTTTCGG-3’ and anti-sense primer XbaI-ACPR, 5’-

172 TCTAGATTATCATCGCCAGCCACC-3’, which directionally inserted the ACPR within the

173 MCS downstream of the CMV promoter. Secondly, the murine homolog (Gα15) of the

174 human promiscuous G protein, Gα16, which indiscriminately couples a wide variety of

175 GPCRs to calcium signaling (Offermanns and Simon, 1995) was inserted into the MCS

176 downstream of the EF1-α promoter in pBudCE4.1 using the sense primer NotI-Gα15, 5’-

177 GCGGCCGCCACCATGGCCCGGTCCCTGAC-3’ and anti-sense primer BglII-Gα15, 5’-

178 AGATCTTCACAGCAGGTTGATCTCGTCCAG-3’, which directionally inserted the

179 murine Gα15 within the MCS downstream of the EF1-α promoter.

8 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

180 The AKHR mammalian expression construct used in the current study, specifically AKHR-

181 IA which was the receptor isoform that exhibited the greatest sensitivity to its native AKH

182 ligand, was prepared previously in pcDNA3.1+ (Oryan et al., 2018) and coupled strongly with

183 calcium signaling in the heterologous system so there was no need utilize the dual promoter

184 vector containing the promiscuous Gα15 as was necessary for ACPR. For each construct, an

185 overnight liquid culture of clonal recombinant bacteria containing the appropriate plasmid

186 construct was grown in antibiotic-containing LB media and used to isolate midiprep DNA

187 using the PureLink Midiprep kit (Life Technologies, Burlington, ON). Midiprep samples

188 were then quantified and sent for sequencing as described previously (Oryan et al., 2018;

189 Wahedi and Paluzzi, 2018).

190

191 Cell culture, transfections, and calcium bioluminescence reporter assay

192 Functional activation of the AedaeACPR and AedaeAKHR-IA receptors was carried out

193 following a previously described mammalian cell culture system involving a Chinese hamster

194 ovary (CHO)-K1 cell line stably expressing aequorin (Paluzzi et al., 2012). Cells were grown

195 in DMEM:F12 media containing 10% heat-inactivated fetal bovine serum (Wisent, St. Bruno,

196 QC), 200μg/mL Geneticin, 1x antimycotic-antibiotic to approximately 90% confluency, and

197 were transiently transfected using either Lipofectamine LTX with Plus Reagent or

198 Lipofectamine 3000 transfection systems (Invitrogen, Burlington, ON) following

199 manufacturer recommended DNA to transfection reagent ratios. Cells were detached from the

200 culture flasks at 48-hours post-transfection using 5mM EDTA in Dulbecco’s PBS. Cells were

201 then prepared for the receptor functional assay following a procedure described previously

202 (Wahedi and Paluzzi, 2018). Stocks of peptide analogs of ACP and AKH were used for

203 preparation of serial dilutions all of which were prepared in assay media (0.1% BSA in

204 DMEM:F12) and loaded in quadruplicates into 96-well white luminescence plates (Greiner

9 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

205 Bio-One, Germany). Cells were loaded with an automatic injector into each well of the plate

206 containing different peptide analogs at various concentrations as well as negative control

207 (assay media alone) and positive control wells (50µM ATP). Immediately after injection of

208 the cells, luminescence was measured for 20 seconds using a Synergy 2 Multi-Mode

209 Microplate Reader (BioTek, Winooski, VT, USA). Calculations, including determination of

210 EC50 values, were conducted in GraphPad Prism 7.02 (GraphPad Software, San Diego, USA)

211 from dose-response curves from 3-4 independent biological replicates.

212 Synthetic Peptides and Analogs

213 Peptide analogs based on the native ACP sequence were designed and synthesized by Pepmic

214 Co., Ltd. (Suzhou, China) at a purity of over 90%. All synthetic peptides were initially

215 prepared as stock solutions at a concentration of 1mM in dimethyl sulfoxide. All stocks of

216 AKH analogs were prepared similarly to the ACP synthetic peptide analogs and

217 were commercially synthesized by either Pepmic Co., Ltd. (Suzhou, China) or by Synpeptide

218 Co. (Shanghai, China) at a purity of over 90%. Specific sequence information for each ACP

219 and AKH peptide analog used in this study is provided in Tables 1 and 2, respectively.

220 221

10 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

222 Results 223 ACP synthetic analog activity on mosquito ACPR 224 Analogs of native ACP with single alanine substitutions at each amino acid were synthesized

225 and evaluated by examining their functional activation of the A. aegypti ACP receptor

226 (ACPR) expressed using a heterologous system. Carboxy-terminal amidation of ACP was

227 found to be critical for bioactivity since the free acid ACP analog elicited no activation of

228 ACPR (Figure 1A) detectable up to the highest tested concentration of 10uM (Table 1). In

229 addition, analogs replacing either of the two aromatic residues of ACP, including the

230 phenylalanine in the fourth position or tryptophan in the eighth position, were similarly

231 inactive on the ACPR (Figure 1A) with no detectable response by any of the tested

232 concentrations (Table 1). Further examination of single residue alanine substitutions revealed

233 that the next most critical residue for retaining ACPR functional activation was the polar

234 threonine in the third position, which upon replacement with alanine resulted in a ~551-fold

235 reduced activity compared to native ACP (Figure 1A, Table 1). A comparable reduction of

236 activity, approximately 411-fold, was measured with an ACP analog that had the first residue,

237 the blocked pyroglutamate in the native ACP, substituted by alanine. If the substitution,

238 however, was also made with a blocked compound, N-Acetyl-alanine instead of pGlu, the

239 analog was 10-times more active as the non-blocked one, thus the reduction in activity was

240 only about 44-fold (Figure 1A). Interestingly, substitutions with analogs containing alanine at

241 the second (valine), fifth (serine), sixth (arginine) and seventh (aspartic acid) positions, had

242 noticeably less impact on the reduction of activity (Figure 1B), by ~70-fold, ~45-fold, ~23-

243 fold, and ~16-fold, respectively. Unexpectedly, the analog containing an alanine substitution

244 for the asparagine in position nine resulted in a more potent ACPR activation, leading to a

245 ~5-fold greater activity compared to the native ACP peptide (Figure 1B). C-terminally

246 truncated analogs, which lacked either the naturally occurring alanine residue in the tenth

247 residue position or the last two C-terminal residues including asparagine and alanine in

11 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

248 positions nine and ten, were almost completely inactive, demonstrating ~434-fold and ~410-

249 fold reduced ACPR activation compared to native ACP (Figure 1C). A C-terminally

250 extended analog ending with a glycine on the C-terminus elicited a small decrease of activity

251 by about ~6-fold (Figure 1C). Lastly, the internally truncated ACP analog lacking a valine in

252 the second position also was without any effect, similar to the two aromatic residue

253 substituted analogs or the non-amidated ACP analog (Figure 1C).

254 Activity of natural arthropod AKH analogs on mosquito AKHR-IA 255 Taking advantage of the variety of natural AKH analogs in arthropods, we examined a subset

256 of AKH peptides (see Table 2) with unique substitutions compared to the endogenous

257 mosquito AKH peptide and determined their activity on the AKH receptor, A. aegypti

258 AKHR-IA. From this analysis, the least effective analog tested was the C-terminally

259 truncated peptide, Lacol-AKH-7mer, whose activity was nearly completely abolished with

260 only ~7% AKHR-IA activation at the highest tested concentration (10μM), which represents

261 a reduced efficacy of at least five orders of magnitude (Figure 2A). The next least active

262 AKH analog was the cricket peptide, Grybi-AKH, which has five residues substituted over

263 the length of the octapeptide compared to the mosquito AKH (see Table 2). Specifically,

264 leucine and threonine in positions two and three are replaced by valine and asparagine,

265 respectively. Further, threonine, proline and serine in positions five to seven are replaced

266 with serine, threonine and glycine, respectively. As a result of these combined substitutions,

267 Grybi-AKH activity on the A. aegypti AKHR-IA was compromised by over three orders of

268 magnitude (1032-fold reduced activity) compared to the endogenous mosquito AKH (Table 2

269 and Figure 2A). We next tested the AKH from the dragonfly, Libellula auripennis, which

270 allowed us to isolate a subset of the substitutions of Grybi-AKH, specifically positions two

271 and three alone. The activity of Libau-AKH was improved nearly two-fold over Grybi-AKH;

272 however, the activity of this peptide on the A. aegypti AKHR-IA was still reduced by 548-

12 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

273 fold compared to Aedae-AKH (Table 2). The next AKH analog examined was that from the

274 firebug Pyrrhocoris apterus (Pyrap-AKH), which has two substitutions in positions three and

275 seven, both asparagine, for the natural threonine and serine, respectively, found in A. aegypti

276 AKH. The activity of Pyrap-AKH was found to be improved compared to Libau-AKH and

277 Grybi-AKH by 1.6 and 3-fold, respectively. However, these substitutions resulted in the

278 activity of Pyrap-AKH on the A. aegypti AKHR-IA being compromised by 337-fold relative

279 to the native AKH (Figure 2A). To better understand the importance of N-terminal residues,

280 we next tested a second AKH from Odonata, this analog from Erythemis simplicicollis

281 (Erysi-AKH) where the second position valine found in Libau-AKH is instead leucine, which

282 is conserved with the A. aegytpi AKH. Thus, the activity of this AKH analog having only a

283 single amino acid substitution in the third position (threonine to asparagine; see Table 2)

284 resulted in Erysi-AKH eliciting activity on the A. aegypti AKHR-IA which was 310-fold less

285 compared to native mosquito AKH (Figure 2A). When testing an AKH analog having a

286 single equivalent substitution in only position seven (serine to asparagine), as is naturally

287 available in the American cockroach peptide Peram-CAH-II, this substitution was completely

288 tolerated since effectively no difference in the activity of this analog on the AKHR-IA

289 relative to the A. aegypti AKH was detected (Figure 2B). Indeed, we could next verify the

290 criticalness of the third position threonine residue by examining the activity of another

291 cockroach AKH analog, V2-Peram-CAH-II (a synthetic analog and not naturally occuring in

292 P. americana), which has a second position substitution of valine for leucine along with the

293 seventh position substitution of asparagine for serine (Table 2). The activity of V2-Peram-

294 CAH-II was found to be much improved, albeit still 69-fold less effective on the AKHR-IA

295 compared to the native A. aegypti AKH (Figure 2B). Since it was demonstrated that the C-

296 terminal tryptophan was fundamental for bioactivty, with truncated analogs having little or no

297 activity (see above for Lacol-AKH-7mer), we were interested to examine if extension of the

13 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

298 C-terminus may also equally impact the activity of the AKH analog on the mosquito A.

299 aegypti AKHR-IA. To do this, we utilized the AKH analog from the domestic silkmoth,

300 Bombyx mori (Bommo-AKH), which contains an extended C-terminus comprised of a

301 glycine and glutamine residue (resulting in a decapeptide). Additionally, Bommo-AKH has a

302 seventh position substitution of glycine for serine; however, considering the extension on the

303 C-terminus as well as the substitution in position seven, this analog was surprinsingly quite

304 tolerated with only 11-fold reduced activity on the A. aegypti AKHR-IA (Figure 2B). Finally,

305 examining naturally occuring analogs of this peptide family containing single substitutions in

306 the C-terminal region, but sparing the eighth critical tryptophan residue, the findings

307 demonstrate that these changes are quite tolerated. It was noted above that the cockroach

308 AKH, Peram-CAH-II, which has a seventh position serine substituted with asparagine, had

309 near identical efficacy to the native mosquito AKH. Similarly, an AKH analog from the

310 sphingid moth, eson (Hipes-AKH-I), which has a serine substitution for proline in

311 the sixth position, also was well tolerated with only a 8-fold reduced activity on the A.

312 aegypti AKHR-IA (Figure 2B). Lastly, using the AKH analog from the black horse fly,

313 Tabanus atratus (Tabat-AKH), the results demonstrated that another substitution in the

314 seventh position, in this case a glycine for the naturally occurring serine in Aedae-AKH, is

315 indeed well tolerated since the activity of this analog matched and partially exceeded (~30%

316 improved efficacy) the activity of the native mosquito AKH (see Table 2; Figure 2B).

317

14 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

318 Table 1. Primary structure of the synthetic ACP analogs designed and tested on the A. 319 aegypti ACP receptor (AedaeACPR). Using heterologous expression of ACPR, the half 320 maximal effective concentration for each of the analogs was determined as well as the 321 corresponding reduction in activity, highlighting how a particular residue substitution or other 322 modification is tolerated in this system. ND = represents no determined EC50 value due to 323 minimal or no detectable activity of the analog when tested up to 10μM. 324 Fold-reduction Activity on A. in activity Peptide/Analog: Primary structure of analog: aegypti ACPR relative to (EC ) 50 AedaeACP AedaeACP pQVTFSRDWNAamide 2.394E-08 1.0 AedaeACP (A1) AVTFSRDWNAamide 1.576E-06 411.1 AedaeACP (A2) pQATFSRDWNAamide 7.052E-07 69.8 AedaeACP (A3) pQVAFSRDWNAamide 2.26E-05 550.8 AedaeACP (A4) pQVTASRDWNAamide ND >1000 AedaeACP (A5) pQVTFARDWNAamide 1.038E-06 44.9 AedaeACP (A6) pQVTFSADWNAamide 6.872E-07 23.3 AedaeACP (A7) pQVTFSRAWNAamide 3.863E-07 15.9 AedaeACP (A8) pQVTFSRDANAamide ND >1000 AedaeACP (A9) pQVTFSRDWAAamide 4.96E-09 0.2 AedaeACP (-OH) pQVTFSRDWNA-OH ND >1000 AedaeACP (A1b) Acetyl-AVTFSRDWNAamide 5.272E-07 43.6 AedaeACP (1-9) pQVTFSRDWNamide 4.153E-06 434.2 AedaeACP (1-8) pQVTFSRDWamide 2.156E-06 410.1 AedaeACP (1,3-10) pQTFSRDWNAamide ND >1000 AedaeACP (+G) pQVTFSRDWNAGamide 1.113E-07 5.6 325 326

15 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

327 Figure 1.

A1 A1b A3 A ACP A4 A8 OH

1001.00

0.75 75

0.50 50 Response 0.25 25

Normalized Luminescent 0.00 0 -10 -9 -8 -7 -6 -5 Peptide concentration (log M)

ACP A2 A5 B A6 A7 A9

1001.00

0.75 75

0.50 50 Response 0.25 25

Normalized Luminescent 0.00 0 -10-9-8-7-6-5 Peptide concentration (log M)

1-8 1-9 C ACP 1,3-10 +G

1001.00

0.75 75

0.50 50 Response 0.25 25

Normalized Luminescent 0.00 0 -10-9-8-7-6-5 Peptide concentration (log M) 328

16 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

329 Table 2. Primary structure of select naturally occurring AKH analogs from insects and their 330 activity on the A. aegypti AKHR-IA receptor (AKHR-IA). Using heterologous expression of 331 AKHR-IA, the half maximal effective concentration for each of the analogs was determined 332 as well as the corresponding reduction in activity, highlighting how a particular single residue 333 substitution, a combination of substitutions or other modifications are tolerated in this system. 334 ND = represents no determined EC50 value due to minimal or no detectable activity of the 335 analog when tested up to 10μM. 336 Activity on Fold-reduction in Primary structure of A. aegypti Peptide/Analog: Species origin: activity relative to analog: AKHR-IA Aedae-AKH-I (EC50) Aedae-AKH-I Aedes aegypti pQLTFTPSWamide 2.159E-09 1.0 Lacol-AKH-7mer Laconobia oleracea pQLTFTSS-amide ND >10,000 Grybi-AKH Gryllus bimaculatus pQVNFSTGWamide 2.23E-06 1031.5 Libau-AKH Libellula auripennis pQVNFTPSWamide 1.18E-06 547.9 Pyrap-AKH Pyrrhocoris apterus pQLNFTPNWamide 7.269E-07 336.7 Erythemis Erysi-AKH simplicicollis pQLNFTPSWamide 6.69E-07 309.9 *Periplaneta V2-PeramCAH-II americana pQVTFTPNWamide 1.485E-07 68.8 Bommo-AKH Bombyx mori pQLTFTPGWGQamide 2.362E-08 10.9 Hipes-AKH-I Hippotion eson pQLTFTSSWamide 1.647E-08 7.6 Periplaneta Peram-CAH-II americana pQLTFTPNWamide 2.226E-09 1.0 Tabat-AKH Tabanus atratus pQLTFTPGWamide 1.522E-09 0.7 * Denotes a synthetic analog that does not occur in Periplaneta americana.

337

338 339

17 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

340 Figure 2.

A Aedae AKH Libau AKH Lacol AKH 7-mer Pyrap AKH Grybi-AKH Erysi AKH

100

75

50 Response 25

Normalized Luminescent Normalized 0 -11 -10 -9 -8 -7 -6 -5 Peptide concentration (log M)

B Aedae AKH Hipes AKH-I V2 Peram CAH-II Peram CAH-II Bommo AKH Ta ba t A KH

100

75

50 Response 25

Normalized Luminescent 0 -11 -10 -9 -8 -7 -6 -5 Peptide concentration (log M) 341 342 343

18 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

344 Discussion 345 In light of the high fidelity of the three GnRH-related systems (Hansen et al., 2010; Oryan et

346 al., 2018; Wahedi and Paluzzi, 2018), the current study set out to examine for the first time

347 the ligand structure-activity relationship for an insect ACP receptor. This is of utmost

348 importance in order to gain insight on the specific structural features of the ACP system

349 given that extensive studies have been carried out previously on AKH receptors in a variety

350 of species using natural and synthetic analogs of this neuropeptide family (Caers et al., 2012;

351 Fox and Reynolds, 1991; Gäde et al., 2000; Gäde, 1992; Gäde, 1993; Gäde et al., 2016;

352 Keeley et al., 1991; Lee and Goldsworthy, 1996; Lee et al., 1996; Lee et al., 1997; Marco and

353 Gäde, 2015; Marco and Gäde, 2019b; Poulos et al., 1994; Velentza et al., 2000; Ziegler et al.,

354 1991; Ziegler et al., 1998). By examining the activity of a variety of ACP analogs on the

355 activation of the A. aegytpi ACP receptor (ACPR), including single alanine substitutions

356 along with truncated or extended analogs, a few observations can be highlighted. Firstly, the

357 overall charge of the peptide does not appear to play a role with regards to its influence on

358 receptor activation since substitution of the basic (arginine) or acidic (aspartic acid) residues

359 did not lead to a highly detrimental effect. Specifically, one would assume that the ACP

360 receptor may prefer a neutral ligand given that the native ACP has no net charge (i.e. is

361 neutral) since it has both a single basic and acidic residue. However, the modified analogs

362 which replace the basic (position 6) or acidic (position 7) residues with alanine, create

363 negatively or positively charged molecules, respectively, with both being quite active and so

364 neither proved detrimental substitutions. Secondly, with the exception of the aromatic

365 tryptophan residue in the eighth position, the C-terminal residues of ACP do not appear to be

366 highly important since alanine substitutions in positions five through seven only marginally

367 impacted activity between ~16 to 45-fold whereas alanine substitution in position nine

368 resulted in this analog having five-fold increased activity. Amidation of the C-terminal

19 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

369 residue was critical since the free acid analog exhibited no activity; however, it is noteworthy

370 that the C-terminally extended peptide having an amidated glycine was also quite effective

371 with ACPR activation compromised by only 6-fold. Previous studies examining free acid

372 analogs of AKH (including both in vitro and in vivo assays) have demonstrated that in some

373 cases the absence of a blocked C-terminus is less critical (Caers et al., 2012; Caers et al.,

374 2016; Lee et al., 1996), whereas other investigations have revealed the presence of an

375 amidated C-terminus to be very important with poor or no responses with analogs lacking this

376 feature (Gäde and Hayes, 1995; Gäde, 1990). These observed differences in response to key

377 structural elements, including the normally amidated C-terminus of AKH, may reflect the

378 occurrence of inter- and intra-species receptor subtypes which may be differentially sensitive

379 to these modifications (Caers et al., 2016; Gäde and Hayes, 1995), and could also point to the

380 well documented phenomenon that some species have AKH receptors that are more

381 promiscuous whereas others are very strict in the structural characteristics of their particular

382 ligand (Caers et al., 2012; Gäde, 1990; Marchal et al., 2018). This latter scenario is more in-

383 line with observations for the A. aegypti ACP system where only a single functional receptor

384 variant occurs (Wahedi and Paluzzi, 2018), which we show herein does not tolerate the

385 absence of the C-terminal amidation and is thus of high importance for the ACP system.

386 Thirdly, both aromatic residues were essential for activity on the ACPR while N-terminal

387 residues of the peptide, including mainly threonine and pyroglutamic acid, were also highly

388 critical since alanine substitutions were not tolerated leading to significantly reduced activity

389 of these analogs relative to the native ACP peptide. However, alternatively blocking this

390 alanine substitution on the N-terminus through acetylation improved the activity of this

391 analog by 10-fold. Similar observations have been made with AKH analogs tested using both

392 in vivo and in vitro assays whereby complete removal of the N-terminal pyroglutamate

393 abolishes activity of the AKH analog whereas the alternatively blocked N-terminus (i.e. N-

20 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

394 Acetyl-Ala), or even simply glutamate or glutamine that can spontantenously undergo

395 cyclization (Schilling et al., 2008), were shown to have improved activity, albeit lower than

396 the native AKH (Caers et al., 2016; Gäde and Hayes, 1995; Marco and Gäde, 2019b). Lastly,

397 although alanine substitution for valine in the second position led to a relatively minor effect

398 on ACPR activation (~70-fold reduced activity), the internally truncated analog omitting

399 valine within the N-terminal region elicited no activation of the ACPR, indicating the spacing

400 between the more critical residues, namely pyroglutamate and polar threonine, is essential for

401 ACP peptide activity on the ACPR. Thus, perhaps in common with AKH structural features,

402 the switch between adjacent hydrophobic and hydrophilic residues is necessary in order to

403 properly ‘fit’ and bind with its particular receptor (Gäde and Hayes, 1995) so that not only is

404 the ACP analog shorter by one amino acid (a nonapeptide), but moreover, the entire structural

405 configuration of the peptide is disrupted and the whole sequence is no longer in sync with its

406 receptor. This corroborates with studies that have examined the conservation of the ACP

407 primary structures across insects from different orders (including Diptera, ,

408 Coleoptera, Hymenoptera and Hemiptera) as well as from various species within the same

409 order (seventeen species of Coleoptera), where the N-terminal pentamer sequence, when the

410 ACP system is present, is nearly completely conserved across these various species with a

411 consensus motif of pQVTFS-, whereas significant sequence variability occurs within the C-

412 terminus of the ACP sequence with deca-, nona- and even dodecapeptides reported (Hansen

413 et al., 2010; Veenstra, 2019).

414 Considering the resultant activation of the mosquito AKH receptor in response to the various

415 mostly naturally-occurring insect AKH analogs examined in the current study, these findings

416 indicate that these bioanalogs can be grouped into distinct categories based on their

417 determined potencies on the A. aegypti AKHR-IA. The first group includes natural or

418 synthetic AKH analogs having substitutions in critical positions, including the absence of the

21 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

419 C-terminal tryptophan that occurs in the heptapeptide, Lacol-AKH-7mer, which is not

420 tolerated whatsoever since activity is completely abolished and is therefore of greatest

421 importance for AKH receptor activation. Observations correlating with these findings have

422 been made in various studies utilizing in vitro heterologous assays as well as in vivo

423 bioassays monitoring lipid- and/or carbohydrate-mobilizing actions of AKH analogs.

424 Specifically, removal or replacement of the critical aromatics that are archetypal features of

425 AKH neuropeptides found in positions four and eight demonstrates the absolute requirement

426 of these features for receptor activation in heterologous systems or in vivo biological activity

427 (Gäde, 1992; Gäde, 1993; Marco and Gäde, 2015; Marco and Gäde, 2019b).

428 The second group in the current study includes analogs that were similar, but had a higher

429 efficacy, to the tested C-terminally truncated peptide lacking the eighth position tryptophan.

430 In particular, additional residues of major importance were identified in both the amino and

431 carboxyl region of the octapeptide, since Grybi-AKH, which has substitions in positions 2-3

432 as well as 5-7 compared to the native A. aegypti AKH, demonstrated a large reduction in

433 activity by over three orders of magnitude. Interestingly, this implies that despite Grybi-

434 AKH retaining the prototypical features of an AKH, including the pyroglutamic acid in the

435 first position, phenylalanine in position four, tryptophan in position eight and an amidated C-

436 terminus, the A. aegypti AKH receptor is quite specific compared to other insect AKH

437 receptors including, for example, the locust, Schistocerca gregaria or the pea aphid,

438 Acyrthosiphon pisum, whose receptors responded similarly to both the mosquito AKH and

439 Grybi-AKH (Marchal et al., 2018). Similar findings were reported for the AKH receptor

440 from another dipteran, namely the fruit fly D. melanogaster, which is also highly selective

441 and was not strongly responsive to non-dipteran AKH analogs (Marchal et al., 2018). In

442 order to determine whether the substitutions in positions 2-3 versus 5-7 were of greatest

443 importance, we examined the activity of a natural AKH analog from L. auripennis (Libau-

22 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

444 AKH), whose sequence matches that of native A. aegypti AKH except for positions 2-3

445 (valine and asparagine) that are instead shared with Grybi-AKH. This allowed us to identify

446 that the substitutions in this N-terminal region are much more critical and are not well

447 tolerated by the AKH receptor system in A. aegypti since this analog diplayed only a

448 marginal improvement (i.e. nearly two-fold) compared to Grybi-AKH. This finding is

449 corroborated by recent analysis of select dipteran AKH analogs on the A. aegypti AKH

450 receptor, which showed that amino acid changes at positions five (serine in both Glossina

451 morsitans AKHs and in D. melanogaster AKH compared to threonine in A. aegypti AKH) or

452 seven (glycine in G. morsitans AKH-I and aspartic acid in D. melanogaster AKH compared

453 to serine in A. aegypti AKH) are indeed well tolerated with these dipteran AKH analogs

454 being similarly active to the endogenous mosquito AKH (Marchal et al., 2018). It was also

455 previously shown in other dipteran AKH systems where structure-activity relationships were

456 investigated, that the least tolerated substitutions were localized to the N-terminal region,

457 with substitutions to residues in positions two through five from the N-terminus proving

458 crucial for activation of D. melanogaster and A. gambiae AKH receptors, which the authors

459 described was due to this core region of the peptide forming a predicted β -strand necessary

460 for receptor interaction (Caers et al., 2012). Comparatively, although only fruit fly AKH

461 analogs were tested, substitutions to position seven alone and positions six plus seven for the

462 D. melanogaster and A. gambiae AKH receptors, respectively, were far more tolerated with

463 little or no influence on the activation of the corresponding receptor (Caers et al., 2012). This

464 supports our conclusions that the C-terminal region of the AKH octapeptide, excluding the

465 critical tryptophan in the eighth position, is not as critical for activity. As a putative reason for

466 this, it is argued that the side chains of a subset of these amino acids may be buried in the

467 assumed β-turn formed by the residues in the C-terminal region of AKH neuropeptides

468 (Caers et al., 2012; Gäde, 1992; Gäde and Hayes, 1995), which is now supported by

23 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

469 molecular modeling and nuclear magnetic resonance spectroscopy studies that have

470 confirmed the β-turn configuration in the AKH secondary structure (Jackson et al., 2014;

471 Nair et al., 2000; Nair et al., 2001; Zubrzycki and Gäde, 1994).

472 Interestingly, structure-activity studies on the two tsetse fly (G. morsitans) AKH receptors

473 using synthetic analogs of the pair of endogenous AKHs identified the most critical sites

474 being the aromatic residues in positions four and eight, but the next most important features

475 included positions 1-3 from the N-terminus with activity of these analogs eliciting only 20-

476 50% activity compared to the two endogenous AKH peptides (Caers et al., 2016). On the

477 other hand, the region of the G. morsitans AKH diplaying increased permissiveness for

478 substitutions included positions six and seven, where glycine substitutions of these sites

479 resulted in these analogs having between 65-97% activity compared to the efficacy of the

480 native AKHs. Interestingly, however, substitutions in the fifth position of the octapeptide

481 were not tolerated, since activity of these analogs was reduced to only 11-13% that of the

482 natural AKHs, which nearly matches the effect of substitutions in the aromatic residue sites at

483 positions four and eight (Caers et al., 2016). Supporting the notion that substitutions in

484 positions 5-7 are more permissive in the mosquito AKH receptor system, we examined the

485 activity of two additional natural AKH analogs, which in comparison to A. aegypti AKH,

486 have substitutions in either the third position alone or combined with a substitution of the

487 seventh position residue. In particular, the firebug P. apterus AKH (Pyrap-AKH) has the

488 third position threonine found in A. aegypti AKH replaced with asparagine while the seventh

489 position serine is also replaced by asparagine, resulting with this analog having ~337-fold

490 reduced activity relative to the native mosquito AKH. Next, by comparing the tolerance of

491 these two substitutions in Pyrap-AKH with that occurring in the AKH from E. simplicicollis

492 (Erysi-AKH), which has only a single substitution in the third position with threonine

24 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

493 replaced by asparagine, reveals this AKH analog has similar activity as Pyrap-AKH with

494 ~310-fold reduced activity relative to the endogenous A. aegypti AKH.

495 The third group includes analogs with activities similar to, or in some cases, stronger than the

496 endogenous A. aegypti AKH that contained substitutions or insertions in the C-terminal

497 region, while sparing the critical tryptophan residue. Specifically, we next tested two natural

498 AKH analogs with substitutions in the C-terminal region of the peptide, maintaining the core

499 5-6 residues on the N-terminus. Specifically, the domestic silkmoth AKH (Bommo-AKH),

500 has a glycine substitution in an equivalent location to the serine in position seven and also has

501 an extended C-terminus containing glycine and glutamine that follows tryptophan in the

502 eighth position, forming a decapeptide. Despite these changes within the C-terminal region,

503 this AKH analog retained strong activity on the A. aegypti AKHR-IA. Similarly, analysis of

504 Hipes-AKH-I from the hawk moth, Hippotion eson, which along with another member of the

505 genus, each produce the greatest number of AKH analogs so far reported from a single insect

506 (Gäde et al., 2013), revealed this analog to be very active, with only a marginal reduction in

507 activity on the mosquito AKHR-IA. Notably, Hipes-AKH-I shares an identical sequence to

508 the Lacol-AKH-7mer with the one exception that the latter lacks the terminal tryptophan

509 residue. This indicates that serine for proline substitution in the sixth position is in fact well

510 tolerated and is not the root cause of the abolished activity of the Lacol-AKH-7mer, which

511 instead is most certainly due to the absence of the C-terminal amphipathic tryptophan residue

512 with its large indole side chain.

513 Considering the resulting activity of these various AKH analogs, the findings support that

514 substitutions in positions 5-7 are more permissive, whereas comparatively, substitutions in

515 positions 2-3 are not well tolerated by the A. aegypti AKH receptor system. Therefore, given

516 the importance of substitutions in positions 2-3, we aimed to discern which of these two sites

517 were the most critical for eliciting AKHR-IA functional activation. To achieve this aim, we

25 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

518 tested a synthetic analog of a cockroach P. americana AKH, V2-PeramCAH-II, which has a

519 second position valine substituted for the naturally occurring leucine in A. aegypti AKH but

520 also contains a substitution in the seventh position with asparagine replacing serine. The

521 results from analysis of this analog confirm the valine substitution for leucine in the second

522 position is far better tolerated compared to the asparagine substitution for threonine in the

523 third position, since V2-Peram-CAH-II displayed nearly five-fold improved activity on the A.

524 aegypti AKHR-IA compared to Pyrap-AKH (which has asparagine substitutions in both

525 positions three and seven). Thus, although both of these residues are important for functional

526 activation of the mosquito AKHR-IA, substitutions of the second position proved this site to

527 be the most critical. One additional matter to consider is the seventh position substitution,

528 which is an asparagine in Pyrap-AKH, V2-Peram-CAH-II and is the only substitution in the

529 natural AKH analog, Peram-CAH-II. The activity of Peram-CAH-II was equivalent to the

530 native mosquito AKH, which confirms that this seventh position is quite permissive to

531 substitutions and, more importantly, that the markedly reduced activity of the AKH analogs

532 Pyrap-AKH and V2-Peram-CAH-II is solely a result of the substitutions in the second and

533 third positions. Finally, to our surprise, the survey of natural AKH analogs revealed that

534 Tabat-AKH, which has a single glycine substitution in position seven replacing the natural

535 serine of A. aegypti AKH, may serve as a potential lead substance for design of a

536 superagonist since this natural AKH analog elicited 30% improved activity relative to the

537 native mosquito AKH. It appears that the more flexible and non-polar Gly residue allows the

538 ligand a tighter binding to the receptor than the polar and neutral Ser. This is particularly

539 interesting given the permissiveness of substitutions in the vicinity of the C-terminus

540 (excluding the eighth position tryptophan) established in the current study and previously for

541 the A. aegypti AKHR-IA (Marchal et al., 2018) as well as homologous receptors from other

542 dipterans (Caers et al., 2012; Marchal et al., 2018), which may allow a greater variety of lead

26 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

543 compounds to be designed and tested that could interfere with this neuropeptide system and

544 perturb its normal functioning vital for mobilizing energy substrates in insects.

545 Integrating the novel data from the present study analyzing structural characteristics of the

546 two most recently diverged GnRH-related family members in arthropods, we can infer with

547 some degree of confidence how these ligands act only upon their respective receptors. The

548 ACP sequence in many insect species often contain charged residues in positions 6, 7 or 9,

549 which in instances where only a single basic or acidic residue occurs, provide an overall

550 charge to the peptide (Hansen et al., 2010; Veenstra, 2019). However, our results revealed

551 that the overall charge of the ACP analog did not largely compromise activity since

552 substitutions of these charged residues with alanine were quite tolerated by the ACPR.

553 Comparatively, AKH receptors do not like charges associated with their ligands since no

554 basic residues occur and only a few natural AKH family members containg an acidic residue

555 (e.g. aspartic acid at position 7), which have been shown to be not well tolerated by AKH

556 receptors in L. migratoria and P. americana (Gäde, 1991; Gäde, 1993; Gäde et al., 1990).

557 Another feature differentiating the AKH and ACP peptides in A. aegypti is the presence of an

558 alternating pattern of hydrophobic and hydrophilic residues over the entire length of the

559 AKH, which is shared with other AKH family members (Gäde and Hayes, 1995), whereas

560 the ACP contains a motif of three hydrophilic residues at its core residues 5-7, which may

561 lead to incompatability with the AKHR-IA. Additionally, clearly the valine substitution in

562 position 2 is not tolerated well by the A. aegypti AKHR-IA, and so this residue which is

563 present in position 2 of the ACP sequence may also contribute towards the inactivity of this

564 peptide on the AKHR-IA. On the other hand, since the AKH peptide from B. mori (a

565 decapeptide) was quite active on the A. aegypti AKHR-IA, the longer length of ACP is

566 unlikely to be the cause of its inactivity on the AKHR-IA.

27 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

567 Our findings also set the framework towards understanding why the A. aegypti AKH has no

568 activity on the ACP receptor. As we saw with the AKHR-IA, substitution of valine for

569 leucine in position 2 led to nearly a 70-fold reduction in receptor activity. On the other hand,

570 the removal of valine and its replacement with alanine in the ACP sequence led to a similar

571 70-fold reduction in activity on the ACPR. Thus, the presence of valine in position two is

572 needed for optimal activity for ACPR whereas its absence in the same position (but on the

573 AKH peptide) is necessary for optimal activation of AKHR-IA. Again, as discussed earlier,

574 the importance of charged residues for optimal ACPR activity that are often found in insect

575 ACPs (Hansen et al., 2010; Veenstra, 2019), whereas charged residues are unusually found

576 on AKH peptides and are not well tolerated by AKH receptors that have been studied either

577 in vivo or in vitro (Gäde, 1991; Gäde, 1993; Gäde et al., 1990). Finally, one clear finding

578 from these studies is that ACP analogs that are too short (nonapeptide or octapeptide), even if

579 they contain all the hallmark features, as demonstrated by the C-terminally truncated analogs

580 for example, fail to activate ACPR. Thus, the length of the ligand is most relevant and the

581 binding pocket of ACPR requires at least a decapeptide leaving the A. aegypti AKH, an

582 octapeptide, too short for ACPR binding and activation.

583 584

28 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

585 Author Contributions

586 GG and JP designed the synthetic analogs and wrote the manuscript. AW performed all the

587 experiments and AW, JP and GG analyzed the data. All authors have contributed towards the

588 revisions and have granted approval of the final manuscript submitted for publication.

589 Funding

590 This work was funded by a Natural Sciences and Engineering Research Council of Canada

591 (NSERC) Discovery Grant and an Ontario Ministry of Research, Innovation and Science

592 Early Researcher Award to JP and Incentive Grant from the National Research Foundation

593 (Pretoria, South Africa; grant number 85768 [IFR13020116790] to GG) and by the

594 University of Cape Town (block grant to GG).

595 Conflict of Interest Statement

596 The authors declare that the research was conducted in the absence of any commercial or

597 financial relationships that could be construed as a potential conflict of interest.

598 599

29 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

600 Figure captions 601

602 Figure 1. Dose–response curves of the native A. aegypti ACP along with synthetic analogs

603 that contain single substitutions or other modifications, which were tested for their activity on

604 the ACP receptor (ACPR) using a cell-based bioluminescent assay. (A) Single amino acid

605 substitutions with alanine or modifications to the normally blocked N- or C-termini resulting

606 in a significant reduction to their activity on the ACPR. (B) Single amino acid substitutions

607 with alanine to the native ACP sequence having minimal effects to the activity of the tested

608 analogs indicating these substitutions are well tolerated. (C) Analogs with internal or terminal

609 truncations or extension (i.e. +Gly analog) relative to the native ACP sequence conferring

610 significant reductions to their activity on the ACPR. The half maximal effective

611 concentration (EC50) for each analog along with the corresponding change in activity relative

612 to native A. aegypti ACP is provided in Table 1. Luminescence is plotted relative to the

613 maximal response achieved when 10-5M Aedae-ACP was applied to the ACPR. Data

614 represent mean +/- standard error of three independent biological replicates.

615

616 Figure 2. Dose–response curves of the native A. aegypti AKH along with mostly naturally

617 occurring analogs from a variety of insects that contain substitutions or other modifications,

618 which were tested for their activity on the AKHR-IA receptor using a cell-based

619 bioluminescent assay. (A) Naturally occurring insect AKH analogs having critical

620 substitutions that are not well tolerated by the AKHR-IA, with moderately compromised

621 activity relative to the native A. aegypti AKH. (B) Naturally occurring insect AKH analogs,

622 or a synthetic analog (V2-Peram-CAH-II), having substitutions that are well tolerated by the

623 AKHR-IA, with marginally compromised or slightly improved activity relative to the native

624 A. aegypti AKH. The half maximal effective concentration (EC50) for each analog along with

30 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

625 the corresponding change in activity relative to native A. aegypti AKH is provided in Table 2.

626 Luminescence is plotted relative to the maximal response achieved when 10-5M Aedae-AKH

627 was applied to the AKHR-IA. Data represent mean +/- standard error of three independent

628 biological replicates.

31 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

629 References:

630 Audsley, N. and Down, R. E. (2015). G protein coupled receptors as targets for next 631 generation pesticides. Insect Biochem. Mol. Biol. 67, 27–37. 632 Baud, D., Gubler, D. J., Schaub, B., Lanteri, M. C. and Musso, D. (2017). An update on 633 Zika virus infection. Lancet 390, 2099–2109. 634 Caers, J., Peeters, L., Janssen, T., De Haes, W., Gäde, G. and Schoofs, L. (2012). 635 Structure–activity studies of Drosophila adipokinetic hormone (AKH) by a cellular 636 expression system of dipteran AKH receptors. Gen. Comp. Endocrinol. 177, 332–337. 637 Caers, J., Janssen, T., Van Rompay, L., Broeckx, V., Van Den Abbeele, J., Gäde, G., 638 Schoofs, L. and Beets, I. (2016). Characterization and pharmacological analysis of two 639 adipokinetic hormone receptor variants of the tsetse fly, Glossina morsitans morsitans. 640 Insect Biochem. Mol. Biol. 70, 73–84. 641 Fox, A. M. and Reynolds, S. E. (1991). The pharmacology of the lipid-mobilizing response 642 to adipokinetic hormone family peptides in the moth, Manduca sexta. J. Insect Physiol. 643 37, 373–381. 644 Gäde, G., Lee, M. J., Goldsworthy, G. J. and Kellner, R. (2000). Potencies of naturally- 645 occurring AKH/RPCH peptides in Locusta migratoria in the acetate uptake assay in 646 vitro and comparison with their potencies in the lipid mobilisation assay in vivo. Acta 647 Biol. Hung. 51, 369–377. 648 Gäde, G. (1991). A unique charged tyrosine-containing member of the adipokinetic 649 hormone/red-pigment-concentrating hormone peptide family isolated and sequenced 650 from two beetle species. Biochem. J. 275, 671–677. 651 Gäde, G. (1992). Structure-activity relationships for the carbohydrate-mobilizing action of 652 further bioanalogues of the adipokinetic hormone/red pigment-concentrating hormone 653 family of peptides. J. Insect Physiol. 38, 259–266. 654 Gäde, G. (1993). Structure-activity relationships for the lipid-mobilizing action of further 655 bioanalogues of the adipokinetic hormone/red pigment-concentrating hormone family of 656 peptides. J. Insect Physiol. 39, 375–383. 657 Gäde, G. (2004). Regulation of intermediary metabolism and water balance of insects by 658 neuropeptides. Annu. Rev. Entomol. 49, 93–113. 659 Gäde, G. and Auerswald, L. (2003). Mode of action of neuropeptides from the adipokinetic 660 hormone family. Gen. Comp. Endocrinol. 132, 10–20. 661 Gäde, G. and Hayes, T. K. (1995). Structure-activity relationships for Periplaneta 662 americana hypertrehalosemic hormone I: The importance of side chains and termini. 663 Peptides 16, 1173–1180. 664 Gäde, G., Wilps, H. and Kellner, R. (1990). Isolation and structure of a novel charged 665 member of the red–pigment-concentrating hormone-adipokinetic hormone family of 666 peptides isolated from the corpora cardiaca of the blowfly Phormia terraenovae 667 (Diptera). Biochem. J. 269, 309–313. 668 Gäde, G., Šimek, P. and Marco, H. G. (2011). An invertebrate [hydroxyproline]-modified 669 neuropeptide: Further evidence for a close evolutionary relationship between insect

32 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

670 adipokinetic hormone and mammalian gonadotropin hormone family. Biochem. 671 Biophys. Res. Commun. 414, 592–597. 672 Gäde, G., Šimek, P., Clark, K. D. and Marco, H. G. (2013). Five functional adipokinetic 673 peptides expressed in the corpus cardiacum of the moth genus Hippotion (Lepidoptera, 674 ). Regul. Pept. 184, 85–95. 675 Gäde, G., Šimek, P. and Marco, H. G. (2016). Identification and bioactivity evaluation of 676 the first neuropeptide from the lesser-known insect order Embioptera (webspinner). 677 Amino Acids 48, 1677–1684. 678 Gäde, G. (1990). Structure–function studies on hypertrehalosaemic and adipokinetic 679 hormones: activity of naturally occurring analogues and some N- and C-terminal 680 modified analogues. Physiol. Entomol. 15, 299–316. 681 Gospocic, J., Shields, E. J., Glastad, K. M., Lin, Y., Penick, C. A., Yan, H., Mikheyev, A. 682 S., Linksvayer, T. A., Garcia, B. A., Berger, S. L., et al. (2017). The neuropeptide 683 corazonin controls social behavior and caste identity in ants. Cell 170, 748-759.e12. 684 Hansen, K. K., Stafflinger, E., Schneider, M., Hauser, F., Cazzamali, G., Williamson, 685 M., Kollmann, M., Schachtner, J. and Grimmelikhuijzen, C. J. P. P. (2010). 686 Discovery of a novel insect neuropeptide signaling system closely related to the insect 687 adipokinetic hormone and corazonin hormonal systems. J. Biol. Chem. 285, 10736–47. 688 Hauser, F. and Grimmelikhuijzen, C. J. P. (2014). Evolution of the 689 AKH/corazonin/ACP/GnRH receptor superfamily and their ligands in the Protostomia. 690 Gen. Comp. Endocrinol. 209, 35–49. 691 Hill, C. A., Sharan, S. and Watts, V. J. (2018). Genomics, GPCRs and new targets for the 692 control of insect pests and vectors. Curr. Opin. Insect Sci. 30, 99–106. 693 Jackson, G. E., Gamieldien, R., Mugumbate, G. and Gäde, G. (2014). Structural studies 694 of adipokinetic hormones in water and DPC micelle solution using NMR distance 695 restrained molecular dynamics. Peptides 53, 270–277. 696 Jackson, G. E., Pavadai, E., Gäde, G., Timol, Z. and Andersen, N. H. (2018). Interaction 697 of the red pigment-concentrating hormone of the crustacean Daphnia pulex , with its 698 cognate receptor, Dappu-RPCHR: A nuclear magnetic resonance and modeling study. 699 Int. J. Biol. Macromol. 106, 969–978. 700 Kaufmann, C., Merzendorfer, H. and Gäde, G. (2009). The adipokinetic hormone system 701 in Culicinae (Diptera: Culicidae): Molecular identification and characterization of two 702 adipokinetic hormone (AKH) precursors from Aedes aegypti and Culex pipiens and two 703 putative AKH receptor variants from A. aegypti. Insect Biochem. Mol. Biol. 39, 770– 704 781. 705 Keeley, L. L., Hayes, T. K., Bradfield, J. Y. and Sowa, S. M. (1991). Physiological actions 706 by hypertrehalosemic hormone and adipokinetic peptides in adult Blaberus discoidalis 707 cockroaches. Insect Biochem. 21, 121–129. 708 Kim, Y.-J., Spalovská-Valachová, I., Cho, K.-H., Zitnanova, I., Park, Y., Adams, M. E. 709 and Zitnan, D. (2004). Corazonin receptor signaling in ecdysis initiation. Proc. Natl. 710 Acad. Sci. U. S. A. 101, 6704–6709. 711 Kodrík, D. (2008). Adipokinetic hormone functions that are not associated with insect flight. 712 Physiol. Entomol. 33, 171–180.

33 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

713 Kodrík, D., Bednářová, A., Zemanová, M. and Krishnan, N. (2015). Hormonal regulation 714 of response to oxidative stress in insects—an update. Int. J. Mol. Sci. 16, 25788–25816. 715 Lee, M. J. and Goldsworthy, G. J. (1996). Modified adipokinetic peptides containing two 716 tryptophan residues and their activities in vitro and in vivo in Locusta. J. Comp. Physiol. 717 B 166, 61–67. 718 Lee, M. J., Goldsworthy, G. J., Poulos, C. P. and Velentza, A. (1996). Synthesis and 719 biological activity of adipokinetic hormone analogues modified at the C-terminus. 720 Peptides 17, 1285–1290. 721 Lee, M. J., Cusinato, O., Luswata, R., Wheeler, C. H. and Goldsworthy, G. J. (1997). N- 722 terminal modifications to AKH-I from Locusta migratoria: assessment of biological 723 potencies in vivo and in vitro. Regul. Pept. 69, 69–76. 724 Li, S., Hauser, F., Skadborg, S. K., Nielsen, S. V., Kirketerp-Møller, N. and 725 Grimmelikhuijzen, C. J. P. (2016). Adipokinetic hormones and their G protein- 726 coupled receptors emerged in Lophotrochozoa. Sci. Rep. 6, 32789. 727 Marchal, E., Schellens, S., Monjon, E., Bruyninckx, E., Marco, H., Gäde, G., Vanden 728 Broeck, J. and Verlinden, H. (2018). Analysis of peptide ligand specificity of different 729 insect adipokinetic hormone receptors. Int. J. Mol. Sci. 19, 542. 730 Marco, H. G. and Gäde, G. (2015). Structure–activity relationship of adipokinetic hormone 731 analogs in the striped hawk moth, Hippotion eson. Peptides 68, 205–210. 732 Marco, H. and Gäde, G. (2019a). Adipokinetic hormone: a hormone for all seasons? In 733 Advances in Invertebrate (Neuro)Endocrinology: A Collection of Reviews in the Post- 734 Genomic Era (ed. Saleuddin, S., Lange, A. B., and Orchard, I.), Volume 2, Chapter 3. 735 Apple Academic Press. 736 Marco, H. G. and Gäde, G. (2019b). Five neuropeptide ligands meet one receptor: how does 737 this tally? A structure-activity relationship study using adipokinetic bioassays with the 738 Sphingid moth, Hippotion eson. Front. Endocrinol. (Lausanne). 10, 231. 739 Marco, H. G., Verlinden, H., Vanden Broeck, J. and Gäde, G. (2017). Characterisation 740 and pharmacological analysis of a crustacean G protein-coupled receptor: the red 741 pigment-concentrating hormone receptor of Daphnia pulex. Sci. Rep. 7, 6851. 742 Mugumbate, G., Jackson, G. E., van der Spoel, D., Kövér, K. E. and Szilágyi, L. (2013). 743 Anopheles gambiae, Anoga-HrTH hormone, free and bound structure – A nuclear 744 magnetic resonance experiment. Peptides 41, 94–100. 745 Nair, M. M., Jackson, G. E. and Gäde, G. (2000). NMR study of insect adipokinetic 746 hormones. Spectrosc. Lett. 33, 875–891. 747 Nair, M. M., Jackson, G. E. and Gäde, G. (2001). Conformational study of insect 748 adipokinetic hormones using NMR constrained molecular dynamics. J. Comput. Aided. 749 Mol. Des. 15, 259–70. 750 Nene, V., Wortman, J. R., Lawson, D., Haas, B., Kodira, C., Tu, Z. J., Loftus, B., Xi, Z., 751 Megy, K., Grabherr, M., et al. (2007). Genome sequence of Aedes aegypti, a major 752 arbovirus vector. Science 316, 1718–1723. 753 Offermanns, S. and Simon, M. I. (1995). G alpha 15 and G alpha 16 couple a wide variety 754 of receptors to phospholipase C. J Biol Chem 270, 15175–15180.

34 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

755 Oryan, A., Wahedi, A. and Paluzzi, J.-P. V. (2018). Functional characterization and 756 quantitative expression analysis of two GnRH-related peptide receptors in the mosquito, 757 Aedes aegypti. Biochem. Biophys. Res. Commun. 497, 550–557. 758 Paluzzi, J.-P. V., Young, P., Defferrari, M. S., Orchard, I., Carlini, C. R. and O’Donnell, 759 M. J. (2012). Investigation of the potential involvement of eicosanoid metabolites in 760 anti-diuretic hormone signaling in Rhodnius prolixus. Peptides 34, 127–134. 761 Paluzzi, J. P. V., Haddad, A. S., Sedra, L., Orchard, I. and Lange, A. B. (2015). 762 Functional characterization and expression analysis of the myoinhibiting peptide 763 receptor in the Chagas disease vector, Rhodnius prolixus. Mol. Cell. Endocrinol. 399, 764 143–153. 765 Poulos, C., Karagiannis, K., Lee, M. and Goldsworthy, G. (1994). Synthesis and 766 biological activity of locust AKH-I and its analogues with modifications at the threonine 767 residues. Int. J. Pept. Protein Res. 44, 589–593. 768 Predel, R., Neupert, S., Russell, W. K., Scheibner, O. and Nachman, R. J. (2007). 769 Corazonin in insects. Peptides 28, 3–10. 770 Predel, R., Neupert, S., Garczynski, S. F., Crim, J. W., Brown, M. R., Russell, W. K., 771 Kahnt, J., Russell, D. H. and Nachman, R. J. (2010). Neuropeptidomics of the 772 mosquito Aedes aegypti. J Proteome Res 9, 2006–2015. 773 Price, D. P., Schilkey, F. D., Ulanov, A. and Hansen, I. A. (2015). Small mosquitoes, large 774 implications: Crowding and starvation affects gene expression and nutrient 775 accumulation in Aedes aegypti. Parasites and Vectors 8, 252. 776 Roch, G. J., Busby, E. R. and Sherwood, N. M. (2011). Evolution of GnRH: Diving 777 deeper. Gen. Comp. Endocrinol. 171, 1–16. 778 Saiz, J.-C., Vázquez-Calvo, Á., Blázquez, A. B., Merino-Ramos, T., Escribano-Romero, 779 E. and Martín-Acebes, M. A. (2016). Zika virus: the latest newcomer. Front. 780 Microbiol. 7, 496. 781 Schilling, S., Wasternack, C. and Demuth, H.-U. (2008). Glutaminyl cyclases from 782 and plants: a case of functionally convergent protein evolution. Biol. Chem. 389, 783 983–991. 784 Tanaka, Y., Hua, Y. J., Roller, L. and Tanaka, S. (2002). Corazonin reduces the spinning 785 rate in the silkworm, Bombyx mori. J. Insect Physiol. 48, 707–714. 786 Tawfik, A. I., Tanaka, S., De Loof, A., Schoofs, L., Baggerman, G., Waelkens, E., Derua, 787 R., Milner, Y., Yerushalmi, Y. and Pener, M. P. (1999). Identification of the 788 gregarization-associated dark-pigmentotropin in locusts through an albino mutant. Proc. 789 Natl. Acad. Sci. 96, 7083–7087. 790 Tian, S., Zandawala, M., Beets, I., Baytemur, E., Slade, S. E., Scrivens, J. H. and 791 Elphick, M. R. (2016). Urbilaterian origin of paralogous GnRH and corazonin 792 neuropeptide signalling pathways. Sci. Rep. 6, 28788. 793 Veenstra, J. A. (1989). Isolation and structure of corazonin, a cardioactive peptide from the 794 American cockroach. FEBS Lett. 250, 231–234. 795 Veenstra, J. A. (2019). Coleoptera genome and transcriptome sequences reveal numerous 796 differences in neuropeptide signaling between species. PeerJ 7, e7144.

35 bioRxiv preprint doi: https://doi.org/10.1101/699645; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

797 Velentza, A., Spiliou, S., Poulos, C. P. and Goldsworthy, G. J. (2000). Synthesis and 798 biological activity of adipokinetic hormone analogues with modifications in the 4–8 799 region. Peptides 21, 631–637. 800 Verlinden, H., Vleugels, R., Zels, S., Dillen, S., Lenaerts, C., Crabbé, K., Spit, J. and 801 Vanden Broeck, J. (2014). Receptors for neuronal or endocrine signalling molecules as 802 potential targets for the control of insect pests. In Advances in Insect Physiology, pp. 803 167–303. 804 Wahedi, A. and Paluzzi, J.-P. (2018). Molecular identification, transcript expression, and 805 functional deorphanization of the adipokinetic hormone/corazonin-related peptide 806 receptor in the disease vector, Aedes aegypti. Sci. Rep. 8, 2146. 807 Zandawala, M., Tian, S. and Elphick, M. R. (2018). The evolution and nomenclature of 808 GnRH-type and corazonin-type neuropeptide signaling systems. Gen. Comp. 809 Endocrinol. 264, 64–77. 810 Zhou, Y. J., Fukumura, K. and Nagata, S. (2018). Effects of adipokinetic hormone and its 811 related peptide on maintaining hemolymph carbohydrate and lipid levels in the two- 812 spotted cricket, Gryllus bimaculatus. Biosci. Biotechnol. Biochem. 82, 274–284. 813 Ziegler, R., Eckart, K., Jasensky, R. D. and Law, J. H. (1991). Structure-activity studies 814 on adipokinetic hormones in Manduca sexta. Arch. Insect Biochem. Physiol. 18, 229– 815 237. 816 Ziegler, R., Cushing, A. ., Walpole, P., Jasensky, R. . and Morimoto, H. (1998). Analogs 817 of Manduca adipokinetic hormone tested in a bioassay and in a receptor-binding assay. 818 Peptides 19, 481–486. 819 Zubrzycki, I. Z. and Gäde, G. (1994). Conformational study on an insect neuropeptide of 820 the AKH/RPCH-family by combined 1H-NMR spectroscopy and molecular mechanics. 821 Biochem. Biophys. Res. Commun. 198, 228–235.

36