Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Neuropeptide Receptor Transcriptome Reveals Unidentified Neuroendocrine Pathways

  • Naoki Yamanaka,

    Current address: Department of Genetics, Cell Biology and Development, Howard Hughes Medical Institute, University of Minnesota, Minneapolis, Minnesota, United States of America

    Affiliation Department of Integrated Biosciences, Graduate School of Frontier Sciences, The University of Tokyo, Chiba, Japan

  • Sachie Yamamoto,

    Affiliation Department of Integrated Biosciences, Graduate School of Frontier Sciences, The University of Tokyo, Chiba, Japan

  • Dušan Žitňan,

    Affiliation Institute of Zoology, Slovak Academy of Sciences, Bratislava, Slovakia

  • Ken Watanabe,

    Affiliation Department of Integrated Biosciences, Graduate School of Frontier Sciences, The University of Tokyo, Chiba, Japan

  • Tsuyoshi Kawada,

    Affiliation Suntory Institute for Bioorganic Research, Osaka, Japan

  • Honoo Satake,

    Affiliation Suntory Institute for Bioorganic Research, Osaka, Japan

  • Yu Kaneko,

    Affiliation Faculty of Agriculture and Life Sciences, Hirosaki University, Aomori, Japan

  • Kiyoshi Hiruma,

    Affiliation Faculty of Agriculture and Life Sciences, Hirosaki University, Aomori, Japan

  • Yoshiaki Tanaka,

    Affiliation National Institute of Agrobiological Sciences, Ibaraki, Japan

  • Tetsuro Shinoda ,

    shinoda@affrc.go.jp (TS); kataoka@k.u-tokyo.ac.jp (HK)

    Affiliation National Institute of Agrobiological Sciences, Ibaraki, Japan

  • Hiroshi Kataoka

    shinoda@affrc.go.jp (TS); kataoka@k.u-tokyo.ac.jp (HK)

    Affiliation Department of Integrated Biosciences, Graduate School of Frontier Sciences, The University of Tokyo, Chiba, Japan

Abstract

Neuropeptides are an important class of molecules involved in diverse aspects of metazoan development and homeostasis. Insects are ideal model systems to investigate neuropeptide functions, and the major focus of insect neuropeptide research in the last decade has been on the identification of their receptors. Despite these vigorous efforts, receptors for some key neuropeptides in insect development such as prothoracicotropic hormone, eclosion hormone and allatotropin (AT), remain undefined. In this paper, we report the comprehensive cloning of neuropeptide G protein-coupled receptors from the silkworm, Bombyx mori, and systematic analyses of their expression. Based on the expression patterns of orphan receptors, we identified the long-sought receptor for AT, which is thought to stimulate juvenile hormone biosynthesis in the corpora allata (CA). Surprisingly, however, the AT receptor was not highly expressed in the CA, but instead was predominantly transcribed in the corpora cardiaca (CC), an organ adjacent to the CA. Indeed, by using a reverse-physiological approach, we purified and characterized novel allatoregulatory peptides produced in AT receptor-expressing CC cells, which may indirectly mediate AT activity on the CA. All of the above findings confirm the effectiveness of a systematic analysis of the receptor transcriptome, not only in characterizing orphan receptors, but also in identifying novel players and hidden mechanisms in important biological processes. This work illustrates how using a combinatorial approach employing bioinformatic, molecular, biochemical and physiological methods can help solve recalcitrant problems in neuropeptide research.

Introduction

Diverse biological processes in metazoans such as development, reproduction and behavior are regulated by neuropeptides. One of the most effective approaches for analyzing their physiological functions is to identify their receptors, most of which belong to the G protein-coupled receptor (GPCR) superfamily. In insects, Drosophila melanogaster has been used as a model organism for such an approach. To date, some forty GPCRs are categorized as neuropeptide GPCRs in Drosophila, and about two-thirds of these receptors have been “deorphanized” (i.e. their endogenous ligands have been found) [1], [2].

In spite of this success in flies, receptors for several key neuropeptides in insect development are yet to be identified. Among such neuropeptides is allatotropin (AT), which is the only neuropeptide widely thought to stimulate juvenile hormone (JH) biosynthesis [3], [4]. JH is involved in various aspects of insect physiology such as development, reproduction and polyphenism [5][7], but the regulatory mechanisms that control its biosynthesis are not well understood in part due to the lack of information on several key molecules such as the AT receptor. The identities of receptors for prothoracicotropic hormone and eclosion hormone, two other important players in the regulation of insect development, also remain uncharacterized.

Considering that the target organs of these key neuropeptides are known, comprehensive tissue expression analyses of the remaining orphan receptors seems likely to be informative in identifying their receptors. With just a few exceptions, however [8], [9], such spatial expression analyses have been hampered due to the small size of the Drosophila. In contrast, larger lepidopteran insects such as Manduca sexta and Bombyx mori allowed detailed tissue expression analyses of several neuropeptide GPCRs, which have helped elucidate novel functions of their ligands [10][12]. Now that a draft sequence of the Bombyx genome has been published [13], [14], a comprehensive analysis of an entire neuropeptide GPCR transcriptome is feasible for the first time among lepidopterans.

Using Bombyx, we report here the first comprehensive cloning and thorough analyses of neuropeptide GPCRs conducted in any organism. This approach, combined with biochemical and physiological studies, led to the molecular characterization of the first AT receptor, as well as the identification of novel allatoregulatory peptides. Our results clearly show the effectiveness of the systematic analysis of a receptor transcriptome in neuropeptide research, and have enabled us to formulate a novel hypothesis for how JH biosynthesis is regulated by AT.

Results

Comprehensive Cloning of Bombyx Neuropeptide GPCR Genes

When a newly released genome is screened for a certain family of genes, homology searching using the family members from other well-characterized species is the most powerful and reliable approach [1], [2], [15][19]. In order to thoroughly identify neuropeptide GPCR genes encoded in Bombyx genome (Bombyx neuropeptide GPCR genes or BNGRs), therefore, we sought to identify all the homologs of previously categorized D. melanogaster neuropeptide GPCRs (see Figure 1 for the whole identification process). Based on earlier reports, 40 Drosophila GPCRs were listed as neuropeptide receptors (Table 1) [1], [2], [20]. Using the amino acid sequences of these Drosophila neuropeptide GPCRs as queries, Bombyx whole-genome shotgun sequence contigs were screened (TBLASTN analysis) by using KAIKOBLAST (http://kaikoblast.dna.affrc.go.jp/; first screening). This resulted in the identification of 195 contigs, whose E-values against at least one Drosophila neuropeptide GPCR are less than 0.1 (Table S1). Since the gene fragments in these contigs may encode other types of proteins whose sequences are similar to the neuropeptide GPCRs, all the contigs were further screened against Drosophila proteome (BLASTX analysis) using FlyBase BLAST (http://flybase.net/blast/; second screening). This second screening yielded 139 contigs, whose predicted amino acid sequences showed the highest similarities to one of the 40 Drosophila neuropeptide GPCRs (Table S1). These contigs were termed putative BNGR contigs, and were subjected to further analyses.

thumbnail
Figure 1. Flow chart for the comprehensive identification of BNGRs.

Rectangles indicate nucleotide sequences. Light gray denotes homologous sequence to Drosophila neuropeptide receptors, while dark gray denotes identified BNGR sequence.

https://doi.org/10.1371/journal.pone.0003048.g001

Considering that the number of BNGRs is likely to be similar to that of Drosophila neuropeptide GPCRs, the large number of putative BNGR contigs clearly suggests that most BNGRs are fragmented into several contigs, making it impossible to predict the whole structure of a BNGR from the published genome sequence. Therefore, the best and perhaps the only way to identify all the BNGRs from this incomplete genome information is to perform rapid amplification of cDNA ends (RACE) from each putative gene fragments. Although this is quite laborious, cloned cDNA sequences are even more reliable than the predicted ones, as evidenced by the repeated correction of Drosophila GPCR gene sequences every when they are actually cloned [21][24].

In Bombyx, 8 genes have been previously reported to encode neuropeptide GPCRs [11], [12], [25][30]. BLAST searching for the contigs containing the fragments of these genes revealed that fragments of all the 8 genes are included in 23 putative BNGR contigs, yielding an average of 2.9 putative BNGR contigs for each gene (Table S2). This ensures the comprehensiveness of the putative BNGR contig repertoire, because a single fragment of a gene is enough to determine the whole structure of the cDNA by RACE.

After removing the contigs attributed to the reported genes, RACE was applied to each of the remaining 116 contigs. Based on the obtained cDNA sequences, the whole coding region of each BNGR was cloned into pME18S, a mammalian expression vector. The cloned BNGR was sequenced, and the predicted amino acid sequence was submitted to TMHMM (http://www.cbs.dtu.dk/services/TMHMM/) [31] and CD-Search (http://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) [32], respectively, in order to check if 7 transmembrane regions and other characteristic features of GPCRs are conserved in each individual BNGR. The certified BNGR coding sequence was then submitted to KAIKOBLAST (BLASTN analysis), to find all the contigs containing the fragments of the cloned BNGR. These identified contigs were removed from the list of putative BNGR contigs, and the next RACE was started from one of the residual contigs. This process was repeated until all the contigs were identified, which resulted in the identification of 39 novel BNGRs, with two splicing variants for BNGR-A6 (Table S3). Among putative BNGR contigs, contig 227253 was shown to contain a fragment of a gene which has high similarity to mammalian melatonin receptors (MTNR). Contig 198123 and 280624 contained fragments of a pseudogene which has multiple stop codons within the conserved domain of GPCRs. Based on the predicted amino acid sequences, phylogenetic analysis was performed for Bombyx and Drosophila neuropeptide GPCRs (Figure 2).

thumbnail
Figure 2. Neighbor-joining phylogenetic trees for neuropeptide GPCRs in Bombyx and Drosophila.

The trees are rooted with DHR and BAR as outgroups for family A (rhodopsin-like) and family B (secretin receptor-like; shown in inset) GPCRs, respectively. Bootstrap support is indicated with an open circle for 50–70%, gray circle for 70–90%, and filled circle for 90–100%. Bombyx GPCRs are in bold type: AKHR, adipokinetic hormone receptor; BAR, Bombyx allatostatin (A-type) receptor; BMSR, Bommo-myosuppressin receptor; BRFaR, Bommo-FMRFamide receptor; DHR, diuretic hormone receptor; DpHR, diapause hormone receptor; PBANR, pheromone biosynthesis-activating neuropeptide receptor; SIFR, SIFamide receptor. The endogenous Drosophila ligands are given in parentheses: AKH, adipokinetic hormone; CCAP, crustacean cardioactive peptide; DH, diuretic hormone; ETH, ecdysis-triggering hormone; NPF, neuropeptide F; PDF, pigment-dispersing factor.

https://doi.org/10.1371/journal.pone.0003048.g002

Quantitative Expression Analysis of BNGRs

As a first step toward elucidating neuropeptide signaling pathways, quantitative tissue expression analysis was performed on two larval stages (Figure 3). Among the previously characterized neuropeptide GPCRs, adipokinetic hormone receptor, diuretic hormone receptor and Bommo-myosuppressin (BMS) receptor were highly expressed in fat body, Malpighian tubules and the prothoracic gland of the last instar larvae, respectively. These expression patterns correspond well to the target tissues of their ligands [11], [33], [34], suggesting that this type of expression analysis is useful for identifying regulatory circuitry. As for the BMS receptor, significant expression was observed not only in the prothoracic gland, but also in the midgut and Malpighian tubules, as shown in the previous report [11]. Indeed, the tissue expression analysis showed that many receptors are expressed in numerous tissues, suggesting multifunctional features of their ligands.

thumbnail
Figure 3. Quantitative tissue expression analyses of Bombyx neuropeptide GPCRs on two larval stages.

Transcript levels of the GPCR genes are presented as arbitrary units. See Figure 2 for the abbreviated GPCR names. IV-HCS, fourth instar head capsule slippage; V-2, fifth instar day 2; CC-CA, corpora cardiaca-corpora allata complex; PG, prothoracic gland; BR, brain; FB, fat body; MG, midgut; EP, epidermis; MU, muscle; MP, Malpighian tubule; SiG, silk gland; SG, salivary gland; TE, testis; OV, ovary.

https://doi.org/10.1371/journal.pone.0003048.g003

Functional Characterization of CC-CA-Expressed Receptors

For detailed studies, we then focused on the six receptors highly expressed in the CC-CA complex: BNGR-A1, A6-A, A10, A11, A16 and B3. There have been two neuropeptides reported to regulate JH production by lepidopteran CA: allatostatin (Manduca-type or C-type; AST-C) and allatotropin (AT) [3], [4], [35]. In spite of the pivotal role of JH in numerous physiological processes, functions of these regulatory peptides have not been sufficiently analyzed, in part due to the lack of information of their receptors [4].

To begin a characterization of these six BNGRs, each was functionally analyzed using a heterologous expression system. HEK293 cells expressing BNGR-A1 responded to AST-C in a dose-dependent manner (Figure 4A). This result is reasonable, since the Drosophila homologs of BNGR-A1 (CG7285 and CG13702; see Figure 2) have been identified as receptors for AST-C-like peptide [36]. Moreover, BNGR-A16 responded to AT specifically and dose-dependently (Figure 4B), showing that this receptor is the functional AT receptor. No clear homolog of BNGR-A16 was found in the Drosophila genome as expected (Figure 2), since there is no AT neuropeptide gene in the fly [1], [37], [38]. The identification of the long-sought AT receptor clearly shows the efficiency of our approach in identifying unknown ligands for orphan receptors.

thumbnail
Figure 4. Functional characterization of CC-CA-expressed receptors.

(A–C) Ca2+ imaging analysis using HEK293 cells expressing BNGR-A1 (A), BNGR-A16 (B) and BNGR-A6-A (C). Dose-response curves for AST-C (A), AT (B), ETH and pre- ecdysis-triggering hormone (PETH; C) are shown. Vertical bars represent S.E. Each datum point was calculated from the responses of 123-165 cells in three independent experiments. (D) Summary of the ligands for the CC-CA-expressed BNGRs analyzed by Ca2+ imaging analysis.

https://doi.org/10.1371/journal.pone.0003048.g004

Reverse-Physiological Identification of Novel Allatostatic Peptides

Another advantage of using Bombyx is that biochemical approaches like peptide purification from organ extracts and physiological techniques like organ culture can be combined with functional analyses of the receptors. We therefore sought to purify unknown ligands for the rest of BNGRs expressed in the CC-CA complex, in pursuit of other regulators of JH biosynthesis. Based on the homology with Drosophila GPCRs, BNGR-A6-A was revealed to be the receptor for ecdysis-triggering hormone (ETH; Figure 4C), while BNGR-A10 and A11 are likely the receptors for short neuropeptide F (sNPF). ETH, which is produced in Inka cells of epitracheal glands and released into the hemolymph during ecdysis, is one of the best-characterized peptides in insects [9], [39]. On the other hand, the physiological functions of sNPF are not well characterized [37], [40], [41], and Bombyx sNPF remains to be identified.

Therefore we purified two sNPF peptides in Bombyx from brain extracts, originally prepared for the purification of BMS and Bommo-FMRFamide (BRFa) [11], [12], using an ELISA system. Four HPLC purification steps were required to purify two sNPF peptides into homogeneity, sufficient for MALDI-TOF MS analysis (Figure 5A). Aliquots of the isolated fractions were subjected to MALDI-TOF MS, which yielded the sequences TPVRLRFamide and SMRLRFamide (Figure S1). Based on these amino acid sequences, the cDNA encoding the Bombyx sNPF precursor was obtained and three putative mature peptides were found in the precursor sequence (Figure 5B; the purified SMRLRFamide seems to be the truncated form of sNPF-3). Functional analyses of BNGR-A10 and A11 revealed that these are indeed the sNPF receptors (Figure 4D and Figure 5C). Although these receptors also responded to BMS and BRFa, their sensitivity to sNPFs was more than 1,000 times higher than that to BMS and BRFa (Figure 5C and data not shown).

thumbnail
Figure 5. Reverse-physiological identification of sNPFs as novel allatostatic peptides.

(A) Purification of sNPFs (arrowheads). The first (upper panel) and the fourth (lower two panels) HPLC separation steps are shown. Solid lines denote UV absorbance, while dashed lines denote the acetonitrile gradient. Fraction No. 22 in the upper panel was originally subjected to purification of BMS, and FMRFamide-related peptide (FaRP) content in this fraction was presented as the amount of purified BMS. (B) The deduced amino acid sequence of the sNPF precursor. The putative site of the signal peptide cleavage is indicated with an arrowhead, and the mature peptides are highlighted. For each mature peptide, C-terminal glycine residue is necessary for amidation. (C) Ca2+ imaging analysis using HEK293 cells expressing BNGR-A10. Relative response was expressed as a percentage of the highest response (induced by 10−8 M sNPF-1–3). For each ligand, each datum point was calculated from the responses of 80–122 cells in three independent experiments. (D) The effect of sNPF mixture (1 µM each) on JH biosynthesis by the fifth instar day 0 CA. Vertical bars represent S.E. n = 6 repeats. ***P<0.001 from Student's t-test.

https://doi.org/10.1371/journal.pone.0003048.g005

We then investigated the putative regulatory role of sNPFs in JH biosynthesis using in vitro CA culture. Addition of sNPFs to the incubation medium resulted in clear inhibition of JH biosynthesis (Figure 5D), suggesting they likely have allatostatic activity in vivo. The identification of these novel allatostatins again demonstrates the effectiveness of our approach in the investigation of unknown neuropeptide functions.

Identification of Unexpected Signaling Network within CC-CA Complex

Surprisingly, in the course of detailed expression analysis of the six BNGRs, we found that BNGR-A16 is mainly expressed in the CC, not in the CA (Figure 6A). Considering that many previous studies on AT activity in vitro were conducted using the CC-CA complex, it is of interest to hypothesize that the CC affects JH biosynthesis in the CA, and AT regulates JH production indirectly via the CC. The CC is well known as the source of a peptide hormone, adipokinetic hormone [42], so it is plausible that the CC is producing other peptide ligands for the BNGRs expressed in the CA. Remarkably, RT-PCR and in situ hybridization analyses of sNPF revealed its expression mainly in the CC (Figure 6B). This result was also confirmed by direct MS analysis, which showed the existence of all three mature sNPF peptides in the CC-CA (Figure S2). Finally, double staining using immunohistochemistry and in situ hybridization revealed AT receptor expression in the sNPF-producing cells in the CC (Figure 6C). These unexpected findings suggest that AT regulates the production and/or the release of sNPFs from the CC, and at least some of the AT functions on JH biosynthesis may be mediated by sNPF (Figure 6D).

thumbnail
Figure 6. Expression of the AT receptor in sNPF-producing CC cells.

(A) Expression analysis of 6 BNGRs within CC-CA complex on the fourth instar head capsule slippage. Vertical bars represent S.E. n = 3. Arbitrary unit is the same to Figure 3. (B) Expression analysis of sNPF on the fourth instar head capsule slippage. Relative expression was shown as a percentage of the expression level in the CC. For tissue names, see Figure 3. Inset, in situ hybridization of sNPF. Scale bar, 100 µm. (C) Double staining of the CC-CA with immunohistochemistry for sNPF (upper) and in situ hybridization for BNGR-A16 (lower). Scale bar, 100 µm. (D) Schematic illustration of the regulatory pathways of JH biosynthesis identified in this study. Putative role of AT in the regulation of sNPF production in the CC remains to be elucidated.

https://doi.org/10.1371/journal.pone.0003048.g006

Discussion

In this study, we conducted a comprehensive cloning and the systematic expression analysis of neuropeptide GPCRs in the silkworm Bombyx mori. These analyses led to the following major findings: 1) identification and functional characterization of the first AT receptor, 2) purification and identification of novel allatostatic peptides produced in the CC, and 3) elucidation of an unexpected signaling network within CC-CA complex, potentially mediated by the above two factors.

Our finding that the AT receptor is predominantly expressed not in the CA but in the sNPF-producing cells of the CC raises an intriguing possibility that the allatotropic activity of AT may be mediated by the CC (Figure 6D). In support of this idea, previous assays for AT activity have been mostly done with CC-CA complexes, rather than the CA alone. If AT indeed has any effect on the production and/or the release of sNPF from the CC, such effect is most likely inhibitory, because sNPFs have clear allatostatic activity (Figure 5D). Thus, AT could inhibit the allatostatic effect of sNPFs produced in the CC, thereby exerting its indirect allatotropic effect by “derepression”. Our preliminary results that the larval CA incubated alone secretes more JH than CC-CA complex in vitro (Y.K. and K.H., unpublished) further support this idea.

However, this hypothesis must be carefully considered from several aspects. First, some allatoregulatory peptides are known to change their activity during development [43]. Detailed physiological analyses of sNPF and AT along developmental stages, as well as the improvement of a reliable in vitro CC-CA co-culture system, are thus necessary to examine the above hypothesis. Second, we cannot fully rule out the possibility that the low, but still detectable expression of the AT receptor in the CA (Figure 6A) may mediate the allatotropic activity of its ligand, though we could not detect any allatotropic activity of AT using our in vitro CA culture (data not shown). Considering the varying activity of AT during development in many insects [4], [44], the temporal expression analysis of the AT receptor within the CC-CA complex, as well as the other receptors expressed in the CA, may help clarify such a possibility. Thus, both of these alternative possibilities require the introduction of the “fourth”, temporal axis to our three-dimensional, spatial expression analysis. Although this is beyond the scope of this study, our work provides a firm foundation upon which such further analyses can be based.

Considering that sNPF has been reported in most insects examined [37], it is possible that sNPF has a conserved role in the regulation of JH biosynthesis. In relation to this, it is interesting to note that several physiological events related to sNPF in other insects (ovarian growth, body size control and adult diapause) are all reminiscent of its regulatory role in JH biosynthesis [37], [40], [41], [45]. Together with the recently reported role of sNPF in the regulation of insulin expression [46], investigation of the functional conservation of sNPF in other insect species is clearly warranted.

Identification of sNPFs as CC-produced allatoregulatory peptides may also provide a direct link between JH biosynthesis and energy balance. JH acts as a signal for starvation-induced growth inhibition, which suggests a tight link between JH biosynthesis and energy balance [47]. On the other hand, CC is well known for its glucose-sensing capacity, which regulates energy homeostasis in insects [42]. Investigation of the sNPF secretion from the CC under a variety of physiological conditions will test the validity of this possibility.

The high expression of the ETH receptor in the CA (Figure 6A) is another unexpected finding of this study. This result was further verified by in situ hybridization (Figure S3), suggesting an unknown physiological role of ETH in the regulation of JH biosynthesis. ETH is produced by Inka cells of epitracheal glands and released into the hemolymph during ecdysis [9], [39], so the putative allatoregulatory role of ETH may explain the transient rise of JH titer at ecdysis reported in many insects [48][50]. Together with the as-yet-unknown ligand of BNGR-B3, elucidation of the physiological significance of these putative regulators of JH biosynthesis will elucidate a number of unexpected features of this important endocrine pathway.

All of the above results clearly suggest that the regulatory mechanisms of JH biosynthesis are as complex as those of ecdysteroidogenesis, in which multiple factors are involved [51]. Indeed, most tissues investigated in this study express various neuropeptide GPCRs (Figure 3), indicating the necessity of a systematic approach to reveal regulatory mechanisms of their functions. The present study provides a solid basis for such an approach, which leads to our thorough understanding of the neuropeptide circuitry that extends beyond individual tissues to form a system-wide network in multicellular organisms.

Materials and Methods

Comprehensive identification of BNGRs

BNGR genes were identified by a cloning-based approach combining homology searching and RACE, using the SMART RACE cDNA Amplification Kit (Clontech Laboratories, Mountain View, CA) according to the manufacturer's instructions. Whole coding regions of all the identified BNGR genes were cloned, sequenced, and confirmed to have seven transmembrane regions and full conserved domains for GPCRs. See Results and Tables S1, S2, S3 for the detailed procedure. The primers used for RACE and cloning are listed in Table S4 and S5, respectively.

Phylogenetic analysis

Deduced amino acid sequences of Bombyx and Drosophila neuropeptide receptors were aligned and rooted neighbor-joining trees were generated using ClustalW. Bootstrap analyses of 1,000 replications were conducted to assess the relationships.

Quantitative RT-PCR (Q-RT-PCR) analysis of the neuropeptide GPCR genes

cDNAs from various tissues were synthesized and Q-RT-PCR was carried out essentially as described [52]. After PCR, the absence of unwanted byproducts was confirmed by automated melting curve analysis. Serial dilutions of pME18S plasmids containing coding sequences of the neuropeptide GPCR genes were used for standards. Rp49 was chosen as a reference gene, and serial dilutions of a pCR2.1 plasmid containing 340 bp of Rp49 cDNA were used for standards. The molar amounts of the GPCR genes and Rp49 were calculated, and the transcript levels of the GPCR genes were normalized with Rp49 levels in the same samples and presented as arbitrary units. The primers used for Q-RT-PCR are listed in Table S6.

Cloning of AST-C

Using the amino acid sequence of Manduca sexta allatostatin (GenBank accession no. P42559) as a query, BLAST analysis (TBLASTN) was performed against Bombyx whole-genome shotgun sequence contigs using KAIKOBLAST. Based on the putative coding sequence of the obtained contig (contig 454046, GenBank accession no. BAAB01095856), RACE primers were designed as follows: sense primer, 5′-ATGGCAGATTGGTGCGACCGTGGAGGG-3′; sense nested primer, 5′-CCGGCAGTGCTATTTCAATCCCATCTCC-3′; antisense primer, 5′-ATAGGGCGAAAGCCGGGGTAAAAGTCG-3′; antisense nested primer, 5′-GGGTAAAAGTCGTGTTATTCACTTGCGG-3′. To obtain the whole cDNA sequence of AST-C (GenBank accession no. AB330418), RACE was performed by using the above primers and the SMART RACE cDNA Amplification Kit. First-strand cDNA prepared from larval brain total RNA was used as a template.

Purification of sNPF-2 and 3

Residual fractions of the first purification step of Bombyx pupal brain extracts, which were previously used for the purification of BMS [11], were applied to the FMRFamide ELISA as previously described [12]. One of the immunoreactive fractions (No. 16 in Figure 5A) was loaded onto a PEGASIL-300 C4P column (4.6×250 mm, Senshu Kagaku Co. Ltd., Tokyo, Japan) using a Waters 2695 Separations Module (Waters, Milford, MA). Elution was performed with a linear gradient of 10–40% acetonitrile in 0.1% trifluoroacetic acid (TFA) over 60 min at the flow rate of 1 ml/min. This resulted in the detection of two FMRFamide-positive fractions, each of which was then applied onto a PEGASIL-300 ODS-II column (4.6×250 mm, Senshu Kagaku Co. Ltd.). The column was eluted with a linear gradient of 10–40% acetonitrile in 0.1% TFA over 60 min at a flow rate of 0.5 ml/min. Each immunoreactive fraction was further purified by passing over a PEGASIL-300 ODS-II column (4.6×250 mm) eluting it with a linear gradient of 10–40% acetonitrile in 0.1% heptafluorobutyric acid over 60 min at the flow rate of 0.5 ml/min.

MALDI-TOF MS

Mass spectra and postsource decay (PSD) mass spectra of peptides were obtained with a AXIMA-CFR (Shimadzu Corporation, Kyoto, Japan) mass spectrometer. Re-crystallized α-cyano-4-hydroxycinnamic acid (α-CHCA) was used as the matrix with a concentration of 5 mg/ml of acetonitrile/water (1∶1, by volume) containing 0.1% TFA. All spectra were measured in reflector mode.

Cloning of sNPF

Using the amino acid sequence of one of the purified peptide [TPVR(L/I)RFG] as a query, BLAST analysis (TBLASTN) was performed against Bombyx whole-genome shotgun sequence contigs using KAIKOBLAST. Based on the putative coding sequence of the obtained contig (contig 432136, GenBank accession no. BAAB01085626), 5′-RACE was performed using the following primer: antisense primer, 5′-CGGAGAGAGAGCAGCTCATCCATCTCG-3′. Based on the obtained sequence of the 5′-RACE product, 3′-RACE was further performed using the following two primers: sense primer, 5′-GATCGCCATCCCGACGCCTGCGTTTCG-3′; sense nested primer, 5′-TGCGTTTCGGCCGCCGATCAGATCCCG-3′. First-strand cDNA prepared from larval brain total RNA and the SMART RACE cDNA Amplification Kit were used.

Q-RT-PCR analysis of sNPF

Q-RT-PCR of sNPF (GenBank accession no. AB330419) was performed as described above for the neuropeptide GPCR genes. Specific primers used are as follows: sense primer, 5′-AAGTCTGTTCGATCGCCATCCCGAC-3′; antisense primer, 5′-GAGAGAGAGCAGCTCATCCATCTCG-3′. The transcript levels of the sNPF were normalized with Rp49 levels in the same samples.

In situ hybridization and immunohistochemistry

In situ hybridization and immunohistochemistry was performed as previously described [9], [12]. Specific primers used for the production of a digoxigenin-labeled probe were as follows: sNPF sense primer, 5′-GGAGGGCTCTATGCTCTTCT-3′; sNPF antisense primer, 5′-TACATCTAGGTTGCGACGAG-3′; BNGR-A16 sense primer 1, 5′-GAGAATGAAACCTGCGTTGGAGATCC-3′; BNGR-A16 sense primer 2, 5′-ACGCTGACTTTCATCTCTGTGGACCG-3′; BNGR-A16 antisense primer 1, 5′-TGGTATGTCCTGGCTTGAAGGATGCC-3′; BNGR-A16 antisense primer 2, 5′-ACGCAGTTGTGTCAGTTCTGCAGTGG-3′; BNGR-A6-A sense primer 1, 5′-CGCGAGAGCTAGAAAACAGG-3′; BNGR-A6-A sense primer 2, 5′-CCGTCGTGAATCAATGTCTG-3′; BNGR-A6-A antisense primer 1, 5′-TCTTTCCGAATCAACGCTCT-3′; BNGR-A6-A antisense primer 2, 5′-GCCACAACGGAATCAACTTT-3′. For sNPF immunohistochemistry, rabbit polyclonal antiserum to RFamide [12] was used.

Direct MS analysis of CC-CA

CC-CA complex was dissected in sterile saline [0.85% NaCl (w/v)], and direct MS analysis was performed as previously described [12].

Synthetic peptides

AST-C, AT, BMS, BRFa-3 and sNPF peptides were all custom synthesized and purified by HPLC. Synthetic AST-C was converted to the disulfide form by iodine oxidation and further purified by HPLC.

HEK293 cell expression and Ca2+ imaging analysis of BNGRs

BNGRs cloned into pME18S vector were transfected into HEK293 cells with the promiscuous G protein Gα15 as described [53]. Ca2+ imaging analysis was performed as reported [53]. Briefly, the transfected cells were cultured for 24 h, and loaded with 2.5 µM fura-2 AM (Invitrogen, Carlsbad, CA) for 30 min at 37°C. Neuropeptides were applied sequentially to the cells for 20 s with a peristaltic pump at a flow rate of 1.5 ml/min. Fluorescence at 510 nm (excitation at 340 or 380 nm with a xenon lamp) was monitored with an intensified CCD camera. An Argus-HiSCA calcium-imaging system (Hamamatsu Photonics, Shizuoka, Japan) was used to determine the fluorescence ratio at 340 and 380 nm.

In vitro radiochemical assay for JH biosynthesis

JH biosynthesis by the CA was measured using an in vitro radiochemical assay as previously described [54], except that the CA was used instead of the CC-CA complex. The mixture of three sNPFs (1 µM each) was added to the incubation medium.

Supporting Information

Table S1.

Summary of the two-way screening of Bombyx whole-genome shotgun sequence contigs. Query Drosophila GPCRs are aligned at the top, and the 195 contigs obtained through the first screening were listed on the left. Figures in the table indicate the BLAST scores of the first screening. Right two columns show the result of the second screening, which extracted 139 putative BNGR contigs (red).

https://doi.org/10.1371/journal.pone.0003048.s001

(0.06 MB XLS)

Table S2.

List of contigs containing the fragments of previously reported Bombyx neuropeptide GPCR genes. Putative BNGR contigs are in red letters.

https://doi.org/10.1371/journal.pone.0003048.s002

(0.02 MB XLS)

Table S3.

List of the newly-identified GPCRs and their corresponding contigs. Putative BNGR contigs are in red letters. Tissue sources for the cDNA clones are indicated on the right: FB, fat body; BR, brain; MG, midgut; OV, ovary; TE, testis; PG, prothoracic gland.

https://doi.org/10.1371/journal.pone.0003048.s003

(0.02 MB XLS)

Table S4.

List of the primers used for the RACE of BNGRs.

https://doi.org/10.1371/journal.pone.0003048.s004

(0.03 MB XLS)

Table S5.

List of the primers used for cloning of BNGRs. Each primer contains a restriction site indicated in the primer name, while all the sense primers incorporate Kozak sequence (“GCCACC”).

https://doi.org/10.1371/journal.pone.0003048.s005

(0.02 MB XLS)

Table S6.

List of the primers used for the Q-RT-PCR analysis.

https://doi.org/10.1371/journal.pone.0003048.s006

(0.03 MB XLS)

Figure S1.

Postsource decay mass spectra of TPVRLRFamide (A) and SMRLRFamide (B). For each peptide, upper panel is the purified sample and the lower panel is the synthetic sample. The b- and y-type ion signals are indicated at the top. Signals of the indicated mass range are magnified 10 times to facilitate comparison. The final confirmation of the leucine residue was carried out based on the cDNA sequence.

https://doi.org/10.1371/journal.pone.0003048.s007

(0.39 MB TIF)

Figure S2.

Direct MS analysis of CC-CA complex. The identified peaks in (A) are summarized in (B). Asterisks in (A) denote sodium adducts for the peptides 5, 9, 12 and 13.

https://doi.org/10.1371/journal.pone.0003048.s008

(0.97 MB TIF)

Figure S3.

In situ hybridization analysis of BNGR-A6-A on the CC-CA. Scale bar, 100 micrometer.

https://doi.org/10.1371/journal.pone.0003048.s009

(7.87 MB TIF)

Acknowledgments

We thank Drs. Yusuke Suzuki, Minoru Suzuki and Akemi Suzuki for MS analysis, Dr. Michael E. Adams for providing synthetic ETH and PETH, Drs. Ivana Daubnerová, Laco Roller and Young-Joon Kim for designing and preparing in situ probes, and Dr. Michael B. O'Connor for critical reading of the manuscript.

Author Contributions

Conceived and designed the experiments: NY KH YT TS HK. Performed the experiments: NY SY D KW YK KH YT TS. Analyzed the data: NY. Contributed reagents/materials/analysis tools: TK HS. Wrote the paper: NY D KH TS.

References

  1. 1. Hewes RS, Taghert PH (2001) Neuropeptides and neuropeptide receptors in the Drosophila melanogaster genome. Genome Res 11: 1126–1142.
  2. 2. Hauser F, Williamson M, Cazzamali G, Grimmelikhuijzen CJ (2006) Identifying neuropeptide and protein hormone receptors in Drosophila melanogaster by exploiting genomic data. Brief Funct Genomic Proteomic 4: 321–330.
  3. 3. Kataoka H, Toschi A, Li J, Carney RL, Schooley DA, et al. (1989) Identification of an allatotropin from adult Manduca sexta. Science 243: 1481–1483.
  4. 4. Gilbert LI, Granger NA, Roe RM (2000) The juvenile hormones: historical facts and speculations on future research directions. Insect Biochem Mol Biol 30: 617–644.
  5. 5. Riddiford LM (1994) Cellular and molecular actions of juvenile hormone. I.General considerations and premetamorphic actions. Advances in Insect Physiology 24: 213–274.
  6. 6. Wyatt GR, Davey KG (1996) Cellular and molecular actions of juvenile hormone. II. Roles of juvenile hormone in adult insects. Advances in Insect Physiology 26: 1–155.
  7. 7. Truman JW, Riddiford LM (2007) The morphostatic actions of juvenile hormone. Insect Biochem Mol Biol 37: 761–770.
  8. 8. Wu Q, Wen T, Lee G, Park JH, Cai HN, et al. (2003) Developmental control of foraging and social behavior by the Drosophila neuropeptide Y-like system. Neuron 39: 147–161.
  9. 9. Kim YJ, Zitnan D, Galizia CG, Cho KH, Adams ME (2006) A command chemical triggers an innate behavior by sequential activation of multiple peptidergic ensembles. Curr Biol 16: 1395–1407.
  10. 10. Kim YJ, Spalovska-Valachova I, Cho KH, Zitnanova I, Park Y, et al. (2004) Corazonin receptor signaling in ecdysis initiation. Proc Natl Acad Sci U S A 101: 6704–6709.
  11. 11. Yamanaka N, Hua YJ, Mizoguchi A, Watanabe K, Niwa R, et al. (2005) Identification of a novel prothoracicostatic hormone and its receptor in the silkworm Bombyx mori. J Biol Chem 280: 14684–14690.
  12. 12. Yamanaka N, Zitnan D, Kim YJ, Adams ME, Hua YJ, et al. (2006) Regulation of insect steroid hormone biosynthesis by innervating peptidergic neurons. Proc Natl Acad Sci U S A 103: 8622–8627.
  13. 13. Mita K, Kasahara M, Sasaki S, Nagayasu Y, Yamada T, et al. (2004) The genome sequence of silkworm, Bombyx mori. DNA Res 11: 27–35.
  14. 14. Xia Q, Zhou Z, Lu C, Cheng D, Dai F, et al. (2004) A draft sequence for the genome of the domesticated silkworm (Bombyx mori). Science 306: 1937–1940.
  15. 15. Vassilatis DK, Hohmann JG, Zeng H, Li F, Ranchalis JE, et al. (2003) The G protein-coupled receptor repertoires of human and mouse. Proc Natl Acad Sci U S A 100: 4903–4908.
  16. 16. Metpally RP, Sowdhamini R (2005) Cross genome phylogenetic analysis of human and Drosophila G protein-coupled receptors: application to functional annotation of orphan receptors. BMC Genomics 6: 106.
  17. 17. Fredriksson R, Schioth HB (2005) The repertoire of G-protein-coupled receptors in fully sequenced genomes. Mol Pharmacol 67: 1414–1425.
  18. 18. Hauser F, Cazzamali G, Williamson M, Blenau W, Grimmelikhuijzen CJ (2006) A review of neurohormone GPCRs present in the fruitfly Drosophila melanogaster and the honey bee Apis mellifera. Prog Neurobiol 80: 1–19.
  19. 19. Bjarnadottir TK, Gloriam DE, Hellstrand SH, Kristiansson H, Fredriksson R, et al. (2006) Comprehensive repertoire and phylogenetic analysis of the G protein-coupled receptors in human and mouse. Genomics 88: 263–273.
  20. 20. Brody T, Cravchik A (2000) Drosophila melanogaster G protein-coupled receptors. J Cell Biol 150: F83–88.
  21. 21. Iversen A, Cazzamali G, Williamson M, Hauser F, Grimmelikhuijzen CJ (2002) Molecular identification of the first insect ecdysis triggering hormone receptors. Biochem Biophys Res Commun 299: 924–931.
  22. 22. Iversen A, Cazzamali G, Williamson M, Hauser F, Grimmelikhuijzen CJ (2002) Molecular cloning and functional expression of a Drosophila receptor for the neuropeptides capa-1 and -2. Biochem Biophys Res Commun 299: 628–633.
  23. 23. Egerod K, Reynisson E, Hauser F, Cazzamali G, Williamson M, et al. (2003) Molecular cloning and functional expression of the first two specific insect myosuppressin receptors. Proc Natl Acad Sci U S A 100: 9808–9813.
  24. 24. Cazzamali G, Hauser F, Kobberup S, Williamson M, Grimmelikhuijzen CJ (2003) Molecular identification of a Drosophila G protein-coupled receptor specific for crustacean cardioactive peptide. Biochem Biophys Res Commun 303: 146–152.
  25. 25. Ha SD, Kataoka H, Suzuki A, Kim BJ, Kim HJ, et al. (2000) Cloning and sequence analysis of cDNA for diuretic hormone receptor from the Bombyx mori. Mol Cells 10: 13–17.
  26. 26. Secher T, Lenz C, Cazzamali G, Sorensen G, Williamson M, et al. (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.
  27. 27. Staubli F, Jorgensen TJ, Cazzamali G, Williamson M, Lenz C, et al. (2002) Molecular identification of the insect adipokinetic hormone receptors. Proc Natl Acad Sci U S A 99: 3446–3451.
  28. 28. Hull JJ, Ohnishi A, Moto K, Kawasaki Y, Kurata R, et al. (2004) Cloning and characterization of the pheromone biosynthesis activating neuropeptide receptor from the silkmoth, Bombyx mori. Significance of the carboxyl terminus in receptor internalization. J Biol Chem 279: 51500–51507.
  29. 29. Jorgensen LM, Hauser F, Cazzamali G, Williamson M, Grimmelikhuijzen CJ (2006) Molecular identification of the first SIFamide receptor. Biochem Biophys Res Commun 340: 696–701.
  30. 30. Homma T, Watanabe K, Tsurumaru S, Kataoka H, Imai K, et al. (2006) G protein-coupled receptor for diapause hormone, an inducer of Bombyx embryonic diapause. Biochem Biophys Res Commun 344: 386–393.
  31. 31. Krogh A, Larsson B, von Heijne G, Sonnhammer EL (2001) Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J Mol Biol 305: 567–580.
  32. 32. Marchler-Bauer A, Anderson JB, Cherukuri PF, DeWeese-Scott C, Geer LY, et al. (2005) CDD: a Conserved Domain Database for protein classification. Nucleic Acids Res 33: D192–196.
  33. 33. Oda Y, Uejuma M, Iwami M, Sakurai S (2000) Involvement of adipokinetic hormone in the homeostatic control of hemolymph trehalose concentration in the larvae of Bombyx mori. Arch Insect Biochem Physiol 45: 156–165.
  34. 34. Audsley N, Kay I, Hayes TK, Coast GM (1995) Cross reactivity studies of CRF-related peptides on insect Malpighian tubules. Comp Biochem Physiol A Physiol 110: 87–93.
  35. 35. Kramer SJ, Toschi A, Miller CA, Kataoka H, Quistad GB, et al. (1991) Identification of an allatostatin from the tobacco hornworm Manduca sexta. Proc Natl Acad Sci U S A 88: 9458–9462.
  36. 36. Kreienkamp HJ, Larusson HJ, Witte I, Roeder T, Birgul N, et al. (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.
  37. 37. Nassel DR (2002) Neuropeptides in the nervous system of Drosophila and other insects: multiple roles as neuromodulators and neurohormones. Prog Neurobiol 68: 1–84.
  38. 38. Liu F, Baggerman G, D'Hertog W, Verleyen P, Schoofs L, et al. (2006) In silico identification of new secretory peptide genes in Drosophila melanogaster. Mol Cell Proteomics 5: 510–522.
  39. 39. Zitnan D, Kim YJ, Zitnanova I, Roller L, Adams ME (2007) Complex steroid-peptide-receptor cascade controls insect ecdysis. Gen Comp Endocrinol 153: 88–96.
  40. 40. De Loof A, Baggerman G, Breuer M, Claeys I, Cerstiaens A, et al. (2001) Gonadotropins in insects: an overview. Arch Insect Biochem Physiol 47: 129–138.
  41. 41. Lee KS, You KH, Choo JK, Han YM, Yu K (2004) Drosophila short neuropeptide F regulates food intake and body size. J Biol Chem 279: 50781–50789.
  42. 42. Kim SK, Rulifson EJ (2004) Conserved mechanisms of glucose sensing and regulation by Drosophila corpora cardiaca cells. Nature 431: 316–320.
  43. 43. Stay B, Zhang JR, Tobe SS (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.
  44. 44. Elekonich MM, Horodyski FM (2003) Insect allatotropins belong to a family of structurally-related myoactive peptides present in several invertebrate phyla. Peptides 24: 1623–1632.
  45. 45. Huybrechts J, De Loof A, Schoofs L (2004) Diapausing Colorado potato beetles are devoid of short neuropeptide F I and II. Biochem Biophys Res Commun 317: 909–916.
  46. 46. Lee KS, Kwon OY, Lee JH, Kwon K, Min KJ, et al. (2008) Drosophila short neuropeptide F signalling regulates growth by ERK-mediated insulin signalling. Nat Cell Biol 10: 468–475.
  47. 47. Truman JW, Hiruma K, Allee JP, Macwhinnie SG, Champlin DT, et al. (2006) Juvenile hormone is required to couple imaginal disc formation with nutrition in insects. Science 312: 1385–1388.
  48. 48. Plantevin G, De Reggi M, Nardon C (1984) Changes in ecdysteroid and juvenile hormone titers in the hemolymph of Galleria mellonella larvae and pupae. Gen Comp Endocrinol 56: 218–230.
  49. 49. Baker FC, Tsai LW, Reuter CC, Schooley DA (1987) In vivo fluctuation of JH, JH acid, and ecdysteroid titer, and JH esterase activity, during development of fifth stadium Manduca sexta. Insect Biochem 17: 989–996.
  50. 50. Sakurai S, Niimi S (1997) Development changes in juvenile hormone and juvenile hormone acid titers in the hemolymph and in-vitro juvenile hormone synthesis by corpora allata of the silkworm, Bombyx mori. J Insect Physiol 43: 875–884.
  51. 51. Truman JW (2006) Steroid hormone secretion in insects comes of age. Proc Natl Acad Sci U S A 103: 8909–8910.
  52. 52. 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.
  53. 53. Katada S, Nakagawa T, Kataoka H, Touhara K (2003) Odorant response assays for a heterologously expressed olfactory receptor. Biochem Biophys Res Commun 305: 964–969.
  54. 54. Kinjoh T, Kaneko Y, Itoyama K, Mita K, Hiruma K, et al. (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.
  55. 55. Garczynski SF, Brown MR, Shen P, Murray TF, Crim JW (2002) Characterization of a functional neuropeptide F receptor from Drosophila melanogaster. Peptides 23: 773–780.
  56. 56. Cazzamali G, Grimmelikhuijzen CJ (2002) Molecular cloning and functional expression of the first insect FMRFamide receptor. Proc Natl Acad Sci U S A 99: 12073–12078.
  57. 57. Meeusen T, Mertens I, Clynen E, Baggerman G, Nichols R, et al. (2002) Identification in Drosophila melanogaster of the invertebrate G protein-coupled FMRFamide receptor. Proc Natl Acad Sci U S A 99: 15363–15368.
  58. 58. Birgul N, Weise C, Kreienkamp HJ, 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.
  59. 59. Larsen MJ, Burton KJ, Zantello MR, Smith VG, Lowery DL, et al. (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.
  60. 60. Li XJ, Wu YN, North RA, Forte M (1992) Cloning, functional expression, and developmental regulation of a neuropeptide Y receptor from Drosophila melanogaster. J Biol Chem 267: 9–12.
  61. 61. Park Y, Kim YJ, Dupriez V, Adams ME (2003) Two subtypes of ecdysis-triggering hormone receptor in Drosophila melanogaster. J Biol Chem 278: 17710–17715.
  62. 62. Park Y, Kim YJ, Adams ME (2002) Identification of G protein-coupled receptors for Drosophila PRXamide peptides, CCAP, corazonin, and AKH supports a theory of ligand-receptor coevolution. Proc Natl Acad Sci U S A 99: 11423–11428.
  63. 63. Monnier D, Colas JF, Rosay P, Hen R, Borrelli E, et al. (1992) NKD, a developmentally regulated tachykinin receptor in Drosophila. J Biol Chem 267: 1298–1302.
  64. 64. Johnson EC, Bohn LM, Barak LS, Birse RT, Nassel DR, et al. (2003) Identification of Drosophila neuropeptide receptors by G protein-coupled receptors-beta-arrestin2 interactions. J Biol Chem 278: 52172–52178.
  65. 65. Kubiak TM, Larsen MJ, Burton KJ, Bannow CA, Martin RA, et al. (2002) Cloning and functional expression of the first Drosophila melanogaster sulfakinin receptor DSK-R1. Biochem Biophys Res Commun 291: 313–320.
  66. 66. Johnson EC, Garczynski SF, Park D, Crim JW, Nassel DR, et al. (2003) Identification and characterization of a G protein-coupled receptor for the neuropeptide proctolin in Drosophila melanogaster. Proc Natl Acad Sci U S A 100: 6198–6203.
  67. 67. Egerod K, Reynisson E, Hauser F, Williamson M, Cazzamali G, et al. (2003) Molecular identification of the first insect proctolin receptor. Biochem Biophys Res Commun 306: 437–442.
  68. 68. Mertens I, Meeusen T, Huybrechts R, De Loof A, Schoofs L (2002) Characterization of the short neuropeptide F receptor from Drosophila melanogaster. Biochem Biophys Res Commun 297: 1140–1148.
  69. 69. Li XJ, Wolfgang W, Wu YN, North RA, Forte M (1991) Cloning, heterologous expression and developmental regulation of a Drosophila receptor for tachykinin-like peptides. Embo J 10: 3221–3229.
  70. 70. Johnson EC, Bohn LM, Taghert PH (2004) Drosophila CG8422 encodes a functional diuretic hormone receptor. J Exp Biol 207: 743–748.
  71. 71. Rosenkilde C, Cazzamali G, Williamson M, Hauser F, Sondergaard L, et al. (2003) Molecular cloning, functional expression, and gene silencing of two Drosophila receptors for the Drosophila neuropeptide pyrokinin-2. Biochem Biophys Res Commun 309: 485–494.
  72. 72. Cazzamali G, Torp M, Hauser F, Williamson M, Grimmelikhuijzen CJ (2005) The Drosophila gene CG9918 codes for a pyrokinin-1 receptor. Biochem Biophys Res Commun 335: 14–19.
  73. 73. Radford JC, Davies SA, Dow JA (2002) Systematic G-protein-coupled receptor analysis in Drosophila melanogaster identifies a leucokinin receptor with novel roles. J Biol Chem 277: 38810–38817.
  74. 74. Cazzamali G, Saxild N, Grimmelikhuijzen C (2002) Molecular cloning and functional expression of a Drosophila corazonin receptor. Biochem Biophys Res Commun 298: 31–36.
  75. 75. Mertens I, Vandingenen A, Johnson EC, Shafer OT, Li W, et al. (2005) PDF receptor signaling in Drosophila contributes to both circadian and geotactic behaviors. Neuron 48: 213–219.
  76. 76. Lear BC, Merrill CE, Lin JM, Schroeder A, Zhang L, et al. (2005) A G protein-coupled receptor, groom-of-PDF, is required for PDF neuron action in circadian behavior. Neuron 48: 221–227.
  77. 77. Hyun S, Lee Y, Hong ST, Bang S, Paik D, et al. (2005) Drosophila GPCR Han is a receptor for the circadian clock neuropeptide PDF. Neuron 48: 267–278.
  78. 78. Johnson EC, Shafer OT, Trigg JS, Park J, Schooley DA, et al. (2005) A novel diuretic hormone receptor in Drosophila: evidence for conservation of CGRP signaling. J Exp Biol 208: 1239–1246.