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

Identification of candidate chemosensory genes by transcriptome analysis in Loxostege sticticalis Linnaeus

  • Hong-Shuang Wei,

    Affiliation State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, China

  • Ke-Bin Li,

    Affiliation State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, China

  • Shuai Zhang,

    Affiliation State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, China

  • Ya-Zhong Cao,

    Affiliation State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, China

  • Jiao Yin

    jyin@ippcaas.cn

    Affiliation State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, China

Abstract

Loxostege sticticalis Linnaeus is an economically important agricultural pest, and the larvae cause great damage to crops, especially in Northern China. However, effective and environmentally friendly chemical methods for controlling this pest have not been discovered to date. In the present study, we performed HiSeq2500 sequencing of transcriptomes of the male and female adult antennae, adult legs and third instar larvae, and we identified 54 candidate odorant receptors (ORs), including 1 odorant receptor coreceptor (Orco) and 5 pheromone receptors (PRs), 18 ionotropic receptors (IRs), 13 gustatory receptors (GRs), 34 odorant binding proteins (OBPs), including 1 general odorant binding protein (GOBP1) and 3 pheromone binding proteins (PBPs), 10 chemosensory proteins (CSPs) and 2 sensory neuron membrane proteins (SNMPs). The results of RNA-Seq and RT-qPCR analyses showed the expression levels of most genes in the antennae were higher than that in the legs and larvae. Furthermore, PR4, OR1-4, 7–11, 13–15, 23, 29–32, 34, 41, 43, 47/IR7d.2/GR5b, 45, 7/PBP2-3, GOBP1, OBP3, 8 showed female antennae-biased expression, while PR1/OBP2, 7/IR75d/CSP2 showed male antennae-biased expression. However, IR1, 7d.3, 68a/OBP11, 20–22, 28/CSP9 had larvae enriched expression, and OBP15, 17, 25, 29/CSP5 were mainly expressed in the legs. The results shown above indicated that these genes might play a key role in foraging, seeking mates and host recognition in the L. sticticalis. Our findings will provide the basic knowledge for further studies on the molecular mechanisms of the olfactory system of L. sticticalis and potential novel targets for pest control strategies.

Introduction

The beet webworm, Loxostege sticticalis L. (Lepidoptera: Pyralidae), a worldwide distributed and migratory pest in North China, causes serious economic damage every year [1, 2]. L. sticticalis seems to be polyphagous in its larval stage, but it has been reported to have obvious host-plant selection for crops (sugar beet, potato and soybean) and pastures [35]. This has been associated with its highly developed olfactory system to detect and distinguish the host-plant volatiles [5, 6].

Chemical sensing by olfaction can regulate insect behaviors, including seeking food, choosing mates, locating suitable oviposition sites, and avoiding natural enemies [7, 8, 9]. Insects discern chemical signals by olfactory receptor neurons (ORNs) in the olfactory sensilla [8]. The ORNs located at the sensilla root are the primary units of olfaction in the insect antennae which include the odorant binding proteins (OBPs), chemosensory proteins (CSPs), odorant receptors (ORs), ionotropic receptors (IRs), and the sensory neuron membrane proteins (SNMPs) [8, 10]. OBPs dissolved in the sensilla lymph are some kinds of acidic proteins with a pattern of six conserved cysteine residues [11]. Insect OBPs were mainly expressed in the antennae of both sexes, which allows the insect to identify odor molecules in environment and plays an important role in the process of insect host location [12, 13]. Two subfamilies of OBPs, general odorant-binding proteins (GOBPs) and pheromone binding proteins (PBPs), are respectively responsible for recognizing and transporting host-plant volatiles and pheromones to ORs to protect them from odorant-degrading enzymes (ODEs) [1416]. Same as OBPs, other soluble proteins named CSPs are also secreted in the sensillum lymph [16]. Although the functions of CSPs reported in previous articles are analogous to OBPs, they are still poorly understood. SNMPs with two transmembrane domains, the accepting stations of odorant ligands located in the dendritic membranes of pheromone-sensitive neurons, play a role in capturing pheromone molecules in coordination with ORs [1719].

There are two types of olfactory receptor (ORs and IRs) proteins and one type of gustatory receptors (GRs) in insects. The conventional ORs binding the ligand molecules released by OBPs are also trans-membrane proteins with seven conservative transmembrane domains [20]. Pheromone sensilla primarily located on the antennae can perceive the pheromone molecules at the periphery of the olfactory system, and pheromone molecules transported to the dendritic membranes of ORNs are recognized by pheromone receptors (PRs), which are a subclass of insect ORs [21]. Beyond that, the odorant receptor coreceptor (Orco) was proved to be heteromeric ligand-gated ion channels and cyclic-nucleotide-activated cation channels with the capacity for transforming chemical signals to electric signals [2225]. Compared to ligands (esters and alcohols) binding to ORs, IRs are narrowly tuned for amine and acid ones [26, 27, 28]. Furthermore, IRs are more standard ion acceptors compared with the ORs [26, 27]. A family proteins of sense of taste expressed in the antennae, proboscis and palps, GRs, were still exposed that they were adjusted for CO2 detection and responsible for selecting brooding spots [29, 30].

In the Lepidoptera, the antenna is a specialized organ for insect sensing, especially for olfaction, and many olfactory genes in some moths have been studied by antennal transcriptome analysis [31, 32]. However, the legs that also have a special olfaction sense though less sensitive than olfaction in the antennae [33, 34], its olfactory gene database seems incomplete for the L. sticticalis. In this study, we sequenced and analyzed integral transcriptomes of L. sticticalis adult antennae, adult legs and third instar larvae using Illumina sequencing platform. Our aims were to identify chemosensory genes of L. sticticalis and report the results including sequencing, gene annotation, GO annotation and specifically, identification and expression pattern of ORs, IRs, GRs, OBPs, CSPs and SNMPs.

Materials and methods

Insect rearing and RNA preparation

The beet webworms were acquired from a laboratory population at the Institute of Plant Protection, Chinese Academy of Agricultural Sciences (Beijing, China). The insects were fed an artificial diet at a temperature of 22 ± 1°C with 70% ± 10% relative humidity under a photoperiod of 16L: 8D (Light, Dark). When the larvae grew up to the third instar, 20 third instar larvae were picked and frozen in liquid nitrogen for conservation. Male and female pupae were placed into separate cages for eclosion. The adult moths were fed with a 5% honey solution after emergence. The antennae and legs from the male and female individuals were excised at 1 to 3 days after eclosion, immediately frozen and stored in liquid nitrogen until the RNA extraction.

The total RNAs were isolated from 100 adult male antennae, 100 adult female antennae, 24 adult legs (male: female = 1:1) and 2 third instar larvae respectively. Three biological replicates were prepared for each pilot part. Total RNA was extracted using Trizol reagent (Invitrogen, Shanghai, China), following the manufacturer’s instructions. The integrity of the RNA samples was detected by gel electrophoresis, and a NanoDrop 2000 spectrophotometer (NanoDrop, Wilmington, DE, USA) was used to determine RNA quantity. Before sequencing, the RNA samples were stored at -80°C.

cDNA library construction, and Illumina sequencing

The cDNA library construction and Illumina sequencing of our RNA samples were performed at Biomarker technologies CO., LTD., Beijing, China. First, the NanoDrop 2000, Qubit 2.0(Invitrogen, Carlsbad, CA, USA) and Agilent 2100 (Agilent Technologies, Santa Clara, CA, USA) methods were used respectively to detect the purity, concentration and integrity of each RNA sample (10ug). Second, Oligo (dT) magnetic beads were used to gather mRNA (poly-A RNA). Using a fragmentation buffer, the mRNA of each sample was broken into short fragments randomly at 94°C for 5 min. Third, The first-strand cDNA were synthesized using N6 random primers and mRNA templates and the second strand cDNA were synthesized using buffer, dNTPs, RNase H and DNA polymerase I. The synthetic cDNA was purified using AMPure XP Beads (Beckman Coulter, Inc.). These dual-strand DNA samples were treated with T4 DNA polymerase and T4 polynucleotide kinase, respectively, for end-repairing and dA-tailing, followed by adaptor ligation to the dA tail of the dsDNA using T4 DNA ligase. Then, suitable fragments were selected with AMPure XP beads (Beckman Coulter, Inc.). Finally, the products were amplified by PCR and purified using the QIAquick PCR Purification Kit (Qiagen, Valencia, CA, USA) to create a cDNA library. The libraries were sequenced on an Illumina HiSeq 2500 platform, and paired-end reads were generated using a PE125 strategy (paired-end reads of 125 base pairs per read).

De novo assembly and function annotation

High-quality clean reads were obtained from the raw reads by removing reads containing either an adapter or poly-N sequence and reads that were in low-quality. Transcriptome de novo assembly was performed with the short read assembly program Trinity [35]. Then, the Trinity outputs were clustered by TGICL [36]. The consensus cluster sequences and singletons compose the unigene dataset. The annotation of unigenes was performed by NCBI BLASTx against a pooled database of non-redundant (nr) and Swiss-Prot protein sequences with e-values < 1e-5. The Blast results were then imported into the Blast2GO [37] pipeline for GO Annotation. Protein coding region prediction was performed by OrfPredictor [38] according to the blast results.

Sequence analysis

The sequence analysis methods used in this paper were as previously described [33]. First, the open reading frames (ORFs) of chemosensory genes in L. sticticalis were predicted online using ORF finder (http://www.ncbi.nlm.nih.gov/gorf/gorf.html). Second, similarity searches were performed with the NCBI-BLAST network server (http://blast.ncbi.nlm.nih.gov/). Then, N-terminal signal peptides of putative LstiOBPs and LstiCSPs were predicted by the SignalP 4.0 server (http://www.cbs.dtu.dk/services/SignalP/). The transmembrane domains of the candidate LstiORs, LstiIRs, LstiGRs and LstiSNMPs were predicted with the TMHMM Server Version 2.0 (http://www.cbs.dtu.dk/services/TMHMM). The nucleotide sequences of all identified olfactory gene are listed in supporting information (S1 Table).

Phylogenetic tree analysis

Multiple alignments of the L. sticticalis amino acid sequences of the chemosensory genes were performed by ClustalX 2.0 [39]. The phylogenetic trees were constructed by MEGA 6.0 [40] using the neighbor-joining method [41] with a p-distance model and a pairwise deletion of gaps. Bootstrap support was assessed by a boot strap procedure based on 1000 replicates. The data sets of chemosensory gene sequences, which were chosen from other Lepidopteran species, are listed in supporting information (S2 Table).

RT-qPCR analysis

Using real-time quantitative PCR (RT-qPCR), we measured the expression profiles of chemosensory genes in different parts (male antennae, female antennae, legs and third instar larvae). The primers used for the RT-qPCR were designed using the Primer Premier 5.0, which are listed in supporting information (S3 Table). The RT-qPCR was performed by ABI 7500 Detection System (Applied Biosystems, Carlsbad, CA, USA). Before transcription, RQ1 RNase-Free DNase (Promega, Madison, USA) was used to remove residual genomic DNA of total RNA. An equal amount of cDNA (150 ng/u l) was synthesized using 1st strand cDNA synthesis kits (TaKaRa, Dalian, China) as the RT-qPCR templates. Each RT-qPCR reaction was conducted in a 25 μ l reaction: 12.5 μ l of 2X SuperReal PreMix Plus (TianGen, Beijing, China), 0.75 μ l of each primer (10 μ M), 2 μ l of sample cDNA, and 9 μ l of sterilized ddH2O. The RT-qPCR was run as follows: 94°C for 2 min, followed by 40 cycles of 95°C for 15 s, 60°C for 30 s, 60°C for 1 min, heated to 95°C for 30 s and cooled to 60°C for 15 s to measure the melting curve.

RT-qPCR data analyses were performed using the 2-ΔΔCT method [42]. Data of relative expression levels in various tissue were subjected to one-way analysis of variance (ANOVA), followed by a least significant difference test (Tukey) for mean comparison. The data were analyzed directly by SPSS 9.20 software (SPSS Inc., Chicago, IL, USA). Differences were considered significant at p < 0.05. The RT-qPCR data were analyzed and exported as TIF files by Graphpad Prism 5.0 (GraphPad Software, La Jolla, CA, USA).

Results

Transcriptome assembly of L. sticticalis

Using the Illumina HiSeq 2500 platform, we performed next-generation sequencing on a cDNA library constructed from L. sticticalis. A total of 869.3 million clean reads (86.93 Gb) were obtained. Q30 bases were more than 85.01% in all the samples. After de novo assembly, we assembled 3,266,885 contigs with a mean length of 68.57 nt and an N50 length of 63 nt, 148,291 transcripts with a mean length of 971.37 nt and an N50 length of 1828 nt and identified 80,761 unigenes with a mean length of 722.82 nt and an N50 length of 1495 nt (Table 1). The size distribution analysis of the unigenes indicated that 14,484 unigenes were larger than 1000 nt in length, which represented 17.93% of all unigenes (S1 Fig). All of the clean data used in this study were uploaded to SRA with the accession number SRS1782539 to SRS1782550 (male antennae: SRS1782539, SRS1782546 and SRS1782548; female antennae: SRS1782540, SRS1782545 and SRS1782550; legs: SRS1782541, SRS1782544 and SRS1782547; larvae: SRS1782542, SRS1782543 and SRS1782549). Most assembled unigene sequences were uploaded to GeneBank with the accession number GFCJ01000001 to GFCJ01079039. The accession numbers of 131 candidate chemosensory genes identified in this study were listed in supporting information (S4 Table).

thumbnail
Table 1. Transcriptome assembly summary of L. sticticalis.

https://doi.org/10.1371/journal.pone.0174036.t001

Nr homology analysis and Gene Ontology (GO) annotation

Of the 80,761 unigenes, the results of annotation by NCBI BLASTx showed that 30,581 (37.87%) unigenes matched to known proteins. The remaining unigenes failed to match any sequence, with an e-value < 1e-5, in neither the Nr nor the Swiss-Prot databases. Among the Nr homology annotated unigenes, 49.62% of the homologous species had best blast match to Lepidopteran sequences. The highest match percentage (28.12%) was to Bombyx mori sequences followed by Danaus plexippus (20.09%) and Papilio xuthus (1.41%) (S2 Fig). Of the Nr annotated unigenes, 62.01% of the unigenes showed strong homology, with an e-value < 1e-45.

Gene ontology (GO) annotation of the unigenes was acquired using the Blast2GO pipeline according to the BLASTx search against Nr, which was used to classify transcripts into functional groups according to the GO category. Of the 80,761 unigenes, 16,899 (20.92%) unigenes were assigned to the various GO terms. Among the 16,899 GO annotated unigenes, the unigenes were allocated to the biological process terms more than the molecular function terms or the cellular component terms. In the molecular function category, the genes expressed in the antennae were mostly enriched for molecular binding activity (e.g., nucleotide, ion and odorant binding) and catalytic activity (e.g., hydrolase and oxidoreductase). In the biological process category, cellular, metabolic and single-organism processes were the most represented. In the cellular component category, cell, cell part and organelle were the most abundant groups (Fig 1). These results are comparable to the reported Chilo suppressalis transcriptional profile [21].

thumbnail
Fig 1. Gene ontology classified annotation of the L. sticticalis unigenes.

https://doi.org/10.1371/journal.pone.0174036.g001

Identification and expression of candidate ORs of L. sticticalis

In this study, we identified 54 candidate ORs in L. sticticalis by bioinformatics analysis. Of these, 38 unigenes had full-length ORFs that encoded 325 to 474 amino acids, and 16 unigenes were partial sequences by the NCBI BLASTp analysis. The 54 OR sequences had a BLASTx best hit to Lepidopteran sequences, with an e-value < 1e-5 (Table 2). Using the TMHMM Server v. 2.0, we also detected 54 candidate OR sequences with 0–8 transmembrane domains (TMDs).

The unigene C57376.g0 was named LstiOrco due to the high level of identity with the conserved Orco proteins of other insect species in Lepidoptera, which was clustered into the Orco clades of Lepidoptera in the phylogenetic tree (Fig 2). Among the 54 candidate LstiORs, LstiOrco showed the highest expression levels in the antennae in both RNA-Seq and RT-qPCR analysis (Fig 3).

thumbnail
Fig 2. Phylogenetic tree of candidate LstiORs with known lepidopteran ORs.

Csup: C. suppressalis, Bmor: B. mori, Harm: H. armigera, Hass: H. assulta. The clade in blue indicates the PR gene clade; the clade in pink indicates the Orco clade.

https://doi.org/10.1371/journal.pone.0174036.g002

thumbnail
Fig 3. Expression pattern of L. sticticalis ORs by RT-qPCR.

Legs (male: female = 1:1). β-actin was used as an internal reference gene to test the integrity of each cDNA template. The standard error is represented by the error bar, and the different letters (a, b, c) above each bar represent significant differences (p < 0.05).

https://doi.org/10.1371/journal.pone.0174036.g003

Five unigenes, named “LstiPRm” (m = 1 to 5), were considered to be pheromone receptors (PRs) because they shared considerable similarity with previously characterized Lepidopteran PRs and were clustered together into one subgroup in the phylogenetic tree (Fig 2). For the relatively conserved PR genes, LstiPR1 and LstiPR2 were clustered together with PR 1, 2, 3 and 4 in C. suppressalis. LstiPR3, 4 and 5 were not closely grouped with the Pyralidae PRs but clustered with the B. mori, H. armigera and H. assulta PR clade with high bootstrap support (Fig 2). The five LstiPRs showed higher expression in the antennae of both sexes than in the legs and larvae (p < 0.05) (Fig 3).

The remaining 48 LstiOR unigenes were highly divergent, which is common for insect olfactory receptor genes. These unigenes were named “LstiORn” (n = 1 to 48), followed by a numeral, in descending order in accordance with their female antennal expression levels. The RT-qPCR results showed that 47 candidate LstiORs had antennae-enriched expression, and 33 candidate LstiORs (OR1-23, OR25, OR27, OR29, OR30, OR32, OR34, OR41, OR43, OR45 and OR47) had female antennae-biased expression, especially for LstiOR7 being female specific. But, the putative LstiOR40 was richly expressed in the antennae and larvae (Fig 3).

Identification and expression of candidate IRs and GRs of L. sticticalis

Based on bioinformatic analysis, we identified 18 candidate IR sequences in L. sticticalis. Ten sequences contained full-length open reading frames (ORFs), and the remaining 8 sequences were marked as incomplete because they lacked a complete 5' or 3' terminus. Seventeen putative IRs in L. sticticalis were predicted to have 1–4 TMDs by TMHMM Server v. 2.0 (Table 3).

A phylogenetic tree of the LstiIRs was constructed based on the amino acid sequences from L. sticticalis, Drosophila melanogaster, B. mori and S. littoralis (Fig 4). The neighbor-joining tree analysis showed a clear segregation between Dmel ionotropic glutamate receptors (iGluRs) and insect IRs, and 18 LstiIR candidates were clustered to antennal IRs and the IR25a/IR8a clades, but did not belong to DmeliGluRs. According to their BLASTx best hits to Lepidopteran IRs and their positions in the phylogenetic tree, the 18 candidate IRs were given names consistent with the number and suffix of the Dmel/Bmor/Slit IR orthologs in the same clade (Table 3).

thumbnail
Fig 4. Phylogenetic tree of candidate LstiIRs with known lepidopteran IRs and iGluRs.

Dmel: D. melanogaster, Bmor: B. mori, Slit: S. littoralis. The clade in blue indicates the iGluR gene clade; the clade in pink indicates the IR8a and IR25a clade.

https://doi.org/10.1371/journal.pone.0174036.g004

Of the 18 named LstiIR candidates, the RT-qPCR results showed 10 putative LstiIRs (7d.2, 21a, 40a, 41a, 64a, 75p, 75p.1, 75q.2, 87a, and 93a) showed antennae specific expression, and expression levels of 8a, 25a, 75d and 76b were higher in the antennae than in the legs and larvae (p < 0.05). But the LstiIR1 showed larvae specific expression, LstiIR7d.3 and 68a in the larvae and LstiIR7g in the legs had higher expression than in the antennae (Fig 5).

thumbnail
Fig 5. Expression pattern of L. sticticalis IRs by RT-qPCR.

The details were same as mentioned in Fig 3.

https://doi.org/10.1371/journal.pone.0174036.g005

In total, we identified 13 GR candidates in L. sticticalis, including 3 unigenes with full-length ORFs and 10 unigenes with partial sequences. Thirteen putative GRs were predicted to have 1–7 transmembrane domains (Table 3). Of the 13 putative LstiGRs, 11 sequences were named based on their clustering into the clades of Dmel/Bmor/Hass/Harm GRs in the phylogenetic tree (Fig 6). Two unigenes (C52834.g1 and C3705.g0) had low bootstrap values and were unable to be placed on the phylogenetic with confidence and were named LstiGR6 and LstiGR7, respectively. The RT-qPCR results showed that 13 candidate LstiGRs were enriched in the antennae and the expression amounts of LstiGR63a.1 in the male antennae was the highest. Interestingly, the putative LstiGR6 was sex-specific expressed in the female antennae, but also expressed in the larvae (Fig 7).

thumbnail
Fig 6. Phylogenetic tree of candidate LstiGRs with known lepidopteran GRs.

Dmel: D. melanogaster, Bmor: B. mori, Harm: H. armigera and Hass: H. assulta.

https://doi.org/10.1371/journal.pone.0174036.g006

thumbnail
Fig 7. Expression pattern of L. sticticalis GRs by RT-qPCR.

The details were same as mentioned in Fig 3.

https://doi.org/10.1371/journal.pone.0174036.g007

Identification and expression of putative OBPs of L. sticticalis

In the process of identification of putative OBPs, we used not only keyword searching by PSI-BLAST, but also motif scanning to detect the conserved six cysteine residue pattern, which is C1-X5-39-C2-X3-C3-X21-44-C4-X7-12-C5-X8-C6 [19], in the sequence of OBPs. In all, we identified 34 candidate OBPs in L. sticticalis, including 3 PBPs and 1 GOBP. The results of the sequence analysis showed 23 unigenes with full–length ORFs and the remaining 11 unigenes corresponding partial sequences. Among the 34 putative LstiOBPs, 22 unigenes were predicted to have signal peptides by SignalP 4.1 Server analysis. These 34 OBP sequences had a BLASTx best hits to Lepidopteran sequences with an e-value < 1e-5 (Table 4).

Four unigenes (C59843.g0 C52747.g0, C52060.g0 and C58964.g0) were clustered into the PBP and GOBP clades of Lepidoptera in the phylogenetic tree (Fig 8) and were named LstiPBP1, LstiPBP2, LstiPBP3 and LstiGOBP1, respectively. The remaining 30 sequences were named LstiOBP1-30 on the basis of the similarity to known Lepidopteran OBPs and female antennal expression levels. OBPs usually were classified into three phylogenetic families. Classic OBPs, which include the PBP-GOBP group, are characterized by the conserved 6 cysteine residue pattern. The Minus-C class has lost cysteine residues, which are generally C2 and C5, and lysine can replace the position of the lost C2 [15]. In contrast, the Plus-C class has 1–2 extra cysteines and one characteristic proline next to the end of the sixth conserved cysteine residue [5]. The results of our sequence analysis showed that 23 complete ORF OBPs of L. sticticalis could be divided into three groups: 17 Classic OBPs (LstiPBP1, PBP3, GOBP1, OBP1, OBP3, OBP4, OBP6, OBP9, OBP12, OBP14, OBP15, OBP16, OBP18, OBP19, OBP21, OBP26 and OBP29), 4 Minus-C OBPs (LstiOBP7, OBP13, OBP17 and OBP28) and 2 Plus-C OBPs (LstiOBP11 and OBP22) (Table 4).

thumbnail
Fig 8. Phylogenetic tree of candidate LstiOBPs with known lepidopteran OBPs.

Csup: C. suppressalis, Bmor: B. mori, Harm: H. armigera, Hass: H. assulta, Cpun: C. punctiferalis. The clade in blue indicates the GOBP gene clade; the clade in pink indicates the PBP clade.

https://doi.org/10.1371/journal.pone.0174036.g008

The RT-qPCR results showed that among the 34 candidate LstiOBPs, 22 LstiOBPs were highly expressed in the antennae, 4 LstiOBPs (OBP15, OBP17, OBP25, and OBP29) were highly enriched in the legs, and 5 LstiOBPs (OBP11, OBP20, OBP21, OBP22, and OBP28) were mainly expressed in the larvae. The expression levels of 3 LstiOBPs (OBP13, OBP19, and OBP26) were not significantly different between the antennae and legs (Fig 9).

thumbnail
Fig 9. Expression pattern of L. sticticalis OBPs by RT-qPCR.

The details were same as mentioned in Fig 3.

https://doi.org/10.1371/journal.pone.0174036.g009

Identification and expression of candidate CSPs and SNMPs of L. sticticalis

CSPs have a conserved cysteine pattern of C1-X6-C2-X18-C3-X2-C4 [11]. Through bioinformatics analysis, we identified 10 candidate CSPs in L. sticticalis. Eight sequences had full-length ORFs, but other unigenes were partial sequences. In addition, the unigenes C50444.g0 and C54133.g0 failed in the SignalP tests (Table 5). The 10 candidate CSPs of L. sticticalis best matched to Lepidopteran sequences, with an e-value < 1e-5 and an identity of more than 55% (Table 5). We named the 10 CSP candidates according to their expression levels in the L. sticticalis female antenna. The 10 CSP sequences in L. sticticalis were clustered with Lepidopteran orthologous genes from L. sticticalis, C. suppressalis, C. punctiferalis, B. mori and H. armigera in the phylogenetic tree (Fig 10). The RT-qPCR results showed that candidate LstiCSP2, LstiCSP7 and LstiCSP10 presented higher expression in the antennae, LstiCSP5 had enriched expression in the legs, and the putative LstiCSP9 was highly expressed in the larvae. In addition, the other 5 LstiCSP candidates (CSP1, CSP3, CSP4, CSP6, and CSP8) were mainly expressed in the antennae and legs (Fig 11).

thumbnail
Fig 10. Phylogenetic tree of candidate LstiCSPs with known lepidopteran CSPs.

Csup: C. suppressalis, Cpun: C. punctiferalis, Bmor: B. mori, Harm: H. armigera.

https://doi.org/10.1371/journal.pone.0174036.g010

thumbnail
Fig 11. Expression pattern of L. sticticalis CSPs by RT-qPCR.

The details were same as mentioned in Fig 3.

https://doi.org/10.1371/journal.pone.0174036.g011

In L. sticticalis, we obtained two SNMPs that were 3'lost and 5'lost sequences, respectively. The two SNMPs separately had a BLASTx best hits to Ostrinia nubilalis SNMP1 (similarity 88%) and SNMP2 (similarity 85%) sequences with an e-value < 1e-05 by NCBI BLASTp (Table 6). LstiSNMP1 and LstiSNMP2 had significantly higher expression in the antennae than in the legs and larvae validated by RT-qPCR analysis (P < 0.05) (Table 5). According to the phylogenetic analysis, LstiSNMP1 and LstiSNMP2 clustered with the known Lepidopteran SNMP groups (Fig 12).

thumbnail
Fig 12. Phylogenetic tree of candidate LstiSNMPs with known lepidopteran SNMPs.

Onub: O. nubilalis, Csup: C. suppressalis, Cmed: C. medinalis, Hvir: Heliothis viresscens, Sexi: S. exigua, Slit: S. litura, Msex: Manduca sexta.

https://doi.org/10.1371/journal.pone.0174036.g012

The protein sequences of the candidate chemosensory genes were listed in supporting information (S5 Table).

Analysis and comparison of RNA-Seq data and RT-qPCR data

We obtained 131 candidate chemosensory genes (54 ORs, 18 IRs, 13 GRs, 34 OBPs, 10 CSPs and 2 SNMPs) in L. sticticalis by Illumina sequencing. The results of RNA-Seq showed that most genes in the antennae had higher FPKM (Fragments per Kb per million reads) than in the legs and larvae (p < 0.05), especially 76 genes with specific expression in the antennae (Fig 13A). Furthermore, the OR7 showed female antennae-specific expression (Fig 13A and 13B). All results analyzed were based on FPKM.

thumbnail
Fig 13. Comparative results of olfactory genes FPKM in the male antennae, female antennae, legs and third instar larvae of L. sticticalis (Venn diagram).

A. comparison among the antennae, legs and larvae. B. comparison between the male and female antennae. Genes in the overlapping intersect show no significant difference among different tissues. Genes outside the intersect show significant difference. Those in the dash-outlined area show specific expression in the tissues.

https://doi.org/10.1371/journal.pone.0174036.g013

To test the result of Illumina sequencing, we investigated the expression patterns of 131 L. sticticalis chemosensory genes with RT-qPCR analyses. The RT-qPCR results showed that the expression levels of these candidate chemosensory genes in different tissues were mostly consistent with the results of RNA-Seq. Most notably, a majority of olfactory genes were predominantly expressed in the antennae. However, the expression levels of several chemosensory genes between the results of RT-qPCR and RNA-Seq have obvious differences. For example, the results of RT-qPCR showed LstiOR28, 29/IR64a, 75P.1/OBP16, 24 in the antennae, LstiOBP29 in the legs and LstiIR1 in the larvae had specific expression (Figs 3, 5 and 9), but these genes in Illumina sequencing analyses only showed higher expression (Fig 13A); on the contrary, the (IR8a, 76b/PBP1-3, GOBP1, OBP1-3, 8, 16/CSP2) only showed higher expression levels in the antennae by RT-qPCR (Figs 5, 9 and 11). These differences in the results need further research for confirmation.

Discussion

At present, the molecular basis of chemoreception in Lepidoptera is well understood compared to other insects, but the research on Pyralidae is relatively scarce. Therefore, we sequenced and analyzed the transcriptome of adult antennae, adult legs and larvae from L. sticticalis and obtained a dataset of 54 ORs, 18 IRs, 13 GRs, 34 OBPs, 10 CSPs and 2 SNMPs. In this study, comparing to the antennal transcriptome in Lepidoptera from C. suppressalis (47 ORs, 20 IRs, 26 OBPs, 21 CSPs and 2 NMPs) [21], C. punctiferalis (62 ORs, 11 IRs, 10 GRs, 15 OBPs, 8 CSP and 2 SNMPs) [43, 44], O. furnacalis (56 ORs, 21 IRs, 5 GRs, 24 OBP, 19 CSP and 2 SNMPs) [45, 46], C. medinalis (29 ORs, 15 IRs, 30 OBPs, 26 CSPs and 2 SNMPs) [9], H. armigera (60 ORs, 19 IRs, 9 GRs, 34 OBPs, 18 CSPs and 2 SNMPs) [33, 47, 48], B. mori (62 ORs, 17 IRs, 69 GRs, 44 OBPs, 18 CSP and 2 SNMPs) [31, 4951] and H. assulta (64 ORs, 19 IRs, 18 GRs, 29 OBPs, 17 CSP and 2 SNMPs) [33, 52], our LstiOR dataset of sequences has no notable difference in the identified gene numbers.

RNA-Seq and RT-qPCR results both showed 54 putative LstiORs were mainly expressed in the antennae, which was similar to the other Lepidopteran results [9, 21, 31, 33, 43, 45]. Studies about B. mori showed that three female-biased ORs (OR19, OR45 and OR47) are capable to respond to host plant volatiles (linalool, benzoic acid, 2-phenylethanol and benzaldehyde) [49, 53]. The 6 female-biased expression LstiORs (OR4, OR23, OR29, OR30, OR32 and OR34) that were clustered with the female-biased ORs from B. mori in the Phylogenetic tree might have similar functions, but further studies were needed. In view of the host selectivity of larvae [3, 4], LstiOR5, OR34 and OR40 that were richly expressed in larvae might play important roles in host-plant selection. Some reports showed that PRs specific expressed in male antennae detected the sex pheromone components of female moths [54, 55, 57, 58]. However, in our study, 5 candidate PRs of L. sticticalis were expressed in the antennae of both sexes, which is consistent with the recent reported results of 6 putative PRs identified in C. suppressalis, 2 PRs (OR6 and OR13) in H. armigera and 2 PRs in S. littoralis [21, 56, 57]. Therefore, the recognition mechanism of LstiPRs to the sex pheromone [59] of the female moth requires further research.

As the complement of ORs, ionotropic receptors were first discovered in D. melanogaster [28] through genomic analyses. Compared to ORs, the IR family is relatively conserved both in sequence and expression pattern. In our study, among the 18 LstiIRs we discovered, 13 sequences have orthologs found in Dmel/Bmor/Slit IRs; the expression levels were not significantly different between male and female antennae, which were similar to the IR expression in S. littoralis [54], C. suppressalis [21] and H. armigera [33]. Lsti76b, as well as LstiIR8a and LstiIR25a, was highly expressed in the antennae, and these genes might also be special subunits of individual odor-specific receptors [60]. The functions of IRs in L. sticticalis are likely to be conserved as IRs in other Lepidoptera, both in terms of the relatively high sequence conservation and the comparability of expression levels.

Gustatory receptors play a critical role in the detection of chemicals, which ultimately influence the insects’ decisions when looking for food, mates and egg deposition sites [32, 62]. Interestingly, our LstiGR4 shared 72% homology with HarmGR4 which were identified as a sugar receptor [47, 61], so LstiGR4 might be a sugar receptor and participate in sugar detection and consumption. GR21a/GR63a that were expressed in CO2-sensing neurons could allow the detection of CO2 concentration in D. melanogaster [6264]. In our study, 5 LstiGRs (GR21a, GR21b, GR63a, GR63a.1, and GR63a.2) were clustered into the clades of DmelGR21a/BmorGR63a in the phylogenetic tree and might be CO2 receptors. However, annotation of these GRs awaits further demonstration.

Of our 34 LstiOBPs, most LstiOBPs were richly expressed in the antennae of both sexes that was similar to other transcriptome analyses in Lepidoptera [9, 21, 33, 43, 44]. As specific OBPs, PBPs usually were considered to have a connection with male moth perception of the sex pheromone components released by female moths [6669]. Our 3 LstiPBPs were closely clustered into the PBP clade of other Lepidoptera in the phylogenetic tree, which suggests that our LstiPBPs might have similar function. Currently, studies also show that OBPs specifically expressed in larvae displayed a high recognition capacity to the major sex pheromone component [65]. Thus, one of the LstiOBPs (OBP11, OBP20, OBP21, OBP22, and OBP28) which specifically expressed in the larvae might play a key role in the perception of female sex pheromone in L. sticticalis.

CSPs are more highly conserved than OBPs across insect species and are widely expressed in different parts of the insect body [31, 70]. Our 10 LstiCSPs were primarily expressed in the legs and antennae of the adults, which was similar to the results of other Lepidoptera [9, 21, 31, 33, 43, 45]. But LstiCSP9 was mainly expressed in larvae. The antennal enriched CSPs might be involved in chemoreception [71], and the CSPs expressed in the legs might participate in other physiological processes beyond chemoreception [72]. However, the function of our putative LstiCSPs requires further research.

Because SNMPs were first identified in Lepidopteran pheromone-sensitive neurons [17, 73], these proteins are believed to be involved in the recognition of insect pheromones. In this study, the expression levels of SNMPs in L. sticticalis were consistent with the reported results that SNMP1 of H. assulta was primarily expressed in the antennae, and SNMP2 of H. assulta was abundantly expressed in the antennae and legs [33]. Previous studies showed that SNMP1 was crucial for the detection of the volatile pheromone 11-cis-vaccenyl acetate in D. melanogaster [18]. SNMP2, in contact with pheromone-sensitive sensilla, was expressed in sensilla support cells [74]. According to the similar expression levels and physiological analysis to other Lepidoptera, we can infer that SNMPs in L. sticticalis might have the same role as in D. melanogaster. However, the general mechanism of SNMPs’ function in insects remains inadequately understood. Therefore, future studies on the function of SNMP1 and SNMP2 in L. sticticalis are necessary.

Conclusion

Our aim of this study was to identify genes potentially involved in olfactory signal detection in L. sticticalis, and this aim was well met by the identification of a repertoire of 54 ORs, 18 IRs, 13 GRs, 34 OBPs, 10 CSPs and 2 SNMPs. Our results not only establish a means to further elucidate the molecular mechanisms of chemosensation, but also provide potential targets for disrupting the chemical communication system in L. sticticalis as a means of pest control.

Supporting information

S1 Fig. Unigene length distribution of L. sticticalis.

https://doi.org/10.1371/journal.pone.0174036.s001

(TIF)

S2 Fig. Distribution of Nr homologous species annotation on L. sticticalis unigenes.

https://doi.org/10.1371/journal.pone.0174036.s002

(TIF)

S1 Table. Nucleotide sequences of all identified candidate olfactory genes.

https://doi.org/10.1371/journal.pone.0174036.s003

(DOCX)

S2 Table. The sequences used for phylogenetic trees of chemosensory genes in L. sticticalis.

https://doi.org/10.1371/journal.pone.0174036.s004

(DOC)

S4 Table. The accession numbers of 131 candidate chemosensory genes in L. sticticalis.

https://doi.org/10.1371/journal.pone.0174036.s006

(DOCX)

S5 Table. The protein sequences of the chemosensory genes (ORs, IRs, OBPs, CSPs, SNMPs) in L. sticticalis.

https://doi.org/10.1371/journal.pone.0174036.s007

(DOC)

Author Contributions

  1. Conceptualization: JY.
  2. Data curation: JY HSW.
  3. Formal analysis: HSW JY.
  4. Investigation: HSW JY.
  5. Methodology: JY HSW.
  6. Resources: YZC KBL.
  7. Software: HSW SZ.
  8. Supervision: JY KBL.
  9. Validation: HSW JY.
  10. Visualization: HSW JY.
  11. Writing – original draft: HSW.
  12. Writing – review & editing: JY HSW SZ KBL YZC.

References

  1. 1. Qu XF, Shao ZR, Wang JQ. Analysis of periodic outbreak of meadow moth in agricultural and pastoral area of North China. Entomol Knowl. 1999; 36(1): 11–14.
  2. 2. Yin J, Feng HL, Sun HY, Xi JH, Cao YZ, Li KB. Functional analysis of general odorant binding protein 2 from the meadow moth, Loxostege sticticalis L. (Lepidoptera: Pyralidae). PLoS One. 2012; 7(3): e33589. pmid:22479417
  3. 3. Yin J, Cao YZ, Luo LZ, Hu Y. Effects of host plants on population increase of meadow moth, Loxostege sticticalis L. J Plant Prot Res. 2004; 31(2): 173–178.
  4. 4. Yin J, Cao YZ, Luo LZ, Hu Y. Oviposition preference of the meadow moth, Loxostege sticticalis L., on different host plants and its chemical mechanism. Acta Ecol Sin. 2005; 25(8): 1844–1852.
  5. 5. Zhang LX, Fan JS, Wang GQ. Research Advances on Loxostege sticticalis L. (Lepidoptera: Pyralidae) in China. Chin Agri Sci Bull. 2010; 26: 215–218.
  6. 6. Yin J, Cao YZ, Luo LZ, Hu Y. Ultra-structure of the antennal sensilla of the meadow moth, Loxostege sticticalis. Entomol Knowl. 2004; 41(1): 56–59.
  7. 7. Kanaujia S, Kaissling KE. Interactions of pheromone with moth antennae: Adsorption, desorption and transport. J Insect Physiol. 1985; 31(1): 71–81.
  8. 8. Sato K, Touhara K. Insect olfaction: receptors, signal transduction, and behavior. Results Probl Cell Differ. 2009; 47: 121–138. pmid:19083129
  9. 9. Zeng FF, Zhao ZF, Yan MJ, Zhou W, Zhang Z, Zhang A, et al. Identification and comparative expression profiles of chemoreception genes revealed from major chemoreception organs of the rice leaf folder, Cnaphalocrocis medinalis (Lepidoptera: Pyralidae). PLoS One. 2015; 10(12): e0144267. pmid:26657286
  10. 10. Vogt RG, Riddiford LM. Pheromone binding and inactivation by moth antennae. Nature. 1981; 293(5828): 161–163. pmid:18074618
  11. 11. Xu YL, He P, Zhang L, Fang SQ, Dong SL, Zhang YJ, et al. Large-scale identification of odorant-binding proteins and chemosensory proteins from expressed sequence tags in insects. BMC genomics. 2009; 10(1): 632.
  12. 12. Qiao HL, Tuccori E, He XL, Gazzano A, Field L, Zhou JJ, et al. Discrimination of alarm pheromone (E)-β-farnesene by aphid odorant-binding proteins. Insect Biochem Mol Biol. 2009; 39(5–6): 414–419. pmid:19328854
  13. 13. Zhou JJ, Robertson G, He XL, Dufour S, Hooper AM, Pickett JA, et al. Characterisation of Bombyx mori odorant-binding proteins reveals that a general odorant-binding protein discriminates between sex pheromone components. J Mol Biol. 2009; 389(3): 529–545. pmid:19371749
  14. 14. Robertson HM, Martos R, Sears CR, Todres EZ, Walden KKO, Nardi JB. Diversity of odourant binding proteins revealed by an expressed sequence tag project on male Manduca sexta moth antennae. Insect Mol Biol. 1999; 8(4): 501–518. pmid:10620045
  15. 15. Blomquist GJ, Vogt RG. Insect pheromone biochemistry and molecular biology: The biosynthesis and detection of pheromones and plant volatiles. London: Elsevier Academic Press; 2003.
  16. 16. Pelosi P, Zhou JJ, Ban L, Calvello M. Soluble proteins in insect chemical communication. Cell Mol Life Sci. 2006; 63(14): 1658–1676. pmid:16786224
  17. 17. Benton R, Vannice KS, Vosshall LB. An essential role for a CD36-related receptor in pheromone detection in Drosophila. Nature. 2007; 450(7167): 289–293. pmid:17943085
  18. 18. Jin X, Ha TS, Smith DP. SNMP is a signaling component required for pheromone sensitivity in Drosophila. Proc Natl Acad Sci U S A. 2008; 105(31): 10996–11001. pmid:18653762
  19. 19. Zhou JJ, He XL, Pickett JA, Field LM. Identification of odorant-binding proteins of the yellow fever mosquito Aedes aegypti: genome annotation and comparative analyses. Insect Mol Biol. 2008; 17(2): 147–163. pmid:18353104
  20. 20. Leal WS. Odorant reception in insects: roles of receptors, binding proteins, and degrading enzymes. Annu Rev Entomol. 2013; 58(1): 373–391.
  21. 21. Cao DP, Liu Y, Wei JJ, Liao XY, Walker WB, Li JH, et al. Identification of candidate olfactory genes in Chilo suppressalis by antennal transcriptome analysis. Int J Biol Sci. 2014; 10(8): 846–860. pmid:25076861
  22. 22. Benton R, Sachse S, Michnick SW, Vosshall LB. Atypical membrane topology and heteromeric function of Drosophila odorant receptors in vivo. PLoS Biol. 2006; 4(2): e20. pmid:16402857
  23. 23. Sato K, Pellegrino M, Nakagawa T, Nakagawa T, Vosshall LB, Touhara K. Insect olfactory receptors are heteromeric ligand-gated ion channels. Nature. 2008; 452(7190): 1002–1006. pmid:18408712
  24. 24. Wicher D, Schäfer R, Bauernfeind R, Stensmyr MC, Heller R, Heinemann SH, et al. Drosophila odorant receptors are both ligand-gated and cyclic-nucleotide-activated cation channels. Nature. 2008; 452(7190): 1007–1011. pmid:18408711
  25. 25. Larsson MC, Domingos AI, Jones WD, Chiappe ME, Amrein H, Vosshall LB. Or83b encodes a broadly expressed odorant receptor essential for Drosophila olfaction. Neuron. 2004; 43(5): 703–714. pmid:15339651
  26. 26. Silbering AF, Rytz R, Grosjean Y, Abuin L, Ramdya P, Jefferis GS, et al. Complementary function and integrated wiring of the evolutionarily distinct Drosophila olfactory subsystems. J Neurosci. 2011; 31(38): 13357–13375. pmid:21940430
  27. 27. Rytz R, Croset V, Benton R. Ionotropic receptors (IRs): chemosensory ionotropic glutamate receptors in Drosophila and beyond. Insect Biochem Mol Biol. 2013; 43(9): 888–897. pmid:23459169
  28. 28. Benton R, Vannice KS, Gomez-Diaz C, Vosshall LB. Variant ionotropic glutamate receptors as chemosensory receptors in Drosophila. Cell. 2009; 136(1): 149–162. pmid:19135896
  29. 29. Montell C. A taste of the Drosophila gustatory receptors. Curr Opin Neurobiol. 2009; 19(4): 345–353. pmid:19660932
  30. 30. Seeta P, Yunjung K, Yun TK, Youngseok L. Gustatory receptors required for sensing umbelliferone in Drosophila melanogaster. Insect Biochem Mol Biol. 2015; 66: 110–118. pmid:26524963
  31. 31. Gong DP, Zhang HJ, Zhao P, Lin Y, Xia QY, Xiang ZH. Identification and expression pattern of the chemosensory protein gene family in the silkworm, Bombyx mori. Insect Biochem Mol Biol. 2007; 37(3): 266–277. pmid:17296501
  32. 32. Gu SH, Sun L, Yang RN, Wu KM, Guo YY, Li XC, et al. Molecular Characterization and Differential Expression of Olfactory Genes in the Antennae of the Black Cutworm Moth Agrotis ipsilon. PLoS One. 2014; 9(8): e103420. pmid:25083706
  33. 33. Zhang J, Wang B, Dong S, Cao D, Dong J, Walker WB, et al. Antennal Transcriptome Analysis and Comparison of Chemosensory Gene Families in Two Closely Related Noctuidae Moths, Helicoverpa armigera and H. assulta. PLoS One. 2015; 10(2): e0117054. pmid:25659090
  34. 34. Ma L, Li ZQ, Bian L, Cai XM, Luo ZX, Zhang YJ, et al. Identification and comparative Study of Chemosensory Genes Related to Host Selection by Legs Transcriptome Analysis in the Tea Geometrid Ectropis obliqua. PLoS One. 2016; 11 (3): e0149591. pmid:26930056
  35. 35. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. Full length transcriptome assembly from RNA Seq data without a reference genome. Nat Biotechnol. 2011; 29(7): 644–652. pmid:21572440
  36. 36. Pertea G, Huang X, Liang F, Antonescu V, Sultana R, Karamycheva S, et al. TIGR Gene Indices clustering tools (TGICL): a software system for fast clustering of large EST datasets. Bioinformatics. 2003; 19(5): 651–652. pmid:12651724
  37. 37. Conesa A, Gotz S, GarciaGomez JM, Terol J, Talon M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics. 2005; 21(18): 3674–3676. pmid:16081474
  38. 38. Min XJ, Butler G, Storms R, Tsang A. OrfPredictor: predicting protein-coding regions in EST-derived sequences. Nucleic Acids Res. 2005; 33(Web Server issue): W677–680. pmid:15980561
  39. 39. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam H, et al. Clustal W and Clustal X version 2.0. Bioinformatics. 2007; 23(21): 2947–2948. pmid:17846036
  40. 40. Saitou N, Nei M. The neighbor-joining method: A new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987; 4(4): 406–425. pmid:3447015
  41. 41. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol Biol Evol. 2013; 30(12): 2725–2729. pmid:24132122
  42. 42. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2 –ΔΔCt method. Methods. 2001; 25(4): 402–408. http://dx.doi.org/10.1006/meth.2001.1262 pmid:11846609
  43. 43. Ge X, Zhang TT, Wang ZY, He KL, Bai SH. Identification of putative chemosensory receptor genes from yellow peach moth Conogethes punctiferalis (Guenée) antennae transcriptome. Sci Rep. 2016; 6: 32636. pmid:27659493
  44. 44. Jia XJ, Wang HX, Yan ZG, Zhang MZ, Wei CH, Qin XC, et al. Antennal transcriptome and differential expression of olfactory genes in the yellow peach moth, Conogethes punctiferalis (Lepidoptera: Crambidae). Sci Rep. 2016; 6: 29067. pmid:27364081
  45. 45. Zhang TT, Coates BS, Ge X, Bai SX, He KL, Wang ZY. Male- and female-biased gene expression of olfactory-related genes in the antennae of asian corn borer, Ostrinia furnacalis (Guenée) (Lepidoptera: Crambidae). PLoS One. 2015; 10(6): e0128550. pmid:26062030
  46. 46. Yang B, Ozaki K, Ishikawa Y, Matsuo T. Identification of Candidate Odorant Receptors in Asian Corn Borer Ostrinia furnacalis. PLoS One. 2015; 10(3): e0121261. pmid:25803580
  47. 47. Jiang XJ, Ning C, Guo H, Jia YY, Huang LQ, Qu MJ, et al. A gustatory receptor tuned to D-fructose in antennal sensilla chaetica of Helicoverpa armigera. Insect Biochem Mol Biol. 2015; 60: 39–46. pmid:25784630
  48. 48. Ning C, Yang K, Xu M, Huang LQ, Wang CZ. Functional validation of the carbon dioxide receptor in labial palps of Helicoverpa armigera moths. Insect Biochem Mol Biol. 2016; 73: 12–19. pmid:27060445
  49. 49. Tanaka K, Uda Y, Ono Y, Nakagawa T, Suwa M, Yamaoka R, et al. Highly Selective Tuning of a Silkworm Olfactory Receptor to a Key Mulberry Leaf Volatile. Curr Biol. 2009; 19(11): 881–890. pmid:19427209
  50. 50. Gong DP, Zhang HJ, Zhao P, Xia QY, Xiang ZH. The odorant binding protein gene family from the genome of silkworm, Bombyx mori. BMC Genomics. 2009; 10: 332. pmid:19624863
  51. 51. Wanner KW, Robertson HM. The gustatory receptor family in the silkworm moth Bombyx mori is characterized by a large expansion of a single lineage of putative bitter receptors. Insect Mol Biol. 2008; 17 (6): 621–629. pmid:19133074
  52. 52. Xu W, Papanicolaou A, Liu NY, Dong SL, Anderson A. Chemosensory receptor genes in the Oriental tobacco budworm Helicoverpa assulta. Insect Mol Biol. 2015; 24(2): 253–263. pmid:25430896
  53. 53. Anderson AR, Wanner KW, Trowell SC, Warr CG, Jaquin-Joly E, Zagatti P, et al. Molecular basis of female-specific odorant responses in Bombyx mori. Insect Biochem Mol Biol. 2009; 39(3): 189–197. pmid:19100833
  54. 54. Sakurai T, Nakagawa T, Mitsuno H, Mori H, Endo Y, Tanoue S, et al. Identification and functional characterization of a sex pheromone receptor in the silkmoth Bombyx mori. Proc Natl Acad Sci U S A. 2004; 101(47): 16653–16658. pmid:15545611
  55. 55. Mitsuno H, Sakurai T, Murai M, Yasuda T, Kugimiya S, Ozawa R, et al. Identification of receptors of main sex-pheromone components of three Lepidopteran species. Eur J Neurosci. 2008; 28(5): 893–902. pmid:18691330
  56. 56. Liu Y, Gu S, Zhang Y, Guo Y, Wang G. Candidate olfaction genes identified within the Helicoverpa armigera Antennal Transcriptome. PloS One. 2012; 7(10): e48260. pmid:23110222
  57. 57. Legeai F, Malpel S, Montagné N, Monsempes C, Cousserans F, Merlin C, et al. An Expressed Sequence Tag collection from the male antennae of the Noctuid moth Spodoptera littoralis: a resource for olfactory and pheromone detection research. BMC Genomics. 2011; 12: 86. pmid:21276261
  58. 58. Patch HM, Velarde RA, Walden KK, Robertson HM. A candidate pheromone receptor and two odorant receptors of the hawkmoth Manduca sexta. Chem Senses. 2009; 34(4): 305–316. pmid:19188280
  59. 59. Struble DL, Lilly CE. An attractant for the beet webworm, Loxostege sticticalis (lepidoptera: pyralidae). Can Ent. 1997; 109(2): 261–266.
  60. 60. Abuin L, Bargeton B, Ulbrich MH, Isacoff EY, Kellenberger S, Benton R. Functional architecture of olfactory ionotropic glutamate receptors. Neuron. 2011; 69(1): 44–60. pmid:21220098
  61. 61. Xu W, Zhang HJ, Anderson A. A sugar gustatory receptor identified from the foregut of cotton bollworm Helicoverpa armigera. J Chem Ecol. 2012; 38(12): 1513–1520. pmid:23224441
  62. 62. Freeman EG, Wisotsky Z, Dahanukar A. Detection of sweet tastants by a conserved group of insect gustatory receptors. Proc Natl Acad Sci U S A. 2014; 111(4): 1598–1603. pmid:24474785
  63. 63. Jones WD, Cayirlioglu P, Kadow IG, Vosshall LB. Two chemosensory receptors together mediate carbon dioxide detection in Drosophila. Nature. 2007; 445(7123): 86–90. pmid:17167414
  64. 64. Kwon JY, Dahanukar A, Weiss LA, Carlson JR. The molecular basis of CO2 reception in Drosophila. Proc Natl Acad Sci U S A. 2007; 104(9): 3574–3578. pmid:17360684
  65. 65. Jin R, Liu NY, Liu Y, Dong SL. A larval specific OBP able to bind the major female sex pheromone component in Spodoptera exigua (Hübner). J Integr Agr. 2015; 14(7): 1356–1366.
  66. 66. Li L, Yang WL, Guo XR, Luo MH, Yuan GH, Qiao Q, et al. Cloning, sequence analysis and spatio- temporal expression of a pheromone binding protein 2 (PBP2) gene from Helicoverpa assulta (Guenee) (Lepidoptera: Noctuidae). Acta Entomol Sin. 2009; 52(11): 199–1205.
  67. 67. Jia XJ, Hao SD, Du YL, Zhang MZ, Qin XC, Wang JZ, et al. cDNA cloning, expression profiling and binding affinity assay of the pheromone binding protein Cpun-PBP1 in the yellow peach moth, Conogethes punctiferalis (Lepidoptera: Crambidae). Acta Entomol Sin. 2015; 58(11): 1167–1176.
  68. 68. Allen JE, Wanner KW. Asian corn borer pheromone binding protein 3, a candidate for evolving specificity to the 12- tetradecenyl acetate sex pheromone. Insect Biochem Mol Biol. 2011; 41(3): 141–149. pmid:21056664
  69. 69. Gu SH, Zhou JJ, Wang GR, Zhang YJ, Guo YY. Sex pheromone recognition and immunolocalization of three pheromone binding proteins in the black cutworm moth Agrotis ipsilon. Insect Biochem Mol Biol. 2013; 43(3): 237–251. pmid:23298680
  70. 70. Pelosi P, Iovinella I, Felicioli A, Dani FR. Soluble proteins of chemical communication: an overview across arthropods. Front Physiol. 2014; 5: 320. pmid:25221516
  71. 71. Zhang YN, Ye ZF, Yang K, Dong SL. Antenna-predominant and male-biased CSP19 of Sesamia inferens is able to bind the female sex pheromones and host plant volatiles. Gene. 2014; 536(2): 279–286. pmid:24361960
  72. 72. Kitabayashi AN, Arai T, Kubo T, Natori S. Molecular cloning of cDNA for p10, a novel protein that increases in the regenerating legs of Periplaneta americana (American cockroach). Insect Biochem Mol Biol. 1998; 28(10): 785–790. pmid:9807224
  73. 73. Rogers ME, Krieger J, Vogt RG. Antennal SNMPs (sensory neuron membrane proteins) of Lepidoptera define a unique family of invertebrate CD36-like proteins. J Neurobiol. 2001; 49(1): 47–61. pmid:11536197
  74. 74. Forstner M, Gohl T, Gondesen I, Raming K, Breer H, Krieger J. Differential expression of SNMP-1 and SNMP-2 proteins in pheromone-sensitive hairs of moths. Chem Senses. 2008; 33(3): 291–299. pmid:18209018