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

Antennal transcriptome analysis of the piercing moth Oraesia emarginata (Lepidoptera: Noctuidae)

  • Bo Feng,

    Affiliation Institute of Health and Environmental Ecology, Wenzhou Medical University, University Town, Wenzhou, Zhejiang, China

  • Qianshuang Guo,

    Affiliation Institute of Health and Environmental Ecology, Wenzhou Medical University, University Town, Wenzhou, Zhejiang, China

  • Kaidi Zheng,

    Affiliation Institute of Health and Environmental Ecology, Wenzhou Medical University, University Town, Wenzhou, Zhejiang, China

  • Yuanxia Qin,

    Affiliation Department of Research and Development, Newcon Inc., Ningbo, Zhejiang, China

  • Yongjun Du

    yongjundu@zju.edu.cn

    Current address: Institute of Pesticide and Environmental Toxicology, Zhejiang University, Hangzhou, Zhejiang, China

    Affiliation Institute of Health and Environmental Ecology, Wenzhou Medical University, University Town, Wenzhou, Zhejiang, China

Abstract

The piercing fruit moth Oraesia emarginata is an economically significant pest; however, our understanding of its olfactory mechanisms in infestation is limited. The present study conducted antennal transcriptome analysis of olfactory genes using real-time quantitative reverse transcription PCR analysis (RT-qPCR). We identified a total of 104 candidate chemosensory genes from several gene families, including 35 olfactory receptors (ORs), 41 odorant-binding proteins, 20 chemosensory proteins, 6 ionotropic receptors, and 2 sensory neuron membrane proteins. Seven candidate pheromone receptors (PRs) and 3 candidate pheromone-binding proteins (PBPs) for sex pheromone recognition were found. OemaOR29 and OemaPBP1 had the highest fragments per kb per million fragments (FPKM) values in all ORs and OBPs, respectively. Eighteen olfactory genes were upregulated in females, including 5 candidate PRs, and 20 olfactory genes were upregulated in males, including 2 candidate PRs (OemaOR29 and 4) and 2 PBPs (OemaPBP1 and 3). These genes may have roles in mediating sex-specific behaviors. Most candidate olfactory genes of sex pheromone recognition (except OemaOR29 and OemaPBP3) in O. emarginata were not clustered with those of studied noctuid species (type I pheromone). In addition, OemaOR29 was belonged to cluster PRIII, which comprise proteins that recognize type II pheromones instead of type I pheromones. The structure and function of olfactory genes that encode sex pheromones in O. emarginata might thus differ from those of other studied noctuids. The findings of the present study may help explain the molecular mechanism underlying olfaction and the evolution of olfactory genes encoding sex pheromones in O. emarginata.

Introduction

Olfaction plays a key role in foraging [13], mating [4,5], and oviposition behaviors [68] of insects. Insect olfaction studies have provided fundamental insights into chemosensory biology and chemical ecology and have provided valuable opportunities for pest management [914]. Lepidopterans are often used for olfaction studies, as these have extensive and sensitive olfactory repertoires. However, molecular studies on olfaction in Lepidopterans lag behind those of other insect models such as fruit fly and mosquitos [15].

Lepidoptera sex pheromones are divided into two main types based on their chemistry [16]. Type I pheromone components have 10- to 18-carbon, even numbered straight chain acetates, aldehydes, and alcohols. Type II pheromones consist of polyunsaturated C17-C23 straight chains, skipped conjugated polyenic hydrocarbons and the corresponding epoxide derivatives [17]. Type I pheromones occur in about 75% of all studied moth species, whereas type II pheromones occur in about 15% of identified Lepidopteran pheromones [17]. These two major types of sex pheromones are produced through distinct pathways that involve different biosynthetic sites, substrates, and enzymes, as well as respectively employ specific endocrine regulatory mechanisms. However, both types of pheromones have the same function in mate recognition and attraction in moths [16,18].

Genes encoding Lepidopteran olfactory proteins have been identified in Bombyx mori [19], and also in the pest species Manduca sexta [20], Heliothis virescens [21], Spodoptera litura [22], S. littoralis [23,24], Agrotis ipsilon [25], and Dendrolimus spp. [26]. Sex pheromones of above species are type I. However, studies on the olfactory genes that encode type II pheromones are limited.

The piercing fruit moth Oraesia emarginata Fabricius (Lepidoptera: Noctuidae) is an important pest of fruits such as citrus, pear, peach, and plum. The larvae feed on plants belonging to the Menispermaceae. Adult moths obtain nutrition from ripe fruits. Mated females lay eggs on Menispermaceae plants (Fig 1) [27]. The electroantennographic and behavioral responses of O. emarginata to volatiles from ripe fruits [28] and the repellency of a volatile compound, sec-butyl β-styryl ketone have been studied [29]. However, little is known about the olfactory mechanism of O. emarginata. Type II pheromones were identified as female sex pheromones in Oraesia species. The major and minor sex pheromone components of the related O. excavate were identified as cis-9,10-epoxy-(Z)-6 –heneicosene and cis-9,10-epoxy-(Z,Z)-3,6- heneicosadiene [30]. Although the sex pheromone of female O. emarginata was not published, it was similar to epoxide components from a preliminary identification (Du et al., unpublished data). In the present study, we achieved significant coverage of olfactory genes with de novo transcriptome and measured gene expression using real-time quantitative reverse transcription PCR analysis (RT-qPCR) for comparison between the sexes. We also discuss the diversification of olfactory genes for the recognition of type I and type II pheromones.

Materials and methods

Insects

O. emarginata larvae were collected from fields in Gannan City of Jiangxi Province, China and reared in the laboratory at 25 ± 1°C and 75 ± 5% relative humidity with a 14-h light/10-h dark photoperiod. Our field collection activities did not impact endangered or protected species. Larvae were fed fresh leaves of Cocculus orbiculatus until pupation. Emergence of males and females was checked every morning, and adults were separately maintained in ventilated wooden cages (35 cm × 35 cm × 50 cm). Emerging adult moths were fed with 10% glucose water soaked into cotton.

Extraction of total RNA from tissues

Antennae of 4-d-old adults were used. A total of 25 adults (males and females separately) were collected after 3.5 h of the dark cycle. Antennae samples from each group were immediately homogenized in TRNzol-A+ (TIANGEN Biotech, Beijing, China) on ice, and total RNA was extracted according to the manufacturer’s instructions. The concentration and purity of the total RNA were determined by using a NanoDrop2000 spectrophotometer (ThermoFisher, Waltham, MA, USA). RNA with an A260/A280 ratio between 1.75–2.05, an A260/A230 ratio > 1, and a concentration > 400 ng/μL was used for the experiments. Total RNA was treated with DNase I (Takara, Kusatsu, Shiga, Japan) to remove any genomic DNA. RNA extractions were performed in triplicate.

De novo transcriptome analysis

The same amount of RNA collected from male and female antennae was pooled for transcriptome analysis. The cDNA library for transcriptome analysis was prepared using a TruSeq SBS Kit v3-HS (Illumina, San Diego, CA, USA), following the manufacturer’s recommendations. The library was sequenced using Illumina HiSeq 2000 (Illumina, San Diego, CA, USA) with a 90-bp read length for the paired-end reads by BGI (Shenzhen, Guangdong, China). Dirty reads containing adapters and unknown or low-quality bases were discarded from the raw reads to obtain clean reads for analysis. De novo transcriptome assembly was conducted with the short reads assembly program, Trinity (r20140413p1, min_kmer_cov:2) [31]. BLASTx (v2.2.28+) alignment (E value < 0.00001) between unigenes and protein databases (NCBI non-redundant protein database, Swiss-Prot, Kyoto Encyclopedia of Genes and Genome (KEGG), and Clusters of Orthologous Groups (COG)) was successively performed. Gene ontology (GO) annotations of the unigenes were determined using Blast2go (http://www.blast2go.org/) [32].

Olfactory gene analysis

The candidate olfactory gene was manually obtained from gene annotation. In addition, a 50% ORF length cutoff was used in identifying putative genes to prevent a gene from being counted twice. The candidate OBPs and CSPs were searched for the presence of N-terminal signal peptides using SignalP4.0 (http://www.cbs.dtu.dk/services/SignalP/) using default parameters [33]. The signal peptides likely contained significant phylogenetic information and were included in the phylogenetic analyses of OBPs and CSPs [34]. Amino acid sequence alignment was performed using CLUSTALX2.1 using default parameters [35]. For phylogenetic analysis, known amino acid sequences of olfactory genes from other insects were downloaded (S1 File). Phylogenetic analyses were conducted using the maximum likelihood method of MEGA 6.0, which was based on the Jones-Taylor-Thornton (JTT) substitution model, partial deletion gaps with 95% site coverage cutoff, a nearest neighbor interchanges (NNI) heuristic search, and other default parameters [36]. Node support for the phylogenetic tree was assessed using the bootstrap method with 1,000 bootstrap replicates.

Profiling analysis of gene expression based on the antennal transcriptome

Gene expression levels were calculated using the fragments per kb per million fragments (FPKM) method based on the results of antennal transcriptome analysis. The number of fragments that uniquely aligned to a gene was divided by the total number of fragments that uniquely aligned to all genes and by the base number in the CDS of that gene [37]. The FPKM method can eliminate the influence of different gene lengths and sequencing levels on the calculation of gene expression.

RT-qPCR analysis of olfactory gene expression in the antennae

Single-stranded cDNAs were synthesized from 1 μg of total RNA using the ReverTra Ace qPCR RT Kit (Toyobo, Kita-ku, Osaka, Japan) following the manufacturer’s recommendations. RT-qPCR was performed with SsoFast EvaGreen® Supermix (Bio-Rad, Hercules, CA, USA), following the manufacturer’s protocols, in a CFX-96 PCR Detection System (Bio-Rad). The cycling conditions were an initial cycle at 95°C for 30 s, followed by 39 cycles of 95°C for 5 s and 60°C for 5 s. Dissociation curves with 0.3°C/s melt rates were used to check for the presence of non-specific dsDNA SYBR Green hybrids. Only primers with a single PCR amplification product were used in the subsequent analyses. The amplification efficiency of each primer was calculated from the slope of the standard curve [38]. The PCR primers used are listed in S1 Table. Ubiquinol-cytochrome c reductase (UCCR) and arginine kinase (AK) were used as reference genes. The difference in gene expression was measured by using the 2-ΔΔCt algorithm [39]. Differential gene expression between females and males was measured, with the female antennae used as reference. Expression levels of target genes were normalized independent of each reference gene with the algorithm, and then averaged. When the gene expression of the female antennae was very low, the gene expression of the male antennae was used as control. RNA extraction was repeated three times for each sample, and two or more RT-qPCR replicates were prepared for each sample.

Data analysis

Data analysis was conducted using SPSS 17.0 (SPSS Inc., Chicago, IL, USA). The significance of the difference between means was determined using the student’s t-test. The critical P value for each test was set at 0.05.

Results

De novo antennal transcriptome assembly

Using the Illumina HiSeq 2000 sequencing system, 117,410,034 raw reads were obtained from the antennal samples. After removing low-quality (< Q20) adaptor and contaminating sequence reads, 103,301,292 (a total of 9,297,116,280 bp) clean reads were generated from antennae, and 42,992 unigenes were assembled (N50 = 1,098), with a mean length of 713 bp. More than 58% (24,954) of the unigenes were aligned to sequences in various protein databases. GO annotation was performed to obtain information on their molecular function, biological process, and cellular location (S1 Fig). The raw sequence of the transcriptome has been deposited to the National Center for Biotechnology Information (NCBI) (GenBank Accession Number PRJNA358570; https://www.ncbi.nlm.nih.gov/bioproject/PRJNA358570).

Analysis of olfactory genes

The 35 candidate OR genes encoding an olfactory receptor co-receptor (OemaORco), OemaOR18, 7 candidate pheromone receptors (PRs, OemaOR3, 4, 21, 26, 28, 29, and 30) and 26 general OR genes were identified from O. emarginata antennae (Table 1, Fig 2). Candidate PRs of O. emarginata were clustered together with previously reported PRs in the phylogenetic tree. Eight general ORs (OemaOR11, 14, 17, 19, 20, 25, 27, and 32) were clustered with OfurOR34, MsexOR42, and AdisOR9 into a specific group, with a bootstrap support value of 87 (Fig 2). Two general OR genes (OemaOR24 and 35) were not clustered with any reported ORs from Lepidopteran species with sufficient bootstrap values (bootstrap values <50). Full open reading frame (ORF) of 8 OR genes (OemaOR5, 9, 19, 22, 26, 29, 35 and ORco) were obtained, with the mean length of 435 aa.

thumbnail
Table 1. BLASTp results of candidate olfactory receptors of O. emarginata.

https://doi.org/10.1371/journal.pone.0179433.t001

thumbnail
Fig 2. Phylogenetic analysis of putative OR gene sequences of O. emarginata (black circle).

The tree was rooted with Orco lineage (pink color). Bootstrap values < 50% are not shown. Color legend: Orange = PR group, yellow = OR18 group, green = OemaORs group, and blue = other general OR groups. Adis, Athetis dissimilis, Aips, A. ipsilon, Bmor, B. mori, Hvir, H. virescens, Msex, M. sexta, Oema, O. emarginata, Ofur, O. furnacalis, Slitu, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g002

The 41 candidate odorant-binding protein (OBP) genes were identified from O. emarginata antennae. and these encoded 34 OBPs, 2 general odorant-binding proteins (GOBPs), 3 pheromone-binding proteins (PBPs), an antennal-binding protein (OemaABPX), and OemaOBP25 (DmelOBP73a analogue) (Table 2, Fig 3). All OemaOBPs were clustered with those of Lepidopteran species with sufficient bootstrap values (bootstrap values > 60). Seven OemaOBP genes (OemaOBP4, 11, 13, 18, 23, 27, and 35) were clustered with AipsOBP4, SlitABP1, SlitOBP12, SexiABP1, HvirABP2, HarmOBP7, and HarmOBP7.2 with a bootstrap support value of 61, and the latter 7 OBPs were clustered into a subgroup with a bootstrap support value of 99 (Fig 3). The mean length of the OBPs was 166 aa, and the full ORF of the 37 OBP genes were obtained. Thirty-three OBPs were a classic group with six conserved cysteines, 3 OBPs (OemaOBP9, 28, and 30) were of the minus-C group with C2 and C5 missing, and 5 OBPs (OemaOBP3, 12, 20, 29 and 33) were of the plus-C OBP group with more than six conserved cysteines (Fig 4).

thumbnail
Table 2. BLASTp results of candidate odorant-binding proteins of O. emarginata.

https://doi.org/10.1371/journal.pone.0179433.t002

thumbnail
Fig 3. Phylogenetic analysis of putative OBP gene sequences of O. emarginata (black circle), other moth species (black lines), and Dipteran species (green lines).

The tree was rooted with the Lepidopteran GOBP-PBP group (green color). Bootstrap values < 50% are not shown. Color legend: Orange = conserved OBP groups, pink = expanded OemaOBPs group, green = Lepidopteran GOBP-PBP group, and blue = other general OBP groups. Adis, A. dissimilis, Agam, Anopheles gambiae, Aips, A. ipsilon, Bmor, B. mori, Cpun, Conogethes punctiferalis, Dmel, Drosophila melanogaster, Dple, D. plexippus, Gmol, Grapholita molesta, Harm, H. armigera, Hvir, H. virescens, Msex, M. sexta, Ofur, O. furnacalis, Oema, O. emarginata, Sexi, S. exigua, Slit, S. littoralis, Slitu, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g003

thumbnail
Fig 4. Aligned putative full ORF of OBP gene sequences of O. emarginata.

Six conserved cysteines are highlighted in blue.

https://doi.org/10.1371/journal.pone.0179433.g004

A total of 20 candidate chemosensory protein (CSP) genes were identified in O. emarginata, with a mean length of 128 aa. The full ORF of the 16 CSP genes were obtained (Table 3, Fig 5). In the phylogenetic tree, OemaCSP9 and OemaCSP16 were clustered the homologous genes of other insect species into two conserved groups (Fig 5). The bootstrap values of 5 CSPs (OemaCSP1, 2, 7, 8, and 10) were < smaller than 50%, although these were clustered with studied CSPs of the Lepidopteran species. Four conserved cysteines were found in all CSP genes, but OemaCSP16 differed from the other CSPs in terms of the number of amino acids (Fig 6).

thumbnail
Table 3. BLASTp results of candidate chemosensory proteins of O. emarginata.

https://doi.org/10.1371/journal.pone.0179433.t003

thumbnail
Fig 5. Phylogenetic analysis of putative CSP gene sequences of O. emarginata (black circles), other moth species (black lines) and Dipteran species (green lines).

Bootstrap values < 50% are not shown. Agam, A. gambiae, Aips, A. ipsilon, Bmor, B. mori, Dmel, D. melanogaster, Gmol, G. molesta, Oema, O. emarginata, Slit, S. littoralis, Slitu, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g005

thumbnail
Fig 6. Aligned putative full ORF of CSP gene sequences of O. emarginata.

Four conserved cysteines are highlighted in blue.

https://doi.org/10.1371/journal.pone.0179433.g006

Six candidate ionotropic receptor (IR) genes and 2 sensory neuron membrane protein (SNMP) genes were identified in O. emarginata, and their mean lengths were 535 aa and 522 aa, respectively (Tables 4 and 5). All O. emarginata IRs and SNMPs were clustered with Lepidopteran IRs and SNMPs, respectively, with the bootstrap values > 80% (Figs 7 and 8). The full ORF of 2 SNMP genes was obtained.

thumbnail
Table 4. BLASTp results of candidate ionotropic receptors of O. emarginata.

https://doi.org/10.1371/journal.pone.0179433.t004

thumbnail
Table 5. BLASTp results of candidate SNMP genes of O. emarginata.

https://doi.org/10.1371/journal.pone.0179433.t005

thumbnail
Fig 7. Phylogenetic analysis of putative IR gene sequences of O. emarginata (black circles).

The tree is rooted with IR25a and IR8a lineages. Bootstrap values < 50% are not shown. Bmor, B. mori, Dmel, D. melanogaster, Harm, H. armigera, Msex, M. sexta, Oema, O. emarginata, Slitu, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g007

thumbnail
Fig 8. Phylogenetic analysis of putative SNMP gene sequences of O. emarginata (black circles), D. melanogaster (black lines), other moth species (purple lines), and Hymenopteran species (green lines).

Bootstrap values < 50% are not shown. Amel, Apis mellifera, Apol, Antheraea polyphemus, Bmor, B. mori, Dmel, D. melanogaster, Harm, H. armigera, Hvir, H. virescens, Mbra, M. brassicae, Msex, M. sexta, Nvit, Nasonia vitripennis, Oema, O. emarginata, Scyn, Samia ricini, Slitu, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g008

Expression of olfactory genes with RNA sequences

The FPKM values of the chemosensory receptors were < 60, and OemaORco showed the highest FPKM value (Tables 1 and 4). The FPKM value of OemaOR29 was higher, but those of the other candidate PRs were lower than the general ORs, including OemaOR14, 25, 27, and 32 (Table 1). The FPKM values of OemaIR75p and OemaIR21a were larger than those of the co-receptors OemaIR25a and OemaIR8a (Table 4). In contrast to chemosensory receptors, 39.0% of the OBP and 52.4% of the CSP genes showed FPKM values > 300, including 3 candidate PBPs (Tables 2 and 3). OemaPBP1 showed the highest FPKM value among all OBPs, and OemaCSP19 had the highest FPKM value among all chemosensory genes. The FPKM value of OemaSNMP1 was < 20, but that of OemaSNMP2 was > 500 (Table 5).

Expression of all olfactory genes between male and female antennae

Five candidate PRs (OemaOR3, 21, 26, 28, and 30), OemaOR13, OemaOR16, OemaOR30, OemaORco, 2 GOBPs, 7 OBPs (OemaOBP4, 9–11, 26, 27, and 29), and OemaSNMP1 were expressed at significantly higher levels in females, and OemaOR26, OemaOR28, OemaOR13, and OemaOBP10 were specifically expressed in females (Fig 9). Two candidate PRs (OemaOR29 and 4), OemaOR18, 4 general ORs (OemaOR8, 15, 20, and 25), 2 PBPs (OemaPBP1 and 3), 3 OBPs (OemaOBP6, 13, and 21), 6 CSPs (OemaCSP1, 5, 6, 9, 10, and 19), OemaIR21a, and OemaSNMP2 were expressed at significantly higher levels in males compared to that in females, and OemaOR29, OemaOR4, OemaOR18, OemaOR15, OemaPBP1, and OemaPBP3 were specifically expressed in males (Fig 9).

thumbnail
Fig 9. Expression levels of olfactory genes in male and female antennae as measured by RT-qPCR analysis.

Gene expression was calculated relative to the reference genes, UCCR and AK. The expression in female antennae was arbitrarily defined as 1 for all genes and was used in the normalization of gene expression of the male antennae. A, Expression levels of CSP, IR, and SNMP genes. B, Expression levels of the OBP genes. C, Expression levels of OR genes.

https://doi.org/10.1371/journal.pone.0179433.g009

Phylogeny of pheromone recognition genes of types I and II pheromones

In the phylogenetic tree, 4 orthologous PRs clusters for type I pheromones were obtained (Cluster PRI-PRIV), and candidate PRs of the noctuid species (excluding O. emarginata) formed subclusters of these 4 clusters, with high bootstrap support (≥ 89, Fig 10). OemaOR29 and ObruOR1 (the only identified pheromone receptor for type II sex pheromones from the geometrid O. brumata) belonged to cluster PRIII (Fig 10). Other candidate PRs of O. emarginata were not grouped with any of these 4 clusters, but 5 (OemaOR3, 4, 21, 26, and 28) were clustered, with a bootstrap support of 78 (Fig 10).

thumbnail
Fig 10. The phylogeny of Lepidopteran PRs.

The tree was rooted with Orco lineage (yellow color). Bootstrap values < 50% are not shown. Genes of O. emarginata, O. brumata, and other noctuid species are indicated by black circles, black triangles, and diamonds, respectively. Clusters PRI—PRIV for type I pheromones are indicated in red, green, purple, and blue, respectively. Aseg, A. segetum, Atra, Amyelois transitella, Bmor, B. mori, Harm, H. armigera, Hvir, H. virescens, Obru, O. brumata, Oema, O. emarginata, Onub, O. nubilalis, Pxyl, P. xylostella, Sexi, S. exigua, Slit, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g010

The PBPs and GOBPs of all test species were clustered into 3 (Cluster PBPI-PBPIII) and 2 (Cluster GOBPI-II) apparent clusters, with good bootstrap support (≥ 52), respectively (Fig 11). OemaPBP3 and OemaGOBP1 were clustered with orthologous PBPs and GOBP1s of the other noctuids for type I pheromones, respectively (bootstrap support ≥ 56) (Fig 11). However, OemaPBP1, OemaPBP2, and OemaGOBP2 were not clustered within PBPs and GOBP2s from other noctuid species for type I pheromones. OemaPBP2 was clustered with MsexPBP2, with a bootstrap value of 74 (Fig 11).

thumbnail
Fig 11. The phylogeny of Lepidopteran PBPs.

The tree was rooted with GOBP lineage. Bootstrap values < 50% are not shown. Genes of O. emarginata, other species with type II pheromones, and the other noctuid species are indicated by black circles, black triangles, and diamonds, respectively. Clusters PBPI—PBPIII are indicated by orange, purple, and blue colors, respectively. Acon, Argyresthia conjugella, Aips, A. ipsilon, Apol, A. polyphemus, Asel, Ascotis selenaria cretacea, Bmor, B. mori, Cpun, C. punctiferalis, Csup, C. suppressalis, Ehip, Eogystia hippophaecolus, Harm, H. armigera, Hass, H. assulta, Gmol, G. molesta, Ldis, Lymantria dispar, Msex, M. sexta, Obru, O. brumata, Oema, O. emarginata, Ofur, O. furnacalis, Onub, O. nubilalis, Pxyl, P. xylostella, Sexi, S. exigua, Sinf, S. inferens, Slit, S. litura.

https://doi.org/10.1371/journal.pone.0179433.g011

Discussion

The unique life history of O. emarginata might have driven the increase in the number of chemosensory genes

O. emarginata has a unique life history. The larvae feed on Menispermaceae plants, but adults suck on the juices of ripe fruits. Mating behavior is mediated by female sex pheromones. Mated females oviposit on Menispermaceae plants. Odorant classes from different species might thus be different [52]. Moths of O. emarginata must recognize a range of different odors with diverse chemical structures emitted from conspecifics, fruits, or orchard background and larval host plants. The olfactory acuity and discriminatory power in O. emarginata may have evolved to fulfill its ecological needs. We found 104 candidate olfactory genes in the antennae of O. emarginata, including 35 ORs, 41 OBPs, 20 CSPs, 6 IRs, and 2 SNMPs. In these 104 olfactory genes, 2 ORs (OemaOR24 and 35) and 5 CSPs (OemaCSP1, 2, 7, 8, and 10) were not effectively clustered with those of other Lepidopterans (bootstrap values < 50) in the phylogenetic analysis. In addition, 8 OemaORs (OemaOR11, 14, 17, 19, 20, 25, 27, and 32) were clustered into the clade of OfurOR34, MsexOR42, and AdisOR9 (bootstrap value = 87) (Fig 2), and 7 OemaOBPs (OemaOBP4, 11, 13, 18, 23, 27, and 35) were clustered with AipsOBP4, SlitABP1, SlitOBP12, SexiABP1, HvirABP2, HarmOBP7, and HarmOBP7.2 (bootstrap value = 61) in the phylogenetic trees (Fig 3). Some of those genes might be species-specific to O. emarginata and used to recognize the odors produced by the Menispermaceae and fruits.

The number of chemosensory binding proteins (including OBPs and CSPs) was slightly smaller than in B. mori, which included the whole genome, but larger than in other moth species studied using the same protocol (antennal transcriptome). These other species included polyphagous insects such as S. litura (Table 6). The larger number of chemosensory binding proteins might be due to the life history of O. emarginata and the larger database in our study. We found a total of 103,301,292 reads that were assembled into 2,202,660 contigs, and compared to 55,288,304 reads assembled into 105,971 contigs in S. litura [51]. However, the number of chemosensory receptors was lower than in most other moths (Table 6). The low expression level of chemosensory receptor genes (FPKM < 60) and short read length (250 bp) of the transcriptome analysis might have resulted in short sequences for many chemosensory receptor genes. However, the long sequence of the chemosensory receptor genes (about 400 aa and 800 aa for OR and IR, respectively) [53,54] and the criterion of 50% ORF length cutoff might have excluded numerous chemosensory receptors with short sequences. No gustatory receptor gene was identified in the antennae, which suggests that the antennae of O. emarginata are not major taste organs. The proboscis, which harbors considerably fewer sensilla than antennae, are believed to specialize in taste reception in some moths [37,55]. In addition, the long sequence of gustatory receptor genes (about 400 aa) and the criterion of 50% ORF length cutoff might have excluded some gustatory receptors with short sequences.

Olfactory genes with sex-specific expression

We identified 2 candidate PRs (OemaOR29 and 4) and 2 candidate PBPs (OemaPBP1 and 3) that showed male-biased expression and might be involved with female sex pheromone recognition in O. emarginata. Our results were consistent with the study on the sex pheromone recognition in a sibling speciesm O. excavate, which produces two sex pheromone compounds at the ratio of 86:14[30]. OemaOR29 was clustered with ObruOR1 and AsegOR3 in the phylogenetic tree, which recognized the pheromonal tetraene of O. brumata, 3Z,6Z,9Z-19:H and the triene 3Z,6Z,9Z-21:H separately [56]. OemaPBP1 and OemaPBP3 were ranked in the clusters PBPI and PBPIII in the phylogenetic analysis, respectively, which showed an equally consistent association with male-specific pheromone sensitive sensilla [57]. Orthologous genes in the clusters PBPI and PBPIII play critical and minor roles in female sex pheromone perception, respectively [5861]. OemaOR29 and OemaPBP1 showed the highest FPKM values in all ORs and OBPs, respectively, and might be used to recognize the main sex pheromone component. OemaOR4 and OemaPBP3 might be involved in the recognition of the minor sex pheromone component. Further studies are needed to verify the function of these genes.

Five candidate pheromone receptor genes (OemaOR3, 21, 26, 28, and 30) showed female-biased expression, and OemaOR26, and OemaOR28 were specifically expressed in females. The function of these genes is unknown, but these might be used by females to recognize male pheromones. Production of short-range pheromones has been reported in male butterflies [62]; these function in female mate selection, act as an aphrodisiac, and arrest female departure [63,64].

Besides the candidate PR genes, some genes with sex-specific expression were detected; for example, OemaOR13 was female-specific. These genes might also be correlated with sex specific behaviors such as the recognition of oviposition cues by females [6567].

Diversification of olfactory recognition to sex pheromones

Type II pheromones have mainly been found in the moth superfamilies Geometroidea and Noctuoidea [17], but olfactory genes for type II pheromones were only identified in the geometrids A. selenaria cretacea [68,69] and O. brumata [56] and the erebids L. dispar [7072] and Hyphantria cunea [73]. The sex pheromone of female O. emarginata was not published, but it was similar to the epoxide components of a preliminary identification (Du et al., unpublished data). In addition, cis-9,10-epoxy-(Z)-6 -heneicosene and cis-9,10-epoxy-(Z, Z)-3,6-heneicosadiene were identified as the major and minor sex pheromone components from a sibling species, O. excavate [30]. In the present study, 7 candidate PRs and 3 candidate PBPs were obtained from the noctuid O. emarginata using antennal transcriptome analysis.

The diversification of olfactory recognition to sex pheromones has been verified for type I pheromones in noctuids such as A. segetum, H. armigera, and S. litura, and the phylogeny of moth PRs and PBPs for type I pheromone identified several apparent orthologous clusters (cluster PRI—PRIV for PRs and cluster PBPI—PBPIII for PBPs). PRs and PBPs from different clusters specifically respond to different type I sex pheromone components [59,74]. Although the functions of PRs for type II pheromone recognition were not identified, phylogenetic analysis clustered 3 candidate PRs of H. cunea [73] and 7 candidate PRs of O. emarginata into three groups. These findings are indicative of the diversification in olfactory recognition to type II pheromones.

Phylogenetic analysis did not separate the PRs and PBPs for types I and II pheromones, thereby suggesting that PRs and PBPs for types I and II pheromones evolved from a common ancestor. However, type I pheromones differed from type II pheromones in its chemical characteristics. OemaOR29 and ObruOR1 belonged to cluster PRIII of type I pheromone recognition, which is under strong purifying selection (a very small dN/dS values), and did not respond to any type I sex pheromone components [75]. On the contrary, ObruOR1 was verified to specifically recognize the pheromonal tetraene of O. brumata, 3Z,6Z,9Z-19:H, and the orthologous receptor AsegOR3 responded strongly to the triene 3Z,6Z,9Z-21:H instead of any female sex pheromone of A. segetum [56]. Cluster III might be specialized in the recognition type II sex pheromone components. In addition, 6 other candidate PRs of O. emarginata were not grouped within any of the four PR clusters of type I sex pheromones, but 5 of these were grouped into a specific cluster, with a bootstrap support value of 78. The candidate main sex pheromone-binding protein OemaPBP1 was not clustered into the subgroup of PBP1 genes from other noctuid species in the phylogenetic tree. These results indicate that the olfactory genes for sex pheromones in O. emarginata might differ from those of other noctuid species, and the diversification of pheromone recognition genes for types I and II sex pheromones might exist in noctuid species.

Conclusions

A total of 104 candidate olfactory genes, including 7 candidate PRs and 3 candidate PBPs were identified from the noctuid O. emarginata. Seven olfactory genes of O. emarginata were not effectively clustered with those of other Lepidoptera, and OemaORs and OemaOBPs in 2 clusters were strongly expanded. These changes in olfactory genes in O. emarginata might correlate with its unique life history. Most candidate PRs and PBPs (except for OemaOR29 and OemaPBP3) of O. emarginata were not clustered with other noctuid species. OemaOR29 was grouped into cluster PRIII of type I pheromones, which recognized type II pheromones instead of type I pheromones. Noctuid species might thus have undergone diversification of the pheromone recognition gene for types I and II sex pheromones. Our results increase our understanding of the molecular mechanism of O. emarginata olfaction and the evolution of olfactory genes associated with sex pheromones.

Supporting information

S1 File. Amino acid sequences of the olfactory genes used in the phylogenetic analysis.

https://doi.org/10.1371/journal.pone.0179433.s003

(TXT)

Acknowledgments

We are grateful to Caroline Du (University of California Irvine) for final English correction and enhancement. The Special Fund for Agro-scientific Research in the Public Interest in China (Grant No. 201203036) to YD and Ningbo Science and Technology Funds (Grant No. 2013C1025) to YQ supported this study.

Author Contributions

  1. Conceptualization: YD BF.
  2. Data curation: BF QG.
  3. Formal analysis: BF KZ YD.
  4. Funding acquisition: YD YQ.
  5. Investigation: BF QG YQ.
  6. Methodology: BF KZ.
  7. Project administration: YD.
  8. Resources: BF KZ.
  9. Software: BF.
  10. Supervision: YD.
  11. Validation: BF YD.
  12. Visualization: BF YD.
  13. Writing – original draft: BF YD.
  14. Writing – review & editing: YD.

References

  1. 1. Riffell JA, Shlizerman E, Sanders E, Abrell L, Medina B, Hinterwirth AJ, et al. (2014) Sensory biology. Flower discrimination by pollinators in a dynamic chemical environment. Science 344: 1515–1518. pmid:24970087
  2. 2. Libert S, Zwiener J, Chu X, Vanvoorhies W, Roman G, Pletcher SD (2007) Regulation of Drosophila life span by olfaction and food-derived odors. Science 315: 1133–1137. pmid:17272684
  3. 3. Goyret J, Markwell PM, Raguso RA (2008) Context- and scale-dependent effects of floral CO2 on nectar foraging by Manduca sexta. Proc Natl Acad Sci U S A 105: 4565–4570. pmid:18212123
  4. 4. Ando T, Inomata S, Yamamoto M (2004) Lepidopteran sex pheromones. Top Curr Chem 239: 51–96. pmid:22160231
  5. 5. Renou M (2014) Pheromones and General Odor Perception in Insects. In: Mucignat-Caretta C, editor. Neurobiology of Chemical Communication. Boca Raton (FL).
  6. 6. Bruce TJA, Wadhams LJ, Woodcock CM (2005) Insect host location: a volatile situation. Trends Plant Sci 10.
  7. 7. Braks MA, Leal WS, Carde RT (2007) Oviposition responses of gravid female Culex quinquefasciatus to egg rafts and low doses of oviposition pheromone under semifield conditions. J Chem Ecol 33: 567–578. pmid:17252215
  8. 8. Stelinski LL, Rodriguez-Saona C, Meyer WL (2009) Recognition of foreign oviposition-marking pheromone in a multi-trophic context. Naturwissenschaften 96: 585–592. pmid:19151965
  9. 9. El-Sayed AM, Suckling DM, Wearing CH, Byers JA (2006) Potential of mass trapping for long-term pest management and eradication of invasive species. J Econ Entomol 99: 1550–1564. pmid:17066782
  10. 10. Cook SM, Khan ZR, Pickett JA (2007) The use of push-pull strategies in integrated pest management. Annu Rev Entomol 52: 375–400. pmid:16968206
  11. 11. Witzgall P, Stelinski L, Gut L, Thomson D (2008) Codling moth management and chemical ecology. Annu Rev Entomol 53: 503–522. pmid:17877451
  12. 12. Suckling DM, Stringer LD, Bunn B, El-Sayed AM, Vander Meer RK (2010) Trail pheromone disruption of red imported fire ant. J Chem Ecol 36: 744–750. pmid:20549330
  13. 13. Carey AF, Carlson JR (2011) Insect olfaction from model systems to disease control. Proc Natl Acad Sci U S A 108: 12987–12995. pmid:21746926
  14. 14. Ding BJ, Hofvander P, Wang HL, Durrett TP, Stymne S, Lofstedt C (2014) A plant factory for moth pheromone production. Nat Commun 5: 3353. pmid:24569486
  15. 15. Leal WS (2013) Odorant reception in insects: roles of receptors, binding proteins, and degrading enzymes. Annu Rev Entomol 58: 373–391. pmid:23020622
  16. 16. Wang HL, Zhao CH, Millar JG, Cardé RT, Löfstedt C (2010) Biosynthesis of Unusual Moth Pheromone Components Involves Two Different Pathways in the Navel Orangeworm, Amyelois transitella. J Chem Ecol 36: 535–547. pmid:20393784
  17. 17. Millar JG (2000) Polyene hydrocarbons and epoxides: A Second Major Class of Lepidopteran Sex Attractant Pheromones. Annu Rev Entomol 45: 575–604. pmid:10761590
  18. 18. Ando T, Kawai T, Matsuoka K (2008) Epoxyalkenyl sex pheromones produced by female moths in highly evolved groups: biosynthesis and its endocrine regulation. J Pestic Sci 33: 17–20.
  19. 19. Krieger J, Grosse-Wilde E, Gohl T, Breer H (2005) Candidate pheromone receptors of the silkmoth Bombyx mori. Eur J Neurosci 21: 2167–2176. pmid:15869513
  20. 20. Grosse-Wilde E, Kuebler LS, Bucks S, Vogel H, Wicher D, Hansson BS (2011) Antennal transcriptome of Manduca sexta. Proc Natl Acad Sci U S A 108: 7449–7454. pmid:21498690
  21. 21. Vogel H, Heidel AJ, Heckel DG, Groot AT (2010) Transcriptome analysis of the sex pheromone gland of the noctuid moth Heliothis virescens. BMC Genomics 11: 29. pmid:20074338
  22. 22. Feng B, Lin X, Zheng K, Qian K, Chang Y, Du Y (2015) Transcriptome and expression profiling analysis link patterns of gene expression to antennal responses in Spodoptera litura. BMC Genomics 16: 269. pmid:25887537
  23. 23. Jacquin-Joly E, Legeai F, Montagne N, Monsempes C, Francois MC, Poulain J, et al. (2012) Candidate chemosensory genes in female antennae of the noctuid moth Spodoptera littoralis. Int J Biol Sci 8: 1036–1050. pmid:22904672
  24. 24. Legeai F, Malpel S, Montagne N, Monsempes C, Cousserans F, Merlin C, et al. (2011) 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 12: 86. pmid:21276261
  25. 25. Gu SH, Sun L, Yang RN, Wu KM, Guo YY, Li XC, et al. (2014) Molecular characterization and differential expression of olfactory genes in the antennae of the black cutworm moth Agrotis ipsilon. PLoS ONE 9: e103420. pmid:25083706
  26. 26. Zhang S, Zhang Z, Wang H, Kong X (2014) Antennal transcriptome analysis and comparison of olfactory genes in two sympatric defoliators, Dendrolimus houi and Dendrolimus kikuchii (Lepidoptera: Lasiocampidae). Insect Biochem Mol Biol 52: 69–81. pmid:24998398
  27. 27. Feng B, Hu WX, Pan H, Du YJ (2013) Morphology, life history and circadian rhythm of the fruit-piercing moth, Oraesia emarginata (Lepidoptera: Noctuidae). Acta Entomol Sinica 56: 1440–1451.
  28. 28. Izumi Y, Tian R, Sonoda S, Imayoshi Y, Iwabuchi H, Miyashita Y, et al. (2015) Analysis of peach fruit headspace volatiles and response by the fruit-piercing moth Oraesia excavata (Lepidoptera: Noctuidae). Appl Entomol Zool 50: 231–238.
  29. 29. Tian R, Izumi Y, Sonoda S, Yoshida H, Fukumoto T, Saito T, et al. (2007) Estimation of repellency of a volatile compound, sec-butyl β-styryl ketone, against fruit-piercing moths. Appl Entomol Zool 42: 433–437.
  30. 30. Ohmasa Y, Wakamura S, Kozai S, Sugie H, Horiike M, Hiran C, et al. (1991) Sex Pheromone of the Fruit-Piercing Moth,Oraesia excavata(BUTLER)(Lepidoptera:Noctuidae):Isolation and Identification. Appl Entomol Zool 26: 55–62.
  31. 31. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. (2011) Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotech 29: 644–652.
  32. 32. Conesa A, Gotz S, Garcia-Gomez JM, Terol J, Talon M, Robles M (2005) Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 21: 3674–3676. pmid:16081474
  33. 33. Petersen TN, Brunak S, von Heijne G, Nielsen H (2011) SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat Meth 8: 785–786.
  34. 34. McKenzie SK, Oxley PR, Kronauer DJ (2014) Comparative genomics and transcriptomics in ants provide new insights into the evolution and function of odorant binding and chemosensory proteins. BMC Genomics 15: 718. pmid:25159315
  35. 35. Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 25: 4876–4882. pmid:9396791
  36. 36. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6: molecular evolutionary genetics analysis version 6.0. Mol Biol Evol 30: 2725–2729. pmid:24132122
  37. 37. Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B (2008) Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 5: 621–628. pmid:18516045
  38. 38. Kubista M, Andrade JM, Bengtsson M, Forootan A, Jonák J, Lind K, et al. (2006) The real-time polymerase chain reaction. Mol Aspects Med 27: 95–125. pmid:16460794
  39. 39. Schmittgen TD, Livak KJ (2008) Analyzing real-time PCR data by the comparative CT method. Nat Protocols 3: 1101–1108. pmid:18546601
  40. 40. Wanner KW, Robertson HM (2008) 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 17: 621–629. pmid:19133074
  41. 41. Tanaka K, Uda Y, Ono Y, Nakagawa T, Suwa M, Yamaoka R, et al. (2009) Highly Selective Tuning of a Silkworm Olfactory Receptor to a Key Mulberry Leaf Volatile. Curr Biol 19: 881–890. pmid:19427209
  42. 42. Cao D, Liu Y, Wei J, Liao X, Walker WB, Li J, et al. (2014) Identification of Candidate Olfactory Genes in Chilo suppressalis by Antennal Transcriptome Analysis. Int J Biol Sci 10: 846–860. pmid:25076861
  43. 43. Bengtsson JM, Trona F, Montagné N, Anfora G, Ignell R, Witzgall P, et al. (2012) Putative Chemosensory Receptors of the Codling Moth, Cydia pomonella, Identified by Antennal Transcriptome Analysis. PLoS ONE 7: e31620. pmid:22363688
  44. 44. Walker WB Iii, Gonzalez F, Garczynski SF, Witzgall P (2016) The chemosensory receptors of codling moth Cydia pomonella—expression in larvae and adults. Sci Rep 6: 23518. pmid:27006164
  45. 45. Zhang S, Zhang Z, Wang H, Kong X (2014) Antennal transcriptome analysis and comparison of olfactory genes in two sympatric defoliators, Dendrolimus houi and Dendrolimus kikuchii (Lepidoptera: Lasiocampidae). Insect Biochemistry and Molecular Biology 52: 69–81. pmid:24998398
  46. 46. Zhang J, Wang B, Dong S, Cao D, Dong J, Walker WB, et al. (2015) Antennal Transcriptome Analysis and Comparison of Chemosensory Gene Families in Two Closely Related Noctuidae Moths, Helicoverpa armigera and H. assulta. PLoS ONE 10: e0117054. pmid:25659090
  47. 47. Yang B, Ozaki K, Ishikawa Y, Matsuo T (2015) Identification of Candidate Odorant Receptors in Asian Corn Borer Ostrinia furnacalis. PLoS ONE 10: e0121261. pmid:25803580
  48. 48. Zhang T, Coates BS, Ge X, Bai S, He K, Wang Z (2015) Male- and Female-Biased Gene Expression of Olfactory-Related Genes in the Antennae of Asian Corn Borer, Ostrinia furnacalis (Guenee) (Lepidoptera: Crambidae). PLoS ONE 10: e0128550. pmid:26062030
  49. 49. Zhang YN, Jin JY, Jin R, Xia YH, Zhou JJ, Deng JY, et al. (2013) Differential Expression Patterns in Chemosensory and Non-Chemosensory Tissues of Putative Chemosensory Genes Identified by Transcriptome Analysis of Insect Pest the Purple Stem Borer Sesamia inferens (Walker). PLoS ONE 8: e69715. pmid:23894529
  50. 50. Poivet E, Gallot A, Montagné N, Glaser N, Legeai F, Jacquin-Joly E (2013) A Comparison of the Olfactory Gene Repertoires of Adults and Larvae in the Noctuid Moth Spodoptera littoralis. PLoS ONE 8: e60263. pmid:23565215
  51. 51. Feng B, Lin X, Zheng K, Qian K, Chang Y, Du Y (2015) Transcriptome and expression profiling analysis link patterns of gene expression to antennal responses in Spodoptera litura. BMC Genomics 16: 269. pmid:25887537
  52. 52. Carey AF, Wang G, Su C-Y, Zwiebel LJ, Carlson JR (2010) Odorant reception in the malaria mosquito Anopheles gambiae. Nature 464: 66–71. pmid:20130575
  53. 53. Krieger J, Raming K, Dewer YME, Bette S, Conzelmann S, Breer H (2002) A divergent gene family encoding candidate olfactory receptors of the moth Heliothis virescens. Eur J Neurosci 16: 619–628. pmid:12270037
  54. 54. Olivier V, Monsempes C, François MC, Poivet E, Jacquin-Joly E (2011) Candidate chemosensory ionotropic receptors in a Lepidoptera. Insect Mol Biol 20: 189–199. pmid:21091811
  55. 55. Reiter S, Campillo Rodriguez C, Sun K, Stopfer M (2015) Spatiotemporal Coding of Individual Chemicals by the Gustatory System. J Neurosci 35: 12309–12321. pmid:26338341
  56. 56. Zhang DD, Wang HL, Schultze A, Froß H, Francke W, Krieger J, et al. (2016) Receptor for detection of a Type II sex pheromone in the winter moth Operophtera brumata. Sci Rep 6: 18576. pmid:26729427
  57. 57. Vogt RG, Große-Wilde E, Zhou JJ (2015) The Lepidoptera Odorant Binding Protein gene family: Gene gain and loss within the GOBP/PBP complex of moths and butterflies. Insect Biochem Mol Biol 62: 142–153. pmid:25784631
  58. 58. Liu NY, He P, Dong SL (2012) Binding properties of pheromone-binding protein 1 from the common cutworm Spodoptera litura. Comp Biochem Phys B 161: 295–302.
  59. 59. Liu NY, Liu CC, Dong SL (2013) Functional differentiation of pheromone-binding proteins in the common cutworm Spodoptera litura. Comp Biochem Phys A 165: 254–262.
  60. 60. Zhang T-T, Mei X-D, Feng J-N, Berg BG, Zhang Y-J, Guo Y-Y (2012) Characterization of three pheromone-binding proteins (PBPs) of Helicoverpa armigera (Hübner) and their binding properties. J Insect Physiol 58: 941–948. pmid:22549127
  61. 61. Zhu GH, Xu J, Cui Z, Dong XT, Ye ZF, Niu DJ, et al. (2016) Functional characterization of SlitPBP3 in Spodoptera litura by CRISPR/Cas9 mediated genome editing. Insect Biochem Mol Biol 75: 1–9. pmid:27192033
  62. 62. Nieberding CM, Fischer K, Saastamoinen M, Allen CE, Wallin EA, Hedenström E, et al. (2012) Cracking the olfactory code of a butterfly: the scent of ageing. Ecol Lett 15: 415–424. pmid:22390373
  63. 63. Nieberding CM, de Vos H, Schneider MV, Lassance J-M, Estramil N, Andersson J, et al. (2008) The Male Sex Pheromone of the Butterfly Bicyclus anynana: Towards an Evolutionary Analysis. PLoS ONE 3: e2751. pmid:18648495
  64. 64. Delle-Vedove R, Frérot B, Hossaert-McKey M, Beaudoin-Ollivier L (2014) Courtship Behavior of the Castniid Palm Borer, Paysandisia archon: Potential Roles of Male Scents and Visual Cues in a Day-Flying Moth. J Insect Sci 14: 52. pmid:25373199
  65. 65. Wanner KW, Anderson AR, Trowell SC, Theilmann DA, Robertson HM, Newcomb RD (2007) Female-biased expression of odourant receptor genes in the adult antennae of the silkworm, Bombyx mori. Insect Mol Biol 16: 107–119. pmid:17257213
  66. 66. Iatrou K, Biessmann H (2008) Sex-biased expression of odorant receptors in antennae and palps of the African malaria vector Anopheles gambiae. Insect Biochem Mol Biol 38: 268–274. pmid:18207086
  67. 67. Anderson AR, Wanner KW, Trowell SC, Warr CG, Jaquin-Joly E, Zagatti P, et al. (2009) Molecular basis of female-specific odorant responses in Bombyx mori. Insect Biochem Mol Biol 39: 189–197. pmid:19100833
  68. 68. Watanabe H, Tabunoki H, Miura N, Matsui A, Sato R, Ando T (2009) Identification of a New Pheromone-Binding Protein in the Antennae of a Geometrid Species and Preparation of Its Antibody to Analyze the Antennal Proteins of Moths Secreting Type II Sex Pheromone Components. Biosci Biotech Biochem 73: 1443–1446.
  69. 69. Watanabe H, Tabunoki H, Miura N, Sato R, Ando T (2007) Analysis of odorant-binding proteins in antennae of a geometrid species, Ascotis selenaria cretacea, which produces lepidopteran Type II sex pheromone components. Invertebr Neurosci 7: 109–118.
  70. 70. Yu Y, Ma F, Cao Y, Zhang J, Zhang Y, Duan S, et al. (2012) Structural and Functional Difference of Pheromone Binding Proteins in Discriminating Chemicals in the Gypsy Moth, Lymantria Dispar. Int J Biol Sci 8: 979–991. pmid:22904666
  71. 71. Yu Y, Plettner E (2013) Enantiomer and conformer recognition of (+) and (−)-disparlure and their analogs by the pheromone binding proteins of the gypsy moth, Lymantria dispar. Bioorg Med Chem 21: 1811–1822. pmid:23434366
  72. 72. Sanes JT, Plettner E (2016) Gypsy moth pheromone-binding protein-ligand interactions: pH profiles and simulations as tools for detecting polar interactions. Arch Biochem Biophys 606: 53–63. pmid:27431057
  73. 73. Zhang LW, Kang K, Jiang SC, Zhang YN, Wang TT, Zhang J, et al. (2016) Analysis of the Antennal Transcriptome and Insights into Olfactory Genes in Hyphantria cunea (Drury). PLoS ONE 11: e0164729. pmid:27741298
  74. 74. Zhang D-D, Löfstedt C (2015) Moth pheromone receptors: gene sequences, function, and evolution. Front Ecol Evol 3.
  75. 75. Zhang D-D, Löfstedt C (2013) Functional Evolution of a Multigene Family: Orthologous and Paralogous Pheromone Receptor Genes in the Turnip Moth, Agrotis segetum. PLoS ONE 8: e77345. pmid:24130875