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

Functional Characterization of Chemosensory Proteins in the Scarab Beetle, Holotrichia oblita Faldermann (Coleoptera: Scarabaeida)

  • Hongyan Sun ,

    Contributed equally to this work with: Hongyan Sun, Li Guan

    Affiliations State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, P.R. China, College of Plant Science, Jilin University, Changchun, Jilin Province, P.R. China

  • Li Guan ,

    Contributed equally to this work with: Hongyan Sun, Li Guan

    Affiliations State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, P.R. China, College of Plant Science, Jilin University, Changchun, Jilin Province, P.R. China

  • Honglin Feng,

    Affiliations State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, Chinese Academy of Agricultural Sciences, Beijing, P.R. China, College of Plant Science, Jilin University, Changchun, Jilin Province, P.R. China

  • Jiao Yin,

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

  • Yazhong Cao,

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

  • Jinghui Xi,

    Affiliation College of Plant Science, Jilin University, Changchun, Jilin Province, P.R. China

  • Kebin Li

    likebin54@163.com

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

Abstract

Chemosensory proteins (CSPs) play important roles in chemical communication by insects, as they recognize and transport environmental chemical signals to receptors within sensilla. In this study, we identified HoblCSP1 and HoblCSP2 from a cDNA library of Holotrichia oblita antennae, successfully expressed them in E. coli and purified them by Ni ion affinity chromatography. We then measured the ligand-binding specificities of HoblCSP1 and HoblCSP2 to 50 selected ligands in a competitive binding assay. These results demonstrated that HoblCSP1 and HoblCSP2 have similar ligand-binding spectra. Both proteins displayed the highest affinity for β-ionone, α-ionone and cinnamaldehyde, indicating that they prefer binding to odorants other than sex pheromones. Additionally, immuno-localization revealed that HoblCSP1 is highly concentrated in sensilla basiconica, while HoblCSP2 is specifically localized to sensilla placodea. In conclusion, HoblCSP1 and HoblCSP2 are responsible for binding to general odorants with slightly different specificities due to their different in vivo environments.

Introduction

The scarab beetle, Holotrichia oblita Faldermann (Coleoptera: Scarabaeida), is a prominent underground pest that causes great economic loss throughout its life cycle [1]. This insect is a polyphagous pest that feeds on a range of plants, such as peanut, soybean, wheat, and potato [2]. Studies have demonstrated that H. oblita responds differentially to different plant leaf odors, indicating that H. oblita can find hosts using plant volatiles as cues [3][5].

Insects develop an elaborate, sensitive, and specific olfactory system to perceive chemical signals in the environment. This system confers the capacity to communicate with mates, locate foods, oviposit, and avoid natural enemies [3], [6], [7]. The insect olfactory system is primarily composed of an antennae lobe in the brain and morphologically distinct sensilla [8]. Most sensilla locate in antennae and/or maxillary palps, which are rich in olfactory receptor neurons (ORN). Odorants pass through a specific channel in the cuticle to the ORN lymph, where they stimulate the odorant receptors (ORs). After the ORs are activated, the odorants are degraded by odorant-degrading enzymes (ODEs) [9].

Chemosensory proteins (CSPs), also known as olfactory specific-D like (OS-D like) proteins or sensory appendage proteins (SAPs), make up one of the most important sensor protein groups in insect chemoreceptors [10]. Drosophila melanogaster OS-D protein [11] and A-10 [12] were the first two reported CSPs. CSPs are generally acidic, soluble proteins that are approximately 13 kDa with 100–115 amino acids. All CSPs contain 4 conserved cysteine residues, which form 2 disulfide bonds (S-S) with their neighboring sulfur side chains. Each of these S-S bonds forms a ring, one by linking 8 surrounding amino acids, the other by linking 4 amino acids [13], [14]. Phylogenetic analysis of 180 CSPs from seven different insect orders demonstrated that CSPs are highly conserved with a N-terminal signature sequence, “YTTKYDN[VI][ND][LV]DEIL” [15], [16] and several α-helix domains in the secondary structures [17][20]. For example, Bombyx mori pheromone binding protein contains six α-helices, forming a cavity for binding pheromones [21]. The primary and secondary structures of CSPs are highly conserved across all insects [17], [19], [20]. This specific structure allows CSPs to interact with linear-chain compounds such as oleamide, which is an endogenous ligand of locust CSPs [22]. However, only a few three-dimensional CSP structures have been reported, including only Mamestra brassicae CSP-A6 (MbraCSP-A6) [23], Schistocerca gregaria CSP4 (SgrCSP4) [24], and B. mori CSP1 (BmorCSP1) [25].

CSPs exist in both male and female organisms. They are not only found in sensilla within antennae [13], [26], [27], proboscises [28], maxillary palps [29], labial palps [30], and tarsus [31], but they are also observed in the abdomen, truncus, cuticle [32], legs [33], wings [22], and pheromone glands [34]. This non-confined localization would allow them to function in a wide spectrum of signaling events [35]. Furthermore, the expression of CSPs in insect pheromone glands indicates that these molecules are involved in formation and transportation of sex pheromones, such as lipocalin in the uterus and/or saliva of mammals [19]. The expression of SgreCSP1-5 and MbraCSP6 in the sensillum lymph indicates that they function in odorant sensing in the same manner as odorant binding proteins (OBPs) [36]. Additionally, differential expression of CSPs throughout development indicates that CSPs functions are also regulated temporally [37]. For example, Apis mellifera CSP5 (AmelCSP5) was found in the embryonic ectoderm of the Italian honeybee, a result was thought to be relevant to the embryo casings’ formation [38].

CSPs present in insects across different orders, including Diptera (4 CSPs in D. melanogaster, 8 CSPs in Anopheles gambiae [39]), Lepidoptera (10 CSPs in M. brassicae [28] and 16 CSPs in B. mori) and so on. Combining the recent discovery of 20 CSPs in Tribolium castaneum, these indicated the importance of CSPs and their potential as targets for pest control. Here, we constructed a cDNA library from antennae of H. oblita and found two CSPs, HoblCSP1 and HoblCSP2. Using a competitive binding assay with the fluorescent probe 1-NPN (N-phenyl-1-naphthylamine), we characterized the ligand-binding specificities and inspected the spatial localizations of these HoblCSPs as a means to improve our understanding of CSPs roles in insects.

Materials and Methods

1 Ethics statement

All animal experiments in this study were performed in strict accordance with the guidelines developed by the State Key Laboratory for Biology of Plant Diseases and Insect Pests, Institute of Plant Protection, and the Chinese Academy of Agricultural Science (IPP, CAAS). The protocol was approved by the committee on the Ethics of Animal Experiments of the IPP, CAAS. The Approval ID is SYXK (Beijing) and the Permit Number is 2008-008.

2 Insects

Adult scarab beetles (H. oblita) were collected from the Hefei Experimental Base of the Institute of Plant Protection, Anhui Academy of Agriculture Science, Hefei, Anhui Province, China, and maintained in the lab at 28°C under a standard photoperiod (L/D: 16 h/8 h). Antennae from adult females were excised and immediately frozen in liquid nitrogen.

3 Construction of an H. oblita antennae cDNA library

Total RNA from 100 female antennae was extracted with Trizol (Invitrogen, USA). Full-length double-stranded cDNA (ds cDNA) with blunt cDNA ends was synthesized and amplified using the Creator™ SMART™ cDNA Library Construction Kit (Clontech, USA). Synthesized ds cDNA was then incubated with 0.08 µg/µl proteinase K at 45°C for 20 min to inactivate the DNA polymerase. After size fractionation using CHROMA SPIN™ columns, the cDNA was incorporated into SfiI-digested λTripIE vector. The recombinant phage vector was transduced into E. coli XL1-Blue (TaKaRa Co., China). The plaques were counted to calculate the phage titer (pfu/ml), and the recombination efficiency was estimated by calculating the ratio of white (recombinant) to blue (non-recombinant) plaques. Fragments >350 bp were sequenced.

4 Identification and sequence analysis of HoblCSP1 and HoblCSP2

Using contig alignment coupled with NCBI BLAST, HoblCSP1 and HoblCSP2 were identified from the antennae cDNA library. The full-length sequences of HoblCSP1 and HoblCSP2 were cloned and verified by fishing with sequence-specific primers. The primers for HoblCSP1 were: forward 5′-GAAAAGAAAAACGATAACGAA-3′ and reverse 5′-CACAATTTTACGTTGGAAGAT-3′, while the primers for HoblCSP2 were: forward 5′-AGATATACAACAAAATATGATAA-3′ and reverse 5′-CAATGTATGCAACAGTGTCCAAG-3′. The signal peptides were predicted through SignalP 3.0 [40], and the molecular weights were calculated using the SWISS-PROT (ExPASy server) program “Compute pI/Mw.” The hydrophobicity of each predicted protein was analyzed at http://us.expasy.org/cgi-bin/protscale.pl. And, BLAST and Mult-Alin were used for homology searches and the alignment of nucleotide and/or amino acid sequences.

The evolutionary history of insect CSPs was inferred by using the Maximum Likelihood method based on the JTT matrix-based model [41] in MEGA 6.06 [42]. Briefly, the CSP cDNAs were aligned in ClustalW2 [43] and the alignment was improved by removing the spurious sequences and poorly aligned regions in TrimAl [44] by setting the gap threshold to 0.25. The tree with the highest log likelihood (−11948.7261) is shown. The analysis involved 109 amino acid sequences. All positions with less than 95% site coverage were eliminated. That is, fewer than 5% alignment gaps, missing data, and ambiguous bases were allowed at any position.

5 Prokaryotic expressions of HoblCSP1 and HoblCSP2

Recombinant pET30a(+)/CSP1 and pET30a(+)/CSP2 were generated by ligating the sticky ends of the designed HoblCSP1 and HoblCSP2 constructs into the expression vector pET30a (Novagen, Germany). pET30a(+)/CSP1 and pET30a(+)/CSP2 were then transformed into E. coli BL21 (DE3) pLysS cells. The expressions of recombinant HoblCSP1 and HoblCSP2 were induced for 4 h by 0.5 mM IPTG following a 3 h pre-incubation. The cells were harvested by centrifugation and then homogenized in phosphate-buffered saline (PBS, 0.04 M, pH 7.0). After centrifugation at 12,000×g for 20 min at 4°C, the supernatants were purified by Ni ion affinity chromatography (GE-Healthcare). Recombinant HoblCSP1 and HoblCSP2 were identified by western blot analysis with antibodies designed against the His-tag (Abcam, USA). To prevent confounding effects in the subsequent experiments, the His-tag was removed by recombinant enterokinase (rEK) (Bio Basic Inc.) and NaCl was removed by dissolving the proteins in dH2O and filtering with a 10 kDa Amicon Ultra-0.5 Device (Millipore, USA). Finally the recombinant proteins were stored at −80°C, until required.

6 Fluorescence competition assays

Fluorescence binding assay was based on method previously described by Yin et al. [45]. Briefly, fifty compounds (Sigma-Aldrich, Germany) with chemical purities ≥97% were tested in a Lengguang 970 CRT spectrofluorimeter (Shanghai Jingmi, China). Assuming the proteins were 100% active, the binding affinities for N-phenyl-1-naphthylamine (1-NPN) were determined by adding aliquots of 1-NPN into a 2 µM protein solution for final concentrations of 2∼24 µM. The dissociation constants of the binding competitors were calculated from IC50 according to Campanacci et al: Ki = [IC50]/(1+[1-NPN]/K(1-NPN)), where [1-NPN] represents the concentration of unbound 1-NPN and K(1-NPN) is the dissociation constant of 1-NPN [46]. Binding data for each ligand was collected from 3 measurements.

7 Preparation of anti-HoblCSP1 and anti-HoblCSP2 antibodies

Purified full-length HoblCSP1 and HoblCSP2 were injected into New Zealand white rabbits following a standard immunization protocol for antibody production. Briefly, 100 µg of recombinant CSP was injected with an equal volume of Freund’s complete adjuvant, followed by three additional injections of 500 µg, with one each on the 21st, 35th, and 49th day. The antiserum was then tested using an enzyme-linked immunosorbent assay (ELISA) and used without further purification. The pre-injected rabbit serum was used as a negative control.

8 Spatial localizations of HoblCSP1 and HoblCSP2 in antennae of H. oblita

Antennal lamellae from both males and females were excised from adult H. oblita and fixed in a mixture of 4% paraformaldehyde (Thermo Scientific, USA) and glutaraldehyde (2%) in 0.1 M PBS (pH 7.4). They were then embedded in LR White resin (TAAB, UK) after dehydration in a graded ethanol series. Ultrathin sections (500–700 nm) were cut on a microtome with glass blades and then incubated with primary anti-HoblCSP1 (1∶2000) and anti-HoblCSP2 (1∶2000) antibodies at 4°C overnight. After incubation, the sections were washed in 2 times in PBGT then incubated with anti-rabbit IgG secondary antibody (1∶20), coupled to 10-nm colloidal gold, for 90 min at room temperature. Gold granules were size-increased by silver intensification for 15 min in the dark, followed by incubation with 2% uranyl acetate for 15 min to increase the contrast. The samples were imaged by transmission electron microscopy (HitachiH-7500).

Results

Characteristics of the HoblCSP1 and HoblCSP2 sequences

From the antennae cDNA library, full-length HoblCSP1 (GenBank: HQ683720) and HoblCSP2 (GenBank: HQ688991) genes were obtained and verified. The open reading frame (ORF) of HoblCSP1 contained 399 nucleotides, encoding 132 amino acids. The predicted molecular weight of HoblCSP1 was 15.55 kDa. The ORF of HoblCSP2 was comprised of 390 nucleotides, encoding 129 amino acids, and the predicted molecular weight was 14.78 kDa. At their N-termini, HoblCSP1 and HoblCSP2 contain signal peptides of 16 and 18 residues, respectively. Both proteins contained 4 conserved cystine residues (Fig. 1A, 1B), consistent with the model Cys-X6–8-Cys-X16–21-CysX2–4-Cys, and also contained a hydrophobic domain. The isoelectric points (PI) of HoblCSP1 and HoblCSP2 were dramatically different at 4.93 and 8.14, respectively. Phylogenetic analysis based on the deduced amino acid sequence also revealed that HoblCSP1 and HoblCSP2 are highly divergent, with HoblCSP1 and HoblCSP2 in two separated Coleopteran mono-phylogenetic groups (Figure 1C).

thumbnail
Figure 1. Characterization and phylogenetic tree of HoblCSP1 and HoblCSP2.

(A–B) Nucleotide and putative amino acid sequence analysis of HoblCSP1 (A) and HoblCSP2 (B). The predicted signal peptides are underlined (Generated from: http://bioinformatics.leeds.ac.uk/prot_analysis/Signal.html). The four conserved cysteine residues are highlighted in pink, and the stop codons are marked as dots in both sequences. (C) Molecular phylogenetic analysis by Maximum Likelihood method. CSPs used include HoblCSP1, HoblCSP2 and 20 CSPs from T. castaneum, 11 from D. ponderosae [53], 6 from Ips typographus [53], 8 from Diabrotica virgifera [51], 1 from D. abbreviates [51], 2 from L. decemlineata [51], 7 from M. alternatus [52], 7 from Dastarcus helophoroides [52], 3 from B. horsfieldi [50], 16 from B. mori [69], 3 from Heliothis virescens [70], 6 from Apis mellifera [16], 2 from Locusta migratoria [22], 5 from A. gambiae [57], 2 from D. melanogaster [15], 10 from Acyrthosiphon pisum [71]. HoblCSP1 and HoblCSP2 are marked with solid dark circles and all other CSPs from Coleopteran are marked with open circles. All sequences are available from the NCBI database.

https://doi.org/10.1371/journal.pone.0107059.g001

Procaryotic expression of HoblCSP1 and HoblCSP2

The recombinant proteins pET30a(+)/CSP1 and pET30a(+)/CSP2 were successfully expressed in BL21(DE3) PlysS cells. For both CSPs, a specific band less than 24 kDa (including His-tag) was observed by western blot analysis (Figure 2), which was consistent with the molecular weight deduced from their predicted amino acid sequences. HoblCSP1 and HoblCSP2 were purified at concentrations of 1.1 mg/ml and 1.2 mg/ml, respectively.

thumbnail
Figure 2. Prokaryotic expression and purification of HoblCSP1 and HoblCSP2.

The purified fusion proteins pET30a(+)/CSP1 and pET30a(+)/CSP2 are shown in separate lanes labeled CSP1 and CSP2, respectively. M: protein molecular weight marker (top to bottom: 90, 62, 40, 30, 24, 16 kDa). HoblCSP1 and HoblCSP2 were identified by using antibodies designed against the His-tag.

https://doi.org/10.1371/journal.pone.0107059.g002

Binding specificities of HoblCSP1 and HoblCSP2 largely overlapped

Based on the dissociation constants of CSP1/1-NPN (2.53 µM) and CSP2/1-NPN (2.93 µM) calculated from the binding curves (Figure 3A), fifty potential odorant compounds were selected for a fluorescence competition assay with 1-NPN. These molecules included Ricinus communis leaf volatiles that attract H. oblita [47]; volatiles isolated from H. oblita host plant, Ulmus pumila Linnaeus [48], [49]; putative sex pheromones from closely related beetles; and previously reported compounds (Table 1). The inhibition constants Ki (for each CSP/ligand combination) are summarized in Table 1. The binding curves of a few representative fluorescence competition assays are presented in figure 3B. These binding curves, coupled with the Ki values, demonstrated that HoblCSP1 and HoblCSP2 displayed similar spectra of binding activity. Of the 50 selected compounds, HoblCSP1 preferred 22, while HoblCSP2 preferred 18. Within these groups, 15 compounds were covered in the binding spectra of both HoblCSP1 and HoblCSP2 (Table 1). Furthermore, both HoblCSP1 and HoblCSP2 bound most strongly to β-ionone, followed by α-ionone and cinnamaldehyde. Other ligands, however, were unique to each of the proteins. For example, HoblCSP1 was able to bind camphene, albeit with a high Ki value, while HoblCSP2 could not bind camphene at all (Figure 3B).

thumbnail
Figure 3. Fluorescence competition assay of the recombinant HoblCSP1 and HoblCSP2.

(A) Binding curves of 1-NPN with HoblCSP1 and HoblCSP2 and their relative Scatchard plots. The dissociation constant of HoblCSP1 is 2.53 µM and that of HoblCSP2 is 2.93 µM. (B) Representative competitive binding curves of the recombinant HoblCSP1 to a series selected ligands (see Table 1). All proteins used were diluted to a fixed concentration of 2 µM, while the concentration of 1-NPN varied with the respective dissociation constant of HoblCSPs/1-NPN. The mixed solution was then titrated with 1 mM of each competing ligand to final concentrations of 0–28 µM. Fluorescence intensities are plotted as percent of the initial fluorescence in the absence of ligands. The molecular formulas of representative ligands are shown here as well. The calculated dissociation constants for all of the ligands are listed in Table 1.

https://doi.org/10.1371/journal.pone.0107059.g003

thumbnail
Table 1. Affinities of selected HoblCSP1 and HoblCSP2 pure odorant ligands.

https://doi.org/10.1371/journal.pone.0107059.t001

Spatial localizations of HoblCSP1 and HoblCSP2 in antennae of H. oblita

Generally, both HoblCSP1 and HoblCSP2 were found in the antenna of both male and female adult H. oblita. These two proteins were primarily distributed in the outer sensillum lymph, with different concentrations in different types of sensilla. HoblCSP1 was concentrated in sensilla placodea of both males and females (Figure 4B, 4B’, 4D and 4D’), and more highly concentrated in male sensilla basiconica (Figure 4A, 4C). However, HoblCSP2 was found to be sensilla-specific, as it was highly expressed in sensilla placodea but rarely found in sensilla basiconica (Figure 5).

thumbnail
Figure 4. Spatial localization of HoblCSP1 in olfactory sensilla basiconica and sensilla placodea of adult H. oblita antennae.

All samples are labeled with an anti-HoblCSP1 antibody. HoblCSP1 proteins are shown as black dots. (A–A’) Longitudinal (A) and Cross (A’) sections of male sensilla basiconica; (B–B’) Longitudinal (B) and Cross (B’) sections of male sensilla placodeum; (C–C’) Longitudinal (C) and Cross (C’) sections of female sensilla basiconica; (D–D’) Longitudinal (D) and Cross (D’) sections of female sensilla placodeum. All antibodies used were diluted to 1∶5000. Each treatment was repeated more than 3 times. Scale bar = 500 nm.

https://doi.org/10.1371/journal.pone.0107059.g004

thumbnail
Figure 5. Spatial localization of HoblCSP2 in olfactory sensilla basiconica and sensilla placodea of adult H. oblita antennae.

All samples are labeled with an anti-HoblCSP2 antibody. HoblCSP2 proteins are shown as black dots. (A) Longitudinal section of male sensilla basiconica; (B–B’) Longitudinal (B) and cross (B’) sections of male sensilla placodeum; (C) Longitudinal section of female sensilla basiconica; (D–D’) Longitudinal (D) and cross (D’) sections of female sensilla placodeum. All antibodies used were diluted to 1∶10000. Each treatment was repeated more than 3 times. Scale bar = 1 µm.

https://doi.org/10.1371/journal.pone.0107059.g005

Discussion

In this study, we characterized two CSPs, HoblCSP1 and HoblCSP2, from a H. oblita antennal cDNA library. Similar numbers of CSPs were also found in other beetles, like 3 CSPs in Batocera horsfieldi [50], 2 CSPs in Leptinotarsa decemlineata, and 1 CSP in Diaprepes abbreviatus [51]. In contrast, 12, 11, and 20 CSPs were found in Monchamus alternatus [52], Dendroctonus ponderosae [53], and T. casteneum [54] respectively. Different numbers of CSPs may be required to distinguish different host plants due to the varieties of their ligand-binding specificities. However, further investigations are required to verify this. HoblCSP1 and HoblCSP2 are highly divergent in deduced amino acid sequences. Additionally, phylogenetic analysis demonstrated that HoblCSP1 and HoblCSP2 are not in a recent mono-phylogeny group (Figure 1C), indicating that they diverged from each other a long time ago. Coupled with the markedly different PIs of HoblCSP1 (4.93) and HoblCSP2 (8.14), we predicted that HoblCSP1 and HoblCSP2 function differently. Surprisingly, we found that the binding spectra of HoblCSP1 and HoblCSP2 largely overlapped. Among the 50 selected compounds, β-ionone and its isoform α-ionone and cinnamaldehyde displayed the highest affinities to both HoblCSP1 and HoblCSP2. This is consistent with the result from the electroantennogram (EAG) examinations that showed α-ionone and cinnamaldehyde can elicit strong electrophysiological responses in H.oblita antennae [1]. Also, a previous study showed cinnamaldehyde displays a great attraction to H.oblita [55]. β-ionone and α-ionone are general plant volatiles [56] that are found in H.oblita’s host plant, potato [2], while cinnamaldehyde is a main volatile from H.oblita’s host plant, castor [2], [55]. These three chemicals showed the best affinities to HoblCSPs, illustrating the underlying mechanism via which H.oblita recognizes their host plants. These chemicals would be the best candidates to distract H.oblita from recognizing their host plants, which in turn can be applied to successfully control them in the field.

Also, HoblCSP1 and HoblCSP2 exhibited medium affinities to trans-2-hexenal, geraniol, myrcene, and benzaldehyde, and limited affinities to sex pheromones such as L-proline methyl ester, Glycine ethyl ester, and L-isoleucine methyl ester. Both proteins bound to very few alkanes, alcohols, and aldehydes.

These data demonstrated the discriminatory power of insect olfactory systems, with the ability to distinguish different isomers of the same compound [45]. Particularly, HoblCSP1 and HoblCSP2 preferred benzene rings in a ligand structure (Figure 3B). These preferential binding affinities of HoblCSP1 and HoblCSP2 indicated that they play important roles in odorant binding beyond simply the sex pheromone response, although HoblCSP1 and HoblCSP2 are also found in locations other than the antennae (unpublished data).

However, HoblCSP1 and HoblCSP2 reserved unique binding affinities to other compounds. HoblCSP1 displayed higher affinities to aromatic compounds, including dimererhyl phthalate, eugenol, and methyl salicylate, whereas HoblCSP2 showed higher affinities to green leaf volatiles, such as cis-3-hexen-1-ol, cis-3-hexen-1-ol, and 6-methyl-5-hepten-2-one. Since the binding affinities of HoblCSP1 and HoblCSP2 were tested under the same condition, it is therefore possible that they may display different binding activities due to their different in vivo environments. Interestingly, the homologous CSPs in different species could also have divergent affinities. In our experiments, the best ligand of HoblCSP1 is β-ionone, whereas the best ligands for its homologous AgamCSP3 are 2-pentylcinnamaldehyde, retinal, citronellal, and nonanal [57]. This could be due to the adaptation of insect olfactory systems to the specific odorant of their hosts [58][61].

H. oblita antennae are sexually dimorphic. The sensilla placodea and sensilla basiconica are the most common sensilla in the antennae of beetles [62]. The numbers of sensilla placodea and sensilla basiconica are approximately equal in females, whereas in males there are significantly more (9 times) sensilla basiconica than sensilla placodea. However, their functions remain unknown. The sensilla placodea, rather than sensilla basiconica, was proposed as the organ responsible for responding to sex pheromones in Popillia japonica and Anomala osakana [63]. In Anomala corpulenta, the sensilla diverged for different functions, with the umbilicate sensilla placodea responding to green leaf volatiles and the hidden sensilla placodea primarily interacting with sex pheromones [64]. In H. oblita, scientists have proposed that the sensilla basiconica is the sensillum used for sex pheromone responses [65] because females synthesize sex pheromones and males respond to them, most likely through CSPs in the antennae [66], [67]. Here, we found that HoblCSP1 was highly concentrated in the sensilla basiconica, and HoblCSP2 was densely localized in the sensilla placodea. This result indicated that, due to its higher affinity to odorants, HoblCSP1 may confer sensilla basiconica the ability to respond to sex pheromones. Additionally, the localizations of HoblCSPs excluded the possibility that they function as homo or heterodimers, which is consistent with a previously published fluorescence competition assay [48].

CSPs closely relate to insect behavioral plasticity. Our work here discovered that HoblCSP1 and HoblCSP2 have specialized characteristics and unique localization patterns. These will assist in devising strategies to disrupt the aggregation of H. oblita. Also, 924 OBPs and 300 CSPs had been identified (UniProt) to date [68], especially up to 20 CSPs in a single species [54]. The ligand binding overlap between different CSPs, and between CSPs and OBPs, point to intriguing questions regarding the evolution of insect olfactory systems and the underlying mechanisms of olfactory recognition.

Acknowledgments

We very appreciate that the TEM technical assistance by Prof. Hongjing Hao. We thank anonymous editors of American Journal Experts for providing language editing help. We also appreciate the grammar errors checking by Dr. Damian Caudill, from the Creative English Writing Program, English Department, and the University of Miami.

Author Contributions

Conceived and designed the experiments: JY YC JX KL. Performed the experiments: HS LG HF. Analyzed the data: HS LG HF JY. Contributed reagents/materials/analysis tools: JY YC JX KL. Contributed to the writing of the manuscript: HS HF KL.

References

  1. 1. Deng SS, Yin J, Zhong T, Cao YZ, Li KB (2012) Function and Immunocytochemical Localization of Two Novel Odorant-Binding Proteins in Olfactory Sensilla of the Scarab Beetle Holotrichia oblita Faldermann (Coleoptera: Scarabaeidae). Chem Senses 37: 141–150.
  2. 2. Zhang YL Yuan YH, Yuan GH, Guo XR, Luo MH (2006) A study on the attraction of Holotrichia oblita (Fadermann) to castor leaves. Journal of Henan Agricultural University 40: 53–57.
  3. 3. Field LM, Pickett JA, Wadhams LJ (2000) Molecular studies in insect olfaction. Insect Mol Biol 9: 545–551.
  4. 4. Zubkov S, Gronenborn AM, Byeon IJ, Mohanty S (2005) Structural consequences of the pH-induced conformational switch in Antheraea polyphemus pheromone-binding protein: mechanisms of ligand release. J Mol Biol 354: 1081–1090.
  5. 5. Jiang QY, Wang WX, Zhang Z, Zhang L (2009) Binding specificity of locust odorant binding protein and its key binding site for initial recognition of alcohols. Insect Biochem Mol Biol 39: 440–447.
  6. 6. Broekaert WF, Lambrechts D, Verbelen JP, Peumans WJ (1988) Datura stramonium Agglutinin: Location in the Seed and Release upon Imbibition. Plant Physiol 86: 569–574.
  7. 7. Broekaert WF, Vanparijs J, Allen AK, Peumans WJ (1988) Comparison of Some Molecular, Enzymatic and Antifungal Properties of Chitinases from Thorn-Apple, Tobacco and Wheat. Physiol Mol Plant Path 33: 319–331.
  8. 8. Vosshall LB, Wong AM, Axel R (2000) An olfactory sensory map in the fly brain. Cell 102: 147–159.
  9. 9. Vogt RG (2003) Biochemical diversity of odor detection. OBPs, ODEs, SNMPs. San Diego (CA): Elsevier Academic Press: 392–445.
  10. 10. Robertson HM, Martos R, Sears CR, Todres EZ, Walden KK, et al. (1999) Diversity of odourant binding proteins revealed by an expressed sequence tag project on male Manduca sexta moth antennae. Insect Mol Biol 8: 501–518.
  11. 11. McKenna MP, Hekmat-Scafe DS, Gaines P, Carlson JR (1994) Putative Drosophila pheromone-binding proteins expressed in a subregion of the olfactory system. J Biol Chem 269: 16340–16347.
  12. 12. Pikielny CW, Hasan G, Rouyer F, Rosbash M (1994) Members of a family of Drosophila putative odorant-binding proteins are expressed in different subsets of olfactory hairs. Neuron 12: 35–49.
  13. 13. Angeli S, Ceron F, Scaloni A, Monti M, Monteforti G, et al. (1999) Purification, structural characterization, cloning and immunocytochemical localization of chemoreception proteins from Schistocerca gregaria. Eur J Biochem 262: 745–754.
  14. 14. Pelosi P, Zhou JJ, Ban LP, Calvello M (2006) Soluble proteins in insect chemical communication. Cell Mol Life Sci 63: 1658–1676.
  15. 15. Wanner KW, Willis LG, Theilmann DA, Isman MB, Feng Q, et al. (2004) Analysis of the insect os-d-like gene family. J Chem Ecol 30: 889–911.
  16. 16. Foret S, Wanner KW, Maleszka R (2007) Chemosensory proteins in the honey bee: Insights from the annotated genome, comparative analyses and expressional profiling. Insect Biochem Mol Biol 37: 19–28.
  17. 17. Picimbon JF, Dietrich K, Angeli S, Scaloni A, Krieger J, et al. (2000) Purification and molecular cloning of chemosensory proteins from Bombyx mori. Arch Insect Biochem Physiol 44: 120–129.
  18. 18. Picimbon JF, Dietrich K, Breer H, Krieger J (2000) Chemosensory proteins of Locusta migratoria (Orthoptera: Acrididae). Insect Biochem Mol Biol 30: 233–241.
  19. 19. Lartigue A, Campanacci V, Roussel A, Larsson AM, Jones TA, et al. (2002) X-ray structure and ligand binding study of a moth chemosensory protein. J Biol Chem 277: 32094–32098.
  20. 20. Briand L, Swasdipan N, Nespoulous C, Bezirard V, Blon F, et al. (2002) Characterization of a chemosensory protein (ASP3c) from honeybee (Apis mellifera L.) as a brood pheromone carrier. Eur J Biochem 269: 4586–4596.
  21. 21. Sandler BH, Nikonova L, Leal WS, Clardy J (2000) Sexual attraction in the silkworm moth: structure of the pheromone-binding-protein-bombykol complex. Chem Biol 7: 143–151.
  22. 22. Ban LP, Scaloni A, Brandazza A, Angeli S, Zhang L, et al. (2003) Chemosensory proteins of Locusta migratoria. Insect Mol Biol 12: 125–134.
  23. 23. Campanacci V, Mosbah A, Bornet O, Wechselberger R, Jacquin-Joly E, et al. (2001) Chemosensory protein from the moth Mamestra brassicae. Expression and secondary structure from 1H and 15N NMR. Eur J Biochem 268: 4731–4739.
  24. 24. Tomaselli S, Crescenzi O, Sanfelice D, Ab E, Wechselberger R, et al. (2006) Solution structure of a chemosensory protein from the desert locust Schistocerca gregaria. Biochemistry 45: 10606–10613.
  25. 25. Jansen S, Chmelik J, Zidek L, Padrta P, Novak P, et al. (2007) Structure of Bombyx mori chemosensory protein 1 in solution. Arch Insect Biochem Physiol 66: 135–145.
  26. 26. Jacquin-Joly E, Francois MC, Burnet M, Lucas P, Bourrat F, et al. (2002) Expression pattern in the antennae of a newly isolated lepidopteran Gq protein alpha subunit cDNA. Eur J Biochem 269: 2133–2142.
  27. 27. Gonzalez D, Zhao Q, McMahan C, Velasquez D, Haskins WE, et al. (2009) The major antennal chemosensory protein of red imported fire ant workers. Insect Mol Biol 18: 395–404.
  28. 28. Nagnan-Le Meillour P, Cain AH, Jacquin-Joly E, Francois MC, Ramachandran S, et al. (2000) Chemosensory proteins from the proboscis of Mamestra brassicae. Chem Senses 25: 541–553.
  29. 29. Maleszka R, Stange G (1997) Molecular cloning, by a novel approach, of a cDNA encoding a putative olfactory protein in the labial palps of the moth Cactoblastis cactorum. Gene 202: 39–43.
  30. 30. Jin X, Brandazza A, Navarrini A, Ban L, Zhang S, et al. (2005) Expression and immunolocalisation of odorant-binding and chemosensory proteins in locusts. Cell Mol Life Sci 62: 1156–1166.
  31. 31. Ozaki K, Utoguchi A, Yamada A, Yoshikawa H (2008) Identification and genomic structure of chemosensory proteins (CSP) and odorant binding proteins (OBP) genes expressed in foreleg tarsi of the swallowtail butterfly Papilio xuthus. Insect Biochem Mol Biol 38: 969–976.
  32. 32. Li H, Zhang L, Ni C, Shang H, Zhuang S, et al. (2013) Molecular recognition of floral volatile with two olfactory related proteins in the Eastern honeybee (Apis cerana). Int J Biol Macromol 56: 114–121.
  33. 33. Nomura A, Kawasaki K, Kubo T, Natori S (1992) Purification and Localization of P10, a Novel Protein That Increases in Nymphal Regenerating Legs of Periplaneta-Americana (American Cockroach). Int J Dev Biol 36: 391–398.
  34. 34. Jacquin-Joly E, Vogt RG, Francois MC, Nagnan-Le Meillour P (2001) Functional and expression pattern analysis of chemosensory proteins expressed in antennae and pheromonal gland of Mamestra brassicae. Chem Senses 26: 833–844.
  35. 35. Liu JX, Zhong GH, Xie JJ, Guan S, Hu MY (2005) Recent advances in chemosensory proteins of insects. Acta Entomologica Sinica 48(3): 418–426.
  36. 36. Monteforti G, Angeli S, Petacchi R, Minnocci A (2002) Ultrastructural characterization of antennal sensilla and immunocytochemical localization of a chemosensory protein in Carausius morosus Brunner (Phasmida: Phasmatidae). Arthropod Struct Dev 30: 195–205.
  37. 37. Qiao HL, Deng PY, Li DD, Chen M, Jiao ZJ, et al. (2013) Expression analysis and binding experiments of chemosensory proteins indicate multiple roles in Bombyx mori. J Insect Physiol 59: 667–675.
  38. 38. Danty E, Briand L, Michard-Vanhee C, Perez V, Arnold G, et al. (1999) Cloning and expression of a queen pheromone-binding protein in the honeybee: an olfactory-specific, developmentally regulated protein. J Neurosci 19: 7468–7475.
  39. 39. Sanchez-Gracia A, Vieira FG, Rozas J (2009) Molecular evolution of the major chemosensory gene families in insects. Heredity (Edinb) 103: 208–216.
  40. 40. Bendtsen JD, Nielsen H, von Heijne G, Brunak S (2004) Improved prediction of signal peptides: SignalP 3.0. J Mol Biol. 340(4): 783–795.
  41. 41. Jones DT, Taylor WR, Thornton JM (1992) The rapid generation of mutation data matrices from protein sequences. Comput Appl Biosci 8(3): 275–282.
  42. 42. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evo 30: 2725–2729.
  43. 43. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, et al. (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23: 2947–2948.
  44. 44. Capella-Gutierrez S, Silla-Martinez JM, Gabaldon T (2009) trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25: 1972–1973.
  45. 45. Yin J, Feng H, Sun H, Xi J, Cao Y, et al. (2012) Functional analysis of general odorant binding protein 2 from the meadow moth, Loxostege sticticalis L. (Lepidoptera: Pyralidae). PLoS One 7(3): e33589
  46. 46. Campanacci V, Krieger J, Bette S, Sturgis JN, Lartigue A, et al. (2001) Revisiting the specificity of Mamestra brassicae and Antheraea polyphemus pheromone-binding proteins with a fluorescence binding assay. J Biol Chem 276: 20078–20084.
  47. 47. Zhang Y, Lu H, Bargmann CI (2005) Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis elegans. Nature 438: 179–184.
  48. 48. Wang B, Guan L, Zhong T, Li KB, Yin J, et al. (2013) Potential Cooperations between Odorant-Binding Proteins of the Scarab Beetle Holotrichia oblita Faldermann (Coleoptera: Scarabaeidae). PLoS One 8(12): e84795
  49. 49. Sun F, Lu JH, Li L, Zhao K (2008) Analysis of volatile component of ulmus pumila by solid phase micro extraction coupled with GC-MS. Journal of Northeast Forestry University 36: 10.3969/j.issn.1000–5382.2008.3905.3021.
  50. 50. Li H, Zhang G, Wang MQ (2012) Chemosensory protein genes of batocera horsfieldi (hope): identification and expression pattern. J Appl Entomol 136: 781–792.
  51. 51. Xu YL, He P, Zhang L, Fang SQ, Dong SL, et al. (2009) Large-scale identification of odorant-binding proteins and chemosensory proteins from expressed sequence tags in insects. BMC Genomics 10: 632.
  52. 52. Wang J, Li DZ, Min SF, Mi F, Zhou SS, et al. (2014) Analysis of chemosensory gene families in the beetle Monochamus alternatus and its parasitoid Dastarcus helophoroides. Comp Biochem Physiol, D: Genomics and Proteomics (11): 1–8.
  53. 53. Andersson MN, Grosse-Wilde E, Keeling CI, Bengtsson JM, Yuen MM, et al. (2013) Antennal transcriptome analysis of the chemosensory gene families in the tree killing bark beetles, Ips typographus and Dendroctonus ponderosae (Coleoptera: Curculionidae: Scolytinae). BMC Genomics 14: 198.
  54. 54. Tribolium Genome Sequencing Consortium (2008) The genome of the model beetle and pest Tribolium castaneum. Nature 452: 949–955.
  55. 55. Li WZ, Yang L, Shen XW, Yuan YH, Yuan GH, et al. (2013) Electroantennographic and behavioural responses of scarab beetles to Ricinus communis leaf volatiles. Acta Ecologica Sinica 33: 6895–6903.
  56. 56. Das A, Lee S-h, Hyun TK, Kim S-w, Kim J-y (2013) Plant volatiles as method of communication. Plant Biotechnol Rep 7: 9–26.
  57. 57. Iovinella I, Bozza F, Caputo B, Della Torre A, Pelosi P (2013) Ligand-binding study of Anopheles gambiae chemosensory proteins. Chem Senses 38: 409–419.
  58. 58. Visser JH (1983) Differential Sensory Perceptions of Plant-Compounds by Insects. Acs Symposium Series 208: 215–230.
  59. 59. Tingle FC, Mitchell ER (1991) Effect of Oviposition Deterrents from Elderberry on Behavioral-Responses by Heliothis-Virescens to Host-Plant Volatiles in Flight Tunnel. J Chem Ecol 17: 1621–1631.
  60. 60. Du YJ, Yan FS (1994) The role of plant volatiles in tritrophic interactions amongphytophagous insects, their host plants and natural enemies. Acta Entomology Sinica 31: 233–249.
  61. 61. Yan S, Zhang D, Chi D (2003) Advances of studies on the effects of plant volatiles on insect behavior. J Applied Ecol 14: 310–313.
  62. 62. Kim JY, Leal WS (2000) Ultrastructure of pheromone-detecting sensillum placodeum of the Japanese beetle, Popillia japonica Newmann (Coleoptera: Scarabaeidae). Arthropod Struct Dev 29: 121–128.
  63. 63. Wojtasek H, Hansson BS, Leal WS (1998) Attracted or repelled? A matter of two neurons, one pheromone binding protein, and a chiral center. Biochem Biophys Res Commun 250: 217–222.
  64. 64. Larsson MC, Leal WS, Hansson BS (2001) Olfactory receptor neurons detecting plant odours and male volatiles in Anomala cuprea beetles (Coleoptera : Scarabaeidae). J Insect Physiol 47: 1065–1076.
  65. 65. Wang GL (2006) Reaction to greenery odors and the ultrastructure of olfactory sensilla in Holotrichia oblita Faldermann. [MSc Thesis] [Harbin (China)]: Northeast Forestry University.
  66. 66. Luo SL, Zhuge PP, Wang MQ (2011) Mating behavior and contact pheromones of Batocera horsfieldi (Hope) (Coleoptera: Cerambycidae). Entomological Science 14: 359–363.
  67. 67. Harari A, Ben-Yakir D, Rosen D (1994) Mechanism of aggregation behavior in Maladera matrida Argaman (Coleoptera: Scarabaeidae). J Chem Ecol 20: 361–371.
  68. 68. The UniProt Consortium (2014) Activities at the Universal Protein Resource (UniProt). Nucl. Acids Res. 42(D1): D191–D198
  69. 69. Gong DP, Zhang HJ, Zhao P, Lin Y, Xia QY, et al. (2007) Identification and expression pattern of the chemosensory protein gene family in the silkworm, Bombyx mori. Insect Biochem Mol Biol 37: 266–277.
  70. 70. Picimbon JF, Dietrich K, Krieger J, Breer H (2001) Identity and expression pattern of chemosensory proteins in Heliothis virescens (Lepidoptera, Noctuidae). Insect Biochem Mol Biol 31: 1173–1181.
  71. 71. Zhou JJ, Vieira FG, He XL, Smadja C, Liu R, et al. (2010) Genome annotation and comparative analyses of the odorant-binding proteins and chemosensory proteins in the pea aphid Acyrthosiphon pisum. Insect Mol Biol 19 Suppl 2113–122.