Skip to main content
Advertisement
Browse Subject Areas
?

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

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Identification of Genes Relevant to Pesticides and Biology from Global Transcriptome Data of Monochamus alternatus Hope (Coleoptera: Cerambycidae) Larvae

  • Songqing Wu ,

    Contributed equally to this work with: Songqing Wu, Xiaoli Zhu, Zhaoxia Liu

    Affiliations College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Xiaoli Zhu ,

    Contributed equally to this work with: Songqing Wu, Xiaoli Zhu, Zhaoxia Liu

    Affiliations Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Zhaoxia Liu ,

    Contributed equally to this work with: Songqing Wu, Xiaoli Zhu, Zhaoxia Liu

    Affiliations Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Ensi Shao,

    Affiliations Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Carballar-Lejarazú Rebeca,

    Affiliation Department of Molecular Biology and Biochemistry, University of California Irvine, Irvine, CA, 92697, United States of America

  • Yajie Guo,

    Affiliation College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Yueting Xiong,

    Affiliations College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Yani Mou,

    Affiliations College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Runxue Xu,

    Affiliation Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Xia Hu,

    Affiliation College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Guanghong Liang,

    Affiliation College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Shuangquan Zou,

    Affiliation College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Xiong Guan ,

    fpzhang1@163.com (FPZ); guanxfafu@126.com (XG)

    Affiliations Fujian-Taiwan Joint Center for Ecological Control of Crop Pests, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China, Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

  • Feiping Zhang

    fpzhang1@163.com (FPZ); guanxfafu@126.com (XG)

    Affiliation College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, People’s Republic of China

Abstract

Monochamus alternatus Hope is the main vector in China of the Pine Wilt Disease caused by the pine wood nematode Bursaphelenchus xylophilus. Although chemical control is traditionally used to prevent pine wilt disease, new strategies based in biological control are promising ways for the management of the disease. However, there is no deep sequence analysis of Monochamus alternatus Hope that describes the transcriptome and no information is available about gene function of this insect vector. We used next generation sequencing technology to sequence the whole fourth instar larva transcriptome of Monochamus alternatus Hope and successfully built a Monochamus alternatus Hope transcriptome database. In total, 105,612 unigenes were assigned for Gene Ontology (GO) terms, information for 16,730 classified unigenes was obtained in the Clusters of Orthologous Groups (COGs) database, and 13,024 unigenes matched with 224 predicted pathways in the Kyoto Encyclopedia of Genes and Genome (KEGG). In addition, genes related to putative insecticide resistance-related genes, RNAi, the Bt receptor, intestinal digestive enzymes, possible future insect control targets and immune-related molecules are described. This study provides valuable basic information that can be used as a gateway to develop new molecular tools for Monochamus alternatus Hope control strategies.

Introduction

Pine Wilt Disease is a devastating disease in pine trees caused by the infection of Bursaphelenchus xylophilus and it is commonly known as the cancer of pine trees [1]. Since the discovery of B. xylophilus in Japanese black pines in the Sun Yat-sen Mausoleum in Nanjing City (Jiangsu Province in China) in 1982, Pine Wilt Disease has occurred in a total of 275 county-level administrative regions (excluding Hong Kong and Taiwan) of 17 provinces (autonomous regions and municipalities), causing immense damage to forest resources and having impact in China’s ecological environment [2]. In China, the principal vector for Pine Wilt Disease is the beetle Monochamus alternatus Hope (M. alternatus) larvae as carriers; after emergence, t larvae use pine trees as food and oviposition sources, therefore they are considered the invasive stage of the insect. Effective control of M. alternatus plays an important role in the prophylaxis and treatment of this disease [3].

At present, the principal strategies to control M. alternatus include: trap trees, biological control, silvicultural control and chemical prevention [3, 4]. Among these, the biological control presents unique advantages: (1) It is difficult for pests to become resistant as microorganisms have adapted to the immune systems of insects during the process of concurrent evolution. Therefore insect immunity to pathogenic microorganisms has been kept at a low level; (2) It has high environmental security; microorganisms typically have strong specificity for their targets and they are harmless to vertebrates, which protects the natural predators of those hosts; (3) insects are an ideal medium for various types of pathogens; the proliferation of insect pathogens in vivo can be spread by diseases and pests or the insect’s body; (4) It’s easy to obtain strains that are strongly pathogenic using genetic engineering and transformation techniques [5, 6]. Current biological control techniques for M. alternatus have progressed, including the spreading of effective natural enemies, creation of black lights and trap-trees. Application of the above techniques has successfully controlled Pine Wilt Disease at test locations [5]. Among the methods of natural enemies are the parasitoid beetles Dastarcus helophoroides and Sclerodermus spp. (Hymenoptera: Bethylidae), showing up to 90% of effectiveness [7, 8]. In general, biological control has brought new insights to control M. alternatus forest infestations. Therefore, microbial control of M. alternatus has increasingly gained attention [4]. However, there is currently a lack of knowledge regarding Monochamus alternatus Hope transcripts, gene expression and gene function in this insect vector.

We used next generation sequencing technology to sequence the whole fourth instar larvae transcriptome of M. alternatus and successfully built a M. alternatus Hope transcriptome database. In addition, our data describe genes related to putative insecticide resistance, intestinal digestive enzymes, possible future insect control targets and immune-related molecules. This study provides valuable basic information that can be used as a key point to develop new molecular tools for M. alternatus Hope control strategies.

Results and Discussion

Sequencing and de novo assembly

Illumina sequencing produced 46, 761 and 743 clean reads for the larvae samples, which is equivalent to an accumulated length of 11, 777, 133 and 130 bp (Table 1).

thumbnail
Table 1. Sequence statistics of the Illumina sequencing assembly.

https://doi.org/10.1371/journal.pone.0147855.t001

The average of raw reads length was 252 bp. First, the sequencing reads were broken into K-mers using Trinity software [9]; the small fragments were assembled into 11,433,166 contigs with a mean length of 41.63 bp. The length of the contigs mainly ranged from 0 to 3000 bp, representing 99.61% of the reads, although some contigs were longer than 3000 bp. Size distribution of the contigs is shown in S1 Fig. Finally, using the method of De Bruijn graphing and sequencing read information, we identified 107,259 transcripts with a mean length of 941.77 bp, transcripts ranged in length from ~200–3000 bp, identifying 49,615 transcripts with a length larger than 500 bp. We obtained 73,090 unigenes with a mean length of 693.24 bp. The lengths of 25,718 and 13,668 unigenes were larger than 500 bp and 1000 bp, respectively, while 64.82% of the unigenes had lengths between 0 to 500 bp (Fig 1). This result indicates that the length distribution of the transcripts and unigenes were represented in majority by short sequences with relatively little redundancy, which is similar to transcriptome analysis reported for other insect species using the same technology [1014]. Importantly the longer sequences contribute only for 7.09% of the M. alternatus transcriptome, the majority of transcripts and unigenes were still less than 500 bp after assembly; this is probably due to the capacity of shorter sequences and low coverage of the transcriptome [5, 15]. A large number of assembled sequential data could provide a more deeply transcriptome information for future research, allowing rapid characterization for most of the transcripts and a reference for the genes of interest [15].

thumbnail
Fig 1. Unigenes length distribution.

The y-axis number has been converted into logarithmic scale.

https://doi.org/10.1371/journal.pone.0147855.g001

Annotation of predicted proteins

All assembled unigenes were used as an input for NR, Swiss-Prot, Gene Oontology (GO), Clusters of Orthologous Groups (COG), KOG and KEGG databases. BLAST and HMMER parameter E-values were set at 10−5 and 10−10 respectively, we were able to obtain annotated information for 36,828 unigenes, representing 50.38% of the unigenes. The rest of the unigene sequences (49.62%) had no significant matches in the existing databases. Unigenes comparison with the NR database produced 34,702 hits, the distribution of E-values demonstrated that 26.91% of the mapped sequences have strong homology (smaller than 1.0E-49) with an annotated sequence, and 62.70% of the homolog sequences ranged from 1.0E-5 to 1.0E-49 (Fig 2A). Based on the best species match, we found that M. alternatus sequences have 30.59% and 8.89% matches with sequences from the Tribolium castaneum and Dendroctonus ponderosae, both belonging to Coleopteran order, while only <6% matched to other insects (Fig 2B). Therefore, M. alternatus have the closest evolutionary distance with T. castaneum.

thumbnail
Fig 2. Characteristics of the homology search of Illumina sequences against the NR database.

(A) E-value distribution of the BLAST hits for each unique sequence with a cut-off E-value of 1.0E-5. (B) Species distribution of the BLASTX results. The first hit of each sequence was used for further in silico analysis.

https://doi.org/10.1371/journal.pone.0147855.g002

GO assignments

GO database is an internationally standardized gene function classification system, which provides a suitably updated standard vocabulary to describe the functional attributes of genes and gene products in an organism [16]. Transcript sequences were used as an input in the GO database; using BlastX a total of 105,612 unigenes were assigned to GO terms (Fig 3, S1 Table). According to the standard GO terms and 60 subcategories, differential gene expression and all unigenes from M. alternatus larva were statistically classified into three main GO categories: biological process, cellular component, and molecular function. Biological process represented the majority of GO annotations (52,883, 50.07%), followed by cellular component (29,970, 28.38%) and molecular function (22,759, 21.55%). The metabolic process (20.48%) and cellular process (18.55%) were predominant within the biological process category, indicating that the analyzed tissue has a high degree of metabolic activity; with the following subcategories: single-organism process (15.34%), biological regulation (8.05%), developmental process (5.59%), response to stimulus (5.49%), localization (5.38%), multicellular organismal process (5.12%), cellular component organization or biogenesis (4.26%) and signaling (3.16%). Biological processes contain most major cellular processes, from transportation and cell formation to transcription, translation and supersession. According to the classification of the cellular components, cell part (21.55%), cell (21.39%), and organelle (14.85%) are the most representative subcategories. Most of the annotated unigenes from the cellular component category, correspond to plastids and mitochondria. We also identified genes involved in the synthesis of secondary metabolites, and were grouped into catalytic activity (41.61%), binding (39.45%), transporter activity (5.63%), structural molecule activity (2.86%), molecular transducer activity (2.76%), receptor activity (2.43%) and nucleic acid binding transcription factor activity (2.02%), etc. A previous study reported similar classifications related to metabolic processes for Tomicus yunnanensis transcriptome [15]. GO annotations describe the contour features of the overall gene expression of M. alternatus, and revealed expressed genes encoding diverse structural, stress and regulatory proteins.

thumbnail
Fig 3. Distribution of second level GO of Monochamus alternatus Hope transcriptome.

Distribution of GO categories assigned to the Monochamus alternatus Hope transcriptome. Unigenes were annotated in three categories: cellular components, molecular functions, and biological process. Right y-axis indicates the number of genes in a category; left y-axis indicates the genes percentage in a specific.

https://doi.org/10.1371/journal.pone.0147855.g003

COG classification

In total, information for 16,730 classified unigenes was obtained in the COG database (Fig 4). COG classifications were divided into 25 functional categories. Among these categories, general function prediction (19.90%) was the largest group, followed by translation, ribosomal structure and biogenesis (10.19%), posttranslational modification, protein turnover, chaperones (8.29%) and transcription (7.57%). Secondary metabolites biosynthesis, transport and catabolism represented 2.71%, given the relative importance of secondary metabolic activity for insect resistance. To some extent, COG classifications further reveal the potential specific reactions and the functional participation in molecular processes for genes expressed in M. alternatus.

thumbnail
Fig 4. COG function classification of Monochamus alternatus Hope transcriptome.

All putative proteins were analyzed using the COG database. COG classifications were divided into 25 functional categories; 16,730 classified unigenes were assigned to 25 COG classifications.

https://doi.org/10.1371/journal.pone.0147855.g004

KEGG analysis

The KEGG Pathway database, is a collection of graphical maps representing different cellular processes, to systematically analyze metabolic pathways and functions of gene products in a cell [17]. To identify the represented biological pathways in M. alternatus, 34,302 annotated unigenes were analyzed using KEGG database. The results indicated that 13,024 unigenes matched with 224 KEGG pathways. These pathways are summarized in S2 Table. The top 10 pathways were Protein processing in endoplasmic reticulum (503 members), Ribosome (486 members), Spliceosome (444 members), RNA transport (415 members), Purine metabolism (410 members), Pyrimidine metabolism (321 members), Ribosome biogenesis in eukaryotes (312 members), Endocytosis (305 members), Oxidative phosphorylation (302 members) and Ubiquitin mediated proteolysis (288 members). These annotations provide powerful information for our research related specific biological processes and pathways of M. alternatus.

Putative insecticide resistance-related genes

1. Cytochrome P450 (P450).

Research over the past 10 years has provided clear evidence that in insects, P450 is involved in a number of physiological functions, such as hormone metabolism [1820], adaptability of parasitic plants [21, 22] and resistance to insecticides [23, 24]. Approximately 222 types of P450-related unigenes were identified in M. alternatus transcriptome (S3 Table). We identified 103 P450-related sequences with a length bigger than 600 bp (46.40% of the global pests in the database); approximately half of the unigene sequences were long sequences (600 bp). In addition, the length of 30 P450-related sequences was larger than 1800 bp (13.51%) (Table 2). P450 genes identified from the M. alternatus transcriptome was comparable in number to those from T. yunnanensis transcriptome, although the final number of genes still needs to be verified via gene cloning [15]. Ai Junwen et al. classified the P450 enzyme system into four large families: CYP2, CYP3, CYP4 and mitochondrial P450, based on the phylogenetic analysis of Drosophila and silkworm P450 gene homology [25].

thumbnail
Table 2. Putative P450 genes identified in Monochamus alternatus Hope.

https://doi.org/10.1371/journal.pone.0147855.t002

Previous studies have reported that 17 P450 genes from CYP3 and CYP4 families are associated with plant toxins and pesticide resistance [25,26]. Based in phylogenic analyses; we found evidence that CYP1, CYP3, CYP4 and mitochondrial families are present in M. alternatus transcriptome. We also identified CYP1A1, which belongs to the CYP1 family, and CYP3A4, CYP3A5 and CYP3A7, which belong to the CYP3 family. Interestingly, the P450 genes identified in M. alternatus differ from those reported in T. castaneum and other insect systems.

Recent studies have shown that P450 has an increased expression in insecticide resistant insects; moreover, there is also evidence for P450 gene duplication and amplification in four types of insects in vivo [27]. However, there is no experimental data to support the importance and biological role of P450 complex related to insecticide resistance in M. alternatus.

2. Glutathione S-transferase (GST).

GSTs specifically catalyze glutathione thiol and interact with other electrophilic groups [28, 29]; GST is one of the main detoxification enzymes in insecticide metabolism. GST high levels of expression are related to insecticide resistance mechanisms in insects [27]. We identified 96 GST unigenes in M. alternatus transcriptome. The length of 19 GST-related sequences was greater than 1000 bp (19.79%) (Table 3). GSTs are divided into three major categories according to their cellular location: cytosolic, microsomal, and mitochondrial [30]. Cytosolic matrix GSTs in insects are further divided into at least six classes (delta, epsilon, omega, sigma, theta, and zeta) based on sequence homology of the N-terminus, substrate specificity, immunoreactivity, and sensitivity to different inhibitors [3133]. Delta and epsilon classes are unique in insects [34]. We found delta, omega, and theta types of GST in M. alternatus transcriptome, however we could not identify any gene from the epsilon class. Previous studies have reported the identification of sigma, delta and theta GST classes in Nasonia vitripennis; epsilon, sigma, omega, and delta in T. castaneum; and one delta unigene in T. yunnanensis [35, 36]. The GST genes identified in M. alternatus transcriptome can contribute to a greater understanding of the relationship between GSTs and insecticide resistance in insects.

thumbnail
Table 3. Putative identified GST genes in Monochamus alternatus Hope.

https://doi.org/10.1371/journal.pone.0147855.t003

Insecticide receptors and resistance-related genes

In addition to the identified genes described above, we identified 384 unigenes that represent potential pesticide receptors and insecticide resistance-related genes; including FigCys-loop ligand-gated ion channel (Cys-loop LGIC), carboxylesterase, superoxide dismutase, acetyl-CoA carboxylase, acetylcholinesterase, c-aminobutyric, acid (GABA) receptors, nicotinic acetylcholine receptors, sodium channels, chloride channels and ryanodine receptors (Fig 5). Further study of these genes could uncover potential insecticide receptors and provide the bases to test whether these genes play functional roles in insecticide resistance.

thumbnail
Fig 5. Number of unigenes related to pesticide receptors and resistance-related genes.

x-axis indicates the number of unigenes, y-axis indicates the specific unigenes related to pesticide receptors and resistance-related genes. Seven hundred and fifty-nine insecticide receptors and resistance-related unigenes were identified in M. alternatus transcriptome.

https://doi.org/10.1371/journal.pone.0147855.g005

RNA interference-genes

RNA interference (RNAi) is an important pathway that is used in many organisims to regulate gene expression [37]. RNAi pathways have been indentified in Drosophila melanogaster [3840], T. castaneum [4146] and Bombyx mori [4749].

Forty-two unigenes related to RNAi pathways were identified (Fig 6); we identified two SID-1, 34 scavenger receptors, Figone RNA-dependent RNA polymerase, and five vacuolar H+ ATPase unigenes in M. alternatus transcriptome. These components have previously been reported as part of the RNAi uptake mechanisms in insects [50]. Interestingly, we did not identify any RSD-3 unigenes in the transcriptome, although its participation in RNAi pathways. Among the identified genes, 26 unigenes were larger than 600 bp (61.90%) and 23 were more than 1 kb (54.76%). Moreover, two SID-1-related unigenes were both represented by longer sequences in the transcriptome, at c38753.graph_c0 (2731.42 bp) and c41338.graph_c0 (3066.51 bp), respectively.

thumbnail
Fig 6. Number of unigenes related to RNAi.

x-axis indicates the corresponding number of unigenes and y-axis indicates the specific unigenes related to RNAi. Forty-two unigenes coding for RNAi identified in the database, 26 of them were larger than 600 bp (61.90%) and 23 were more than 1 kb (54.76%).

https://doi.org/10.1371/journal.pone.0147855.g006

SID-1 is a multispan transmembrane protein that is crucial for systemic RNAi pathways in Caenorhabditis elegans, it delivers dsRNAs to cells [50]. However, SID-1 has not been described in D. melanogaster, indicating that the existence or absence of sid-1 gene could play an important part in determining whether systemic RNAi is present in the biome [51]. Although a sid-1 ortholog has also been found in the cotton aphid Aphis gossypii [52], further research is still needed to provide evidence for the molecular basis of systemic RNAi in M. alternatus.

Scavenger receptors can recognize extensive polyanionic ligands, and they play key roles mediating phagocytosis of pathogens in Drosophila [50]. Among the 34 identified unigenes related to scavenger receptors, nine of them had lengths above 1 kb, while the rest ranged from 100 bp to 1 kb. However, we were not able to identify important relevant sequences homologous sequences to the scavenger receptors of mammals [53, 54], but we were able to identify them by homology to the flesh fly and C. elegans.

Finally, we identified five vacuolar H+ ATPase unigenes and one RNA-dependent RNA polymerase-1 unigene. Previous studies have found that vacuolar H+ ATPase-deficient S2 Drosophila cells accumulate dsRNA in endocytic vesicles and showing an endocytosis-based mechanism as a way to disrupt vacuolar H+ ATPase in the S2 cells to induce RNAi silencing [55].

The study of SID-1, scavenger receptors, vacuolar H+ ATPase, RSD-3 and RNA-dependent RNA polymerase genes can deepen our understanding of the biology of defense against parasitic endogenous nucleic acids and exogenous pathogen nucleic acids and provides a basis for the expression of regulatory protein coding genes [56, 57].

Potential Bacillus thuringensis (Bt) receptors

Receptor molecules related to B. thuringensis Cry toxin mechanisms, have been characterized in the the insect midgut; they have been widely studied, most notably: aminopeptidase (APN), alkaline phosphatase (ALP) and cadherin [58, 59]. We identified a total of 448 Bt receptor unigenes in M. alternatus transcriptome, including: ALP, APN, cadherin and the ABC transporter, which are currently the four described Bt insect receptor molecules (Fig 7).

thumbnail
Fig 7. Number of unigenes related to Bt receptors.

x-axis indicates the corresponding number of unigenes and y-axis indicates the specific unigenes related to Bt receptors. Four hundred and forty-eight Bt receptor unigenes were confirmed in the Monochamus alternatus Hope transcriptome through in silico comparisons.

https://doi.org/10.1371/journal.pone.0147855.g007

APN [6062] and cadherin [63] are considered the most important type of receptors within the putative insect Cry receptors. M. alternatus transcriptome revealed 38 APN unigenes and 116 cadherin unigenes, representing both approximately 8% of the total number of related Bti-receptor unigenes. APNs belong to the zinc-binding metalloprotease/peptidase enzymes and are anchored to the midgut membrane through GPI anchors [64]. Crava et al. clustered APNs into seven classes based on phylogenetic analyses of lepidopteran, Hughes found that in intestinal tissue, APN1 class was the most highly expressed within a complex mix of APN expression data [65, 66]. Cadherins are mainly localized in adherens junctions and are involved in cell adhesion [59]; they have been found anchoring midgut epithelial cells of Manduca sexta and Lymantria dispar larvae. We found that the number of identified cadherin unigenes contributed in nearly 26% of the total number of Bt receptor molecules in M. alternatus, second only to ABC transporters (278, ~62%). It has previously been reported that APN and cadherin-like (CAD-like) midgut proteins in lepidotera can interact with Cry1 toxins. In Diptera, homologous APN, CAD-like, and alkaline phosphatase proteins of mosquitoes are also considered as Cry11 and Cry4 receptor proteins. In the longicorn of coleopterous, a cadherin-like protein acts as Cry3Aa receptor; finally, gene silencing has confirmed that APN and CAD-like proteins are the most representative Cry receptors [67]. In insects, ALPs are a major group of Cry-binding proteins; their roles as receptor molecules have extensively studied in Lepidoptera, Coleoptera and Diptera larvae [6871]. Data suggest that the interaction between Cry1Ac and ALP affects the midgut protease activity during the incubation period in Heliothis virescens and M. sexta larvae [62, 68]. Monochanus alternatus transcriptome revealed 16 ALP unigenes, which contribute to only ~3% of the Bt receptor-related unigenes. Using RNAi silencing of APN and ALP genes in M. sexta larvae, Flores-Escobar et al. found that for Cry1Ab toxicity, binding to ALP was more important than APN, however for Cry1Ac, APN was more important, these results suggest that Cry binding receptors have specific affinity for different Cry toxins [72].

In addition to ANP, ALP and cadherins, M. alternatus transcriptome revealed 278 ABC transporter-related unigenes, contributing for 62.05% of the total number of Bt receptor-related unigenes. The ABC family in insects is related to multi-drug resistance [73]. Recently, ATP-binding cassette transporter subfamily C member 2 (ABCC2) was identified as a Cry1 toxin receptor in B. mori larvae [74], with a consistent role in the mechanism of Cry1 toxin resistance [75, 76].

Intestinal digestive enzymes

Midgut proteases play an important role in activating Cry toxins, producing the toxin’s core 3D structure in the insect midgut [77] and the type and abundance of insect proteases are important for toxin specificity [78]. Changes in the content and activity of proteases can lead to resistance [78]. Frederick S. Walters et al. found that mCry3A toxicity to corn rootworm larvae was attributed to a chymotrypsin/cathepsin G site, which enhances cleavage and subsequent binding of the activated toxin to midgut cells [79].

In Lepidoptera and Diptera, serine proteases are the main type of intestinal protease [80, 81]; meanwhile in Coleoptera, cysteine and aspartic proteases are the principal class of digestive enzymes. Cathepsin G serine protease has been considered as the principal enzyme because its activity is a key step towards Cry toxicity. It has been demonstrated that cathepsin activity can be Cry toxin specific, for example, Acyrthosiphon pisum mainly expresses cathepsin L and cathepsin B types [82], but different serine proteases have activity for Cry4A and Cry4B in sensitive Diptera. [80]. Moreover, cathepsins B, L and serine peptidases such as trypsins and chymotrypsins are the major enzyme components in the larval midgut of tenebrionid beetles [83]. In general, the digestive protease activity of Coleopteran insects depends mainly on cysteine proteases [84, 85]. We found 394 protease-related unigenes in M. alternatus transcriptome (Fig 8). Among these unigenes, serine proteases were the most represented group (232, 58.88%), followed by cysteine proteases (79, 20.05%), metalloproteases (77, 19.54%) and aspartic acid proteases (61.52%). Threonine proteases and glutamic acid proteases were not found in our data. Interestingly, we found 212 serine proteases genes (Fig 8). Only 82 serine protease genes had lengths above 1 kb. The unigenes with lengths less than 1 kb were represented by 95 trypsin and 28 chymotrypsin genes. Serine proteases are the main protein digestive enzymes, accounting for 95% of the digestive activity in Lepidoptera [86]. In insects, genes encoding serine proteases (SP) and serine protease homologs (SPH) comprise a large family of proteins involved in digestion, immune defense, development and other process [87]. Cysteine proteases play an important role as virulence factors in Entamoeba histolytica. Furthermore, from invasive trophozoites to parasites in dormant infective cyst stages, cysteine proteases have an important role in parasite morphology [88]. It has been reported that Colorado potato beetle (CPB) midgut membrane metalloproteases participate in the proteolytic processing of Cry3Aa toxin [89]. C. Rausell et al. found that Brush border membrane vesicles (BBMV)-associated metalloproteases can cut Cry3Aa toxin specificity, thus significantly reducing the activity of pore formation [90]. In vertebrates, the four main aspartic proteases are pepsins, cathepsin D, cathepsin E and renins [9193], these enzymes mainly degrade endogenous proteins. Our data confirmed the abundance of protease-related unigenes in M. alternatus transcriptome, which has significant implications for future research regarding the mechanism of action of Cry toxin in M. alternatus. The information provided by insect transcriptomes can contribute in understanding of digestive enzyme functions, in addition to generate new tools to improve the activity of proteases.

thumbnail
Fig 8. Number of unigenes related to intestinal digestive enzymes.

x-axis indicates the corresponding number of unigenes and y-axis indicates the specific unigenes related to intestinal digestive enzymes. Three hundred and ninety-four protease-related unigenes were identified in M. alternatus Hope transcriptome. (I) Number of specific unigenes related to serine proteases in intestinal digestive enzymes. The icon indicates specific unigenes. The number in brackets indicates the corresponding number of unigenes.

https://doi.org/10.1371/journal.pone.0147855.g008

Possible future insect control targets

In addition to the diverse type of genes described above (Fig 9), we also identified candidate unigenes as targets for future insect control strategies, such as cathepsin B, cysteine peptidases, neuropeptides and serine peptidases. Among them, we found serine carboxypeptidase and serine-type endopeptidase, belonging to the serine peptidases (Fig 9). Of these, 117 unigenes had a length above 600 bp, 87 of which were above 1 kb. Moreover, we found 96 neuropeptide unigenes, corresponding to insect neurohormones that signal via G-protein-coupled receptors (GPCRs) to control growth, reproduction, behavior, breeding and other physiological processes [94]. Identifying important molecule-related unigenes for these basic physiological processes in M. alternatus will provide a reference for possible future research into targets for insect control and other applications.

thumbnail
Fig 9. Number of unigenes related to possible future insect control targets.

x-axis indicates the corresponding number of unigenes and y-axis indicates the specific unigenes related to possible future insect control targets. 465 possible future insect control targets unigenes were found in M. alternatus Hope transcriptome. Among these, 117 unigenes had a length above 600 bp and 87 above 1 kb. (I) Number of specific unigenes related to Serine Peptidase in possible future insect control targets. The icon indicates specific unigenes. The number in brackets indicates the corresponding number of unigenes. Serine carboxypeptidase, serine-type endopeptidase and others belongs to the serine peptidases.

https://doi.org/10.1371/journal.pone.0147855.g009

Immune-related molecules

Organisms have unique immune mechanisms against pathogens in the environment. Compared to vertebrates, invertebrates lack of an acquired immune system and rely on an innate immune system; adaptive immunity in insects has yet to be identified [95]. The innate immune system of insects is divided into cellular immunity and humoral immunity. Cellular immunity comprises phagocytosis, melanization and encapsulation. Humoral immunity has three steps:1) pattern recognition protein receptors (PRRs) recognize and bind to pathogen-associated molecular patterns (PAMPs); 2) a series of innate immune responses are activated, and 3) finally triggering the generation of innate effector activities (phagocytosis) and effector molecules (antimicrobial peptides (AMPs)) [96, 97]. RNAi machinery also plays an important role in regulating the innate immune response in insects and other organisms [98]. We identified 478 unigenes related to immune molecules and receptors related to immune activities (Fig 10). This group contains 20 widely recognized immune factors, including serine protease inhibitors (94, ~20%) and being the most abundant, followed by peroxidases (75, 16%). Interestingly, MD2(an extracellular binding partner of Toll-like receptor 4 (TLR4),)-like proteins and ML (MD-2-related lipid-recognition) -related unigenes were not found, although MD2-like gene family encodes for secretory proteins [99]. Therefore, molecules directly related to MLs may not be present in M. alternatus immune system.

thumbnail
Fig 10. Number of unigenes related to immune-related molecules.

x-axis indicates the corresponding number of unigenes and y-axis indicates the specific unigenes related to immune-related molecules. 478 unigenes were identified related to immune molecules and receptors in the transcriptome. This group contains 20 widely recognized immune factors. (I) Number of specific unigenes related to IMD pathway in immune-related molecules. (II) Number of specific unigenes related to Toll pathway in immune-related molecules. The icon indicates specific unigenes. The number in brackets indicates the corresponding number of unigenes.

https://doi.org/10.1371/journal.pone.0147855.g010

1. Antimicrobial peptides.

AMP-related unigenes were represented by approximately 7% of the total identified immune factors. AMPs play important roles in the innate immune system of invertebrates and possess significantly broad antimicrobial activity. AMPs are widespread in nature and mainly distributed in Lepidoptera, Coleoptera, Diptera, Hymenoptera, Hemiptera, Isoptera, Homoptera and Odonata orders. Antimicrobial peptides such as metchnikowin, drosocin, defensin, diptericins, attacins, cecropins, drosomycins, have been characterized in D. melanogaster [100, 101]. We identified 33 antimicrobial peptide-related coding unigenes: antifungal peptides were the most abundant (15, ~ 45%), while attacins, cecropins and defensins represented 15%, 9% and 6%, respectively. Iijima R et al. isolated an anti-fungal peptide (AFP) from Sarcophaga peregrina as the first reported AFP [102], it is composed of 67 amino acid residues and is rich in glycine and histidine. We identified 5 attacin-related unigenes. Attacins only participate in bacteriostasis for some gram-negative bacteria [103]. At present, more than 20 cecropin analogues have been isolated from Lepidoptera and Diptera insects [104]. Cecropines form a voltage-dependent channel that changes the permeability of the bacterial cell membrane. Its antibacterial spectrum includes gram-negative and gram-positive bacteria. Moreover, we found two defensin unigenes (~6%), insect defensins were first isolated in Phormia terranovae, due to their high homology with mammalian defensins and later identified in S. peregrina (sapecins) [105]. Members of insect defensins family were also identified in A. aegypti and A. gambiae [106]. Although fewer in number defensin activities cannot be ignored in M. alternatus immune system. Dobson et al. hypothesized that the highly diverse function of antibacterial peptides have implications in the evolution of pathogen inhibition mechanisms in the host [107].

2. Immune enzymes and inhibitors.

We found 17 unigenes related to caspases (CASPs), a class of aspartic acid proteolytic enzymes containing cysteine and responsible for apoptosis initiation; they play a critical role in regulating cell death in the growth process of an organism. CASPs 4, 5 and 11 have been reported as cytoplasmic receptors for gram-negative lipopolysaccharides (LPS) [30, 108]. LPS are important structural components of gram-negative bacterial outer cell membranes and can activate the innate immune response through TLR4.

Ninety-four serine protease inhibitors (SRPN) unigenes were identified in our data, contributing for ~20% of the total number of immune factor unigenes; which may possess important active defense immune functions in the process of pathogenic microorganism infection. SRPNs can participate in a variety of physiological reactions and are present higher eukaryotes and viruses [109, 110].

3. Oxidative stress molecules.

In insect immune responses, a high level of reactive oxygen species (ROS) is produced. Peroxidase-related unigenes identified in M. alternatus contributed for 16% of the total number of immune factor-related unigenes, we identified 75 peroxidase-related unigenes, 21 unigenes corresponding to including glutathione [GPXs], 8 corresponding to thioredoxin [TPXs], and 30 unigenes corresponding to Haem. Furthermore, we identified 15 catalase-related unigenes (CAT), 16 Lysozyme-related unigenes (LYS), and 6 prophenoloxidases-related unigenes (PPO). CATs can effectively convert H2O2 into water and oxygen, removing hydrogen peroxide from the cell to prevent H2O2 poisoning. Therefore, CATs are key enzymes in the biological defense system. Studies have shown that CAT gene disruption in D. melanogaster quickly leads to death [111]. Moreover, we also identifed superoxide dismutases (SODs) unigenes (5%). SODs can convert •O2-- into hydrogen peroxide (H2O2), which is less toxic.

4. Lectins and galectins.

Seven unigenes related to C-Type lectins (CTLs) and five galectin (GALEs) genes were identified in our data. CTLs are secreted proteins or membrane proteins that depend on calcium ions for their function, they are the largest and most diverse family of animal lectins. Several invertebrate immune responses involving CTLs include opsonization [112, 113], pathogen elimination, hemocytes biosynthesis and activating prophendoxidase to produce melaninization [114116]. GALEs are β-galactoside binding proteins that depend on mercaptans. There is evidence that galectins are related to congenital immunity in Drosophila and An. gambiae [117].

5. Pattern recognition proteins.

Peptidoglycan recognition proteins (PGRP) are pattern recognition receptors that can recognize bacteria peptidoglycans and they play an important role in recognition and adaption of innate immunity. We found 7 PGRP-related unigenes and interestingly, 34 pattern recognition scavenger receptor (SCR), SCR mediates the recognition of pathogen LPS and LTA [118120] and “clean up” after apoptosis [53, 54, 121125]. We hypothesize that SCR-associated molecules in M. alternatus may have similar actions. In addition, we identified 11 fibrinogen-related proteins (FREP) and 30 CLIP-domain serine proteases, contributing for 2% and 6% respectively. CLIP proteases have regulatory roles for immune responses in insects [111, 126132]. We also discovered one unigene that correlated with a Gram-Negative Binding Protein (GNBP). Based on the genetics and biochemistry analysis GNBP1 has been identified as a co-receptor for gram positive microbes and PGRP-SA [133, 134]. Finally, we identified 9 thioester-containing (TEP) protein-related unigenes (~2%). The TEP family is mainly composed of vertebrate complementary factors (C3, C4, C5) and α2-macroglobulin. Anopheles gambiae TEP1 is a typical TEP in insects, which can bind to the surface of pathogenic bacteria and promote the haemocytes of mosquitoes to phagocyte the bacteria or eliminate Plasmodium [135137].

6. IMD, Toll, JAK/STAT and JNK-signaling pathway.

We identifed unigenes that are closely related to the IMD, Toll, JAK/STAT and JNK-signaling pathway, represented by 20, 62, 4 and 9 unigenes, respectively.

Imd pathway.

The Imd pathway is the primary immune pathway that acts against Gram-negative bacteria. We identified 20 unigenes, such as IMD, TAK1, IAP, IAP2, dredd, ikkb and relish (Fig 10). Imd pathway regulates the activity of a third Drosophila NF-kB protein named Relish, controls the expression of most of the Drosophila AMPs and, thus is indispensable for normal immunity [138]. TAK1 is ubiquitin-dependent kinase of IKK [68,139]. IAP2 participates in Relish nuclear localization in Drosophila [140]. IKK enhanced the phosphorylation of Relish, after Relish cleavage by DREDD, causing translocation of the N-terminal end to the nucleus [141].

Toll pathway.

ixty-two identified unigenes were related to the Toll pathway, including spatzle, Toll, MyD88, Pelle, pellino, Cactin and Dorsal/Dif (Fig 10). In Drosophila the Toll receptors are essential for embryonic development and immunity. The induction of the Toll pathway by Gram-positive bacteria or fungi leads to the activation od cellular immunity and the systemic production of AMPs. The Toll receptor is activated when the proteolytically cleaved ligand Spatzle binds the receptor leading to the activation of the NF-kB factors Dorsal-related immunity factor [141, 142]. Cactus is phosphorylated by a complex consisting of MyD88, tube and pelle, causing degradation of cactus and release of Dorsal/Dif [143, 144]. Besides, pellino and Cactin genes showed high expression levels in T. castaneum transcriptome [145].

JAK/STAT-signaling pathway/ JNK-signaling pathway.

two additional signaling pathways have been shown to have immune functions in insects: the JNK and JAK-STAT pathways involved in cell stress or wound response as other immune pathways in insects [139]. Only two genes involved in JAK-STAT signaling pathway were identified in our transcriptome sequencing data, including 3 unigenes related to STAT92E and 1 unigenes related to hopscotch. We could not identified the reaining components of this pathway (unpaired, STAT, JAK, and the receptor Domeless/Master of marelle (DOME/MOM)) [146, 147]. We also identified 9 unigenes related to the JNK-signaling pathway. Five unigenes were related to Kay and 4 to hemipterous.

Conclusion

M. alternatus is known as the cancer of pine trees and is considered a devastating disease, causing serious environmental and economic losses. Vector control strategies are needed to stop and/or control the spreading of this disease [5]. In this work, we sequenced and characterized the transcriptome in the insect vector M. alternatus using Illumina sequencing. We identified a large set of genes related to putative insecticide resistance, intestinal digestive enzymes, possible future insect control targets and immune-related molecules. This study provides valuable information that may serve as key point to develop new control strategies for Pine Wilt Disease.

Materials and Methods

Ethics Statement

There are no specific permits for insect collection in the selected locations. The chosen locations are not privately-owned or natural protected areas. Insects used for the experiments are not considered endangered or protected species, and its collection is legal in China.

Insects

Pinus massoniana (P. massoniana) trunks infested with M. alternatus were selected from trees withered during the first year in the town of GuanTou, LianJiang county and FuJian province (N 26.15046°;E 119.59261°). Trunks were cut into 1 m size with a chain saw in the open field and transported to the isolated laboratory of FuJian Agriculture and Forestry University in sealed canvas bags. Trucks were kept in a rearing cage (1.5 m x 1.0 m in length/width) with 1mm iron mesh to prevent insect scaping. Insects were maintained using an artificial diet and kept for two generations. Twenty-five whole larvae (fourth instar) were collected for RNA extraction.

cDNA library and Illumina sequencing

Total RNA was extracted from 25 whole larvae (fourth instar) using TRIzol Reagent (Invitrogen). Extracted RNA was processed using the E.Z.N.A. ® HP Total RNA Kit (OMEGA RNA, Invitrogen) to eliminate polysaccharides. RNA purity, quality and concentration were determined using Nanodrop, Qubit 2.0 and Agilent 2100 methods. Messenger RNA was extracted from 6 μg (50 ng/ul) of total RNA, using oligo (dT) magnetic beads, fragmentation buffer was added to the beads coated with mRNA and mRNA was broken randomly. mRNA was used to synthetize the first cDNA chain and then subjected to a second amplification to obtain double stranded cDNA. Double stranded cDNA was purified using AMPure XP. Finally, the cDNA library was created by PCR enrichment using HiSeq2500 high-throughput sequencing, the read length for the sequencing was PE125.

Bioinformatic analysis

The cDNA library was sequenced by High-throughput sequencing platform to produce a large number of high quality reads based on Sequencing by Synthesis (SBS) technology. Raw data were cleaned from joint sequence and low-quality reads. Depurated reads were assembled into contigs using Trinity software. The transcript sequences were identified in each fragment collection using the De Bruijn method of graphing the sequencing read information [9]. The BLAST parameter E-value was set at 10−5, and the HMMER parameter E-value was set at 10−10. All unigenes were compared with NR [148], Swiss-Prot [149], GO [150], COG [150], KOG [151] and KEGG [152] databases using BLAST [153] software. GO database was used to determine the function of the identified transcript and to assign Gene Ontology (GO) terms. In addition, metabolic pathways were predicted using the COG and KEGG databases. Amino acid sequence of the unigenes was analyzed by HMMER [154] software and the outputs were searched in the Pfam [155] database to obtain annotated information for the unigenes.

Supporting Information

S1 Table. Gene Ontology of Monochamus alternatus Hope unigenes.

https://doi.org/10.1371/journal.pone.0147855.s002

(XLSX)

S2 Table. KEGG summary of Monochamus alternatus Hope transcriptome.

https://doi.org/10.1371/journal.pone.0147855.s003

(XLSX)

S3 Table. Putative P450 genes identified in Monochamus alternatus Hope transcriptome.

https://doi.org/10.1371/journal.pone.0147855.s004

(XLSX)

S4 Table. Identified genes in Monochamus alternatus Hope transcriptome.

https://doi.org/10.1371/journal.pone.0147855.s005

(XLS)

Acknowledgments

We wish to thank Dr. Brian McGarvey and Dr. Rebecca Carballar for proofreading this paper. This study was supported by the Special Fund for Forestry Research in the Public Interest of China [Grant No. 201304401], Project of Fujian-Taiwan Joint Center for Ecological Control of Crop Pests [Minjiaoke (2013) 51], the United Fujian Provincial Health and Education Project for Tackling Key Research [Grant No. WKJ-FJ-25], National Natural Science Foundation of China [Grant No. 31301724, No. 31201574 and No. 31401802], Education Department Foundation of Fujian Province Distinguished Young Scholars in Universities of Fujian Province [Grant No.JA12092], Science Fund for Distinguished Young Scholars of Fujian Agriculture and Forestry University [Grant No. xjq201203], Key Projects of Fujian Provincial Department of Science and Technology [No. 2011|N5003 and No. 2013N0003], Natural Science Foundation of Fujian Province [Grant No. 2013J01079] and Programme for New Century Excellent Talents in Fujian Province [JA13089], the Special Fund for Grain Research in the Public Interest of China [Grant No. 201313002-3], the Special Fund for Key Construction Project of Fujian Agriculture and Forestry University [Grant No. 6112C035003], National Undergraduate Training Programme for Innovation and Entrepreneurship [Grant No. 201510389013, No.201510389018, No. 201510389023, No. 201510389025, No. 201510389028, No. 201510389101 and No. 201510389173]. Science and Technology Project in Fujian Province Department of Education [JA15161], Central Finance Forestry Science and Technology Demonstration Project [Min (2015) TG017].

Author Contributions

Conceived and designed the experiments: FPZ XG SQW XLZ. Performed the experiments: SQW XLZ ZXL. Analyzed the data: SQW XLZ ZXL ESS CLR YJG YTX RXX YNM XH GHL SQZ. Contributed reagents/materials/analysis tools: SQW XLZ ZXL ESS CLR YJG YTX RXX YNM XH GHL SQZ. Wrote the paper: SQW XLZ.

References

  1. 1. Wu XQ, Yuan WM, Tian XJ, Fan B, Fang X, Ye JR, et al. Specific and functional diversity of endophytic bacteria from pine wood nematode Bursaphelenchus xylophilus with different virulence. Int J Biol Sci. 2013;9(1):34–44. pmid:23289015; PubMed Central PMCID: PMC3535532.
  2. 2. State Forestry Bureau of China (2014) The announcement of Bursaphelenchus xylophilus infected area in 2014. On February 12, 2014. http://wwwforestrygovcn/main/3600/content-657977html.
  3. 3. Zhao BG, Futai K, Sutherland JR, Takeuchi Y. Pine wilt disease: Springer; 2008.
  4. 4. Hajek AE, Glare TR, O'Callaghan M. Use of microbes for control and eradication of invasive arthropods: Springer Science & Business Media; 2008.
  5. 5. Yang ZQ, Wang XY, Zhang YN. Recent advances in biological control of important native and invasive forest pests in China. Biol Control. 2014;68:117–28. WOS:000327821500011.
  6. 6. YanLong Z, ZhongQi Y, XiaoYi W, YiNan Z, ChengJin W, ShengFu M, et al. Functional response of the parasitoid Sclerodermus sp.(Hymenoptera: Bethylidae) to the third instar larvae of host Monochamus alternatus (Coleoptera: Cerambycidae). Acta Entomologica Sinica. 2012;55(4):426–34.
  7. 7. Wang Y-H, Lai Y-X, Cen D-H, Wang J-Y, Yang Z-Q. Behavior and numbers of Dastarcus helophoroides beetle’s laying egg and its first-instar larvae parasitism on larvae of Monochamus alternatus. Journal of Environmental Entomology. 2011;4.
  8. 8. Yuanliang Y, Yang Zq, Wang Xy, Yu Jx, Yan Xw. Biological control of Monochamus alternatus (Coleoptera: Cerambycidae) by releasing eggs and adults of Dastarcus helophoroides (Coleoptera: Bothrideridae). Scientia Sivae Sinicae. 2013;49(3):103–9.
  9. 9. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nature Biotechnology. 2011;29(7):644–U130. WOS:000292595200023. pmid:21572440
  10. 10. Bai X, Zhang W, Orantes L, Jun T-H, Mittapalli O, Mian MAR, et al. Combining next-generation sequencing strategies for rapid molecular resource development from an invasive aphid species, Aphis glycines. PloS one. 2010;5(6):e11370. pmid:20614011
  11. 11. Crawford JE, Guelbeogo WM, Sanou A, Traoré A, Vernick KD, Sagnon NF, et al. De novo transcriptome sequencing in Anopheles funestus using Illumina RNA-seq technology. PloS one. 2010;5(12):e14202. pmid:21151993
  12. 12. Karatolos N, Pauchet Y, Wilkinson P, Chauhan R, Denholm I, Gorman K, et al. Pyrosequencing the transcriptome of the greenhouse whitefly, Trialeurodes vaporariorum reveals multiple transcripts encoding insecticide targets and detoxifying enzymes. BMC genomics. 2011;12:56. pmid:21261962; PubMed Central PMCID: PMC3036619.
  13. 13. Wang XW, Luan JB, Li JM, Bao YY, Zhang CX, Liu SS. De novo characterization of a whitefly transcriptome and analysis of its gene expression during development. BMC genomics. 2010;11:400. pmid:20573269; PubMed Central PMCID: PMC2898760.
  14. 14. Xue J, Bao YY, Li BL, Cheng YB, Peng ZY, Liu H, et al. Transcriptome analysis of the brown planthopper Nilaparvata lugens. PloS one. 2010;5(12):e14233. pmid:21151909; PubMed Central PMCID: PMC2997790.
  15. 15. Zhu JY, Zhao N, Yang B. Global Transcriptome Profiling of the Pine Shoot Beetle, Tomicus yunnanensis (Coleoptera: Scolytinae). PloS one. 2012;7(2). ARTN e32291 WOS:000302916100067. pmid:22384206
  16. 16. Gene Ontology C. The Gene Ontology project in 2008. Nucleic acids research. 2008;36(Database issue):D440–4. pmid:17984083; PubMed Central PMCID: PMC2238979.
  17. 17. Cancino-Rodezno A, Lozano L, Oppert C, Castro JI, Lanz-Mendoza H, Encarnacion S, et al. Comparative proteomic analysis of Aedes aegypti larval midgut after intoxication with Cry11Aa toxin from Bacillus thuringiensis. PloS one. 2012;7(5):e37034. pmid:22615881; PubMed Central PMCID: PMC3353955.
  18. 18. Daimon T, Kozaki T, Niwa R, Kobayashi I, Furuta K, Namiki T, et al. Precocious metamorphosis in the juvenile hormone-deficient mutant of the silkworm, Bombyx mori. PLoS genetics. 2012;8(3):e1002486. pmid:22412378; PubMed Central PMCID: PMC3297569.
  19. 19. Helvig C, Koener JF, Unnithan GC, Feyereisen R. CYP15A1, the cytochrome P450 that catalyzes epoxidation of methyl farnesoate to juvenile hormone III in cockroach corpora allata. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(12):4024–9. WOS:000220472200009. pmid:15024118
  20. 20. Niwa R, Matsuda T, Yoshiyama T, Namiki T, Mita K, Fujimoto Y, et al. CYP306A1, a cytochrome P450 enzyme, is essential for ecdysteroid biosynthesis in the prothoracic glands of Bombyx and Drosophila. Journal of Biological Chemistry. 2004;279(34):35942–9. WOS:000223303400095. pmid:15197185
  21. 21. Niu G, Rupasinghe SG, Zangerl AR, Siegel JP, Schuler MA, Berenbaum MR. A substrate-specific cytochrome P450 monooxygenase, CYP6AB11, from the polyphagous navel orangeworm (Amyelois transitella). Insect Biochem Mol Biol. 2011;41(4):244–53. pmid:21220011.
  22. 22. Schuler MA, Berenbaum MR. Structure and function of cytochrome P450S in insect adaptation to natural and synthetic toxins: insights gained from molecular modeling. J Chem Ecol. 2013;39(9):1232–45. pmid:24036972.
  23. 23. Daborn PJ, Lumb C, Boey A, Wong W, Ffrench-Constant RH, Batterham P. Evaluating the insecticide resistance potential of eight Drosophila melanogaster cytochrome P450 genes by transgenic over-expression. Insect Biochem Mol Biol. 2007;37(5):512–9. pmid:17456446.
  24. 24. Feyereisen R. Evolution of insect P450. Biochem Soc T. 2006;34:1252–5. WOS:000243013500061.
  25. 25. Ai JW, Zhu Y, Duan J, Yu QY, Zhang GJ, Wan F, et al. Genome-wide analysis of cytochrome P450 monooxygenase genes in the silkworm, Bombyx mori. Gene. 2011;480(1–2):42–50. WOS:000291773600006. pmid:21440608
  26. 26. Xia Q, Cheng D, Duan J, Wang G, Cheng T, Zha X, et al. Microarray-based gene expression profiles in multiple tissues of the domesticated silkworm, Bombyx mori. Genome biology. 2007;8(8):R162. pmid:17683582
  27. 27. Bass C, Field LM. Gene amplification and insecticide resistance. Pest management science. 2011;67(8):886–90. WOS:000293327300002. pmid:21538802
  28. 28. Mannervik B, Helena Danielson U, Ketterer B. Glutathione transferases-structure and catalytic activit. Critical Reviews in Biochemistry and Molecular Biology. 1988;23(3):283–337.
  29. 29. Pickett CB, Lu AYH. Glutathione S-Transferases: Gene Structure, Regulation, and Biological Function. Annual Review of Biochemistry. 1989;58(1):743–64. pmid:2673020.
  30. 30. Shi J, Zhao Y, Wang Y, Gao W, Ding J, Li P, et al. Inflammatory caspases are innate immune receptors for intracellular LPS. Nature. 2014;514(7521):187–92. pmid:25119034.
  31. 31. Ranson H, Claudianos C, Ortelli F, Abgrall C, Hemingway J, Sharakhova MV, et al. Evolution of supergene families associated with insecticide resistance. Science. 2002;298(5591):179–81. WOS:000178370500053. pmid:12364796
  32. 32. Ding YC, Ortelli F, Rossiter LC, Hemingway J, Ranson H. The Anopheles gambiae glutathione transferase supergene family: annotation, phylogeny and expression profiles. BMC genomics. 2003;4. Artn 35 WOS:000185828600001. pmid:12914673
  33. 33. Ranson H, Rossiter L, Ortelli F, Jensen B, Wang X, Roth CW, et al. Identification of a novel class of insect glutathione S-transferases involved in resistance to DDT in the malaria vector Anopheles gambiae. Biochem J. 2001;359(2):295–304.
  34. 34. Pemble SE, Taylor JB. An evolutionary perspective on glutathione transferases inferred from class-theta glutathione transferase cDNA sequences. Biochem J. 1992;287 (Pt 3):957–63. pmid:1445253; PubMed Central PMCID: PMC1133100.
  35. 35. Friedman R. Genomic organization of the glutathione S-transferase family in insects. Mol Phylogenet Evol. 2011;61(3):924–32. pmid:21930223.
  36. 36. Gullipalli D, Arif A, Aparoy P. Identification of a developmentally and hormonally regulated Delta-Class glutathione S-transferase in rice moth Corcyra cephalonica. Comparative Biochemistry & Physiology Part B Biochemistry & Molecular Biology. 2010;156(1):33–9.
  37. 37. Burand JP, Hunter WB. RNAi: future in insect management. J Invertebr Pathol. 2013;112 Suppl:S68–74. pmid:22841639.
  38. 38. Roignant JY, Carre C, Mugat B, Szymczak D, Lepesant JA, Antoniewski C. Absence of transitive and systemic pathways allows cell-specific and isoform-specific RNAi in Drosophila. RNA. 2003;9(3):299–308. pmid:12592004; PubMed Central PMCID: PMC1370397.
  39. 39. Bischoff V, Vignal C, Duvic B, Boneca IG, Hoffmann JA, Royet J. Downregulation of the Drosophila immune response by peptidoglycan-recognition proteins SC1 and SC2. PLoS Pathog. 2006;2(2):e14. pmid:16518472; PubMed Central PMCID: PMC1383489.
  40. 40. Miller SC, Brown SJ, Tomoyasu Y. Larval RNAi in Drosophila? Development Genes And Evolution. 2008;218(9):505–10. WOS:000258717800005. pmid:18663472
  41. 41. Arakane Y, Li B, Muthukrishnan S, Beeman RW, Kramer KJ, Park Y. Functional analysis of four neuropeptides, EH, ETH, CCAP and bursicon, and their receptors in adult ecdysis behavior of the red flour beetle, Tribolium castaneum. Mechanisms of Development. 2008;125(11–12):984–95. WOS:000262302100006. pmid:18835439
  42. 42. Fujita K, Shimomura K, Yamamoto K, Yamashita T, Suzuki K. A chitinase structurally related to the glycoside hydrolase family 48 is indispensable for the hormonally induced diapause termination in a beetle. Biochemical and Biophysical Research Communications. 2006;345(1):502–7. WOS:000237877700068. pmid:16684504
  43. 43. Konopova B, Jindra M. Broad-complex acts downstream of Met in juvenile hormone signaling to coordinate primitive holometabolan metamorphosis. Development. 2008;135(3):559–68. WOS:000252254900015. pmid:18171683
  44. 44. Minakuchi C, Namiki T, Shinoda T. Kruppel homolog 1, an early juvenile hormone-response gene downstream of Methoprene-tolerant, mediates its anti-metamorphic action in the red flour beetle Tribolium castaneum. Developmental Biology. 2009;325(2):341–50. WOS:000262589400004. pmid:19013451
  45. 45. Parthasarathy R, Palli SR. Molecular Analysis of Juvenile Hormone Analog Action in Controlling the Metamorphosis of the Red Flour Beetle, Tribolium Castaneum. Archives of Insect Biochemistry and Physiology. 2009;70(1):57–70. WOS:000262205000006. pmid:19072925
  46. 46. Tomoyasu Y, Denell RE. Larval RNAi in Tribolium (Coleoptera) for analyzing adult development. Development genes and evolution. 2004;214(11):575–8. pmid:15365833.
  47. 47. Quan GX, Kanda T, Tamura T. Induction of the white egg 3 mutant phenotype by injection of the double-stranded RNA of the silkworm white gene. Insect molecular biology. 2002;11(3):217–22. pmid:12000640.
  48. 48. Ohnishi A, Hull JJ, Matsumoto S. Targeted disruption of genes in the Bombyx mori sex pheromone biosynthetic pathway. Proc Natl Acad Sci U S A. 2006;103(12):4398–403. pmid:16537410; PubMed Central PMCID: PMC1450183.
  49. 49. Hossain M, Shimizu S, Matsuki M, Imamura M, Sakurai S, Iwami M. Expression of 20-hydroxyecdysone-induced genes in the silkworm brain and their functional analysis in post-embryonic development. Insect Biochem Mol Biol. 2008;38(11):1001–7. pmid:18835445.
  50. 50. Huvenne H, Smagghe G. Mechanisms of dsRNA uptake in insects and potential of RNAi for pest control: a review. Journal of insect physiology. 2010;56(3):227–35. pmid:19837076
  51. 51. Tribolium Genome Sequencing C, Richards S, Gibbs RA, Weinstock GM, Brown SJ, Denell R, et al. The genome of the model beetle and pest Tribolium castaneum. Nature. 2008;452(7190):949–55. pmid:18362917.
  52. 52. Xu W, Han Z. Cloning and phylogenetic analysis of sid-1-like genes from aphids. J Insect Sci. 2008;8:1–6. pmid:20302524; PubMed Central PMCID: PMC3061602.
  53. 53. Pearson A, Lux A, Krieger M. Expression cloning of dSR-CI, a class C macrophage-specific scavenger receptor from Drosophila melanogaster. Proc Natl Acad Sci U S A. 1995;92(9):4056–60. pmid:7732030; PubMed Central PMCID: PMC42101.
  54. 54. Ramet M, Pearson A, Manfruelli P, Li X, Koziel H, Gobel V, et al. Drosophila scavenger receptor CI is a pattern recognition receptor for bacteria. Immunity. 2001;15(6):1027–38. pmid:11754822.
  55. 55. Saleh MC, van Rij RP, Hekele A, Gillis A, Foley E, O'Farrell PH, et al. The endocytic pathway mediates cell entry of dsRNA to induce RNAi silencing. Nature Cell Biology. 2006;8(8):793–802. WOS:000239423000009. pmid:16862146
  56. 56. Tomoyasu Y, Miller SC, Tomita S, Schoppmeier M, Grossmann D, Bucher G. Exploring systemic RNA interference in insects: a genome-wide survey for RNAi genes in Tribolium. Genome Biol. 2008;9(1):R10. pmid:18201385; PubMed Central PMCID: PMC2395250.
  57. 57. Sijen T, Fleenor J, Simmer F, Thijssen KL, Parrish S, Timmons L, et al. On the role of RNA amplification in dsRNA-triggered gene silencing. Cell. 2001;107(4):465–76. pmid:11719187.
  58. 58. Bravo A, Likitvivatanavong S, Gill SS, Soberon M. Bacillus thuringiensis: A story of a successful bioinsecticide. Insect Biochem Mol Biol. 2011;41(7):423–31. pmid:21376122.
  59. 59. Pigott CR, Ellar DJ. Role of receptors in Bacillus thuringiensis crystal toxin activity. Microbiol Mol Biol Rev. 2007;71(2):255–81. pmid:17554045; PubMed Central PMCID: PMC1899880.
  60. 60. Knight PJ, Knowles BH, Ellar DJ. Molecular cloning of an insect aminopeptidase N that serves as a receptor for Bacillus thuringiensis CryIA(c) toxin. J Biol Chem. 1995;270(30):17765–70. pmid:7629076.
  61. 61. Knight PJ, Crickmore N, Ellar DJ. The receptor for Bacillus thuringiensis CrylA(c) delta-endotoxin in the brush border membrane of the lepidopteran Manduca sexta is aminopeptidase N. Mol Microbiol. 1994;11(3):429–36. pmid:7908713.
  62. 62. Sangadala S, Walters FS, English LH, Adang MJ. A mixture of Manduca sexta aminopeptidase and phosphatase enhances Bacillus thuringiensis insecticidal CryIA(c) toxin binding and 86Rb(+)-K+ efflux in vitro. J Biol Chem. 1994;269(13):10088–92. pmid:8144508.
  63. 63. Vadlamudi RK, Weber E, Ji I, Ji TH, Bulla LA Jr. Cloning and expression of a receptor for an insecticidal toxin of Bacillus thuringiensis. J Biol Chem. 1995;270(10):5490–4. pmid:7890666.
  64. 64. Garczynski SF, Adang MJ. Bacillus thuringiensis CryIA (c) δ-endotoxin binding aminopeptidase in the Manduca sexta midgut has a glycosyl-phosphatidylinositol anchor. Insect biochemistry and molecular biology. 1995;25(4):409–15.
  65. 65. Crava CM, Bel Y, Lee SF, Manachini B, Heckel DG, Escriche B. Study of the aminopeptidase N gene family in the lepidopterans Ostrinia nubilalis (Hubner) and Bombyx mori (L.): sequences, mapping and expression. Insect Biochem Mol Biol. 2010;40(7):506–15. pmid:20420910.
  66. 66. Hughes AL. Evolutionary diversification of aminopeptidase N in Lepidoptera by conserved clade-specific amino acid residues. Mol Phylogenet Evol. 2014;76:127–33. pmid:24675701.
  67. 67. Contreras E, Schoppmeier M, Real MD, Rausell C. Sodium Solute Symporter and Cadherin Proteins Act as Bacillus thuringiensis Cry3Ba Toxin Functional Receptors in Tribolium castaneum. Journal of Biological Chemistry. 2013;288(25):18013–21. WOS:000320721900008. pmid:23645668
  68. 68. English LH, Readdy TL. Delta endotoxin inhibits a phosphatase in midgut epithelial membranes of Heliothis virescens. Insect biochemistry. 1989;19(2):145–52.
  69. 69. McNall RJ, Adang MJ. Identification of novel Bacillus thuringiensis Cry1Ac binding proteins in Manduca sexta midgut through proteomic analysis. Insect Biochemistry and Molecular Biology. 2003;33(10):999–1010. WOS:000185958200005. pmid:14505693
  70. 70. Krishnamoorthy M, Jurat-Fuentes JL, McNall RJ, Andacht T, Adang MJ. Identification of novel Cry1Ac binding proteins in midgut membranes from Heliothis virescens using proteomic analyses. Insect Biochem Mol Biol. 2007;37(3):189–201. pmid:17296494.
  71. 71. Bayyareddy K, Andacht TM, Abdullah MA, Adang MJ. Proteomic identification of Bacillus thuringiensis subsp. israelensis toxin Cry4Ba binding proteins in midgut membranes from Aedes (Stegomyia) aegypti Linnaeus (Diptera, Culicidae) larvae. Insect Biochem Mol Biol. 2009;39(4):279–86. pmid:19272330.
  72. 72. Flores-Escobar B, Rodriguez-Magadan H, Bravo A, Soberon M, Gomez I. Differential role of Manduca sexta aminopeptidase-N and alkaline phosphatase in the mode of action of Cry1Aa, Cry1Ab, and Cry1Ac toxins from Bacillus thuringiensis. Appl Environ Microbiol. 2013;79(15):4543–50. pmid:23686267; PubMed Central PMCID: PMC3719532.
  73. 73. Heckel DG. Learning the ABCs of Bt: ABC transporters and insect resistance to Bacillus thuringiensis provide clues to a crucial step in toxin mode of action. Pesticide Biochemistry and Physiology. 2012;104(2):103–10. WOS:000311068900005.
  74. 74. Tanaka S, Miyamoto K, Noda H, Jurat-Fuentes JL, Yoshizawa Y, Endo H, et al. The ATP-binding cassette transporter subfamily C member 2 in Bombyx mori larvae is a functional receptor for Cry toxins from Bacillus thuringiensis. FEBS J. 2013;280(8):1782–94. pmid:23432933.
  75. 75. Atsumi S, Miyamoto K, Yamamoto K, Narukawa J, Kawai S, Sezutsu H, et al. Single amino acid mutation in an ATP-binding cassette transporter gene causes resistance to Bt toxin Cry1Ab in the silkworm, Bombyx mori. Proc Natl Acad Sci U S A. 2012;109(25):E1591–E8. pmid:22635270; PubMed Central PMCID: PMC3382473.
  76. 76. Gahan LJ, Pauchet Y, Vogel H, Heckel DG. An ABC transporter mutation is correlated with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS genetics. 2010;6(12):e1001248. pmid:21187898; PubMed Central PMCID: PMC3002984.
  77. 77. Bravo A, Gomez I, Porta H, Garcia-Gomez BI, Rodriguez-Almazan C, Pardo L, et al. Evolution of Bacillus thuringiensis Cry toxins insecticidal activity. Microbial biotechnology. 2013;6(1):17–26. pmid:22463726; PubMed Central PMCID: PMC3815381.
  78. 78. Deist BR, Rausch MA, Fernandez-Luna MT, Adang MJ, Bonning BC. Bt toxin modification for enhanced efficacy. Toxins. 2014;6(10):3005–27. pmid:25340556; PubMed Central PMCID: PMC4210883.
  79. 79. Walters FS, Stacy CM, Lee MK, Palekar N, Chen JS. An engineered chymotrypsin/cathepsin G site in domain I renders Bacillus thuringiensis Cry3A active against western corn rootworm larvae. Applied and environmental microbiology. 2008;74(2):367–74. pmid:18024675
  80. 80. Borovsky D. Biosynthesis and control of mosquito gut proteases. IUBMB Life. 2003;55(8):435–41. pmid:14609198.
  81. 81. Schnepf E, Crickmore N, Van Rie J, Lereclus D, Baum J, Feitelson J, et al. Bacillus thuringiensis and its pesticidal crystal proteins. Microbiol Mol Biol Rev. 1998;62(3):775–806. pmid:9729609; PubMed Central PMCID: PMC98934.
  82. 82. Carrillo L, Martinez M, Alvarez-Alfageme F, Castanera P, Smagghe G, Diaz I, et al. A barley cysteine-proteinase inhibitor reduces the performance of two aphid species in artificial diets and transgenic Arabidopsis plants. Transgenic Res. 2011;20(2):305–19. pmid:20567901.
  83. 83. Vinokurov KS, Elpidina EN, Oppert B, Prabhakar S, Zhuzhikov DP, Dunaevsky YE, et al. Diversity of digestive proteinases in Tenebrio molitor (Coleoptera: Tenebrionidae) larvae. Comp Biochem Physiol B Biochem Mol Biol. 2006;145(2):126–37. pmid:16859942.
  84. 84. Bown DP, Wilkinson HS, Jongsma MA, Gatehouse JA. Characterisation of cysteine proteinases responsible for digestive proteolysis in guts of larval western corn rootworm (Diabrotica virgifera) by expression in the yeast Pichia pastoris. Insect Biochem Mol Biol. 2004;34(4):305–20. pmid:15041015.
  85. 85. Murdock LL, Brookhart G, Dunn PE, Foard DE, Kelley S, Kitch L, et al. Cysteine digestive proteinases in Coleoptera. Comparative Biochemistry and Physiology Part B: Comparative Biochemistry. 1987;87(4):783–7.
  86. 86. Srinivasan A, Giri AP, Gupta VS. Structural and functional diversities in lepidopteran serine proteases. Cellular & molecular biology letters. 2006;11(1):132–54. pmid:16847755.
  87. 87. Kuwar S, Pauchet Y, Vogel H, Heckel DG. Adaptive regulation of digestive serine proteases in the larval midgut of Helicoverpa armigera in response to a plant protease inhibitor. Insect Biochemistry and Molecular Biology. 2015.
  88. 88. De Cádiz AE, Jeelani G, Nakada-Tsukui K, Caler E, Nozaki T. Transcriptome analysis of encystation in Entamoeba invadens. PloS one. 2013;8(9):e74840. pmid:24040350
  89. 89. Rausell C, Ochoa-Campuzano C, Martínez-Ramírez AC, Bravo A, Real MD. A membrane associated metalloprotease cleaves Cry3Aa Bacillus thuringiensis toxin reducing pore formation in Colorado potato beetle brush border membrane vesicles. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2007;1768(9):2293–9.
  90. 90. Rausell C, Ochoa-Campuzano C, Martinez-Ramirez AC, Bravo A, Real MD. A membrane associated metalloprotease cleaves Cry3Aa Bacillus thuringiensis toxin reducing pore formation in Colorado potato beetle brush border membrane vesicles. Biochimica et biophysica acta. 2007;1768(9):2293–9. pmid:17643388.
  91. 91. Barrett AJ. Cathepsin D: the lysosomal aspartic proteinase. Protein degradation in health and disease. 1980:37–42.
  92. 92. Hamilton KA, Nisbet AJ, Lehane MJ, Taylor MA, Billingsley PF. A physiological and biochemical model for digestion in the ectoparasitic mite, Psoroptes ovis (Acari: Psoroptidae). International journal for parasitology. 2003;33(8):773–85. pmid:12865077
  93. 93. Akov S. Blood digestion in ticks. Physiology of ticks. 1982:197–211.
  94. 94. Richards S, Gibbs RA, Weinstock GM, Brown SJ, Denell R, Beeman RW, et al. The genome of the model beetle and pest Tribolium castaneum. Nature. 2008;452(7190):949–55. pmid:18362917
  95. 95. Brennan CA, Anderson KV. Drosophila: the genetics of innate immune recognition and response. Annual Review of Immunology. 2004;22:457–83. pmid:15032585.
  96. 96. Vallet-Gely I, Lemaitre B, Boccard F. Bacterial strategies to overcome insect defences. Nature Reviews Microbiology. 2008;6(4):302–13. WOS:000253961800012. pmid:18327270
  97. 97. Welchman DP, Aksoy S, Jiggins F, Lemaitre B. Insect immunity: from pattern recognition to symbiont-mediated host defense. Cell Host Microbe. 2009;6(2):107–14. pmid:19683677.
  98. 98. Lindsay MA. microRNAs and the immune response. Trends in immunology. 2008;29(7):343–51. pmid:18515182.
  99. 99. Inohara N, Nunez G. ML—a conserved domain involved in innate immunity and lipid metabolism. Trends Biochem Sci. 2002;27(5):219–21. pmid:12076526.
  100. 100. Hetru C, Troxler L, Hoffmann JA. Drosophila melanogaster antimicrobial defense. J Infect Dis. 2003;187 Suppl 2:S327–S34. pmid:12792847.
  101. 101. Bulet P, Stocklin R, Menin L. Anti-microbial peptides: from invertebrates to vertebrates. Immunol Rev. 2004;198:169–84. pmid:15199962.
  102. 102. Iijima R, Kurata S, Natori S. Purification, characterization, and cDNA cloning of an antifungal protein from the hemolymph of Sarcophaga peregrina (flesh fly) larvae. J Biol Chem. 1993;268(16):12055–61. pmid:8505329.
  103. 103. Carlsson A, Engstrom P, Palva ET, Bennich H. Attacin, an antibacterial protein from Hyalophora cecropia, inhibits synthesis of outer membrane proteins in Escherichia coli by interfering with omp gene transcription. Infect Immun. 1991;59(9):3040–5. pmid:1715318; PubMed Central PMCID: PMC258132.
  104. 104. Hoffmann JA, Kafatos FC, Janeway CA, Ezekowitz RAB. Phylogenetic perspectives in innate immunity. Science. 1999;284(5418):1313–8. pmid:10334979
  105. 105. Matsuyama K, Natori S. Molecular cloning of cDNA for sapecin and unique expression of the sapecin gene during the development of Sarcophaga peregrina. J Biol Chem. 1988;263(32):17117–21. pmid:3182837.
  106. 106. Lowenberger C, Bulet P, Charlet M, Hetru C, Hodgeman B, Christensen BM, et al. Insect immunity: isolation of three novel inducible antibacterial defensins from the vector mosquito, Aedes aegypti. Insect Biochem Mol Biol. 1995;25(7):867–73. pmid:7633471.
  107. 107. Dobson AJ, Purves J, Kamysz W, Rolff J. Comparing selection on S. aureus between antimicrobial peptides and common antibiotics. PloS one. 2013;8(10):e76521. pmid:24204634; PubMed Central PMCID: PMC3799789.
  108. 108. Smith C, Wang X, Yin H. Caspases come together over LPS. Trends in immunology. 2015;36(2):59–61. pmid:25573609; PubMed Central PMCID: PMC4323722.
  109. 109. Gettins PG. Serpin structure, mechanism, and function. Chem Rev. 2002;102(12):4751–804. pmid:12475206.
  110. 110. Potempa J, Korzus E, Travis J. The serpin superfamily of proteinase inhibitors: structure, function, and regulation. J Biol Chem. 1994;269(23):15957–60. pmid:8206889.
  111. 111. Griswold CM, Matthews AL, Bewley KE, Mahaffey JW. Molecular characterization and rescue of acatalasemic mutants of Drosophila melanogaster. Genetics. 1993;134(3):781–8. pmid:8349109; PubMed Central PMCID: PMC1205515.
  112. 112. Jomori T, Natori S. Function of the lipopolysaccharide-binding protein of Periplaneta americana as an opsonin. FEBS letters. 1992;296(3):283–6. pmid:1537405.
  113. 113. Yu XQ, Kanost MR. Manduca sexta lipopolysaccharide-specific immulectin-2 protects larvae from bacterial infection. Dev Comp Immunol. 2003;27(3):189–96. pmid:12590970.
  114. 114. Yu XQ, Gan H, Kanost MR. Immulectin, an inducible C-type lectin from an insect, Manduca sexta, stimulates activation of plasma prophenol oxidase. Insect Biochem Mol Biol. 1999;29(7):585–97. pmid:10436935.
  115. 115. Yu XQ, Kanost MR. Immulectin-2, a lipopolysaccharide-specific lectin from an insect, Manduca sexta, is induced in response to gram-negative bacteria. J Biol Chem. 2000;275(48):37373–81. pmid:10954704.
  116. 116. Chen C, Durrant HJ, Newton RP, Ratcliffe NA. A study of novel lectins and their involvement in the activation of the prophenoloxidase system in Blaberus discoidalis. Biochem J. 1995;310 (Pt 1):23–31. pmid:7646450; PubMed Central PMCID: PMC1135849.
  117. 117. Pace KE, Baum LG. Insect galectins: roles in immunity and development. Glycoconjugate journal. 2004;19(7–9):607–14. pmid:14758086.
  118. 118. Gough PJ, Gordon S. The role of scavenger receptors in the innate immune system. Microbes Infect. 2000;2(3):305–11. pmid:10758407.
  119. 119. Peiser L, Gordon S. The function of scavenger receptors expressed by macrophages and their role in the regulation of inflammation. Microbes Infect. 2001;3(2):149–59. pmid:11251301.
  120. 120. Peiser L, Mukhopadhyay S, Gordon S. Scavenger receptors in innate immunity. Curr Opin Immunol. 2002;14(1):123–8. pmid:11790542.
  121. 121. Agaisse H, Burrack LS, Philips JA, Rubin EJ, Perrimon N, Higgins DE. Genome-wide RNAi screen for host factors required for intracellular bacterial infection. Science. 2005;309(5738):1248–51. pmid:16020693.
  122. 122. Franc NC, Dimarcq JL, Lagueux M, Hoffmann J, Ezekowitz RA. Croquemort, a novel Drosophila hemocyte/macrophage receptor that recognizes apoptotic cells. Immunity. 1996;4(5):431–43. pmid:8630729.
  123. 123. Franc NC, Heitzler P, Ezekowitz RA, White K. Requirement for croquemort in phagocytosis of apoptotic cells in Drosophila. Science. 1999;284(5422):1991–4. pmid:10373118.
  124. 124. Franc NC, White K. Innate recognition systems in insect immunity and development: new approaches in Drosophila. Microbes Infect. 2000;2(3):243–50. pmid:10758400.
  125. 125. Peiser L, De Winther MP, Makepeace K, Hollinshead M, Coull P, Plested J, et al. The class A macrophage scavenger receptor is a major pattern recognition receptor for Neisseria meningitidis which is independent of lipopolysaccharide and not required for secretory responses. Infect Immun. 2002;70(10):5346–54. pmid:12228258; PubMed Central PMCID: PMC128305.
  126. 126. Adema CM, Hertel LA, Miller RD, Loker ES. A family of fibrinogen-related proteins that precipitates parasite-derived molecules is produced by an invertebrate after infection. Proc Natl Acad Sci U S A. 1997;94(16):8691–896. pmid:9238039; PubMed Central PMCID: PMC23082.
  127. 127. Dimopoulos G, Casavant TL, Chang S, Scheetz T, Roberts C, Donohue M, et al. Anopheles gambiae pilot gene discovery project: identification of mosquito innate immunity genes from expressed sequence tags generated from immune-competent cell lines. Proc Natl Acad Sci U S A. 2000;97(12):6619–24. pmid:10841561; PubMed Central PMCID: PMC18678.
  128. 128. Dimopoulos G, Christophides GK, Meister S, Schultz J, White KP, Barillas-Mury C, et al. Genome expression analysis of Anopheles gambiae: responses to injury, bacterial challenge, and malaria infection. Proc Natl Acad Sci U S A. 2002;99(13):8814–9. pmid:12077297; PubMed Central PMCID: PMC124381.
  129. 129. Gokudan S, Muta T, Tsuda R, Koori K, Kawahara T, Seki N, et al. Horseshoe crab acetyl group-recognizing lectins involved in innate immunity are structurally related to fibrinogen. Proc Natl Acad Sci U S A. 1999;96(18):10086–91. pmid:10468566; PubMed Central PMCID: PMC17846.
  130. 130. Wang X, Fuchs JF, Infanger LC, Rocheleau TA, Hillyer JF, Chen CC, et al. Mosquito innate immunity: involvement of beta 1,3-glucan recognition protein in melanotic encapsulation immune responses in Armigeres subalbatus. Mol Biochem Parasitol. 2005;139(1):65–73. pmid:15610820.
  131. 131. Zdobnov EM, von Mering C, Letunic I, Torrents D, Suyama M, Copley RR, et al. Comparative genome and proteome analysis of Anopheles gambiae and Drosophila melanogaster. Science. 2002;298(5591):149–59. pmid:12364792.
  132. 132. Jiang H, Kanost MR. The clip-domain family of serine proteinases in arthropods. Insect Biochem Mol Biol. 2000;30(2):95–105. pmid:10696585.
  133. 133. Gobert V, Gottar M, Matskevich AA, Rutschmann S, Royet J, Belvin M, et al. Dual activation of the Drosophila toll pathway by two pattern recognition receptors. Science. 2003;302(5653):2126–30. pmid:14684822.
  134. 134. Wang L, Weber AN, Atilano ML, Filipe SR, Gay NJ, Ligoxygakis P. Sensing of Gram-positive bacteria in Drosophila: GNBP1 is needed to process and present peptidoglycan to PGRP-SA. EMBO J. 2006;25(20):5005–14. pmid:17024181; PubMed Central PMCID: PMC1618108.
  135. 135. Blandin S, Shiao SH, Moita LF, Janse CJ, Waters AP, Kafatos FC, et al. Complement-like protein TEP1 is a determinant of vectorial capacity in the malaria vector Anopheles gambiae. Cell. 2004;116(5):661–70. pmid:15006349.
  136. 136. Levashina EA, Moita LF, Blandin S, Vriend G, Lagueux M, Kafatos FC. Conserved role of a complement-like protein in phagocytosis revealed by dsRNA knockout in cultured cells of the mosquito, Anopheles gambiae. Cell. 2001;104(5):709–18. pmid:11257225.
  137. 137. Moita LF, Wang-Sattler R, Michel K, Zimmermann T, Blandin S, Levashina EA, et al. In vivo identification of novel regulators and conserved pathways of phagocytosis in A. gambiae. Immunity. 2005;23(1):65–73. pmid:16039580.
  138. 138. Jose AM, Hunter CP. Transport of sequence-specific RNA interference information between cells. Annual review of genetics. 2007;41(1):305–30.
  139. 139. Xia X, Yu L, Xue M, Yu X, Vasseur L, Gurr GM, et al. Genome-wide characterization and expression profiling of immune genes in the diamondback moth, Plutella xylostella (L.). Scientific Reports. 2015;5.
  140. 140. Mcnall RJ, Adang MJ. Identification of novel Bacillus thuringiensis Cry1Ac binding proteins in Manduca sexta midgut through proteomic analysis. Insect Biochemistry & Molecular Biology. 2003;33(10):999–1010.
  141. 141. Li X, Cui Z, Liu Y, Song C, Shi G. Transcriptome analysis and discovery of genes involved in immune pathways from hepatopancreas of microbial challenged mitten crab Eriocheir sinensis. PloS one. 2013;8(7):374-.
  142. 142. Ai J, Yong Z, Duan J, Yu Q, Zhang G, Fei W, et al. Genome-wide analysis of cytochrome P450 monooxygenase genes in the silkworm, Bombyx mori. Gene. 2011;480(1–2):42–50. pmid:21440608
  143. 143. Friedman R. Genomic organization of the glutathione S-transferase family in insects. Molecular Phylogenetics & Evolution. 2011;61(3):924–32.
  144. 144. Susanna V, Jing-Huan W, Mika RM. The Drosophila Toll signaling pathway. Journal of Immunology. 2011;186(2):649–56.
  145. 145. Boran A, Abdelnaser E, Nurper G, Grundler FMW, Andreas V, Heinz-Wilhelm D. Next Generation Sequencing Based Transcriptome Analysis of Septic-Injury Responsive Genes in the Beetle Tribolium castaneum. Plos One. 2013;8(1):e52004. pmid:23326321
  146. 146. Krishnamoorthy M, Jurat-Fuentes JL, Mcnall RJ, Andacht T, Adang MJ. Identification of novel Cry1Ac binding proteins in midgut membranes from Heliothis virescens using proteomic analyses. Insect Biochemistry & Molecular Biology. 2007;37(3):189–201.
  147. 147. Agaisse H, Perrimon N. The roles of JAK/STAT signaling in Drosophila immune responses. Immunological Reviews. 2004;198(1):72–82.
  148. 148. Deng YY, J.Q. L, Wu SF, Zhu YP, Chen YW, He FC. Integrated nr Database in Protein Annotation System and Its Localization. Computer Engineering. 2006;32(5):71–4.
  149. 149. Apweiler R, Bairoch A, Wu CH, Barker WC, Boeckmann B, Ferro S, et al. UniProt: the Universal Protein knowledgebase. Nucleic acids research. 2004;32(Database issue):D115–9. pmid:14681372; PubMed Central PMCID: PMC308865.
  150. 150. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, et al. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nat Genet. 2000;25(1):25–9. pmid:10802651; PubMed Central PMCID: PMC3037419.
  151. 151. Koonin EV, Fedorova ND, Jackson JD, Jacobs AR, Krylov DM, Makarova KS, et al. A comprehensive evolutionary classification of proteins encoded in complete eukaryotic genomes. Genome Biol. 2004;5(2):R7. pmid:14759257; PubMed Central PMCID: PMC395751.
  152. 152. Kanehisa M, Goto S, Kawashima S, Okuno Y, Hattori M. The KEGG resource for deciphering the genome. Nucleic acids research. 2004;32(Database issue):D277–80. pmid:14681412; PubMed Central PMCID: PMC308797.
  153. 153. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic acids research. 1997;25(17):3389–402. pmid:9254694; PubMed Central PMCID: PMC146917.
  154. 154. Finn RD, Bateman A, Clements J, Coggill P, Eberhardt RY, Eddy SR, et al. Pfam: the protein families database. Nucleic acids research. 2014;42(Database issue):D222–30. pmid:24288371; PubMed Central PMCID: PMC3965110.
  155. 155. Eddy SR. Profile hidden Markov models. Bioinformatics. 1998;14(9):755–63. pmid:9918945.