Skip to main content

Global transcriptome profiling and functional analysis reveal that tissue-specific constitutive overexpression of cytochrome P450s confers tolerance to imidacloprid in palm weevils in date palm fields

Abstract

Background

Cytochrome P450-dependent monooxygenases (P450s), constituting one of the largest and oldest gene superfamilies found in many organisms from bacteria to humans, play a vital role in the detoxification and inactivation of endogenous toxic compounds. The use of various insecticides has increased over the last two decades, and insects have developed resistance to most of these compounds through the detoxifying function of P450s. In this study, we focused on the red palm weevil (RPW), Rhynchophorus ferrugineus, the most devastating pest of palm trees worldwide, and demonstrated through functional analysis that upregulation of P450 gene expression has evolved as an adaptation to insecticide stress arising from exposure to the neonicotinoid-class systematic insecticide imidacloprid.

Results

Based on the RPW global transcriptome analysis, we identified 101 putative P450 genes, including 77 likely encoding protein coding genes with ubiquitous expression. A phylogenetic analysis revealed extensive functional and species-specific diversification of RPW P450s, indicating that multiple CYPs actively participated in the detoxification process. We identified highly conserved paralogs of insect P450s that likely play a role in the development of resistance to imidacloprid: Drosophila Cyp6g1 (CYP6345J1) and Bemisia tabaci CYP4C64 (CYP4LE1). We performed a toxicity bioassay and evaluated the induction of P450s, followed by the identification of overexpressed P450s, including CYP9Z82, CYP6fra5, CYP6NR1, CYP6345J1 and CYP4BD4, which confer cross-resistance to imidacloprid. In addition, under imidacloprid insecticide stress in a date palm field, we observed increased expression of various P450 genes, with CYP9Z82, CYP4BD4, CYP6NR1 and CYP6345J1 being the most upregulated detoxification genes in RPWs. Expression profiling and cluster analysis revealed P450 genes with multiple patterns of induction and differential expression. Furthermore, we used RNA interference to knock down the overexpressed P450s, after which a toxicity bioassay and quantitative expression analysis revealed likely candidates involved in metabolic resistance against imidacloprid in RPW. Ingestion of double-stranded RNA (dsRNA) successfully knocked down the expression of CYP9Z82, CYP6NR1 and CYP345J1 and demonstrated that silencing of CYP345J1 and CYP6NR1 significantly decreased the survival rate of adult RPWs treated with imidacloprid, indicating that overexpression of these two P450s may play an important role in developing tolerance to imidacloprid in a date palm field.

Conclusion

Our study provides useful background information on imidacloprid-specific induction and overexpression of P450s, which may enable the development of diagnostic tools/markers for monitoring the spread of insecticide resistant RPWs. The observed trend of increasing tolerance to imidacloprid in the date palm field therefore indicated that strategies for resistance management are urgently needed.

Background

Approximately 700 species of insects cause serious damage to agricultural crops worldwide, both in the field and during storage [1]. It has been almost seven decades since the introduction of the new concept of synthetic organic insecticides for successful insect control, and current pesticide expenditures exceed $10.2 billion globally [2, 3]. However, the recent intensive and repeated use of certain insecticides has led to resistance in many insects worldwide. One of the ways insects evolve to cope with insecticides by invoking internal enzyme systems to break down and excrete these compounds [4,5,6]; these functions are achieved through elevated activity of detoxifying enzymes [7] and xenobiotic transporters [8], respectively. In general, there are three main types of detoxifying enzymes: cytochrome P450-dependent monooxygenases (P450s), general esterases (Ests) and glutathione S-transferases (GSTs); and ATP-binding cassette (ABC) transporters are also involved in the metabolic detoxification and excretion of insecticides [7, 8]. One characteristic of these enzyme systems is increased levels of activity resulting from constitutive overexpression in resistant insects [7, 9,10,11,12,13]. In addition to greater efficiency, these enzyme systems may exhibit a broad spectrum of activity, allowing them to degrade many different insecticides and plant toxins that are lethal to insects. Overall, increased activity and elevated expression of the Est, GST and P450 enzyme systems are associated with insecticide resistance [7], and studies conducted over the past two decades provide evidence that induction and constitutive overexpression are responsible for the detoxification of insecticides in resistant insects [5, 6, 9, 14,15,16,17,18]. For instance, constitutive overexpression of detoxifying enzymes is associated with neonicotinoid resistance in Bemisia tabaci [19, 20], Nilaparvata lugens [21, 22], Leptinotarsa decemlineata [23,24,25,26], Tribolium castaneum [27], Laodelphax striatellus [28] and Bradysia odoriphaga [29]. It has been hypothesized that the induction and constitutive overexpression of detoxification enzymes and their activities in insects are involved in the adaptation of insects to their environment, detoxification of insecticides and development of resistance [5, 9, 23]. Such resistance has serious consequences in agriculture practice, and some of the major concerns include significant yield losses and environmental contamination, which have been estimated to translate to $3–5 billion per year on average [30, 31].

P450s are a common route by which insects become resistant to insecticides [4, 9, 32]. P450s metabolize and detoxify insecticides through oxidative reactions; thus, constitutive upregulation of P450 genes is considered a marker of the development of insecticide resistance [5, 6, 32]. P450 activity is found in many insect tissues, such as the fat body, midgut, Malpighian tubules, nervous system and antennae, with the highest activity usually being associated with gut tissues and the fat body [33,34,35]. Although many studies have identified overexpression of multiple P450 genes in various insects upon neonicotinoid exposure [36,37,38,39,40,41], there have been few studies on the P450 mechanisms of insecticide resistance in insect pests of trees. Here, we present the P450 repertoire of the red palm weevil (RPW), Rhynchophorus ferrugineus (Olivier), which is the most successful pest of palm trees, and we compare P450 expression between field-caught imidacloprid-resistant, laboratory-induced and susceptible weevils. After these analyses, we selected the highly induced P450 transcripts, and functional analysis of these genes was performed through RNA interference (RNAi).

RPW is among the world’s most invasive pest species of palm trees. It is indigenous to South Asia and has spread rapidly during the last three decades, mostly due to the transport of infested planting material to Middle Eastern countries, Africa, and Europe [42,43,44,45,46]. In Middle Eastern countries, this pest has wreaked havoc during the last two decades, with annual losses in the Gulf region due to the eradication of severely infested palms estimated at US $8 million in 2010 [42]. Moreover, R. ferrugineus is the most destructive pest of date palms in Saudi Arabia, where the Agriculture Ministry launched a $31 million national campaign to fight the pest in January 2011. Although various methods of control are still practiced in many countries, insecticide treatment using “trunk injection techniques” provides satisfactory control for infected trees [42, 43]. In Middle Eastern countries, next-generation insecticides belonging to the neonicotinoid (imidacloprid) and phenylpyrazole (fipronil) groups have commonly been used during the last decade as prophylactic and curative applications against R. ferrugineus in date palms [42, 47,48,49]. Although insecticide treatment initially provided satisfactory results, recent studies have shown that RPW continues to attack date palm trees even after many insecticide applications [50]. In this study, we sought to determine the P450 mechanism underlying the connection between the induction of imidacloprid in R. ferrugineus and the identification of a resistant population in the field. We focused on the cytochrome P450 complement (CYPome) because many previous biochemical and physiological studies have demonstrated the involvement of P450-mediated detoxification in imidacloprid resistance in insects [19, 20, 23, 27, 51]. We investigated potential differences in the expression levels of P450s that might be associated with imidacloprid resistance. We then used RNA interference to knock-down the expression of overexpressed P450 genes in RPWs, and toxicity bioassays were conducted to evaluate the phenotypic effects of these genes on imidacloprid resistance mechanisms. We compared the gene expression pattern of silenced resistance-related P450s between field-caught resistant RPWs and laboratory-induced RPWs. Our analyses confirmed that constitutive overexpression of P450s likely plays a role in developing tolerance to imidacloprid in RPWs. The results of this study are expected to contribute to a better understanding of resistance in R. ferrugineus in Middle Eastern countries and to provide more insight into the candidate P450s that are directly involved in neonicotinoid detoxification in R. ferrugineus.

Methods

Insecticide

Imidacloprid (Confidor® 350 SC), which was purchased from Bayer Corporation (Riyadh, Saudi Arabia), is a suspension concentrate containing 350 g.a.i./L imidacloprid (a 35% concentration of the active ingredient, imidacloprid).

Insect tissue collection and RNA extraction

The RPW collections were made with the direct permission of a cooperating landowner [Al-Kharj region (24.1500° N, 47.3000° E) of Saudi Arabia] in 2009; since then, the RPW culture has been maintained in our laboratory on sugarcane stems, as previously described [52]; this RPW strain is considered to be susceptible to imidacloprid. The tolerant strain of RPW was originally collected from a date palm orchard in the Al Qassim (25.8275° N, 42.8638° E) area, where farmers have continuously used imidacloprid-trunk-injection in last 10 years as a pest management strategy.

Total RNA was extracted from RPW fat body and gut tissues and previously reported whole body and antennae transcriptome data [53, 54] were screened and annotated for candidate P450 genes. For tissue specificity and reverse transcription polymerase chain reaction (RT-qPCR) analyses, samples of approximately 30 mg of tissue were collected from the antennae, proboscis, legs, wings, abdomen and thorax of adult males and females, and total RNA was isolated using the RNeasy Plus kit (Qiagen, USA) according to the manufacturer’s instructions. First-strand cDNA was synthesized from total RNA using SuperScript II Reverse Transcriptase (Invitrogen, USA) in accordance with the manufacturer’s instructions, and the quantity and quality of the total RNA and cDNA were validated using a NanoDrop spectrophotometer (Thermo Fisher Scientific, USA) and through PCR with the tubulin gene primer pair RferTubulin F/R (Additional file 3: Table S1) [46, 54, 55], respectively.

Identification and classification of R. ferrugineus P450 genes

Each P450 transcript from the RPW transcriptome dataset [53, 54] was identified via local or web-based searches using the BLASTx and BLASTn programs of the National Center for Biotechnology Information (NCBI) [56]. A quality control step was first performed on the raw sequencing reads using the NGS QC Toolkit [30]. BLAST hits with e-values < 1.0e− 5 were considered to be significant [57], and transcripts were putatively assigned to each contig based on BLASTx hit score values. The BLAST XML files were uploaded to Blast2GO, followed by mapping, transcript annotation, INTERPRO and Kyoto Encyclopedia of Genes and Genomes (KEGG) analyses [58, 59]. Transcripts containing errors leading to incorrect assembly were edited using Geneious v7.1.5 [60], and de novo assembly of the P450 isotigs were performed. The open reading frame (ORF) finder tool (NCBI) was used to determine the ORF of each P450 contig. Reads per kilobase per million (RPKM) values were calculated for the assembled transcripts based on mapping data [61]. We reassembled both the antennal and whole-body transcriptome P450s using Geneious v7.1.5 to generate a reliable transcriptome dataset, and the cytochrome P450 genes and nomenclature were determined based on the guidelines provided by the Committee on Standardized Cytochrome P450 Nomenclature [62, 63], and P450 genes were named by Dr. David Nelson (CYP nomenclature committee). Both Blast2GO and manual annotations were performed based on the nomenclature, and the recommended identity cut-offs of 40% for the CYP family and 55% for the subfamily were strictly followed. The transcripts were categorized into four CYP clans: CYP3, CYP4, CYP2 and mitochondrial clans [14].

Tissue-specific expression analysis

The tissue specificity of the P450 candidates was analyzed in various tissues collected from both male and female R. ferrugineus, including samples of the antennae, proboscis, thorax, abdomen, legs and wings. Total RNA extraction and cDNA synthesis were performed as described above. The gene-specific primers indicated in Additional file 3: Table S1 were used to amplify P450 genes, and tubulin and β-actin were used as internal controls [46, 55, 64, 65]. PCR amplification was performed using the following touchdown PCR program: 95 °C for 5 min, followed by 30 cycles of 95 °C for 1 min, 55 °C (±5 °C) for 30 s and 72 °C for 1 min, with a final cycle at 72 °C for 10 min. The PCR products were separated in 3% agarose gels and visualized with ethidium bromide staining.

Phylogenetic analysis of RPW P450s

The sequence homology and evolutionary aspects of the RPW P450s were analyzed by reconstructing a phylogenetic tree using protein sequences from RPW and closely related beetles or insects. The annotated P450 sequences from the representative insect species T. castaneum (red flour beetle), D. melanogaster (fruit fly), L. decemlineata (Colorado potato beetle), Agasicles hygrophila (alligator weed flea beetle) and Apis mellifera (European honey bee) were obtained from the NCBI database and the cytochrome P450 monooxygenase homepage (http://drnelson.uthsc.edu/CytochromeP450.html). Multiple sequence alignments were performed using MAFFT v.7 [66], with the E-INS-i iterative refinement strategy and default parameters. The LG + I + G + F substitution model was determined as the best-fit models of protein evolution by ProtTest 3.4 [67]. Phylogenetic reconstruction and analysis of the CYP clade distribution were performed via the recommended maximum likelihood method, with the statistical bootstrap support of 1000 replications, using PhyML 3.0 [68].

Relative expression analysis of P450s in the gut and fat body by RT-qPCR

cDNAs were prepared from RNA (~ 1 μg) extracted from the gut and fat body of 20-day-old insects, as noted above. The RT-qPCR experiments were performed according to the Minimum Information Required for Publication of Quantitative Real-Time PCR Experiments (MIQE) Guidelines [69, 70]. RT-qPCR was carried out using SYBR Green PCR Master Mix (Life Technologies, USA) with three biological and three technical replicates, according to the manufacturer’s instructions. The oligonucleotide primers for these analyses were the same as those used in the tissue-specific studies, and tubulin and β-actin (Additional file 3: Table S1) were employed to normalize gene expression. Relative P450 expression levels were measured via the 2−ΔCT method [71]. The following thermal cycling program was used to perform PCR amplification: 50 °C for 20 s (precycling), then 95 °C for 10 min (holding), followed by 40 cycles of 95 °C for 15 s and 55 °C for 40 s; then, a melting curve analysis stage consisting of 95 °C for 15 s, 60 °C for 1 min, 95 °C for 30 s and 60 °C for 15 s.. The RT-qPCR products were examined through 3% agarose gel electrophoresis and visualized via ethidium bromide staining.

Toxicity bioassay - induction of P450s using imidacloprid

Laboratory trials were conducted using a susceptible strain of RPW and field-caught RPWs (21-day-old males and females at a ratio of 1:1) to evaluate the efficiency of insecticide dilutions using a bioassay technique in which food is dipped into the insecticide [49, 72]. The numbers of moribund or dead adults were recorded daily, and moribund adults were counted as dead adults. Mortality was recorded for a period of 11 days. A preliminary laboratory toxicity bioassay was performed to identify a range of concentrations to be used for determining the LC50 values of imidacloprid. Based on the preliminary laboratory toxicity bioassay of Confidor® 350 SC, concentrations of 1, 2, 4, 6 and 8 mM (250, 500, 1000, 1500 and 2000 ppm, respectively), were prepared using dH2O; sugarcane stems treated with dH2O were used as the control. For the determination of LC50 values, the bioassay results were analyzed via probit analysis using the Ldp Line program (http://www.ehabsoft.com/ldpline/). Mortality was corrected using Abbott’s formula for each probit analysis. In this study, we considered 33% of the median lethal value as the sublethal concentration. Biological triplicate samples of RPWs of both the susceptible strain (to obtain an induced strain, hereafter referred to as ind) and field-caught strains (hereafter, res) were exposed to sublethal concentrations of imidacloprid for 11 days. Adults of the susceptible strain of RPW were treated with distilled water to obtain a negative control (hereafter, sus). After 11 days, sus, ind and res weevils were dissected, and the gut and fat body tissues were removed, placed in ice-cold phosphate-buffered saline (PBS) and stored in RNAlater® solution (Ambion, USA). Total RNA was extracted, and cDNA was synthesized as described above.

Overexpression profiling of P450s in sus, ind and res populations of R. ferrugineus

Expression profiling of P450s induced in response to imidacloprid was evaluated by comparing P450 expression in sus, ind and res samples via quantitative real-time PCR (RT-qPCR). Tissues from the major detoxification sites (i.e., the gut and fat body) were removed from the three samples and stored in RNAlater® solution (Ambion, USA) prior to RNA extraction. RNA extraction and cDNA synthesis were conducted as described above, and RT-qPCR was performed with the same primers used for the tissue specificity analysis (Additional file 3: Table S1) with Power SYBR® Green PCR Master Mix (Life Technologies, USA). Of the 77 P450 genes identified in this study, 50 were selected for RT-qPCR expression profiling based on their tissue-specific expression patterns, RPKM values and available support in the literature. Twenty-seven P450s were discarded based on previous analyses and published evidence of a lack of expression in relevant tissues. From literature-supported important CYP families, specifically CYP6, CYP4, CYP9 and CYP345, a total of 50 P450s from R. ferrugineus were selected for RT-qPCR analysis using cDNA from the sus, res and ind populations of RPWs (Additional file 3: Table S1). Thermal cycling was conducted using a 7500 Fast Real-Time PCR system (Life Technologies, USA) at 50 °C for 20 s (precycling), then 95 °C for 10 min (holding), followed by 40 cycles of 95 °C for 15 s and 55 °C for 35 s; then, a melting curve analysis stage consisting of 95 °C for 15 s, 60 °C for 1 min, 95 °C for 30 s and 60 °C for 15 s was conducted. The reactions were performed in triplicate with three biological replicates for each tissue type/sample, and relative quantification of gene expression in each target tissue was calculated using the comparative 2-ΔΔCт method [73] in comparison with the same tissues from sus insects. The RPW tubulin and β-actin genes (Additional file 3: Table S1) were used as endogenous controls [46, 54, 55].

Cluster analysis, expression profiling and statistical analysis

Hierarchical cluster analysis was performed on the RT-qPCR data using Cluster software [74]. The 2-ΔCт values from the RT-qPCR analysis were log transformed, and data from each sample and tissue were clustered on the basis of the average linkage distance between the median values [71, 74]. The significance of differences in P450 transcript expression were calculated between ind and res samples compared to sus using the paired t-test, with an alpha significance level of 0.05 (p < 0.05), using SPSS (v24). Multiple-comparison testing with the least significant difference (LSD) test was performed to assess the differential expression of transcripts within each CYP family (p < 0.05) using SPSS (v24).

P450 silencing through RNA interference (RNAi) and toxicity bioassays

The three most important and differentially expressed P450 genes (CYP9Z82, CYP345J1 and CYP6NR1) were selected for RNAi experiments. We selected CYP9Z82, CYP6NR1 and CYP345J1 for the RNAi experiments because the first two P450s were highly expressed and differentially expressed, respectively, and the remaining P450 was a paralog of Drosophila Cyp6g1 [41, 75, 76] and B. tabaci CYP4C64 [19], which confer cross-resistance to imidacloprid. The PCR amplification reactions were carried out as follows: 95 °C for 5 min, followed by 30 cycles of 95 °C for 1 min, 56 °C for 30 s and 72 °C for 2 min and, finally, one cycle at 72 °C for 10 min.

The full-length amplified PCR products within the expected size range were gel purified with the Wizard® SV Gel and PCR Clean-Up System (Promega, WI, USA), followed by ligation into the pGem-T Easy vector (Promega) and transformation into the JM109 Escherichia coli system. The plasmid products were isolated and sequenced with an ABI 3500 genetic analyzer (Life Technologies) using vector primers M13-F/R (Additional file 3: Table S1). We used plasmids containing the full-length P450s ORF as template DNA to synthesize double-stranded RNA (dsRNA). The vector primer, M13-F and reverse primers targeting the ORFs with a T7 overhang (Additional file 3: Table S1) were used to amplify and linearize the ORFs, which were rechecked via direct sequencing (ABI 3500, Life Technologies, USA). dsRNA synthesis was performed using the MEGAscript RNAi Kit (Life Technologies, USA) according to the manufacturer’s instructions, and the results were quantified using a NanoDrop 2000 (Thermo Scientific, DE, USA). The integrity and efficiency of dsRNA duplex formation were examined through 1% agarose gel electrophoresis. The RNAi experimental design was set up into three separate controls which were universal negative dsRNA control (Integrated DNA Technologies, Leuven, Belgium) (hereafter, negative control), no-injection (hereafter, NI) and nuclease free water-injected (hereafter, NFW) group. We selected 10-day-old R. ferrugineus pupae for the RNAi experiments, and 100 ng/μL dsRNA (in 20 μL) was injected, after which the dsRNA-injected and RPW pupae were maintained as previously described [55, 64]. The emerged adults were transferred to a separate box containing a piece of fresh sugarcane, and after 7 days were further subjected to toxicity assays using sublethal concentrations of imidacloprid (~ 1.1 mM) for 11 days. Seven days after injection, cDNAs prepared from RNA extracted from the fat body of each individual insect and the gene expression pattern of RPW adults test sample (CYP9Z82, CYP345J1 and CYP6NR1) was compared with that of dsRNA control, NI and NFW weevils to obtain the percentage knockdown of each P450 gene. RT-qPCR reactions were carried out using SYBR Green PCR Master Mix (Life Technologies, USA) according to the manufacturer’s instructions, with three biological and three technical replicates. Eleven days after toxicity assays, all surviving adults were exposed to a temperature of − 20 °C until completely immobilized (approximately 10 min), at which time RNA extraction and cDNA synthesis was carried out as described below, from the gut and fat body tissues for each experimental group. Thereafter, an analysis of phenotypic effects and comparison of P450 gene expression pattern between silenced weevils and res and ind RPWs were carried out as described above. The toxicity assays were carried out with two separate controls, P450 dsRNA-injected RPW adults were treated with nuclease-free water (hereafter, n-trt) or were left uninjected (same as sus). P450 dsRNA-injected RPW adults were subjected to toxicity assays (hereafter, trt) and were compared with the sus, ind and res groups to determine the gene-silencing effect on CYP gene induction and constitutive overexpression.

Phenotypic effects and gene silencing validation via RT-qPCR

During the toxicity assay, conducted over a period of 11 days, the adult weevil survival rate was monitored daily, and Kaplan-Meier survival analysis with the log rank test was performed using SPSS (v24) to determine the survival function in the treated and control groups. cDNAs prepared from RNA extracted from the gut and fat body of each individual insect (surviving) in the experimental (n-trt and survived adult-trt) and sus, ind and res RPW groups were also used as templates for RT-qPCR. Reactions were carried out using SYBR Green PCR Master Mix (Life Technologies, USA) according to the manufacturer’s instructions, with six biological and three technical replicates. Tubulin and β-actin primers were used to normalize gene expression (Additional file 3: Table S1). The relative expression levels of P450s in the silenced vs. control groups were measured via the 2−ΔΔCT method [71, 73]. PCR amplification, data analysis and gel evaluation were performed as described above. Difference analysis was performed by using Student’s t test, followed by Duncan’s multiple comparison test, with SPSS (v24) software. A value of p < 0.05 was considered significant.

Results

Identification of putative RPW P450 candidates

The RPW transcriptome dataset sequence assembly and annotation are summarized in Additional file 4: Table S2. Our extensive search for P450s in the RPW antennal and whole-body transcriptome datasets resulted in the identification of 101 putative candidates, which were further screened for duplications through a BLASTx homology search within the Blast2GO annotation program (Additional file 5: Table S3). The 24 removed transcripts included hypothetical proteins as well as repeated transcripts from the antenna and whole-body databases. After combined assembly of the antennal and whole-body transcriptome datasets of R. ferrugineus, we obtained 77 transcripts encoding P450s (Additional file 5: Table S3). The adopted nomenclature was according to the guidelines provided by the Committee on Standardized Cytochrome P450 Nomenclature [63] (Additional file 5: Table S3), and the transcripts were categorized into four CYP clans, namely, the CYP3, CYP4, CYP2 and mitochondrial clans [14, 62, 63, 77, 78]. The official nomenclature obtained from Dr. David R. Nelson (CYP nomenclature committee) is presented in Additional file 5: Table S3. The number of R. ferrugineus representatives in each P450 clan and the BLASTx results used for nomenclature are provided in Table 1 and Additional file 5: Table S3, respectively. A total of 16 CYP families were identified, including CYP6 (42.8%) and CYP4 (23%) within the total identified CYPs. The CYP9 family and mitochondrial clan accounted for only 10% of the total R. ferrugineus CYPs (Additional file 5: Table S3). The largest families included CYP6, with 33 genes, followed by CYP4, with 14 genes (Table 1). The CYP2 clan contained fewer genes and exhibited less evolutionary differentiation than the other clans.

Table 1 Numbers of genes in P450 clans and families identified in R. ferrugineus

P450 phylogenetic analysis

Phylogenetic analysis was conducted based on amino acid sequences using R. ferrugineus and four other insects (T. castaneum, L. decemlineata, A. hygrophila, B. tabaci and Drosophila melanogaster) [20, 23, 75, 76, 79] to identify gene orthologs and paralogs. RPW contains highly expanded CYP3 (5 families, 43 individual genes) and CYP4 (8 families, 23 individual genes) clans, most notably in family 4, family 6 and family 9 (Table 1). These results show that there is a high frequency of species-specific CYP genes in the CYP4 and CYP6 family, as reported in other coleopterans, T. castaneum [79] and L. decemlineata [23]. We selected a best fit-model of P450 protein evolution, and maximum-likelihood trees were constructed for CYP2, CYP3, CYP4 and mitochondrial clans.

In R. ferrugineus CYP4 clan, 8 families (CYP4, CYP341, CYP433, CYP412, CYP15, CYP326, CYP349 and CYP410) and 16 subfamilies were found (Table 1, Fig. 1). Members of the CYP4 clan show wide diversity in both their sequences and functions (Fig. 1). In Clan 4, CYP4 family is the largest gene family and we identified CYP4G, CYP4BD, CYP4Q, CYP4AA, CYP4LE, CYP4KX, CYP4BQ and CYP4D (Fig. 1). Most RPW P450s in the CYP4 clan had gene orthologs in L. decemlineata and T. castaneum, or exhibited gene expansions with close phylogenetic relationships. Gene expansions were noticed within the CYP4 clan, particularly the cluster of CYP4BD (Fig. 1). R. ferrugineus CYP4 tree also revealed that CYP4G clade belongs to a group containing CYP4G isoforms from T. castaneum, L. decemlineata, D. melanogaster and B. tabaci, with > 90% bootstrap support (Fig. 1). Except in CYP433 and CYP4KX, all other CYP genes in R. ferrugineus was clustered with their related genes in other insects. Interestingly, CYP4Q6 and CYP4BQ9from R. ferrugineus were found to be clustered with L. decemlineata LdCYP4Qb and LdCYP4e, respectively, which has been reported to be more than five-fold overexpressed in imidacloprid resistance (Fig. 1) [23]. The phylogenetic analysis identified an ortholog of L. decemlineata (CYP4Qc/CYP4Q10) that is involved in imidacloprid resistance [23] in R. ferrugineus. Similarly, the phylogenetic analysis identified a paralog of B. tabaci CYP4C64 (GenBank JX144366) that is involved in imidacloprid resistance [20] in R. ferrugineus (CYP4LE1), with 67% bootstrap support (Fig. 1). In the L. decemlineata CYP4Q subfamily, two genes (CYP4Qc and CYP4Qb) are overexpressed upon imidacloprid induction [23], and we identified two candidates, RferCYP4Q6 and RferCYP4KX2, that were similar in R. ferrugineus (Fig. 1). The CYP4BD subfamily exhibited the greatest increase in diversity compared with other insects, possibly reflecting adaptation to environmental xenobiotics.

Fig. 1
figure 1

Maximum likelihood consensus tree of CYP4 clan. CYP4 clan sequences from Tribolium castaneum, Bemisia tabaci, Leptinotarsa decemlineata and Drosophila melanogaster P450s were used as references to classify R. ferrugineus CYP4 (red). Multiple sequence alignments were performed using MAFFT v.7 with the E-INS-i iterative refinement strategy and default parameters, and the LG + I + G + F substitution model was determined as the best-fit model of protein evolution by ProtTest 3. The maximum likelihood (ML) analysis was computed using PhyML 3.0, with 1000 bootstrap replications and the bootstrap values are indicated at the nodes. P450s from different species were marked with different colors. The phylogenetic tree was visualized using FigTree (http://tree.bio.ed.ac.uk/software/figtree/) and branch appearance was colored based on the bootstrap values. Scale = 0.5 amino acid substitutions per site

In R. ferrugineus, we identified four CYP families (CYP6, 9, 324, 345, 347) and 20 subfamilies in the CYP3 clan and the CYP6 family is the largest family that has 35 individual genes. R. ferrugineus CYP6 family had 13 subfamilies, including CYP6AB, CYP6AE, CYP6BW17, CYP6BX5, CYP6CR10, CYP6DG2, CYP6HY3, CYP6MS2, CYP6NR1, CYP6NS1, CYP6NT1, CYP6NU1 and CYP6NV1. Genes in the CYP6 -family showed multiple expansions in R. ferrugineus, resulting in species-specific clusters and beetle-specific clusters (Fig. 2). During the construction of the phylogenetic tree, we used CYP3 clan from T. castaneum and L. decemlineata to identify beetle-specific expansion of CYP6 genes, with D. melanogaster as an outgroup. These results suggest that lineage-specific expansions of CYPs have occurred in Coleoptera and Diptera [6] (Fig. 2). Significant species-specific expansion and divergence may also have occurred in R. ferrugineus, particularly in the clusters containing CYP6BW, CYP6AB, CYP6BX, and CYP6DG (Additional file 5: Table S3; Fig. 2). The CYP6BW subfamily showed the greatest increase in diversity compared with other insects, possibly reflecting adaptation to environmental xenobiotics. The CYP6 genes CYP6EH1 CYP6FA1 and CYP6BH from L. decemlineata have been reported to be overexpressed in the presence of imidacloprid [23], and we identified orthologs in R. ferrugineus (CYP6CR, CYP6fra1 and CYP6DG, respectively), with sound bootstrap support (> 90%) (Fig. 2). Similarly, as shown in Fig. 2, one CYP6 gene from L. decemlineata, CYP6EZ1 [23], which is associated with imidacloprid resistance, forms a clade with two R. ferrugineus CYP6 candidates, CYP6CR10 and CYP6CR11, with 90% bootstrap support. Phylogenetic analysis identified a paralog of D. melanogaster CYP6g1 (GenBank: AAF58557.1), a gene with an established association with imidacloprid resistance [80], in R. ferrugineus (CYP345J1); this gene is also similar to T. castaneum CYP354A1 (GenBank: EFA12856.1) [27] (Fig. 2). The CYP6 gene CYP6AY1 from Nilaparvata lugens [22] has been reported to be overexpressed in the presence of imidacloprid, and we identified a paralog in R. ferrugineus (CYP6ABf11), with 82% bootstrap support, also forms a clade with R. ferrugineus CYP6ABf6, CYP6ABf9, and CYP6ABf10. Close to this clade is represented by a highly expressed CYP6CM1vQ from B. tabaci (51% bootstrap support), which is associated with imidacloprid resistance (GenBank: ACA51846.1) [19] (Fig. 2). All CYP354 group together in one cluster, with 84% bootstrap support (Fig. 2). This gene appears to undergo expansion by gene duplication forming allelic variants with novel functions.

Fig. 2
figure 2

Maximum likelihood consensus tree of CYP3 clan. The phylogenetic tree reconstruction was performed using the same methods mentioned in Fig. 1, with LG + I + G + F substitution model determined as the best-fit model of protein evolution. P450s from Tribolium castaneum, Bemisia tabaci, Leptinotarsa decemlineata and Drosophila melanogaster P450s were used in the analysis to classify R. ferrugineus CYP3 clan. Additionally, Clan 3 candidates with reported imidacloprid resistance were added from Laodelphax striatellus and Nilaparvata lugens. P450s from different species were marked with different colors. The CYP6g1 clade is highlighted in yellow. The phylogenetic tree was visualized using FigTree (http://tree.bio.ed.ac.uk/software/figtree/) and branch appearance was colored based on the bootstrap values. Scale = 0.5 amino acid substitutions per site

In the clan 3, CYP9 family was the second biggest in R. ferrugineus that consisted of three subfamilies CY9Z, CYP9CS and CYP9AN. The phylogenetic analysis of CYP9 identified two major clades: one containing CYP9Z subfamily genes from L. decemlineata, T. castaneum and R. ferrugineus (96% bootstrap support); and a second containing CYP9A subfamily genes from T. castaneum, L. decemlineata, R. ferrugnieus and D. melanogaster (100% bootstrap support), with a high degree of species-specific expansion (Fig. 2). Among them, all CYP9Z group together in one cluster, also includes CYP9CS with 99% bootstrap support (Fig. 2). The CYP9 families are relatively highly conserved across insect species and show diversity in terms of both sequence and function.

Phylogenetic analyses was performed for the R. ferrugineus mitochondrial clan and clan 2 with P450s identified in T. castaneum, D. melanogaster, L. decemlineata, A. hygrophila and Apis mellifera. R. ferrugineus CYP2 and mitochondrial clans showed a high degree of 1:1 orthology with those from other insect species. Within the clan 2, CYP303 A1showed high level of 1:1:1:1 orthologies. Three P450s were found in clan 2, belonging to CYP18, CYP305 and CYP306 families (Fig. 3). Of these, CYP306A1 was encoded by the Phantom gene from other insects, showed high sequence similarities with. T. castaneum and A. hygrophila (Fig. 3). In the clan 2, CYP18, CYP305 and CYP306 families, the tree branching showed that all coleopterans, R. ferrugineus, T. castaneum, L. decemlineata and A. hygrophila were grouped together and evolutionally separated from A. mellifera and D. melanogaster (Fig. 3). The major difference between the R. ferrugineus and T. castaneum in the clan 2 was possible loss of the CYP303, CYP304, CYP307 and CYP15 paralogs or orthologs in the R. ferrugineus.

Fig. 3
figure 3

Maximum likelihood consensus tree of CYP2 and Mitochondrial clans. The phylogenetic tree reconstruction was performed using the same methods mentioned in Fig. 1, with LG + I + G + F substitution model determined as the best-fit model of protein evolution. P450s from Tribolium castaneum, Bemisia tabaci, Leptinotarsa decemlineata and Drosophila melanogaster P450s were used in the analysis to classify R. ferrugineus CYP2 and Mitochondrial clans. Additionally, candidates from Agasicles hygrophila and Apis mellifera were added for classification purposes. P450s from different species were marked with different colors. The phylogenetic tree was visualized using FigTree (http://tree.bio.ed.ac.uk/software/figtree/) and branch appearance was colored based on the bootstrap values. Scale = 2.0 amino acid substitutions per site

In the mitochondrial clan in R. ferrugineus, eight P450s were found in six families of CYP12, CYP49, CYP301, CYP314, CYP315 and CYP334, and P450s, CYP314A1 and CYP315A1, encode orthologs of the Halloween genes Shade (shd) and Shadow (sad) [81], respectively from other insect species (Additional file 5: Table S3). In the mitochondrial clan, R. ferrugineus P450s were closely related to the coleopteran clade, namely CYP12, CYP301, CYP314 and CYP315 (Fig. 3). The major difference between the R. ferrugineus and T. castaneum in the mitochondrial clan was possible loss of the CYP302 and CY353 paralogs or orthologs in the R. ferrugineus. R. ferrugineus CYP301, CYP314 and CYP315 shared high sequence similarities with corresponding proteins in D. melanogaster, A. mellifera, T. castaneum, L. decemlineata and A. hygrophila.

Tissue-specificity analysis

The tissue-specific expression analysis indicated that most P450s are expressed in multiple tissues, with notable exceptions such as leg-specific CYP6ABfra9 and proboscis-specific CYP305F1 (Fig. 4). CYP306A1 showed high expression levels in antenna, abdomen and thorax tissues. CYP433A3 was found to be expressed exclusively in tissues such as the antenna, proboscis and thorax, and CYP4fra, from clade 4, showed localized expression in the antenna, proboscis and abdomen (Fig. 4). In contrast, low expression of CYP6fra2 was observed in antennal tissues, and notably, a few transcripts with otherwise ubiquitous expression were absent from the wings (Fig. 4). The mitochondrial P450s were ubiquitously expressed in all tissues analyzed, except for CYP12V3, which was not expressed in the legs and wings (Fig. 4).

Fig. 4
figure 4

Relative tissue-specific expression of P450 transcripts, as analyzed via PCR and visualized in 3% agarose gels. Antenna Male (AM), Antenna Female (AF), Proboscis (S), Legs (L), Thorax (Tx), Abdomen (Ab) and Wings (W) indicate the source tissues. The expression of all P450s in the gut and fat body (FB) tissues was quantified via RT-qPCR, and the mean fold-changes in gene expression compared to tubulin and β-actin are provided under RQ. The oligonucleotide primer details, annealing temperature and amplification size (bp) are given in Additional file 1: Table S1. The 3 kb DNA ladder was used as a marker

Relative expression analysis of P450s in gut and fat body tissues

The expression of all P450s in the sus R. ferrugineus gut and fat body tissues was quantitatively measured, and the relative quantification (RQ) values are provided in Fig. 4. Based on the RQ data, CYP9Z82, CYP6BX6, CYP6DG7, CYP6BW18, CYP6fra7, CYP6BW21, CYP6NU1 and CYP6BW19 are highly expressed P450s in the fat body of R. ferrugineus. Compared to all CYPs, CYPZ82, CYP6DG7 and CYP6BX6 were found to be highly expressed in the fat body (Fig. 4). Other candidate genes showing high expression in the gut were CYP6CR10, CYP6BW21, CYP9Z82, and CYP18A1. One proboscis-specific candidate gene also displayed moderate expression (mean-fold-change normalized by tubulin and β-actin gene expression), as shown in Fig. 4.

Toxicity bioassay: induction of P450s using imidacloprid

The probit analysis of different Confidor® 350 SC treatments (1–8 mM) against susceptible and field-caught adult RPWs over 4 days of exposure is presented in Additional file 6: Table S4. The action of imidacloprid against RPWs was comparatively rapid at the 8 mM concentration, where 75% mortality was observed for the susceptible RPWs within 1 d, whereas 41.67% mortality was observed for the field-caught adult RPWs. In susceptible RPWs, 100% mortality was observed within 4 d at the 8 mM concentration, whereas field-caught RPWs exhibited 66.67% mortality. In susceptible RPWs, probit analysis identified 3.44 mM as the LC50 value for 4 days of exposure (Additional file 6: Table S4), whereas in field-caught RPWs, 5.97 mM was identified as the LC50 value for 4 days of exposure (Additional file 6: Table S4). Thus, we selected ~ 1.14 mM, which represents 33% of the median lethal value, as the sublethal concentration for subsequent experiments. Biological triplicate samples of RPWs of both susceptible and field-caught strains (to obtain res) were exposed to an ~ 1.14 mM concentration of imidacloprid for 11 days to obtain induced (ind) and res strains, respectively. The susceptible strain of RPW adults was treated with double distilled water (hereafter, sus) for 11 days to obtain a negative control. No RPWs died during the experimental period. After 11 days, the adult weevils were classified as sus (negative control), ind (induced) and res (positive control); the tissues were dissected; and CYP expression profiling was carried out.

Overexpression profiling of P450s in sus, ind and res populations of R. ferrugineus

To identify P450s that might be induced in the sus, ind and res populations, we generated a relative expression profile of 50 P450s belonging to the CYP4, CYP6, CYP9 families and CYP2 clan. Candidates were selected based on tissue-specific expression profiles and RPKM and RQ values. Nine P450 candidates showing no expression in abdomen tissues (Fig. 4) and eight candidates that had not been previously associated with insecticide resistance, such as CYP349 (2), CYP341 (2), CYP305 (2) CYP410 (1) and CYP (1), were excluded (Additional file 5: Table S3). The induction of CYP genes was validated via quantitative PCR (RT-qPCR) on independent biological samples. The results showed a differential P450 induction pattern in gut and fat body tissues.

Induction in gut tissues

In gut tissues, only two CYP transcripts (CYP9Z82 and CYP4BD13) were found to be induced with a more than five-fold increase in expression compared to sus strains (Fig. 5). Apart from these CYP9 and CYP4 family candidates, CYP6NT1 is another candidate from the CYP6 family that is induced (Additional file 1: Figure S1). Interestingly, we observed that the same CYPs were overexpressed in res insects, which highlights the importance of these P450s in the detoxification function in the gut (Fig. 5). Apart from these three genes, four additional P450s formed the CYP4 clan; specifically, CYP4BQ9, CYP4BG7, CYP4BD4fra1 and CYP433A3 were overexpressed in res insects. The remaining P450s showed similar expression to that in the sus insects, exhibiting a mean fold-change in expression of less than one compared to sus (Fig. 5).

Fig. 5
figure 5

P450s differentially expressed in the gut and fat body tissues of ind (treated) and res (field-caught) beetles compared to sus insects are provided with colour codes indicating the mean fold change in expression

Induction in fat body tissues

Similarly, induction in fat body tissues was also calculated as the mean fold-change increase with respect to the sus group. In ind insects, four P450s (CYP9Z82, CYP6fra5, CYP6fra7 and CYP6NR1) were highly induced by more than 50-fold compared to the sus group. Five P450s (CYP4BD4, CYP6NT1, CYP6fra1 and CYP345J1) showed a more than 10-fold increase in expression compared to the sus group (Fig. 5). Apart from these 9 candidate genes, the expression of the remaining P450s was similar to the sus group, presenting mean fold-changes ranging from 0.01 to 5. In the res group, eight P450s were found to be overexpressed with a more than 100-fold increase in expression compared to the sus group, which were CYP9Z82, CYP4BD4, CYP6fra5, CYP6fra7, CYP345J1, CYP6fra1, CYP6fra3 and CYP6NR1. Ten additional P450s showed a more than 10-fold increase in expression compared to the sus strains (Fig. 5), and the remaining P450s exhibited a mean fold-change of less than five compared to the sus strains. Interestingly, we observed common P450s between the ind and res groups sharing transcripts from all three CYP9, CYP4 and CYP6 families and a significantly high number (18) of P450s induced in fat body tissues compared to only two in the gut. The variation in expression between the ind and res gut tissues was significantly different (p = 0.015). Comparison between the gut tissues of different groups revealed a significant difference between the res and sus RPW gut samples, with a p value of 0.017.

Differential induction of P450s

We found all three major CYP families (CYP6, CYP4 and CYP9) to be differentially induced in gut and fat body tissues upon treatment with imidacloprid. Induction in the gut was no higher than10-fold, whereas in fat body tissues, 100-fold induction of some P450 genes was observed compared to the sus insects. Two P450s (CYP9Z82 and CYP4BD13) were induced in gut tissues, whereas up to 25 P450s were induced in fat body tissues (Fig. 5), including CYP9Z82, CYP6fra5, CYP6fra7, and CYP6NR1, as highly induced genes. Apart from the ubiquitous overexpression of CYP9Z82, most of the P450s induced in fat body tissues were CYP4 and CYP6 candidates (Figs. 5 and 6). Additionally, three P450 genes (CYP9Z82, CYP6NT1 and CYP4BD4fra1) were common to both tissues of ind and res R. ferrugineus.

Fig. 6
figure 6

Venn diagram representing the overlap of P450 overexpression in induced (ind) and field-caught (res) RPW tissue samples: ind GUT, res GUT, ind fat body (ind FB) and res FB

In the ind gut, five CYP4 transcripts, two CYP6 transcripts and one CYP9 were overexpressed. In the res fat body, 4 CYP6 transcripts, two CYP4s and one CYP9 were overexpressed. We also identified CYP6fra8 as being exclusively induced in the gut tissue (Fig. 6).

P450s induced in different tissue samples were detected using RT-qPCR, and the Venn diagram presented in Fig. 6 depicts overlapping P450s in the four tissue types. CYP9Z82, CYP6NT1 and CYP4BD4fra1 were the three P450s overlapping in all four samples examined. Among the P450s overexpressed in gut tissues, four P450s were common between the ind and res samples, including CYP4BD13, CYP4BG7, CYP4BQ9 and CYP433A3. Between the res and ind fat body tissues, six P450s were common, which comprised CYP6fra5, CYP6fra7, CYP6NR1, CYP6fra1, CYP345J1 and CYP4KX2. We identified three P450s that were overexpressed in all tissues compared, which were CYP9Z82, CYP6NT1 and CYP4BD4fra1 (Fig. 6).

Expression profiling reveals P450 clusters

Cluster analysis was performed on the RT-qPCR data (2- ΔCt) from gut and fat body tissues of the sus, ind and res groups revealed possible synergy between P450 transcripts. The result (Fig. 7) shows six clusters with characteristic expression patterns. Cluster one comprised 12 P450s, mainly induced in fat body tissues and includes a subgroup of seven P450s with overexpression in both ind and res fat body tissues (CYP6NT1, CYP345J1, CYP6fra7, CYP6fra5 and CYP6fra1, CYP4BD4f1, CYP9CS2). Cluster 2 includes five P450s induced in the sus fat body (Fig. 7) and cluster 3 includes 7 candidates overexpressed in the res gut. Cluster 4 shows six interesting candidates with overexpression in both ind and res gut tissues, which were CYP4BG7, CYP4BD13, CYP6BW3, CYP306, CYP4BQ9 and CYP433A3. Cluster 5 includes seven candidates induced exclusively in the sus fat body tissues, and 6th cluster includes 11 P450s induced mainly in sus gut tissues (Fig. 7). A cluster analysis conducted separately on gut and fat body tissues is provided as supporting information (Additional file 2: Figure S2). Additional file 2: Figure S2A shows three clusters identified in gut tissues with overexpression in the res gut and ind gut, while the remaining candidates presented low expression. In fat body tissues Additional file 2: Figure S2B, two clusters were identified as overexpressed cluster 1 in res insects and the remaining less induced P450s in another cluster (Additional file 2: Figure S2B).

Fig. 7
figure 7

Expression profile of P450s in the gut and fat body tissues of sus, ind and res beetles. The relative expression of P450s in each tissue compared to tubulin and β-actin gene expression (2-ΔCт) is represented; red indicates overexpression, while green represents low expression, and black represents moderate expression. The six major clusters identified are indicated in the tree

RNAi-based gene silencing of P450s

We selected CYP9Z82, CYP6NR1 and CYP345J1 for the RNAi experiments because the first two P450s were highly expressed and differentially expressed, respectively, and the remaining P450 was a paralog of Drosophila Cyp6g1 [41, 75, 76, 82, 83] and B. tabaci CYP4C64 [19], which confer cross-resistance to imidacloprid. The RT-qPCR results revealed 79.42, 54.99 and 69.76% reduction in the 2-ΔΔCt values respectively in CYP6NR1, CYP345J1 and CYP9Z82 RPW adults compared to the dsRNA negative control, and a significant difference to values for NFW and NI RPWs (Fig. 8). Results from one-way ANOVA showed that injecting dsRNA for the P450 genes resulted in a significant reduction in the mRNA levels of all genes [CYP9Z82 (p = 0.008), CYP6NR1 (p = 0.029) and CYP345J1 (p = 0.001)] compared to the dsRNA negative control RPWs (Fig. 8).

Fig. 8
figure 8

RNAi-based silencing of the selected P450s CYP9Z82, CYP345J1 and CYP6NR1. Mean fold change in P450s expression in fat body tissues estimated by RT-qPCR, from different samples (NI: non-injected, NFW: Nuclease free water- injected, negative control dsRNA-injected and respective P450 dsRNA-injected). (*) represents the statistical significance measured at p < 0.05 and error bars represents SEM. All P450s showed a significant reduction in expression compared to the dsRNA negative control. Representative visual band gel images are provided in the insight. (p < 0.05; one-way ANOVA with LSD)

Phenotypic effects and validation of gene silencing by RT-qPCR

The percentages of mortality observed in CYP9Z82-, CYP6NR1- and CYP345J1-silenced insects were determined to be 25, 33 and 50%, respectively, upon exposure to a sublethal concentration of imidacloprid for 11 days. The Kaplan-Meier survival analysis between the silenced insects showed no significant difference in survival functions between the silenced groups (p = 0.054, Mantel-Cox Log Rank test) (Fig. 9). However, the 50% mortality observed in CYP345J1-silenced insects upon imidacloprid treatment demonstrates the importance of this gene in imidacloprid detoxification.

Fig. 9
figure 9

The Kaplan-Meier survival analysis provides the survival functions from three groups of P450-silenced insects, represented as CYP9Z82 (blue), CYP345J1 (green), CYP6NR1 (yellow), and control (violet) line diagrams. The experiment was censored after 11 days and the data represents comparison of 8 days, till the survival rate of silenced insects reached 0. The survival functions calculated based on the Mantel-Cox Log Rank test showed no significant difference in survival functions between the silenced groups (p = 0.054, Mantel-Cox Log Rank test)

The expression status of P450s was further analyzed in the gut and fat body tissues of the silenced insects, which allowed us to compare the effect of imidacloprid treatment in injected, treated and surviving insects. There were significant differences (p = 0.010) in each group compared to the sus group, especially in fat body tissues. The comparison showed that the ind group presented the greatest differences between the genes (p ≤ 0.001), followed by the n-trt (p = 0.042), trt (P = 0.059), sus (p = 0.011) and res (p = 0.015) groups. For CYP9Z82, which displayed the highest expression in the ind and res strains (> 100-fold), showed no significant increase in expression after treatment in either gut or fat body tissues. CYP345J1, with > 10-fold and > 50-fold expression in the ind and res strains, respectively, also showed low expression in trt samples. Likewise, CYP6NR1 showed a similar expression pattern with high expression in the gut ind and res strains (> 100-fold and > 10-fold, respectively) and low expression in trt samples. These results show that the silencing of these genes was lower in all treated samples, and thus, mortality rates can be considered to be a factor that represents the importance of all three P450s in contributing to imidacloprid resistance (Fig. 10).

Fig. 10
figure 10

The figure represents relative expression of the CYP9Z82, CYP345J1 and CYP6NR1 genes in the gut and fat body tissues of different experiment groups, silenced and imidacloprid treated (trt), silenced and non-treated (n-trt), non-silenced and imidacloprid treated (ind), and field (res), compared to non-silenced and non-treated (sus) group. The bars represent mean fold change in expression compared to sus group. Each sample group was color coded as n-trt (orange), trt (yellow), ind (green) and res (brown). Error bars represent SEM. The statistical significance was measured using Tukey’s HSD with p < 0.05 as the level of significance

Discussion

The RPW, R. ferrugineus, is the most invasive and globally important quarantine pest [42,43,44,45]. The cause of the high rate of spread of this pest is human intervention via transport of infested young or adult date palm trees and offshoots from contaminated areas into uninfected areas [43]. As the earliest commercial neonicotinoid, imidacloprid has been extensively used for the control of R. ferrugineus in date palm cropping systems in Middle Eastern countries [42]. However, the widespread application of imidacloprid has resulted in the development of resistance, which has emerged as a major problem for R. ferrugineus control [50]. Differential expansion and diversification of P450 subfamilies were observed in R. ferrugineus, and these changes are possible to be linked to adaptations to novel ecological niches. We showed that dsRNA injection and silencing of the CYP345J1 and CYP6NR1 genes resulted in a significant decrease in the survival rate of imidacloprid-treated adult weevils. Hence, based on our RNAi results, we conclude that constitutive overexpression of several P450s and selective detoxification involving at least CYP345J1 and CYP6NR1, along with several other enzyme systems, are involved in imidacloprid detoxification in RPWs in a date palm field.

The current study provides the global transcriptome repertoire of RPW P450s and a comprehensive picture of the R. ferrugineus CYPome, based on the identification of genes and imidacloprid-induced expression profiling. Our tissue-specific expression analysis revealed nearly ubiquitous expression patterns, including some P450s with possible duplications or pseudogenes that are not expressed. Our identification of 77 P450s from R. ferrugineus, which belong to 22 families and 45 subfamilies; is more or less similar to previous reports on other insects, such as Bombyx mori, exhibiting 84 P450s, L. decemlineata, exhibiting 96 P450s [23], and D. melanogaster, exhibiting 85 P450s [77]. However, the total P450s in R. ferrugineus is relatively low compared to the numbers identified in other coleopteran model insects including T. castaneum (143 P450s) [79], as genome sequence data become available for Tribolium.

The phylogenetic analysis of R. ferrugineus P450s with other insect species revealed that insect P450s have diverged over time into different groups leads to species-specific and lineage specific expansions (Figs. 1, 2 and 3), consistent with the earlier studies [6, 14, 84]. R. ferrugineus P450s CYP2 and mitochondrial clans showed a high degree of sequence homology with other insects, indicating functional conservation of these P450s. Notably, CYP18A1, CYP305A1 and CYP306A1 from the CYP2 clan, and CYP12, CYP302A1, CYP314A1 and CYP315A1 from the mitochondrial clade were clustered into one group with > 90% bootstrap support, indicates that these P450s may be originated from a common ancestor and differentiated during evolution linked to adaptations. Several earlier studies have demonstrated that P450s in the CYP2 and mitochondrial clan were generally associated with the ecdysteroid biosynthesis such as 306A1 encoded by Phantom gene in D. melanogaster and B. mori [85, 86], T. cataneum [79], L. decemlineata [87] and A. mellifera; 18A1 encoded an ecdysteroid 26-hydroxylase/oxidase activities in D. melanogaster [88]; CYP314A1 and CYP315A1 encoded by Halloween genes Shadow and Shade respectively in D. melanogaster [81], and CYP301A1 involved in the ecdysone regulation in D. melanogaster [89]. Whereas, CYP49A1 and CYP18A1 are expressed in the integument during Drosophila development [90, 91], and CYP12 family genes associated xenobiotic metabolism in D. melanogaster and A. gambiae [84]. In the R. ferrugineus CYP2 and mitochondrial clans, we can speculate based on above mentioned studies that CYP305A1, CYP306A1, CYP301A1, CYP314A1, and CYP315A1 may be involved in ecdysone biosynthesis; CYP12 family genes associated xenobiotic metabolism, and CYP49A1 and CYP18A1 in the integument during R. ferrugineus development.

Members of the CYP3 clan are reported to be involved in xenobiotic and insecticide metabolism, and are induced by the exposure of pesticides and natural products [7, 9, 37, 41]. The phylogenetic analysis showed that genes in Clan 3 and 4 appear to undergo multiple expansion and extensive species-specific clustering leads to CYP “blooms”. Most P450s in the CYP3 clan had gene orthologos in T. castaneum [79] and L. decemlineata [23, 87], or gene expansion with close phylogenetic relationships (Fig. 2). P450s in the CYP6 family had several functions. For instance, CYP6G1, CYP6BQ, CYP6CM/CYP6ER1/CYP6AY1 and CYP6D/CYPBG subfamilies are reported to be involved in DDT, deltamethrin, imidacloprid and pyrethroid metabolism respectively in several insects [19, 27, 75, 92]. R. ferrugineus CYP6 phylogenetic analysis clustered with several orthologs CYP6 subfamily genes reported earlier, notably, CYP6G1, CYP6EZ1, CYP6EH1, CYP6AY1 and CYP6CM1, which are associated with imidacloprid resistance (Fig. 2). Our phylogenetic analysis suggesting the potential functions of these R. ferrugineus P450s in detoxification of imidacloprid in the date palm field. However, many other P450s in the CYP6 family genes have been reported to be involved in detoxification of plant allelochemicals and had various functions. For instance, CYP6B subfamily genes reported to be involved in the metabolism of plant toxins furanocoumarins to nontoxic compound in the insect order Papilionidae [93], and CYP6A subfamily genes associated with pheromone-sensing, and its expression level in the olfactory sensilla associated with the social behavior, aggressiveness in D. melalanogastor [94, 95]. The second largest family in the Clan 3 is CYP9, and several P450s in the family is reported to be associated with insecticide and odorant metabolism in insects [96, 97]. In our study, three CYP9 subfamily P450s (CYP9Z, CYP9AN and CYP9CS) displayed close phylogenetic relationship with T. castaneum and L. decemlineata, however, high degree of species-specific expansion reflecting adaptation to environmental xenobiotics. In the Clan 4, CYP4 is the largest gene family identified so far; and our phylogenetic analysis identified several potential P450s orthologs and paralogs associated with biosynthesis of endogeneous compounds [98, 99], pyrethroid resistance [18, 100] and pheromone metabolism [101]. Gene expansion and species specific CYP “bloom” is common in the CYP4 family as reported in other insects [14, 84], particularly in the CYP4BD and CYP4G cluster. The orthologs CYP4, CYPQc and CYP4Qb from L. decemlineata reported to be involved in the imidacloprid resistance [23] were found grouped with R. ferrugineus CYP4Q6 and CYP4KX2, indicating a common ancestor (Fig. 1). CYP4G7 and CYP4G61 are thought to be a strong candidate for enhancing the cypermethrin and pyriproxyfen resistance respectively in T. castaneum [27] and Trialeurodes vaporariorum [102]. Similarly, the CYP4G25 is associated with diapause in Antheraea yamanai [103]. Such diversity in CYP4G and CYP4D subfamily genes indicating a diverse function.

P450s are well known for their role in detoxification mechanisms in a wide range of organisms, including humans. In insects, the major sites of detoxification are the head, gut and fat body, among which the fat body is the most important [22, 23]. We identified three CYP6 candidates and one CYP9 candidates that were highly overexpressed in the fat body of ind RPWs in a laboratory bioassay (Fig. 5). Interestingly, we observed that imidacloprid-induced transcripts in ind insects were also highly overexpressed in field-caught res samples, indicating that they may play a role in the development of resistance in response to the heavy doses of imidacloprid used in the date palm field in Saudi Arabia. Among CYP families, 6, 4 and the CYP345 family have been reported to be involved in imidacloprid resistance in various insects (Additional file 4: Table S2) [19, 23, 27, 80, 104,105,106,107,108], and the current study specifically focused on genes induced in response to imidacloprid treatment.

Interestingly, the candidate (CYP345J1) with the highest RPKM value shows high sequence similarity to CYP6g1, a well-studied DDT and imidacloprid cross-resistance gene from D. melanogaster [75, 80, 109] and D. simulans [82]; this candidate is overexpressed in the fat body of ind and res RPWs (Fig. 5). CYP345A1 from T. castaneum is reported to be paralog of CYP6g1 [27] and shares 52% nucleotide identity with CYP345J1, which was significantly upregulated by imidacloprid. Another candidate, CYP4LE1, is a paralog (53% nucleotide identity) of B. tabaci CYP4C64, which is associated with imidacloprid resistance [19, 20]. R. ferrugineus CYP4BD4 is overexpressed and induced in both res and ind RPW populations (Fig. 7). Notably, CYP4BD4 expression was higher in the RPW fat body than in the gut, indicating that metabolic detoxification and catalytic activity are more prevalent in the fat body. More interestingly, our bioassay revealed imidacloprid-induced transcripts, with 9 more transcripts being overexpressed at a higher level in the ind fat body than in the gut. Our induction data from the ind fat body also highlight the role of multiple P450s involved in the metabolic detoxification process (Fig. 5), which supports the recent identification of the induction of multiple P450s [75] in L. decemlineata [23]. The promising imidacloprid-induced P450 transcripts that may lead to resistance included six common transcripts found in both ind and res fat body tissues as well as four common transcripts found in the ind and res gut. Interestingly, 4 CYP6 family transcripts fell into this category for the fat body, whereas the CYP4, CYP6 and CYP9 families were all considered to be involved in detoxification in the gut (Fig. 6).

Intriguingly, the overlap of P450s induced in different tissues indicates that these enzymes act at different detoxification sites, especially in the gut and fat body. The largest number of P450s was induced (37 of the 50 studied) in the res fat body (Additional file 1: Figure S1). Half of all P450s studied were also found to be induced in the res gut, indicating the possibility of high rates of detoxification in this organ. We identified imidacloprid-induced genes that overlap with overexpressed/resistance genes from res insects [gut 9:25 (1:1.3), fat body 22:37 (1:1.7)]. Moreover, the highly induced candidate gene CYP9Z82, showed higher induction in the fat body compared to the gut. This transcript, with a > 100-fold increase in expression, also exhibits 65% identity with CYP9Zg (GenBank: XP_023014564) from L. decemlineata, a gene reported to be involved in imidacloprid resistance [23]. Similarly, we found three RPW CYP4 candidates (CYP4G185, CYP4G186 and CYP4G187) showing 51–59% nucleotide identity with T. castaneum CYP4G7, a gene reported to be involved in imidacloprid resistance [27]. Among these candidates, only CYP4G186 was overexpressed in the fat body of ind and res RPWs. These results indicate that CYP family members are differentially induced between various strains of beetles and that even closely related insects exhibit limited P450 functional similarity. As only an induction profile for L. decemlineata CYP is available for comparison [23], there is a significant need for more studies on insect CYPomes and induction profiles.

Considering the large number of P450s reported in insects and their possible roles in different biosynthetic and detoxification mechanisms, one approach to identifying P450s involved in imidacloprid detoxification is to develop an induction profile in response to imidacloprid. In the whitefly B. tabaci, increased expression levels of CYP6CM1 and CYP4C64 are reported to be associated with imidacloprid resistance [19, 20], and in N. lugens, overexpression of CYP6ER1 and CYP6AY1 are closely linked to imidacloprid resistance [21, 22]. Furthermore, in the Colorado potato beetle L. decemlineata, upregulation through the induction and constitutive overexpression of multiple genes of the CYP4, 6 and 9 families was found to be associated with imidacloprid resistance [23]. We identified six clusters or expression patterns for P450s that can provide insight into the P450s involved in imidacloprid resistance. The four transcripts from cluster 2 have key detoxification functions in fat body tissues, as supported by 13 transcripts from cluster 1. Fifteen candidates identified in clusters 4 and 5 are also proposed to be involved in the detoxification of insecticides and plant toxins in gut tissues (Additional file 1: Figure S1).

We further evaluated the role of three overexpressed P450s in imidacloprid detoxification by silencing those genes using the RNAi method, followed by imidacloprid bioassays. Our results showed that 79.42 and 54.99% silencing of CYP6NR1 and CYP345J1, respectively, resulted in 33 and 50% mortality in RPW adults in toxicity bioassays. Similar results showing that up to 76% silencing of CYP9e2-like resulted in 35.78% mortality in B. odoriphaga have been reported [29], and silencing of CYP4Q3 resulted in 46.7% mortality in L. decemlineata in an imidacloprid toxicity bioassay [24]. The 50% mortality observed in CYP345J1-silenced insects was also comparable to the 55.1% mortality achieved for CYP6AY1-silenced N. lugens [110]. The sole CYP9 RNAi candidate from RPW, CYP9Z82, was associated with only 25% mortality upon imidacloprid treatment, which was similar to the 25.3% mortality achieved in CYP9AQ2-silenced L. migratoria, although this was determined in a deltamethrin bioassay [111]. Considering the mortality rates achieved, it is reasonable to conclude that the three P450s that we studied play key roles in imidacloprid detoxification in RPW. The 33 to 50% mortality achieved in the current experiment is promising considering the fact that other enzymes, such as UDP-glucosyl transferases (UGTs), GSTs, Ests and ATP-binding cassette transporters, are possibly important in imidacloprid detoxification, as reported recently in L. decemlineata [24]. Considering that nearly equal silencing levels were achieved for all three genes, it is possible to differentiate the importance of P450s based on the mortality rates achieved. Thus, this study confirms the importance of two CYPs (CYP6NR1 and CYP345J1) in imidacloprid detoxification. Three factors that we need to consider further before targeting these genes as a P450-based pest control strategy are the species-specificity of these genes, the role of other detoxification enzymes and the possible synergism between them. Considering the development of resistance as a multistage process with multiple contributing factors, it is also important to study the role of mutations as well as the role of other detoxification enzymes, especially role of GST, as recently reported in house flies [112]. It will be more interesting to study role of regulatory genes, such as those related to xenobiotic factors, in the overexpression P450 genes [113].

Conclusion

Our study presents 77 cytochrome P450s in R. ferrugineus, which belong to 22 families and 45 subfamilies, and the results provide a basis for the further molecular and functional characterization of P450s. The current study demonstrates that multiple P450s are overexpressed and differentially expressed in field-caught RPW populations and in laboratory-induced (ind) strains, which should be considered an indicator of metabolic tolerance to imidacloprid. The important findings of our work included the functional identification of CYP345J1 and CYP6NR1; constitutive overexpression of these genes presumably allows insecticide molecules to be metabolized more efficiently, resulting in enhanced tolerance of RPWs in a date palm field. In the present study, we used three highly expressed P450s in the functional analysis; together with other differentially expressed P450s as well as Ests, GSTs and ATP-binding cassette transporters, our data help to explain imidacloprid resistance in RPW, though some components of this mechanism remain to be identified. Insecticide resistance in RPW is a major problem faced in date palm cultivation and is a major reason for the spread of R. ferrugineus around the world. Biochemical assays for detecting metabolic resistance due to P450s are increasingly being used in mosquitoes [40, 114]. Similarly, with the identification of P450s associated with imidacloprid resistance, the development of more specific diagnostic molecular markers is highly recommended for the screening of P450 overexpression in R. ferrugineus that confers metabolic resistance to imidacloprid. RNAi has the potential to be used to target upregulated molecular mechanisms of resistance and could be a practicable option for RPW management and needs to be studied in depth.

Availability of data and materials

All relevant data are within the paper and its supporting additional files. The RPW P450 FASTA nucleotide sequence information is given in Additional file 7: Dataset S1. The P450 nucleotide sequence can be obtained from the Transcriptome Shotgun Assembly project DDBJ/EMBL/GenBank under accession number GDKA00000000. The RPW P450 contig names are provided in Additional file 5: Table S3 and can be assessed at the GenBank (for example, the GenBank acc. no. of Rfer_c29199_CYP6BW4 is GDKA01029199.

Abbreviations

dsRNA control :

Universal negative dsRNA control

FB:

Fat body

ind :

Susceptible strain of RPWs treated with imidacloprid

LC50 :

Lethal concentration 50

NFW :

Nuclease free water-injected RPWs

NI :

No-injection RPWs

n-trt :

Susceptible strain of RPWs injected with P450 dsRNA and treated with water

P450:

Cytochrome P450-dependent monooxygenases

res :

Field-caught strain of RPWs treated with imidacloprid

Rfer:

Rhynchophorus ferrugineus

RPW:

Red palm weevil

sus :

Susceptible strain RPWs treated with water

trt :

Susceptible strain of RPWs injected with dsRNA and treated with imidacloprid

References

  1. Ware GW, Whitacre DM. An introduction to insecticides. In: MeisterPro information resources, a division of Meister media worldwide, 4th edn Willoughby, Ohio, USA; 2004.

    Google Scholar 

  2. Grube A, Donaldson D, Kiely T, Wu L. Pesticides industry sales and usage. Washington, DC: US EPA; 2011.

    Google Scholar 

  3. Whitacre D, Ware G. The pesticide book. Ohio: Meister Media Worldwide; 2004. p. 3–13.

  4. Scott JG, Wen Z. Cytochromes P450 of insects: the tip of the iceberg. Pest Manag Sci. 2001;57(10):958–67.

    Article  CAS  PubMed  Google Scholar 

  5. Feyereisen R. Insect P450 enzymes. Annu Rev Entomol. 1999;44(1):507–33.

    Article  CAS  PubMed  Google Scholar 

  6. Feyereisen R, Lawrence I. 8–insect CYP genes and P450 enzymes. In: Insect molecular biology and biochemistry; 2012. p. 236–316.

    Chapter  Google Scholar 

  7. Li X, Schuler MA, Berenbaum MR. Molecular mechanisms of metabolic resistance to synthetic and natural xenobiotics. Annu Rev Entomol. 2007;52:231–53.

    Article  PubMed  CAS  Google Scholar 

  8. Dermauw W, Wybouw N, Rombauts S, Menten B, Vontas J, Grbić M, Clark RM, Feyereisen R, Van Leeuwen T. A link between host plant adaptation and pesticide resistance in the polyphagous spider mite Tetranychus urticae. Proc Natl Acad Sci. 2013;110(2):E113–22.

    Article  CAS  PubMed  Google Scholar 

  9. Liu N, Li M, Gong Y, Liu F, Li T. Cytochrome P450s – their expression, regulation, and role in insecticide resistance. Pestic Biochem Physiol. 2015;120(Supplement C):77–81.

    Article  CAS  PubMed  Google Scholar 

  10. Yang N, Xie W, Yang X, Wang S, Wu Q, Li R, Pan H, Liu B, Shi X, Fang Y. Transcriptomic and proteomic responses of sweetpotato whitefly, Bemisia tabaci, to thiamethoxam. PLoS One. 2013;8(5):e61820.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Zhu YC, Luttrell R. Altered gene regulation and potential association with metabolic resistance development to imidacloprid in the tarnished plant bug, Lygus lineolaris. Pest Manag Sci. 2015;71(1):40–57.

    Article  CAS  PubMed  Google Scholar 

  12. Pan Y, Peng T, Gao X, Zhang L, Yang C, Xi J, Xin X, Bi R, Shang Q. Transcriptomic comparison of thiamethoxam-resistance adaptation in resistant and susceptible strains of Aphis gossypii Glover. Comp Biochem Physiol Part D Genomics Proteomics. 2015;13:10–5.

    Article  CAS  PubMed  Google Scholar 

  13. Gao Q, Li M, Sheng C, Scott JG, Qiu X. Multiple cytochrome P450s overexpressed in pyrethroid resistant house flies (Musca domestica). Pestic Biochem Physiol. 2012;104(3):252–60.

    Article  CAS  Google Scholar 

  14. Feyereisen R. Evolution of insect P450. Biochem Soc Trans. 2006;34(6):1252–5.

    Article  CAS  PubMed  Google Scholar 

  15. Terriere LC. Induction of detoxication enzymes in insects. Annu Rev Entomol. 1984;29(1):71–88.

    Article  CAS  PubMed  Google Scholar 

  16. Liu N, Scott J. Phenobarbital induction of CYP6D1 is due to atransacting factor on autosome 2 in house flies, Musca domestica. Insect Mol Biol. 1997;6(1):77–81.

    Article  CAS  PubMed  Google Scholar 

  17. Poupardin R, Reynaud S, Strode C, Ranson H, Vontas J, David J-P. Cross-induction of detoxification genes by environmental xenobiotics and insecticides in the mosquito Aedes aegypti: impact on larval tolerance to chemical insecticides. Insect Biochem Mol Biol. 2008;38(5):540–51.

    Article  CAS  PubMed  Google Scholar 

  18. Scharf M, Parimi S, Meinke LJ, Chandler L, Siegfried BD. Expression and induction of three family 4 cytochrome P450 (CYP4) genes identified from insecticide-resistant and susceptible western corn rootworms, Diabrotica virgifera virgifera. Insect Mol Biol. 2001;10(2):139–46.

    Article  CAS  PubMed  Google Scholar 

  19. Karunker I, Benting J, Lueke B, Ponge T, Nauen R, Roditakis E, Vontas J, Gorman K, Denholm I, Morin S. Over-expression of cytochrome P450 CYP6CM1 is associated with high resistance to imidacloprid in the B and Q biotypes of Bemisia tabaci (Hemiptera: Aleyrodidae). Insect Biochem Mol Biol. 2008;38(6):634–44.

    Article  CAS  PubMed  Google Scholar 

  20. Yang X, Xie W, Wang S-l, Wu Q-j, Pan H-p, Li R-m, Yang N-n, Liu B-m, Xu B-y, Zhou X. Two cytochrome P450 genes are involved in imidacloprid resistance in field populations of the whitefly, Bemisia tabaci, in China. Pestic Biochem Physiol. 2013;107(3):343–50.

    Article  CAS  PubMed  Google Scholar 

  21. Bass C, Carvalho R, Oliphant L, Puinean A, Field L, Nauen R, Williamson M, Moores G, Gorman K. Overexpression of a cytochrome P450 monooxygenase, CYP6ER1, is associated with resistance to imidacloprid in the brown planthopper, Nilaparvata lugens. Insect Mol Biol. 2011;20(6):763–73.

    Article  CAS  PubMed  Google Scholar 

  22. Pang R, Chen M, Liang Z, Yue X, Ge H, Zhang W. Functional analysis of CYP6ER1, a P450 gene associated with imidacloprid resistance in Nilaparvata lugens. Sci Rep. 2016;6:34992.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Zhu F, Moural TW, Nelson DR, Palli SR. A specialist herbivore pest adaptation to xenobiotics through up-regulation of multiple cytochrome P450s. Sci Rep. 2016;6:20421.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Kaplanoglu E, Chapman P, Scott IM, Donly C. Overexpression of a cytochrome P450 and a UDP-glycosyltransferase is associated with imidacloprid resistance in the Colorado potato beetle, Leptinotarsa decemlineata. Sci Rep. 2017;7(1):1762.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Clements J, Schoville S, Peterson N, Lan Q, Groves RL. Characterizing molecular mechanisms of imidacloprid resistance in select populations of Leptinotarsa decemlineata in the Central Sands region of Wisconsin. PLoS One. 2016;11(1):e0147844.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Clements J, Schoville S, Peterson N, Huseth AS, Lan Q, Groves RL. RNA interference of three up-regulated transcripts associated with insecticide resistance in an imidacloprid resistant population of Leptinotarsa decemlineata. Pestic Biochem Physiol. 2017;135:35–40.

    Article  CAS  PubMed  Google Scholar 

  27. Liang X, Xiao D, He Y, Yao J, Zhu G, Zhu KY. Insecticide-mediated up-regulation of cytochrome P450 genes in the red flour beetle (Tribolium castaneum). Int J Mol Sci. 2015;16(1):2078–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Elzaki MEA, Miah MA, Wu M, Zhang H, Pu J, Jiang L, Han Z. Imidacloprid is degraded by CYP353D1v2, a cytochrome P450 overexpressed in a resistant strain of Laodelphax striatellus. Pest Manag Sci. 2017;73(7):1358–63.

    Article  CAS  PubMed  Google Scholar 

  29. Chen C, Wang C, Liu Y, Shi X, Gao X. Transcriptome analysis and identification of P450 genes relevant to imidacloprid detoxification in Bradysia odoriphaga. Sci Rep. 2018;8(1):2564.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Pimentel D. Environmental and economic costs of the application of pesticides primarily in the United States. Environ Dev Sustain. 2005;7(2):229–52.

    Article  Google Scholar 

  31. Pimentel D, Burgess M. Environmental and economic costs of the application of pesticides primarily in the United States. In: Integrated pest management. Dordrecht: Springer; 2014. p. 47–71.

    Chapter  Google Scholar 

  32. Scott JG. Cytochromes P450 and insecticide resistance. Insect Biochem Mol Biol. 1999;29(9):757–77.

    Article  CAS  PubMed  Google Scholar 

  33. Ronis M, Hodgson E. Cytochrome P-450 monooxygenases in insects. Xenobiotica. 1989;19(10):1077–92.

    Article  CAS  PubMed  Google Scholar 

  34. Korytko PJ, Scott JG. CYP6D1 protects thoracic ganglia of houseflies from the neurotoxic insecticide cypermethrin. Arch Insect Biochem Physiol. 1998;37(1):57–63.

    Article  CAS  PubMed  Google Scholar 

  35. Wojtasek H, Leal WS. Degradation of an alkaloid pheromone from the pale-brown chafer, Phyllopertha diversa (Coleoptera: Scarabaeidae), by an insect olfactory cytochrome P450. FEBS Lett. 1999;458(3):333–6.

    Article  CAS  PubMed  Google Scholar 

  36. Nauen R, Denholm I. Resistance of insect pests to neonicotinoid insecticides: current status and future prospects. Arch Insect Biochem Physiol. 2005;58(4):200–15.

    Article  CAS  PubMed  Google Scholar 

  37. Liu N. Insecticide resistance in mosquitoes: impact, mechanisms, and research directions. Annu Rev Entomol. 2015;60:537–59.

    Article  CAS  PubMed  Google Scholar 

  38. Liu N, Scott JG. Increased transcription of CYP6D1 causes cytochrome P450-mediated insecticide resistance in house fly. Insect Biochem Mol Biol. 1998;28(8):531–5.

    Article  CAS  PubMed  Google Scholar 

  39. Kasai S, Scott JG. Overexpression of cytochrome P450 CYP6D1 is associated with monooxygenase-mediated pyrethroid resistance in house flies from Georgia. Pestic Biochem Physiol. 2000;68(1):34–41.

    Article  CAS  Google Scholar 

  40. David J-P, Ismail HM, Chandor-Proust A, Paine MJI. Role of cytochrome P450s in insecticide resistance: impact on the control of mosquito-borne diseases and use of insecticides on earth. Phil Trans R Soc B. 2013;368(1612):20120429.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  41. ffrench-Constant RH, Daborn PJ, Le Goff G. The genetics and genomics of insecticide resistance. Trends Genet. 2004;20(3):163–70.

    Article  CAS  PubMed  Google Scholar 

  42. Al-Dosary NMN, Al-Dobai S, Faleiro JR. Review on the management of red palm weevil Rhynchophorus ferrugineus olivier in date palm Phoenix dactylifera L. Emirates J Food Agric. 2016;28(1):34–44.

    Article  Google Scholar 

  43. Faleiro J. A review of the issues and management of the red palm weevil Rhynchophorus ferrugineus (Coleoptera: Rhynchophoridae) in coconut and date palm during the last one hundred years. Int J Trop Insect Sci. 2006;26(03):135–54.

    CAS  Google Scholar 

  44. Dembilio Ó, Jaques JA. Biology and management of Red Palm Weevil. In: Sustainable pest management in Date Palm: current status and emerging challenges. Switzerland: Springer; 2015. p. 13–36.

    Chapter  Google Scholar 

  45. Rugman-Jones PF, Hoddle CD, Hoddle MS, Stouthamer R. The lesser of two weevils: molecular-genetics of pest palm weevil populations confirm Rhynchophorus vulneratus (panzer 1798) as a valid species distinct from R. ferrugineus (Olivier 1790), and reveal the global extent of both. PLoS One. 2013;8(10):e78379.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Antony B, Johny J, Aldosari S, Abdelazim M. Identification and expression profiling of novel plant cell wall degrading enzymes from a destructive pest of palm trees, Rhynchophorus ferrugineus. Insect Mol Biol. 2017;26(4):469–84.

    Article  CAS  PubMed  Google Scholar 

  47. Al-Shawaf AM, Al-Shagag A, Al-Bagshi M, Al-Saroj S, Al-Bather S, Al-Dandan AM, Abdallah AB, Faleiro JR. A quarantine protocol against red palm weevil Rhynchophorus ferrugineus (Olivier)(Coleptera: Curculiondae) in date palm. J Plant Prot Res. 2013;53(4):409–15.

    Article  Google Scholar 

  48. Al–Shawaf AM, Al-Shagagh AA, Al-Bakshi MM, Al-Saroj SA, Al-Badr SM, Al-Dandan AM, Abdallah AB. Toxicity of some insecticides against red palm weevil, Rhynchophorus ferrugineus. Indian J Plant Prot. 2010;38(1):13–6.

    Google Scholar 

  49. Kaakeh W. Toxicity of imidacloprid to developmental stages of Rhynchophorus ferrugineus (Curculionidae: Coleoptera): laboratory and field tests. Crop Prot. 2006;25(5):432–9.

    Article  CAS  Google Scholar 

  50. Al-Ayedh H, Hussain A, Rizwan-ul-Haq M, Al-Jabr AM. Status of insecticide resistance in field-collected populations of Rhynchophorus ferrugineus (Olivier)(Coleoptera: Curculionidae). Int J Agric Biol. 2016;18(1):103–10.

    Article  CAS  Google Scholar 

  51. Alyokhin A, Dively G, Patterson M, Castaldo C, Rogers D, Mahoney M, Wollam J. Resistance and cross-resistance to imidacloprid and thiamethoxam in the Colorado potato beetle Leptinotarsa decemlineata. Pest Manag Sci. 2007;63(1):32–41.

    Article  CAS  PubMed  Google Scholar 

  52. Abdel-Azim MM, Vidyasagar P, Aldosari SA, Mumtaz R. Impact of mating frequency on fecundity, fertility and longevity of red palm weevil, Rhynchophorus ferrugineus (Olivier)(Coleoptera: Curculionidae). J Agric Sci Technol A. 2012;2(4):520–8.

    Google Scholar 

  53. Wang L, Zhang XW, Pan LL, Liu WF, Wang DP, Zhang GY, Yin YX, Yin A, Jia SG, Yu XG. A large-scale gene discovery for the red palm weevil Rhynchophorus ferrugineus (Coleoptera: Curculionidae). Insect Sci. 2013;20(6):689–702.

    Article  CAS  PubMed  Google Scholar 

  54. Antony B, Soffan A, Jakše J, Abdelazim MM, Aldosari SA, Aldawood AS, Pain A. Identification of the genes involved in odorant reception and detection in the palm weevil Rhynchophorus ferrugineus, an important quarantine pest, by antennal transcriptome analysis. BMC Genomics. 2016;17:69.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Soffan A, Antony B, Abdelazim M, Shukla P, Witjaksono W, Aldosari SA, Aldawood AS. Silencing the olfactory co-receptor RferOrco reduces the response to pheromones in the red palm weevil, Rhynchophorus ferrugineus. PLoS One. 2016;11(9):e0162203.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25(17):3389–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Anderson I, Brass A. Searching DNA databases for similarities to DNA sequences: when is a match significant? Bioinformatics. 1998;14(4):349–56.

    Article  CAS  PubMed  Google Scholar 

  58. Conesa A, Götz S, García-Gómez JM, Terol J, Talón M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics. 2005;21(18):3674–6.

    Article  CAS  PubMed  Google Scholar 

  59. Götz S, García-Gómez JM, Terol J, Williams TD, Nagaraj SH, Nueda MJ, Robles M, Talón M, Dopazo J, Conesa A. High-throughput functional annotation and data mining with the Blast2GO suite. Nucleic Acids Res. 2008;36(10):3420–35.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton S, Cooper A, Markowitz S, Duran C, et al. Geneious basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28:1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Mortazavi A, Williams B, McCue K, Schaeffer L, Wold B. Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods. 2008;5(7):621–8.

    Article  CAS  PubMed  Google Scholar 

  62. Nelson DR. Cytochrome P450 nomenclature. In: Methods in molecular biology-Clifton then Totowa, vol. 107; 1998. p. 15–24.

    Google Scholar 

  63. Nelson DR. Cytochrome P450 nomenclature, 2004. In: Cytochrome P450 protocols; 2006. p. 1–10.

    Google Scholar 

  64. Antony B, Johny J, Aldosari SA. Silencing the odorant binding protein RferOBP1768 reduces the strong preference of palm weevil for the major aggregation pheromone compound Ferrugineol. Front Physiol. 2018;9:252.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Hussain A, Rizwan-ul-Haq M, Al-Ayedh H, AlJabr A. Susceptibility and immune defence mechanisms of Rhynchophorus ferrugineus (Olivier)(Coleoptera: Curculionidae) against entomopathogenic fungal infections. Int J Mol Sci. 2016;17(9):1518.

    Article  PubMed Central  CAS  Google Scholar 

  66. Katoh K, Rozewicki J, Yamada KD. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Brief Bioinform. 2017;30:3059.

  67. Darriba D, Taboada GL, Doallo R, Posada D. ProtTest 3: fast selection of best-fit models of protein evolution. Bioinformatics. 2011;27(8):1164–5.

    Article  CAS  PubMed  Google Scholar 

  68. Guindon S, Dufayard J-F, Lefort V, Anisimova M, Hordijk W, Gascuel O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 2010;59(3):307–21.

    Article  CAS  PubMed  Google Scholar 

  69. Bustin SA, Benes V, Garson JA, Hellemans J, Huggett J, Kubista M, Mueller R, Nolan T, Pfaffl MW, Shipley GL. The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem. 2009;55(4):611–22.

    Article  CAS  PubMed  Google Scholar 

  70. Taylor S, Wakem M, Dijkman G, Alsarraj M, Nguyen M. A practical approach to RT-qPCR—publishing data that conform to the MIQE guidelines. Methods. 2010;50(4):S1–5.

    Article  CAS  PubMed  Google Scholar 

  71. Schmittgen TD, Livak KJ. Analyzing real-time PCR data by the comparative CT method. Nat Protoc. 2008;3(6):1101–8.

    Article  CAS  PubMed  Google Scholar 

  72. Ajlan A, Shawir M, Abo-El-Saad M, Rezk M, Abdulslam K. Laboratory evaluation of certain organophosphorus insecticides against the red palm weevil, Rhynchophorus ferrugineus (Olivier). Sci J King Faisal Univ (Basic Appl Sci). 2000;1(1):En15–34.

    Google Scholar 

  73. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2^(−ΔΔCT) method. Methods (San Diego, Calif). 2001;25:402–8.

    Article  CAS  Google Scholar 

  74. de Hoon MJ, Imoto S, Nolan J, Miyano S. Open source clustering software. Bioinformatics. 2004;20(9):1453–4.

    Article  PubMed  CAS  Google Scholar 

  75. Daborn P, Yen J, Bogwitz M, Le Goff G, Feil E, Jeffers S, Tijet N, Perry T, Heckel D, Batterham P, et al. A single P450 allele associated with insecticide resistance in Drosophila. Science. 2002;297(5590):2253–6.

    Article  CAS  PubMed  Google Scholar 

  76. Daborn P, Boundy S, Yen J, Pittendrigh B, ffrench-Constant R. DDT resistance in Drosophila correlates with Cyp6g1 over-expression and confers cross-resistance to the neonicotinoid imidacloprid. Mol Gen Genomics. 2001;266(4):556–63.

    Article  CAS  Google Scholar 

  77. Nelson DR. The cytochrome p450 homepage. Hum Genomics. 2009;4(1):1.

    Article  Google Scholar 

  78. Nelson DR. Cytochrome P450 and the individuality of species. Arch Biochem Biophys. 1999;369(1):1–10.

    Article  CAS  PubMed  Google Scholar 

  79. Zhu F, Moural TW, Shah K, Palli SR. Integrated analysis of cytochrome P450 gene superfamily in the red flour beetle, Tribolium castaneum. BMC Genomics. 2013;14(1):1.

    Article  CAS  Google Scholar 

  80. Joußen N, Heckel DG, Haas M, Schuphan I, Schmidt B. Metabolism of imidacloprid and DDT by P450 CYP6G1 expressed in cell cultures of Nicotiana tabacum suggests detoxification of these insecticides in Cyp6g1-overexpressing strains of Drosophila melanogaster, leading to resistance. Pest Manag Sci. 2008;64(1):65–73.

    Article  PubMed  CAS  Google Scholar 

  81. Rewitz KF, O’Connor MB, Gilbert LI. Molecular evolution of the insect Halloween family of cytochrome P450s: phylogeny, gene organization and functional conservation. Insect Biochem Mol Biol. 2007;37(8):741–53.

    Article  CAS  PubMed  Google Scholar 

  82. Le Goff G, Boundy S, Daborn P, Yen J, Sofer L, Lind R, Sabourault C, Madi-Ravazzi L, ffrench-Constant R. Microarray analysis of cytochrome P450 mediated insecticide resistance in Drosophila. Insect Biochem Mol Biol. 2003;33(7):701–8.

    Article  PubMed  CAS  Google Scholar 

  83. 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.

    Article  CAS  PubMed  Google Scholar 

  84. Ranson H, Claudianos C, Ortelli F, Abgrall C, Hemingway J, Sharakhova MV, Unger MF, Collins FH, Feyereisen R. Evolution of supergene families associated with insecticide resistance. Science. 2002;298(5591):179–81.

    Article  CAS  PubMed  Google Scholar 

  85. Niwa R, Matsuda T, Yoshiyama T, Namiki T, Mita K, Fujimoto Y, Kataoka H. CYP306A1, a cytochrome P450 enzyme, is essential for ecdysteroid biosynthesis in the prothoracic glands of Bombyx and Drosophila. J Biol Chem. 2004;279(34):35942–9.

    Article  CAS  PubMed  Google Scholar 

  86. Warren JT, Petryk A, Marqués G, Parvy J-P, Shinoda T, Itoyama K, Kobayashi J, Jarcho M, Li Y, O’Connor MB. Phantom encodes the 25-hydroxylase of Drosophila melanogaster and Bombyx mori: a P450 enzyme critical in ecdysone biosynthesis. Insect Biochem Mol Biol. 2004;34(9):991–1010.

    Article  CAS  PubMed  Google Scholar 

  87. Wan P-J, Shi X-Q, Kong Y, Zhou L-T, Guo W-C, Ahmat T, Li G-Q. Identification of cytochrome P450 monooxygenase genes and their expression profiles in cyhalothrin-treated Colorado potato beetle, Leptinotarsa decemlineata. Pestic Biochem Physiol. 2013;107(3):360–8.

    Article  CAS  PubMed  Google Scholar 

  88. Bassett MH, McCarthy JL, Waterman MR, Sliter TJ. Sequence and developmental expression of Cyp18, a member of a new cytochrome P450 family from Drosophila. Mol Cell Endocrinol. 1997;131(1):39–49.

    Article  CAS  PubMed  Google Scholar 

  89. Sztal T, Chung H, Berger S, Currie PD, Batterham P, Daborn PJ. A cytochrome p450 conserved in insects is involved in cuticle formation. PLoS One. 2012;7(5):e36544.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Guittard E, Blais C, Maria A, Parvy J-P, Pasricha S, Lumb C, Lafont R, Daborn PJ, Dauphin-Villemant C. CYP18A1, a key enzyme of Drosophila steroid hormone inactivation, is essential for metamorphosis. Dev Biol. 2011;349(1):35–45.

    Article  CAS  PubMed  Google Scholar 

  91. Chung H, Sztal T, Pasricha S, Sridhar M, Batterham P, Daborn PJ. Characterization of Drosophila melanogaster cytochrome P450 genes. Proc Natl Acad Sci. 2009;106(14):5731–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Tomita T, Liu N, Smith F, Sridhar P, Scott J. Molecular mechanisms involved in increased expression of a cytochrome P450 responsible for pyrethroid resistance in the housefly, Musca domestica. Insect Mol Biol. 1995;4(3):135–40.

    Article  CAS  PubMed  Google Scholar 

  93. Cohen MB, Schuler MA, Berenbaum MR. A host-inducible cytochrome P-450 from a host-specific caterpillar: molecular cloning and evolution. Proc Natl Acad Sci. 1992;89(22):10920–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Dierick HA, Greenspan RJ. Molecular analysis of flies selected for aggressive behavior. Nat Genet. 2006;38(9):1023.

    Article  CAS  PubMed  Google Scholar 

  95. Wang L, Dankert H, Perona P, Anderson DJ. A common genetic target for environmental and heritable influences on aggressiveness in Drosophila. Proc Natl Acad Sci. 2008;105(15):5657–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Rose RL, Goh D, Thompson DM, Verma KD, Heckel DG, Gahan LJ, Roe RM, Hodgson E. Cytochrome P450 (CYP) 9A1 in Heliothis virescens: the first member of a new CYP family. Insect Biochem Mol Biol. 1997;27(6):605–15.

    Article  CAS  PubMed  Google Scholar 

  97. Yang Y, Chen S, Wu S, Yue L, Wu Y. Constitutive overexpression of multiple cytochrome P450 genes associated with pyrethroid resistance in Helicoverpa armigera. J Econ Entomol. 2006;99(5):1784–9.

    Article  CAS  PubMed  Google Scholar 

  98. Maıbeche-Coisne M, Monti-Dedieu L, Aragon S, Dauphin-Villemant C. A new cytochrome P450 from Drosophila melanogaster, CYP4G15, expressed in the nervous system. Biochem Biophys Res Commun. 2000;273(3):1132–7.

    Article  PubMed  CAS  Google Scholar 

  99. Sutherland T, Unnithan G, Andersen J, Evans P, Murataliev M, Szabo L, Mash E, Bowers W, Feyereisen R. A cytochrome P450 terpenoid hydroxylase linked to the suppression of insect juvenile hormone synthesis. Proc Natl Acad Sci. 1998;95(22):12884–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Pridgeon JW, Zhang L, Liu N. Overexpression of CYP4G19 associated with a pyrethroid-resistant strain of the German cockroach, Blattella germanica (L.). Gene. 2003;314:157–63.

    Article  CAS  PubMed  Google Scholar 

  101. Maïbèche-Coisne M, Merlin C, François MC, Porcheron P, Jacquin-Joly E. P450 and P450 reductase cDNAs from the moth Mamestra brassicae: cloning and expression patterns in male antennae. Gene. 2005;346:195–203.

    Article  PubMed  CAS  Google Scholar 

  102. Karatolos N, Williamson MS, Denholm I, Gorman K, Bass C. Over-expression of a cytochrome P450 is associated with resistance to pyriproxyfen in the greenhouse whitefly Trialeurodes vaporariorum. PLoS One. 2012;7(2):e31077.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Yang P, Tanaka H, Kuwano E, Suzuki K. A novel cytochrome P450 gene (CYP4G25) of the silkmoth Antheraea yamamai: cloning and expression pattern in pharate first instar larvae in relation to diapause. J Insect Physiol. 2008;54(3):636–43.

    Article  CAS  PubMed  Google Scholar 

  104. Markussen MD, Kristensen M. Cytochrome P450 monooxygenase-mediated neonicotinoid resistance in the house fly Musca domestica L. Pestic Biochem Physiol. 2010;98(1):50–8.

    Article  CAS  Google Scholar 

  105. Puinean AM, Foster SP, Oliphant L, Denholm I, Field LM, Millar NS, Williamson MS, Bass C. Amplification of a cytochrome P450 gene is associated with resistance to neonicotinoid insecticides in the aphid Myzus persicae. PLoS Genet. 2010;6(6):e1000999.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Tiwari S, Gondhalekar A, Mann R, Scharf M, Stelinski L. Characterization of five CYP4 genes from Asian citrus psyllid and their expression levels in Candidatus Liberibacter asiaticus-infected and uninfected psyllids. Insect Mol Biol. 2011;20(6):733–44.

    Article  CAS  PubMed  Google Scholar 

  107. Garrood WT, Zimmer CT, Gorman KJ, Nauen R, Bass C, Davies TG. Field-evolved resistance to imidacloprid and ethiprole in populations of brown planthopper Nilaparvata lugens collected from across south and East Asia. Pest Manag Sci. 2016;72(1):140–9.

    Article  CAS  PubMed  Google Scholar 

  108. Elzaki MEA, Zhang W, Feng A, Qiou X, Zhao W, Han Z. Constitutive overexpression of cytochrome P450 associated with imidacloprid resistance in Laodelphax striatellus (Fallén). Pest Manag Sci. 2015;72(5):1051–8.

    Article  PubMed  CAS  Google Scholar 

  109. Catania F, Kauer M, Daborn P, Yen J, ffrench-Constant R, Schlötterer C. World-wide survey of an accord insertion and its association with DDT resistance in Drosophila melanogaster. Mol Ecol. 2004;13(8):2491–504.

    Article  CAS  PubMed  Google Scholar 

  110. Ding Z, Wen Y, Yang B, Zhang Y, Liu S, Liu Z, Han Z. Biochemical mechanisms of imidacloprid resistance in Nilaparvata lugens: over-expression of cytochrome P450 CYP6AY1. Insect Biochem Mol Biol. 2013;43(11):1021–7.

    Article  CAS  PubMed  Google Scholar 

  111. Guo Y, Zhang X, Wu H, Yu R, Zhang J, Zhu KY, Guo Y, Ma E. Identification and functional analysis of a cytochrome P450 gene CYP9AQ2 involved in deltamethrin detoxification from Locusta migratoria. Pestic Biochem Physiol. 2015;122:1–7.

    Article  CAS  PubMed  Google Scholar 

  112. Reid WR, Sun H, Becnel JJ, Clark AG, Scott JG. Overexpression of a glutathione S-transferase (Mdgst) and a galactosyltransferase-like gene (Mdgt1) is responsible for imidacloprid resistance in house flies. Pest Manag Sci. 2018;75(1):37–44.

    Article  PubMed  CAS  Google Scholar 

  113. Gaddelapati SC, Kalsi M, Roy A, Palli SR. Cap’n’collar C regulates genes responsible for imidacloprid resistance in the Colorado potato beetle, Leptinotarsa decemlineata. Insect Biochem Mol Biol. 2018;99:54–62.

    Article  CAS  PubMed  Google Scholar 

  114. Hemingway J, Hawkes NJ, McCarroll L, Ranson H. The molecular basis of insecticide resistance in mosquitoes. Insect Biochem Mol Biol. 2004;34(7):653–65.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We thank the date palm farmers in the Al Qassim area for support in obtaining red palm weevils and advice in adult weevil collection. We thank the KSU - Deanship of Scientific Research, Research Chair Program, Saudi Arabia. The authors are grateful to Dr. David. R. Nelson (University of Tennessee) for the P450 nomenclature. JJ is thankful for a researcher stipend obtained from the 12-AGR2554-02 project. We also thank Samy M. Mustafa of CDPR, Dr. Asma A. Al-Nujiban of Qassim University, Dr. Saleh. A. Aldosari and Mr. Waheed M. Alraha of CDPR for providing technical support and conducting RPW field collection. We are grateful to the editor and reviewers for their numerous perceptive and constructive comments, which helped us to improve our manuscript.

Funding

Funding for this research (awarded to BA) (Grant no. KACST-NSTIP 12-AGR2554–02) was provided by the National Plan for Science, Technology and Innovation (MAARIFAH) of King Abdul Aziz City for Science and Technology (KACST), Kingdom of Saudi Arabia. The authors are grateful to the Deanship of Scientific Research, King Saud University for funding through the Vice Deanship of Scientific Research Chairs. The transcriptome data were generated using faculty baseline funding from the King Abdullah University of Science and Technology (KAUST) to AP (BAS/1/1020-01-01). The funders had no role in study design, data collection and analysis, and interpretation of data and preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

BA conceived the study. BA acquired the grant, and BA and AP participated in its design and coordination. MMA and MAA participated in RPW collection, rearing and bioassays. BA and JJ carried out the experiments and compiled the data. JJ and BA performed phylogenetic analysis. JeJ and BA carried out the trimming, de novo assembly, and quality control analysis, and BA performed the Blast2GO analysis and local BLASTx searches. BA wrote the paper with contributions from JJ, and all authors have read and approved the final manuscript.

Corresponding author

Correspondence to Binu Antony.

Ethics declarations

Ethics approval and consent to participate

The RPW collections were generated with the direct permission of a cooperating landowner (Al-Kharj and Al-Qassim of Saudi Arabia). We confirm that no field experiments were involved in this study.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Additional files

Additional file 1:

Figure S1. P450s induced in the gut and fat body tissues of ind and res strains represented as the mean fold-change in expression compared to the respective sus gut/fat body tissues (2-ΔΔCт). P450s are arranged in descending order of the mean fold-change in expression from higher (green) to lower (yellow). (TIF 1586 kb)

Additional file 2:

Figure S2. Cluster analysis of P450s in gut (A) and fat body (B) tissues of ind, sus and res RPW strains. The relative expression of P450s in each tissue compared to tubulin and β-actin expression (2-ΔCт) was used in cluster analysis, and expression levels are indicated with red (overexpression), green (low expression), and black (moderate expression). The major clusters identified are marked in the tree. (TIF 910 kb)

Additional file 3:

Table S1. Lists of the oligonucleotide primers with amplicon size, Tm, min-max Ct and primer efficiency. (XLSX 20 kb)

Additional file 4:

Table S2. Rhynchophorus ferrugineus P450 functional annotation and classification. (PDF 12 kb)

Additional file 5:

Table S3. Rhynchophorus ferrugineus P450s nomenclature and blastx homology. (XLSX 15 kb)

Additional file 6:

Table S4. Toxicity assay of Confidor® 350 SC against susceptible and field-caught adult RPWs over 4 days of exposure. (PDF 366 kb)

Additional file 7:

Dataset S1. P450s identified in this study, FASTA format file. (PDF 497 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Antony, B., Johny, J., Abdelazim, M.M. et al. Global transcriptome profiling and functional analysis reveal that tissue-specific constitutive overexpression of cytochrome P450s confers tolerance to imidacloprid in palm weevils in date palm fields. BMC Genomics 20, 440 (2019). https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-019-5837-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-019-5837-4

Keywords