Next Article in Journal
Clinical Significance of Tie-2-Expressing Monocytes/Macrophages and Angiopoietins in the Progression of Ovarian Cancer—State-of-the-Art
Previous Article in Journal
Induction of Cardiac Pathology: Endogenous versus Exogenous Nrf2 Upregulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification and Functional Characterization of the Transcription Factors AhR/ARNT in Dendroctonus armandi

1
State Key Laboratory for Conservation and Utilization of Subtropical Agro-Bioresources, Guangdong Laboratory for Lingnan Modern Agriculture, College of Forestry and Landscape Architecture, South China Agricultural University, Guangzhou 510642, China
2
College of Forestry, Northwest A&F University, Xianyang 712100, China
*
Author to whom correspondence should be addressed.
Cells 2022, 11(23), 3856; https://doi.org/10.3390/cells11233856
Submission received: 29 September 2022 / Revised: 28 November 2022 / Accepted: 29 November 2022 / Published: 30 November 2022
(This article belongs to the Section Cell Signaling)

Abstract

:
The aryl hydrocarbon receptor (AhR) and aryl hydrocarbon receptor nuclear translocator (ARNT) belong to the bHLH-PAS (basic Helix–Loop–Helix–Period/ARNT/Single-minded) family of transcription factors, which participate in the sensing and transmitting stimuli of exogenous and endogenous chemical substances, and subsequently activates genes transcription involved in various detoxification and physiological functions. However, they have not been identified in Dendroctonus armandi, and their roles in the detoxification metabolism are unclear. In the present study, AhR and ARNT of D. armandi were characterized. Spatiotemporal expression profiling indicated that DaAhR and DaARNT were highly expressed in the adult and larval stages of D. armandi and mainly expressed in the midgut and Malpighian tubules of adults. Additionally, the expression of DaAhR and DaARNT significantly increased after exposure to (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene. Silencing DaAhR and DaARNT increased the susceptibility of D. armandi to (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene, and the activities of detoxification enzyme were also remarkably reduced. Moreover, DaCYP6DF1 and DaGSTs2 were significantly down-regulated after injections of dsAhR and dsARNT in the male and female adults, with the expression of DaCYP6DF1 decreasing by higher than 70%. The present study revealed that the transcription factors AhR and ARNT of D. armandi were induced by terpenoids and participated in the regulation of DaCYP6DF1 expression, which was associated with D. armandi’s susceptibility to (−)-𝛽-pinene and (±)-limonene. These results may provide a theoretical basis for the integrated control of D. armandi and improve our comprehension of insect toxicology.

1. Introduction

The Chinese white pine beetle, Dendroctonus armandi Tsai and Li (Coleoptera: Curculionidae: Scolytinae), is an aggressive and destructive pest in coniferous forests in the middle Qinling Mountains of China. It invades not only healthy Pinus armandii but also attracts other pests to the host trees, which has resulted in the destruction of the forest ecological system and caused serious economic losses [1]. An important period of the bark beetle’s life process is the host colonization stage, during which they must resist the host’s defenses to reproduce successfully [2]. The resistance of P. armandii to bark beetles mainly depends on the composition and induced physical and chemical defense, and the induced oleoresin terpenes are the main defense components [3]. Oleoresin is a complex composed of dozens of monoterpenes, diterpenes, and a few sesquiterpenes [4]. According to a previous study, 𝛼-pinene, 𝛽-pinene, limonene, myrcene, and camphene are the main compounds in the volatile oleoresin terpenes, which were derived from the resin of P. armandi [3]. Moreover, they can harm or kill beetles due to their toxic effects [5,6].
In insects, the detoxification of toxic chemicals mainly contains the enhancement of cytochrome P450 monooxygenase, glutathione S-transferase, and esterase activity levels [7]. The expression of detoxification genes directly affects the activity level of detoxification enzymes [8,9]. The increase of detoxification enzyme activity attributed to the upregulation of gene expression is likely to be influenced by transcriptional regulation, including trans-acting factors and cis-acting elements [10,11,12]. The former, also called transcription factors, bind to specific reaction elements on promoters and regulate the expression of target genes, and then activate or inhibit the transcription of related genes [13]. Three transcription factor superfamilies in insects are known as xenobiotic sensors that regulate the expression of detoxification genes, including basic leucine zipper (bZIP) proteins, nuclear receptors, and basic helix–loop–helix/Per–Arnt–Sim (bHLH-PAS) [14]. In Plutella xylostella, a transcription factor FTZ-F1 affiliated with a nuclear receptor regulates the CYP6BG1 expression, possibly improving its resistance to chlorantraniliprole [15]. In addition, as a bZIP transcription factor, CncC modulates the expression of several GST and P450 genes to enhance malathion resistance in Drosophila melanogaster [16].
The family of bHLH-PAS proteins contains several dimeric transcription factors with multiple functions [17]. AHR is a member of the bHLH PAS protein family, which belongs to a ligand-activated transcription factor that regulates multiple xenobiotic responses [18]. In vertebrates, AhR contains AhR1 and AhR2, with AhR1 expressed in almost all vertebrates and AhR2 only expressed in birds and fish. [19]. ARNT, as another bHLH-PAS protein family member, is a constitutive nuclear protein forming heterodimers with AhR [20]. Heterodimer recognizes and then combines with the xenobiotic response elements of target genes to mediate their expression [21]. The structure of transcription factor AhR/ARNT can be divided into bHLH, PAS, and transcriptional activation domains (TAD) from N- to C termini, the last of which primarily regulate the transcriptional activation of downstream related genes [22]. The AhR and ARNT have only one subtype in insects, which form heterodimers that participate in regulating the related detoxification gene expression [23,24,25]. For instance, AhR/ARNT regulates the CYP6CY3 and CYP6CY4 expression levels to enable Myzus persicae resistance to nicotine [23]. Similarly, the expression of CYP6DA2 is also induced by AhR/ARNT in cotton aphids, which is involved in the resistance to spirotetramat [24]. Moreover, AhR is involved in chlorpyrifos susceptibility in Locusta migratoria by regulating the expression of LmGSTd7 [25].
As we have never used pesticides against D. armandi in the sampling site, the host’s physical and chemical defenses are the main pressure on beetles. Transcription factors AhR and ARNT have not been characterized in D. armandi, and their roles in the metabolic process of detoxification are not fully clear. In the present study, we cloned DaAhR and DaARNT and then analyzed the gene structures. The effects of different terpenoids exposure on the expression level of DaAhR and DaARNT were performed with quantitative real-time PCR. Both genes were knocked down by RNA interference, and the D. armandi susceptibility to (−)-𝛼-pinene, (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene was investigated. The expression of detoxification genes and their respective P450, GST, and CarE activity levels were also measured. Our study indicated the transcription factors DaAhR and DaARNT are involved in the xenobiotic metabolism of D. armandi. These results, which may provide a new perspective for these detoxification enzyme genes, are transcriptionally regulated and a theoretical basis for pest control.

2. Materials and Methods

2.1. Insects and Reagents Preparation

D. armandi were collected and reared as previously described [26]. (−)-𝛼-pinene (98%), (−)-𝛽-pinene (99%), (+)-3-carene (90%), (±)-limonene (95%), and Dimethyl sulfoxide (DMSO) were obtained from the Aladdin Industrial Corporation (Shanghai, China).

2.2. RNA Extraction, cDNA Synthesis, and Reverse Transcription Quantitative PCR (qRT-qPCR)

Total RNA was determined as previously described [27]. The relative expression level of each gene was determined by qRT-PCR. The PCR cycling conditions were performed as previously described [28]. The sequences of β-actin (accession number: KJ507199.1) and CYP4G55 (accession number: JQ855658.1) in D. armandi were used as the reference genes [29,30]. The relative expression levels were analyzed by the 2-ΔΔCt method [31]. All the primers are listed in Table S1.

2.3. Gene Cloning and Bioinformatic Analysis

The specific primers (shown in Table S1) were designed to clone the full-length cDNA sequences of DaAhR and DaARNT. The two obtained sequences were deposited in the GenBank, and their accession numbers are shown in Table 1. In addition, the open reading frames (ORFs) of cDNA sequences were obtained by ORF Finder (https://www.ncbi.nlm.nih.gov/orffinder/, accessed on 20 June 2022). Multiple sequence comparison of proteins was performed with DNAMAN 6.0. Molecular weight (kDa) and Isoelectric points were predicted by ProtParam (http://web.expasy.org/protparam/, accessed on 20 June 2022). MEGA 6.0 was used to construct the phylogenetic trees with the neighbor-joining method [32].

2.4. Developmental- and Tissue-Dependent Expression Profiles of DaAhR and DaARNT

D. armandi larvae were separated into the following stages: early larvae, late larvae, early pupae, late pupae, teneral adults, emerged adults, and feeding adults. The antennae, brain, hindgut, midgut, foregut, fat body, pheromone gland, hemolymph, and Malpighian tubules of emerged adults were collected by dissection and then stored at −80 °C. A total of three independent biological replicates were prepared for gene expression analysis.

2.5. Terpenoids Exposure

Fumigation treatment was performed as previously described [29]. The male and female of emerged adults were divided into six groups and treated with LC50 concentrations of (−)-𝛼-pinene, (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene for 2 h in 20 mL glass vial [33]. The group of DMSO exposure was used as a control. Each treatment contained 40 adults of essentially the same size. After the adults regained their vitality, they were transferred to an artificial climate cabinet. To explore the effect of terpenoids on the expression of DaAhR and DaANRT, the surviving adults were collected at 24 h post-exposure to LC50 of each terpenoid. Meanwhile, the DMSO-treated surviving adults were collected as controls at the same time point.

2.6. Double-Strand RNA (dsRNA) Synthesis and Injection

The synthesis of dsRNA was performed as previously described [26]. Briefly, the T7 Ribo-MAXTM Express RNAi System (Promega, Madison, MI, USA) was used for the synthesis of dsGFP (395 bp), dsAhR (412 bp), dsARNT (387 bp), and dsCYP6DF1 (455 bp). RNAi primers (Table S1) were designed based on the sequences obtained. Injection with dsGFP was used as a control. To prevent off-target effects, we chose specific target fragments to avoid any overlap with other genes, and the sequence specificity of target fragments was tested via NCBI BLAST. The final dsRNA products were diluted to 1000 ng/µL with diethylpyrocarbonate (DEPC)-treated water, then stored at –80 °C and used within 6 months. Before injection, D. armandi were placed in an ice bath for 10 min. The beetles were immobilized on an agarose plate using manual forceps. Afterward, each of the D. armandi emerged adults were microinjected with 0.2 µL dsRNA solution. Each treatment group contained 40 individuals, and 6 individuals from each treatment group were collected 24, 48, and 72 h after injection and then stored at –80 °C until qRT-PCR. Each treatment group contained three biological replicates. In addition, after injection at 48 h, beetles were also used for enzyme determination, and the P450, GST, and CarE activity levels were measured as previously described [7]. At the 48 h after dsRNA injection, the adults were treated with (−)-𝛼-pinene, (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene as described previously, then they were reared in normal conditions for 48 h and the mortality rates were determined. We selected 12 genes from the three classes of D. armandi detoxifying enzymes: 6 CYP genes from the CYP3 clade, 4 GST genes from epsilon and sigma superfamilies, and 2 carboxylesterases, which are involved in the xenobiotic compounds’ detoxification in D. armandi [34,35]. These detoxification gene expression levels (P450, CarE, and GST) after dsRNA injection at 48 h were measured by qRT-PCR. The primers are listed in Table S1. Each treatment group contained three biological replicates.

2.7. Statistical Analysis

SPSS Statistics 19.0 (IBM, Chicago, IL, USA) was performed in all the statistical data analyses. Post-hoc Tukey tests were used to check the difference through one-way ANOVA. The two-sample analyses were performed with Student’s t-test. In addition, Prism 6.0 (GraphPad Software, San Diego, CA, USA) was used to plot graphs.

3. Results

3.1. Sequencing and Bioinformatic Analysis

The full-length D. armandi AhR and ARNT cDNAs were cloned and characterized. The lengths of the coding regions of AhR and ARNT are 2412 bp and 2106 bp, which encode 803 and 701 amino acids, respectively. Moreover, the molecular weights (MW) of proteins of AhR and ARNT are 90.63 and 77.29 kDa, and the isoelectric points (PI) are 7.29 and 6.16, respectively (Table 1). A phylogenetic analysis showed that DaAhR (Figure 1A) and DaARNT (Figure 1B) have a high homology with their counterpart in Dendroctonus ponderosa, and they were also clustered with the Coleoptera group. An amino acid sequence alignment indicated that the bHLH and PAS domains of DaAhR and DaARNT are relatively conserved among different species (Figure S1).

3.2. Spatiotemporal Expression Pattern of DaAhR and DaARNT

AhR and ARNT were expressed in all developmental stages of D. armandi. They were expressed the highest in the adult stage, followed by the larval stage, and the lowest in pupae (Figure 2). Moreover, as for the adult stage, the expression level of AhR in females was significantly higher than that in males in the feeding adults (Figure 2A). While ARNT showed the opposite result in the feeding adults, there was no statistical significance (Figure 2B). In addition, the expression of AhR and ARNT was found at different levels in tissues, and there were gender differences in some tissues. Specifically, AhR and ARNT were expressed predominantly in the midgut and Malpighian tubules (Figure 3). Moreover, the expression of AhR in female adults was higher than in male adults in the midgut (Figure 3).

3.3. Exposure to Terpenoids

As shown in Figure 4 the expression of DaAhR and DaARNT can be induced to varying levels at 48 h after terpenoids treatment in male and female adults. After (−)-𝛼-pinene, (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene exposure, the expression of DaAhR significantly increased by 2.72-, 1.17-, and 3.35-folds in male adults (Figure 4A), and 4.22-, 2.39-, and 5.46-folds in female adults (Figure 4B), respectively, as compared with DMSO-treated controls. DaARNT showed increases by 1.11-, 2.47-, and 2.99-folds in male adults (Figure 4C) and 2.17-, 1.25-, and 1.85-folds in female adults (Figure 4D), respectively. However, there was no significant change after (−)-𝛼-pinene treatment in D. armandi.

3.4. Functional Analysis of DaAhR and DaARNT by RNAi Silencing

To further investigate the role of DaAhR and DaARNT in host chemical defense, the expression of two genes was repressed by RNAi in male and female adults. Compared with the control group, after dsRNA injection, the expression levels of DaAhR and DaARNT of adults were significantly downregulated at 24, 48, and 72 h, except for in male adults at 24 h (Figure 5A,D). In addition, the expression level of DaAhR and DaARNT decreased most at 72 h in male and female adults, reaching 80.3% and 57.0%, 86.0%, and 55.7%, respectively (Figure 5A,D). Moreover, the relative activities of P450, GST, and CarE remarkably reduced compared to the control after the injection of dsRNA (Figure 5B,E). Additionally, knocking down DaAhR and DaARNT increased the beetle’s susceptibility to terpenoids. The mortality of dsAhR and dsARNT-injected beetles elevated by 37.1% and 21.2%, 22.4% and 23.3, and 43.4% and 38.1% in male adults (Figure 5C), respectively, after the (−)-𝛽-pinene (+)-3-carene and (±)-limonene treatment when compared with the control. The female adults were enhanced by 34.3% and 18.7%, 20.3% and 20.1, and 36.4% and 19.5%, respectively (Figure 5F). Nevertheless, there was no significant change in (−)-𝛼-pinene susceptibility (Figure 5C,F).

3.5. Analysis of DaAhR and DaARNT Regulation of Detoxification Genes

To further explore the role of DaAhR and DaARNT in the detoxification metabolism of D. armandi, the expression levels of detoxification-related genes were determined using qRT-PCR when DaAhR and DaARNT were silenced. Among the P450, GST, and CarE genes, CYP6CR2, CYP6DF1, GSTs1, GSTs2, and CarE4 expression levels were significantly downregulated when DaAhR and DaARNT were reduced in male adults (Figure 6A). The expression levels of CYP6BX1, CYP6DE5, CYP6DF1, GSTe1, GSTe4, GSTs2, and CarE3 were significantly inhibited when silencing DaAhR and DaARNT in female adults (Figure 6B). These results indicated that DaAhR and DaARNT regulate the expression levels of detoxification genes and the metabolic detoxification enzyme activity of the corresponding genes participating in D. armandi’s susceptibility to host allelochemicals.
Moreover, the expression of CYP6DF1 and GSTs2 significantly decreased after DaAhR and DaARNT injection, with CYP6DF1 being reduced by higher than 70% (Figure 6). To further reveal the role of CYP6DF1 in the detoxification metabolism, we determined the susceptibility of D. armandi to four terpenoids after silencing CYP6DF1. The expression of DaCYP6DF1 in adults was down-regulated significantly at 24, 48, and 72 h when compared to the control after dsRNA injection. In addition, the expression level of DaCYP6DF1 decreased the most at 72 h in male and female adults, reaching 81.7% and 62.4%, respectively (Figure 7A,B). The mortality significantly elevated in the dsCYP6DF1-injected group after treatment with (−)-𝛽-pinene and (±)-limonene compared with the control, with 38.7% and 37.3% in male adults, 21.4% and 31.3% in female adults, respectively (Figure 7C,D). Nevertheless, there was no significant change in (−)-𝛼-pinene and (+)-3-carene susceptibility.

4. Discussion

During D. armandi attack host trees, they must overcome the induced defense system of the pine trees, which contains a large number of terpenoids [3]. To avoid host toxicity, insects have evolved a complicated regulatory mechanism to solve xenobiotics exposure, including enhanced detoxification metabolisms [36]. Although the relationship between the induction of relevant detoxification genes and insect tolerance to exogenous chemicals has been widely reported, the regulatory cascades of these xenobiotic response genes in D. armandi are still basically unknown. In the present study, we clarified the effects of transcription factors DaAhR and DaARNT on D. armandi downstream-related detoxification genes and relevant regulation metabolism in plant toxin resistance.
Stage-dependent expression patterns of DaAhR and DaARNT were detected in different developmental stages with RT-qPCR, and the results indicated that DaAhR and DaARNT were expressed in all developmental stages of D. armandi. The highest expression level was detected in the larval and adult stage, and the lowest was in the pupae, which may be involved in detoxifying xenobiotics in food. Moreover, the expression of DaAhR and DaARNT was further investigated in various tissues of adults, suggesting that it was expressed predominantly in the midgut and Malpighian tubules compared to other parts. The midgut is not only the place where insects digest and absorb but also plays a key role in resisting exogenous substances [37]. Previous studies have reported that the Malpighian tube also has a function in xenobiotics detoxification [38,39]. It was worth noting that the expression of DaAhR in female adults was higher than in male adults in the feeding adults and midgut, while there was no gender difference in DaARNT. This indicates that DaAhR plays an important role in the location and detoxification of host volatiles in female adults. Therefore, the overexpression of DaAhR and DaARNT in the detoxification organ of D. armandi may play a key role in the detoxification of exogenous substances. Further studies can be performed to show the function of DaAhR and DaARNT in other different developmental stages and tissues of insects.
Under normal circumstances, the increased metabolism in insects arises from the increased detoxification effects attributed to amplification, constitutive overexpression, and mutation of genes [40]. Elements in promoters and transcription factors mediate transcriptional regulation and participate in the expression of detoxification genes. The suppression of hepatocyte nuclear factor 4 by transcription factor results in the resistance to imidacloprid in N. lugens through regulating UDP-glycosyltransferase (UGT) and P450 genes [41]. The promoter region motif may be a candidate region for the transcription of tolerance-related genes [42,43]. A cis-acting element was called a xenobiotic response element (XRE) of flavonoids in Helicoverpa zea, which mediated the expression of CYP321A1 induced by flavonoids [44]. In Spodoptera litura, Nrf2 activated the GSTe1 expression by binding to the cis-acting element in its promoter in response to insecticides and phytochemicals [45]. As vital transcription factors for xenobiotic sensors, AhR and ARNT regulate a large number of metabolic genes in various insect species [10]; however, their roles in D. armandi detoxification remain unclear.
In the present study, the expression of DaAhR and DaARNT significantly increased after treatment with (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene, and silencing DaAhR and DaARNT enhanced the susceptibility of D. armandi, with the decrease of enzyme activity level corresponding to greater sensitivity. This revealed that the AhR/ ARNT signal pathway in D. armandi may be involved in metabolic responses to environmental stresses associated with host allelochemical exposure. In addition, DaAhR and DaARNT are homologous to other species. For instance, the AhR signal pathway is induced when mediating chemical-induced developmental toxicity in zebrafish [46]. Silencing LmAhR enhanced the susceptibility to chlorpyrifos in L. migratoria [25]. Moreover, the expression of the transcription factor SlituCncC is induced after indoxacarb exposure in S. litura [47]. Additionally, silencing NlAhR and NlARNT affects isoprocarb, imidacloprid, and etofenprox susceptibility in N. lugens [48].
The change in detoxification enzyme activity is due to the change in the expression of detoxification genes. In this study, silencing DaAhR and DaARNT resulted in the differential downregulation of some detoxification enzyme genes, with DaCYP6DF1 being downregulated by higher than 70%. These results indicated that DaCYP6DF1 was supposed to be a candidate gene that mainly participated in host chemical susceptibility. Similar results have also been reported in other insects. For example, AhR/ARNT of M. persicae regulated the expression of CYP6CY3 and CYP6CY4 cooperatively, leading to the nicotine adaption to tobacco [23]. The cis-regulatory elements AhR/ARNT and CncC/Maf participated in the expression regulation of some GST detoxification enzyme genes in Spodoptera exigua. Moreover, the constitutive overexpression of CncC, Maf, and AhR promoted the increased expression of several detoxification-related genes and led to chlorpyrifos and cypermethrin resistance [49]. In addition, NlAhR and NlARNT were induced by pesticides and participated in the NlCarE7 regulation, which was involved in the susceptibility to etofenprox and isoprocarb in N. lugens [48]. Previous studies have reported that CYP genes in D. armandi were remarkably increased in response to (−)-𝛼-pinene, (−)-𝛽-pinene, (+)-3-carene, and (±)-limonene treatments [29,34,35]. Because knockdowns of DaCYP6DF1 significantly increased D. armandi susceptibility to (−)-𝛽-pinene and (±)-limonene, we supposed that DaAhR and DaARNT regulate the susceptibility by targeting DaCYP6DF1.

5. Conclusions

These results indicated that the transcription factors DaAhR and DaARNT of D. armandi were induced by terpenoids and involved in mediating the expression level of DaCYP6DF1, which was associated with D. armandi susceptibility to (−)-𝛽-pinene and (±)-limonene. Our study showed the potential roles of DaAhR and DaARNT in the detoxification metabolism of D. armandi. These findings uncovered the molecular mechanisms of host allelochemical detoxification in insects and may provide a theoretical basis for controlling this pest.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11233856/s1, Figure S1: Multiple amino acid sequence alignments of DaAhR (A) and DaARNT (B) with six other insect species. Three characteristic domains (bHLH, PAS-A and PAS-B) of DaAhR and DaARNT are boxed in red. Table S1: Primer sequences used in the research.

Author Contributions

B.L.: Investigation, Data curation, Methodology, Visualization, Validation, Software, Writing—original draft, Writing—review & editing. H.C.: Conceptualization, Funding acquisition, Methodology, Supervision, Writing—original draft, Writing—review & editing, Visualization, Validation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (31870636) and the Laboratory of Lingnan Modern Agriculture Project (NZ2021025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are included in the text and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, H.; Li, Z.; Tang, M. Laboratory evaluation of flight activity of Dendroctonus armandi (Coleoptera: Curculionidae: Scolytinae). Can. Entomol. 2010, 142, 378–387. [Google Scholar] [CrossRef]
  2. Huber, D.P.W.; Robert, J.A. The proteomics and transcriptomics of early host colonization and overwintering physiology in the Mountain pine beetle, Dendroctonus ponderosae Hopkins (Coleoptera: Curculionidae). In Pine Bark Beetles; Tittiger, C., Blomquist, G.J., Eds.; Academic Press: Cambridge, MA, USA, 2016; pp. 101–128. [Google Scholar]
  3. Chen, H.; Tang, M.; Gao, J.; Chen, X.; Li, Z. Changes in the composition of volatile monoterpenes and sesquiterpenes of Pinus armandi, P-tabulaeformis, and P-bungeana in Northwest China. Chem. Nat. Compd. 2006, 42, 534–538. [Google Scholar] [CrossRef]
  4. Bohlmann, J. Pine terpenoid defences in the mountain pine beetle epidemic and in other conifer pest interactions: Specialized enemies are eating holes into a diverse, dynamic and durable defence system. Tree Physiol. 2012, 32, 943–945. [Google Scholar] [CrossRef] [PubMed]
  5. Fernanda Lopez, M.; Cano-Ramirez, C.; Shibayama, M.; Zuniga, G. Alpha-Pinene and myrcene induce ultrastructural changes in the midgut of Dendroctonus valens (Coleoptera: Curculionidae: Scolytinae). Ann. Entomol. Soc. Am. 2011, 104, 553–561. [Google Scholar] [CrossRef] [Green Version]
  6. Chiu, C.C.; Keeling, C.; Bohlmann, J. Toxicity of pine monoterpenes to Mountain pine beetle. Sci. Rep. 2017, 7, 8858. [Google Scholar] [CrossRef]
  7. Lu, K.; Cheng, Y.; Li, Y.; Li, W.; Zeng, R.; Song, Y. Phytochemical flavone confers broad-spectrum tolerance to insecticides in Spodoptera litura by activating ROS/CncC-mediated xenobiotic detoxification pathways. J. Agric. Food Chem. 2021, 69, 7429–7445. [Google Scholar] [CrossRef]
  8. Zhang, X.; Liao, X.; Mao, K.; Yang, P.; Li, D.; Alia, E.; Wan, H.; Li, J. The role of detoxifying enzymes in field-evolved resistance to nitenpyram in the brown planthopper Nilaparvata lugens in China. Crop Prot. 2017, 94, 106–114. [Google Scholar] [CrossRef]
  9. Mao, K.; Zhang, X.; Ali, E.; Liao, X.; Jin, R.; Ren, Z.; Wan, H.; Li, J. Characterization of nitenpyram resistance in Nilaparvata lugens (Stal). Pestic. Biochem. Physiol. 2019, 157, 26–32. [Google Scholar] [CrossRef]
  10. Amezian, D.; Nauen, R.; Le Goff, G. Transcriptional regulation of xenobiotic detoxification genes in insects—An overview. Pestic. Biochem. Physiol. 2021, 174, 104822. [Google Scholar] [CrossRef]
  11. Kalsi, M.; Palli, S.R. Transcription factors, CncC and Maf, regulate expression of CYP6BQ genes responsible for deltamethrin resistance in Tribolium castaneum. Insect Biochem. Mol. Biol. 2015, 65, 47–56. [Google Scholar] [CrossRef]
  12. Kalsi, M.; Palli, S.R. Transcription factor cap n collar C regulates multiple cytochrome P450 genes conferring adaptation to potato plant allelochemicals and resistance to imidacloprid in Leptinotarsa decemlineata (Say). Insect Biochem. Mol. Biol. 2017, 83, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Misra, J.R.; Lam, G.; Thummel, C.S. Constitutive activation of the Nrf2/Keap1 pathway in insecticide-resistant strains of Drosophila. Insect Biochem. Mol. Biol. 2013, 43, 1116–1124. [Google Scholar] [CrossRef] [Green Version]
  14. Palli, S.R. CncC/Maf-mediated xenobiotic response pathway in insects. Arch. Insect Biochem. Physiol. 2020, 104, e21674. [Google Scholar] [CrossRef]
  15. Li, X.; Shan, C.; Li, F.; Liang, P.; Smagghe, G.; Gao, X. Transcription factor FTZ-F1 and cis -acting elements mediate expression of CYP6BG1 conferring resistance to chlorantraniliprole in Plutella xylostella. Pest Manag. Sci. 2018, 75, 1172–1180. [Google Scholar] [CrossRef]
  16. Misra, J.R.; Horner, M.A.; Lam, G.; Thummel, C.S. Transcriptional regulation of xenobiotic detoxification in Drosophila. Genes Dev. 2011, 25, 1796–1806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Saxlund, M.A.; Sadler-Riggleman, I.; Skinner, M.K. Role of basic helix-loop-helix (bHLH) and CREB transcription factors in the regulation of sertoli cell androgen-binding protein expression. Mol. Reprod. Dev. 2004, 68, 269–278. [Google Scholar] [CrossRef]
  18. Vorrink, S.U.; Domann, F.E. Regulatory crosstalk and interference between the xenobiotic and hypoxia sensing pathways at the AhR-ARNT-HIF1 alpha signaling node. Chem. Biol. Interact. 2014, 218, 82–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Jin, H.; Ji, C.; Ren, F.; Aniagu, S.; Tong, J.; Jiang, Y.; Chen, T. AHR-mediated oxidative stress contributes to the cardiac developmental toxicity of trichloroethylene in zebrafish embryos. J. Hazard. Mater. 2019, 385, 121521. [Google Scholar] [CrossRef]
  20. Berg, P.; Pongratz, I. Differential usage of nuclear export sequences regulates intracellular localization of the dioxin (aryl hydrocarbon) receptor. J. Biol. Chem. 2001, 276, 43231–43238. [Google Scholar] [CrossRef] [Green Version]
  21. Kikuchi, Y.; Ohsawa, S.; Mimura, J.; Ema, M.; Takasaki, C.; Sogawa, K.; Fujii-Kuriyama, Y. Heterodimers of bHLH-PAS protein fragments derived from AhR, AhRR, and Arnt prepared by co-expression in Escherichia coli: Characterization of their DNA binding activity and preparation of a DNA complex. J. Biochem. 2003, 134, 83–90. [Google Scholar] [CrossRef] [PubMed]
  22. Kudo, I.; Hosaka, M.; Haga, A.; Tsuji, N.; Nagata, Y.; Okada, H.; Fukuda, K.; Kakizaki, Y.; Okamoto, T.; Grave, E.; et al. The regulation mechanisms of AhR by molecular chaperone complex. J. Biochem. 2017, 163, 223–232. [Google Scholar] [CrossRef] [PubMed]
  23. Pan, Y.; Peng, T.; Xu, P.; Zeng, X.; Tian, F.; Song, J.; Shang, Q. Transcription factors AhR/ARNT regulate the expression of CYP6CY3 and CYP6CY4 switch conferring nicotine adaptation. Int. J. Mol. Sci. 2019, 20, 4521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Peng, T.; Chen, X.; Pan, Y.; Zheng, Z.; Wei, X.; Xi, J.; Zhang, J.; Gao, X.; Shang, Q. Transcription factor aryl hydrocarbon receptor/aryl hydrocarbon receptor nuclear translocator is involved in regulation of the xenobiotic tolerance-related cytochrome P450 CYP6DA2 in Aphis gossypii Glover. Insect Mol. Biol. 2017, 26, 485–495. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, X.; Jie, D.; Liu, J.; Zhang, J.; Zhang, T.; Zhang, J.; Ma, E. Aryl hydrocarbon receptor regulates the expression of LmGSTd7 and is associated with chlorpyrifos susceptibility in Locusta migratoria. Pest Manag. Sci. 2019, 75, 2916–2924. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, B.; Fu, D.; Gao, H.; Ning, H.; Sun, Y.; Chen, H.; Tang, M. Cloning and expression of the neuropeptide F and neuropeptide F receptor genes and their regulation of food intake in the Chinese white pine beetle Dendroctonus armandi. Front. Physiol. 2021, 12, 662651. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, B.; Fu, D.; Ning, H.; Tang, M.; Chen, H. Identification of the short neuropeptide F and short neuropeptide F receptor genes and their roles of food intake in Dendroctonus armandi. Insects 2021, 12, 844. [Google Scholar] [CrossRef]
  28. Liu, B.; Fu, D.; Ning, H.; Tang, M.; Chen, H. Identification and functional characterization of the sulfakinin and sulfakinin receptor in the Chinese white pine beetle Dendroctonus armandi. Front. Physiol. 2022, 13, 927890. [Google Scholar] [CrossRef]
  29. Dai, L.; Ma, M.; Wang, C.; Shi, Q.; Zhang, R.; Chen, H. Cytochrome P450s from the Chinese white pine beetle, Dendroctonus armandi (Curculionidae: Scolytinae): Expression profiles of different stages and responses to host allelochemicals. Insect Biochem. Mol. Biol. 2015, 65, 35–46. [Google Scholar] [CrossRef]
  30. Vandesompele, J.; De Preter, K.; Pattyn, F.; Poppe, B.; Van Roy, N.; De Paepe, A.; Speleman, F. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 2002, 3, research0034.1. [Google Scholar] [CrossRef] [Green Version]
  31. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2–ΔΔCt method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  32. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhao, M.; Kang, X.; An, H.; Liu, B.; Chen, H. Two key volatiles of Chinese white pine (Pinus armandii) (Pinales: Pinaceae: Pinoideae) phloem resist invasion by Chinese white pine beetle (Dendroctonus armandii) (Coleoptera: Curculionidae: Scolytinae). Appl. Ecol. Environ. Res. 2022, 20, 587–600. [Google Scholar] [CrossRef]
  34. Dai, L.; Gao, H.; Chen, H. Expression levels of detoxification enzyme genes from Dendroctonus armandi (Coleoptera: Curculionidae) fed on a solid diet containing pine phloem and terpenoids. Insects 2021, 12, 926. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, B.; Fu, D.; Ning, H.; Tang, M.; Chen, H. Knockdown of CYP6CR2 and CYP6DE5 reduces tolerance to host plant allelochemicals in the Chinese white pine beetle Dendroctonus armandi. Pestic. Biochem. Physiol. 2022, 187, 105180. [Google Scholar] [CrossRef] [PubMed]
  36. Liao, X.; Xu, P.; Gong, P.; Wan, H.; Li, J. Current susceptibilities of brown planthopper Nilaparvata lugens to triflumezopyrim and other frequently used insecticides in China. Insect Sci. 2020, 28, 115–126. [Google Scholar] [CrossRef]
  37. Zhang, X.; Dong, J.; Wu, H.; Zhang, H.; Zhang, J.; Ma, E. Knockdown of cytochrome P450 CYP6 family genes increases susceptibility to carbamates and pyrethroids in the migratory locust, Locusta migratoria. Chemosphere 2019, 223, 48–57. [Google Scholar] [CrossRef]
  38. Dow, J.A.T.; Davies, S.A. The Malpighian tubule: Rapid insights from post-genomic biology. J. Insect Physiol. 2006, 52, 365–378. [Google Scholar] [CrossRef]
  39. Huang, Y.; Shen, G.-M.; Jiang, H.-B.; Jiang, X.-Z.; Dou, W.; Wang, J.-J. Multiple P450 genes: Identification, tissue-specific expression and their responses to insecticide treatments in the oriental fruit fly, Bactrocera dorsalis (Hendel) (Diptera: Tephritidea). Pestic. Biochem. Physiol. 2013, 106, 1–7. [Google Scholar] [CrossRef]
  40. Li, X.; Schuler, M.A.; Berenbaum, M.R. Molecular mechanisms of metabolic resistance to synthetic and natural xenobiotics. Annu. Rev. Èntomol. 2007, 52, 231–253. [Google Scholar] [CrossRef]
  41. Cheng, Y.; Li, Y.; Li, W.; Song, Y.; Zeng, R.; Lu, K. Inhibition of hepatocyte nuclear factor 4 confers imidacloprid resistance in Nilaparvata lugens via the activation of cytochrome P450 and UDP-glycosyltransferase genes. Chemosphere 2020, 263, 128269. [Google Scholar] [CrossRef]
  42. Wittkopp, P.J.; Kalay, G. Cis-regulatory elements: Molecular mechanisms and evolutionary processes underlying divergence. Nat. Rev. Genet. 2012, 13, 59–69. [Google Scholar] [CrossRef]
  43. Lenhard, B.; Sandelin, A.; Carninci, P. Metazoan promoters: Emerging characteristics and insights into transcriptional regulation. Nat. Rev. Genet. 2012, 13, 233–245. [Google Scholar] [CrossRef]
  44. Zhang, C.; Luo, X.; Ni, X.; Zhang, Y.; Li, X. Functional characterization of cis-acting elements mediating flavone-inducible expression of CYP321A1. Insect Biochem. Mol. Biol. 2010, 40, 898–908. [Google Scholar] [CrossRef]
  45. Chen, S.; Lu, M.; Zhang, N.; Zou, X.; Mo, M.; Zheng, S. Nuclear factor erythroid-derived 2-related factor 2 activates glutathione S-transferase expression in the midgut of Spodoptera litura (Lepidoptera: Noctuidae) in response to phytochemicals and insecticides. Insect Mol. Biol. 2018, 27, 522–532. [Google Scholar] [CrossRef]
  46. Billiard, S.M.; Timme-Laragy, A.; Wassenberg, D.M.; Cockman, C.; Di Giulio, R.T. The role of the aryl hydrocarbon receptor pathway in mediating synergistic developmental toxicity of polycyclic aromatic hydrocarbons to zebrafish. Toxicol. Sci. 2006, 92, 526–536. [Google Scholar] [CrossRef] [Green Version]
  47. Shi, L.; Shi, Y.; Liu, M.F.; Zhang, Y.; Liao, X.L. Transcription factor CncC potentially regulates the expression of multiple detoxification genes that mediate indoxacarb resistance in Spodoptera litura. Insect Sci. 2020, 28, 1426–1438. [Google Scholar] [CrossRef]
  48. Wang, Y.; Jin, R.; Liu, C.; Gao, Y.; Deng, X.; Wan, H.; Li, J. Functional characterization of the transcription factors AhR and ARNT in Nilaparvata lugens. Pestic. Biochem. Physiol. 2021, 176, 104875. [Google Scholar] [CrossRef]
  49. Hu, B.; Huang, H.; Wei, Q.; Ren, M.; Mburu, D.K.; Tian, X.; Su, J. Transcription factors CncC/Maf and AhR/ARNT coordinately regulate the expression of multiple GSTs conferring resistance to chlorpyrifos and cypermethrin in Spodoptera exigua. Pest Manag. Sci. 2019, 75, 2009–2019. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic analysis of D. armandi AhR (A) and ARNT (B) with other insect species. The phylogenetic tree was constructed in MEGA 6.0 using the neighbor-joining method. Bootstrap values (1000 replicates) are indicated next to the branches, and GenBank accession numbers are shown in parentheses. The black dot indicates D. armandi AhR and ARNT.
Figure 1. Phylogenetic analysis of D. armandi AhR (A) and ARNT (B) with other insect species. The phylogenetic tree was constructed in MEGA 6.0 using the neighbor-joining method. Bootstrap values (1000 replicates) are indicated next to the branches, and GenBank accession numbers are shown in parentheses. The black dot indicates D. armandi AhR and ARNT.
Cells 11 03856 g001
Figure 2. Relative expression levels of AhR (A) and ARNT (B) in different developmental stages of D. armandi. The relative expression levels were normalized with β-actin and CYP4G55. Different lowercase letters indicate significant differences at p < 0.05. Post-hoc Tukey tests following one-way analysis of variance (ANOVA). All values are mean ± SE, n = 3. EL, early larvae; LL, late larvae; EP, early pupae; LP, late pupae; TA, teneral adult; EA, emerged adult; FA, feeding adult.
Figure 2. Relative expression levels of AhR (A) and ARNT (B) in different developmental stages of D. armandi. The relative expression levels were normalized with β-actin and CYP4G55. Different lowercase letters indicate significant differences at p < 0.05. Post-hoc Tukey tests following one-way analysis of variance (ANOVA). All values are mean ± SE, n = 3. EL, early larvae; LL, late larvae; EP, early pupae; LP, late pupae; TA, teneral adult; EA, emerged adult; FA, feeding adult.
Cells 11 03856 g002
Figure 3. Relative expression levels of emerged adults of AhR (A) and ARNT (B) in different tissues of D. armandi. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between female and male expression levels (* p < 0.05, independent Student’s Test). All values are mean ± SE, n = 3. AN, antennae; BR, brain; FG, foregut; MG, midgut; HG, hindgut; FB, fat body; MT, Malpighian tubules; PG, pheromone gland; HE, hemolymph.
Figure 3. Relative expression levels of emerged adults of AhR (A) and ARNT (B) in different tissues of D. armandi. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between female and male expression levels (* p < 0.05, independent Student’s Test). All values are mean ± SE, n = 3. AN, antennae; BR, brain; FG, foregut; MG, midgut; HG, hindgut; FB, fat body; MT, Malpighian tubules; PG, pheromone gland; HE, hemolymph.
Cells 11 03856 g003
Figure 4. Relative expression levels of AhR and ARNT in male adults (A,C) and female adults (B,D) of D. armandi after stimulation with four terpenoids at an exposure time of 48 h. The relative expression levels were normalized with β-actin and CYP4G55. Different letters indicate significant differences at p < 0.05. Post-hoc Tukey tests following one-way analysis of variance (ANOVA). All values are mean ± SE, n = 3.
Figure 4. Relative expression levels of AhR and ARNT in male adults (A,C) and female adults (B,D) of D. armandi after stimulation with four terpenoids at an exposure time of 48 h. The relative expression levels were normalized with β-actin and CYP4G55. Different letters indicate significant differences at p < 0.05. Post-hoc Tukey tests following one-way analysis of variance (ANOVA). All values are mean ± SE, n = 3.
Cells 11 03856 g004
Figure 5. Functional analyses of AhR and ARNT by RNAi in D. armandi adults. (A,D) Relative expression levels of DaAhR and DaARNT in adults injected with dsRNA at 24, 48, and 72 h post-injection. (B,E) Relative detoxification enzyme activities after the silencing of DaAhR and DaARNT. (C,F) The mortality of adults exposed to four terpenoids was assessed at 48 h after dsRNA injection. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, *** p < 0.001, independent Student’s Test). All values are mean ± SE, n = 3.
Figure 5. Functional analyses of AhR and ARNT by RNAi in D. armandi adults. (A,D) Relative expression levels of DaAhR and DaARNT in adults injected with dsRNA at 24, 48, and 72 h post-injection. (B,E) Relative detoxification enzyme activities after the silencing of DaAhR and DaARNT. (C,F) The mortality of adults exposed to four terpenoids was assessed at 48 h after dsRNA injection. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, *** p < 0.001, independent Student’s Test). All values are mean ± SE, n = 3.
Cells 11 03856 g005
Figure 6. RT-qPCR analyses of the expression of 12 detoxification genes after silencing AhR (A) and ARNT (B) in D. armandi adults. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, independent Student’s Test). All values are mean ± SE, n = 3.
Figure 6. RT-qPCR analyses of the expression of 12 detoxification genes after silencing AhR (A) and ARNT (B) in D. armandi adults. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, independent Student’s Test). All values are mean ± SE, n = 3.
Cells 11 03856 g006
Figure 7. (A,B) Relative expression levels of DaCYP6DF1 in adults injected with dsRNA at 24, 48, and 72 h post-injection. (C,D) The mortality of adults exposed to four terpenoids was assessed at 48 h after dsRNA injection. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, *** p < 0.001, independent Student’s Test). All values are mean ± SE, n = 3.
Figure 7. (A,B) Relative expression levels of DaCYP6DF1 in adults injected with dsRNA at 24, 48, and 72 h post-injection. (C,D) The mortality of adults exposed to four terpenoids was assessed at 48 h after dsRNA injection. The relative expression levels were normalized with β-actin and CYP4G55. The asterisk indicates a significant difference between treatment groups (* p < 0.05, ** p < 0.01, *** p < 0.001, independent Student’s Test). All values are mean ± SE, n = 3.
Cells 11 03856 g007
Table 1. Physicochemical properties of putative D. armandi AhR and ARNT proteins.
Table 1. Physicochemical properties of putative D. armandi AhR and ARNT proteins.
Gene NameAccession NoORF (bp) aAmino Acid Residues aMW (KDa) aIP a
AhROP378567241280390.637.29
ARNTON378568210670177.296.16
Note: a As predicted by BLAST (http://www.ncbi.nlm.nih.gov, accessed on 20 June 2022).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, B.; Chen, H. Identification and Functional Characterization of the Transcription Factors AhR/ARNT in Dendroctonus armandi. Cells 2022, 11, 3856. https://doi.org/10.3390/cells11233856

AMA Style

Liu B, Chen H. Identification and Functional Characterization of the Transcription Factors AhR/ARNT in Dendroctonus armandi. Cells. 2022; 11(23):3856. https://doi.org/10.3390/cells11233856

Chicago/Turabian Style

Liu, Bin, and Hui Chen. 2022. "Identification and Functional Characterization of the Transcription Factors AhR/ARNT in Dendroctonus armandi" Cells 11, no. 23: 3856. https://doi.org/10.3390/cells11233856

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop