Next Article in Journal
Pentastiridius leporinus (Linnaeus, 1761) as a Vector of Phloem-Restricted Pathogens on Potatoes: ‘Candidatus Arsenophonus Phytopathogenicus’ and ‘Candidatus Phytoplasma Solani’
Previous Article in Journal
Symbiotic Bacterial Communities of Insects Feeding on the Same Plant Lineage: Distinct Composition but Congruent Function
Previous Article in Special Issue
Cyantraniliprole and Thiamethoxam Exposure Changes Expression of Transcripts Associated with Small Non-Coding RNA Processing in the Colorado Potato Beetle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

RNA Interference Reveals the Impacts of CYP6CY7 on Imidacloprid Resistance in Aphis glycines

College of Plant Protection, Northeast Agricultural University, Harbin 150030, China
*
Authors to whom correspondence should be addressed.
Insects 2024, 15(3), 188; https://doi.org/10.3390/insects15030188
Submission received: 3 January 2024 / Revised: 9 March 2024 / Accepted: 12 March 2024 / Published: 13 March 2024
(This article belongs to the Special Issue Challenges and Future Trends of RNA Interference in Insects)

Abstract

:

Simple Summary

The soybean aphid Aphis glycines is a globally present agricultural pest that is increasingly developing resistance to pesticides, causing huge economic losses. Cytochrome P450 monooxygenases (CYP450s) are major detoxifying enzymes that metabolize plant toxins and insecticides. In this study, we evaluated the relative expression levels of the CYP6 families in response to imidacloprid treatment using qRT-PCR. The gene CYP6CY7 was significantly increased after treatment induction. Using a transdermal delivery system for nanocarriers to carry siRNAs and detergents that attach to the nota of aphids, this gene was successfully altered, thus significantly decreasing the gene’s P450 activity by up to 21.96%. A further bioassay showed that the mortality of A. glycines due to imidacloprid treatment increased by 14.71%. Overall, our findings indicate that CYP6CY7 might detoxify imidacloprid in A. glycines, providing a theoretical basis for the further study of the mechanisms of action of CYP6s and potential new methods for improving insecticidal efficacy.

Abstract

Cytochrome P450 (CYP) is a group of important detoxification enzymes found in insects related to their resistance to insecticides. To elucidate the CYP6 family genes of P450, which are potentially related to imidacloprid resistance in Aphis glycines, the CYP6 cDNA sequences of A. glycines were studied. The transcriptome of A. glycines was constructed, and the CYP6 cDNA sequences of A. glycines were screened. Their relative expression levels in response to imidacloprid induction were examined through qRT-PCR, and the CYP6s with higher expression levels were used to study the detoxification of imidacloprid through RNA interference and a bioassay. Twelve CYP6s were obtained from the A. glycines transcriptome. These samples were named by the International P450 Nomenclature Committee and registered in GenBank. After 3, 6, 12, 24 and 48 h of induction with LC50 concentrations of imidacloprid, the relative expression levels of these CYP6s increased; the expression level of CYP6CY7 experienced the highest increase, being more than 3-fold higher than that of those of the non-imidacloprid-induced CYP6s. After RNA interference for CYP6CY7, the relative expression level of CYP6CY7 significantly decreased after 3, 6 and 12 h, while the corresponding P450 enzyme activity decreased after 12 and 24 h. The mortality of A. glycines due to imidacloprid treatment increased by 14.71% at 24 h. CYP6CY7 might detoxify imidacloprid in A. glycines. This study provides a theoretical basis for the further study of the mechanism of action of CYP6s and potential new methods for improving insecticidal efficacy.

Graphical Abstract

1. Introduction

The soybean aphid Aphis glycines is a major soybean-damaging pest native to Asia and invasive in North America [1,2]. Not only does A. glycines damage soybean by performing sucking behavior, but it is also an important vector for transferring many viruses that affect soybean quality and production, such as soybean mosaic virus (SMV), alfalfa mosaic virus (AMV), potato Y virus (PVY) and tobacco ringspot virus (TRSV) [3,4,5,6,7]. Additionally, the growth of black sooty mold fungus on honeydew produced by A. glycines affects soybean photosynthesis [8]. Chemical control is the main prevention measure against yield loss inflicted by soybean aphids. However, extensive use of insecticides has led to an increase in production cost, a reduction in natural enemies and resistance of soybean aphid populations.
The invention and development of neonicotinoid insecticides provides valuable, new tools for global agricultural producers to control pests. Imidacloprid is a first-generation neonicotinoid insecticide and is widely used in agriculture [9]. Unlike traditional insecticides, imidacloprid selectively acts on nicotinic acetylcholine receptors (nAChRs) in the insect nervous system by stimulating nAChRs on the post-synaptic membrane, disrupting the normal transmission of the insect central nervous system and paralyzing or even killing insects [10]. In mammals, imidacloprid’s ability to penetrate the blood–brain barrier is poor. Moreover, acetylcholine receptors in the central nervous system and peripheral nervous system have weaker interactions, thus exhibiting low toxicity to mammals [11,12]. Imidacloprid, due to its special mechanism of action, has a great control effect on the piercing–sucking mouthparts of pests such as aphids, planthoppers, leafhoppers and whiteflies. Therefore, it is widely used to control sucking pests affecting crops in China. However, long-term use will lead to increased resistance among A. glycines to imidacloprid [13]. It is vital to study the resistance mechanism of imidacloprid in A. glycines to increase the insecticidal efficacy of imidacloprid in the field.
Pest resistance arises from the use of insecticides. Under strong insecticide selection pressure, agricultural pests experience rapid adaptations and evolution [14]. Insect pests develop insecticide resistance through gene mutation, gene amplification, metabolic enzyme activity change, epidermal penetration change, behavioral change and other mechanisms [15,16,17,18]. The enhanced activities of insects’ detoxifying enzymes are ones involving vital mechanisms through which insects develop pesticide resistance. Enhanced detoxification accelerates the detoxification metabolisms of insecticides in insects [19]. P450s are important metabolic systems involved in the detoxification of xenobiotics (pesticides, plant toxins), and almost all types are present in all insect tissues. P450s represent a type of single-chain protein that binds to heme, and it is named after its maximum absorption peak, occurring at a wavelength of 450 nm when combined with reduced CO [20,21,22,23]. P450s make up a gene superfamily composed of multiple gene families mainly divided into four major clades as follows: CYP2, CYP3, CYP4 and mitochondrial P450s. The CYP6 family belongs to the CYP3 clan [22], and many detoxification issues related to P450 genes have been reported to originate from this family, such as Bemisia tabaci CYP6CM1 [24], CYP6CM1vQ [25], Nilaparvata lugens CYP6AY1, CYP6ER1 [26] and Mamestra brassicae CYP6AB56 [27].
RNA interference technology selectively inhibits the expression of specific genes in target insects; therefore, it has good application prospects in pest control. Recently, under laboratory conditions, the genetic control of various pests has been achieved by interfering with the expression of key genes in pests [28,29]. However, the effectiveness of RNAi highly depends on the efficiency of nucleic acid molecule delivery to the target. Traditionally, dsRNA has been delivered into organisms that use electroporation, gene gun or microinjection, but these methods may injure model organisms. They are difficult to enact on small insects, and these methods cannot be applied to farmland. The feeding method used to introduce dsRNA is simple to operate, minimizes trauma and has a shorter consumption compared to that of the injection method. However, the efficiency of dsRNA molecules in food used to penetrate the insect intestinal membrane barrier to alter target genes is low. Traditional injection and feeding methods struggle to achieve efficient RNA interference under the conditions of species diversity and genetic diversity. Recently, carrying RNA into the insect body through nanocarriers has provided new ideas for RNA interference experiments. He et al. used acylamide nanocarriers to carry dsRNA into the live Asian corn borer, interfering with the normal expression of chitinase genes and thus delaying development, ceasing molting and even causing the death of the pests [30]. This finding is highly encouraging for the development of RNAi-sourced pesticides.
In this study, 12 CYP6 genes potentially related to imidacloprid resistance were identified through transcriptome sequencing. The mRNA relative expression levels of CYP6s after imidacloprid treatment were used to determine whether they were related to imidacloprid induction. The major CYP6 sequences related to imidacloprid induction were further studied. siRNA was transferred into insects’ bodies to interfere with the expression of the corresponding CYP6 genes in A. glycines. The interference efficiency and p450 enzyme activity were detected. Finally, the function of CYP6 in affecting the resistance of A. glycines to imidacloprid was verified through a bioassay.

2. Materials and Methods

2.1. Insect Culture and Insecticide Used

A. glycines adults were collected from the Xiangyang Experimental Station of Northeast Agricultural University and cultured in a climate chamber using a potted soybean seedling as the host for many generations without contact with an insecticide. The temperature in the climate chamber was 25 ± 2 °C, the relative humidity was 70% and the photoperiod was 14L:10D. The insecticide used was that of an imidacloprid (Bayer, Leverkusen, Germany) water-dispersible granule, and the content of the effective ingredient was 70%.

2.2. CYP6 Sequence Cloning and Analysis

The first to fourth instar nymphs and adult aphids were collected and sequenced to identify the transcriptome (Annoroad, Beijing, China). Twelve A. glycines CYP6 sequences were screened from the transcriptome database. Twelve pairs of sequence primers (Table S1) derived from CYP6 sequences were designed to determine the correctness of the sequences obtained from the database used for PCR amplification. The corrected CYP6 sequences were submitted to the P450 International Nomenclature Committee for Names and registered in GenBank.

2.3. CYP6 Expression Pattern Due to Imidacloprid Treatment Induction

A. glycines adults were treated with LC50 imidacloprid (37.4 mg/L) for 3, 6, 12, 24 and 48 h. The bioassay method used was defined as follows: We cut fresh soybean leaves to a size of 2 cm × 2 cm and immersed them in the prepared imidacloprid for 10 s. The residual liquid was then removed using filter paper and streaked onto 1% agar medium. We then gently transferred the adults to the leaves, and the plate was inverted and cultured. Each treatment consisted of 3 replicates. Twenty aphids that survived each treatment were collected in one EP tube and snap-frozen in liquid nitrogen and immediately stored at −80 °C until further use. Each treatment was repeated three times.
Total RNA was extracted from A. glycines treated with different processing times using Trizol reagent® (Invitrogen, Carlsbad, CA, USA) based on the manufacturer’s instruction. RNA was reverse-transcribed into cDNA using the ReverTra Ace qPCR RT Kit (Toyobo, Shanghai, China).
A real-time polymerase chain reaction (qRT-PCR) was performed to determine the mRNA levels of CYP6s using THUNDERBIRD SYBR qPCRMix (Toyobo, Shanghai, China) through the CFX Connect Real-Time PCR Detect System (Bio-Rad, Hercules, CA, USA). The cycle conditions were defined as follows: 94 °C for 3 min, followed by 40 cycles of 94 °C for 30 s, 59 °C for 30 s and 72 °C for 30 s. The relative expression levels were calculated through the double-internal-parameter calculation method [31], in which the geometric mean of the CT values of two internal reference genes was combined with the 2−ΔΔCT method [32]. All primer sequences for qRT-PCR are listed in Table S1. According to Raman Banasl [33], the A. glycines samples’ TBP (TATA-box binding protein, JQ654781) and RPS9 (ribosomal protein S9, JQ654782) were used as reference genes.

2.4. Study of the Function of A. glycines CYP6s during Imidacloprid Treatment

Based on the above results, CYP6CY7 was selected for the detoxification of imidacloprid using RNA interference (RNAi). The siRNAs of the CYP6CY7 and negative control used in the experiments were designed and synthesized by the GenePharma Company (GenePharma, Shanghai, China), and they are listed in Table S1. The negative control was a common siRNA that had no RNAi effects on A. glycines. siRNAs were dissolved using nucleic acid-free DEPC water, and the final dilution concentration was 20 μM.
In order to convert siRNAs into A. glycines, we introduced a transdermal delivery system to carry nucleic acids through nanocarriers and detergents to attach them to the nota of aphids, which smoothly spread around the aphid integument and, finally, penetrated the integument [34]. The SPc was selected as the nanocarrier. SPc and siRNA were mixed at a volume ratio of 1:1 and incubated for 30 min at room temperature. Next, 10% volume of detergent (Liby, Beijing, China) was added to the formulation. Then, 0.1 μL of siRNA/nanocarrier/detergent was applied to the aphid notum for 1, 2, 3, 6, 12, 24 and 48 h. Each treatment was repeated three times. The RNAs of 30 aphids were extracted from each group. The relative expression of mRNA after RNAi was detected through qRT-PCR, and the experiment protocol was the same as that in Section 2.3.
P450 enzyme activities after RNAi were examined using the double-antibody sandwich method based on the insect cytochrome P450 enzyme-linked immunosorbent assay kit (Jiangsu Meibiao Biotechnology Co., Ltd., Yancheng, China). Thirty adult aphids receiving each treatment obtained after RNAi were homogenized on ice in 800 μL of 0.01 M PBS (Beijing Biotopped Technology, Beijing, China) buffer. The homogenate was centrifuged at 4 °C for 10 min at 12,000× g. The supernatant was then used to examine P450 activities at a wavelength of 450 nm based on the manufacturer’s instructions. The total protein concentration was determined at 595 nm using a Bradford assay from the TaKaRa Bradford Protein Assay Kit (TaKaRa, Beijing, China). The final unit of the cytochrome P450 enzyme measured was pmol/g.
The sensitivity of A. glycines after being subjected to RNAi to imidacloprid was tested using the leaf dipping method, as described above. The adult aphids after 1 h of RNAi were further treated with imidacloprid-impregnated leaves at a concentration of LC50 for 12 h and 24 h, and the mortality was assessed. Twenty aphids were included in each group. Each treatment was repeated three times.

2.5. Statistical Analysis

Statistical analyses were performed using SPSS 18.0 software (IBM Corp., Armonk, NY, USA). Differences between groups of means were examined through one-way analysis of variance, followed by Duncan’s multiple range test. A p value < 0.05 was considered to be significant.

3. Results

3.1. Identification of Twelve CYP6 Sequences

Twelve P450 CYP6 sequences of A. glycines were identified, nominated and submitted to GenBank. Their molecular characteristics are listed in Table 1.
The amino acid identity is higher than 40% in the same CYP family [35]. In this experiment, CYP6 amino acids had a sequence identity of 58.97%, indicating that they belonged to the same family. As Figure 1 shows, CYP6CY7 was closer to CYP6CY14 (59.27% identity), followed by CYP6CY16 (55.62% identity) and CYP6CY12 (51.36% identity). CYP6CY48 was closer to CYP6CY20 (65.82% identity), followed by CYP6CY18 (61.91% identity) and CYP6CY8 (56.34% identity).
In general, the insects’ P450 amino acid sequences contained five conserved motifs, namely those of the heme-binding motif (F/YxxGxRxCxG/A), helix-C motif (WxxxR), helix-I motif (AGxxT), helix K motif (ExxR) and “Meander” motif (PxxFxPxxF) [35,36]. As shown in Figure 1, 11 of the 12 CYP6 amino acid sequences in the A. glycines contained all the five conserved motifs, with the exception being CYP6CY12, while CYP6CY12 contained four conserved motifs, with the exception being the helix-I motif (AGxxT).

3.2. Differential Expression of CYP6s in Response to Imidacloprid Treatment Induction

To identify the specific CYP6s that contribute to imidacloprid induction in A. glycines, the relative expression levels of 12 CYP6s from these aphids were evaluated using qRT-PCR. The results showed that the expressions of 12 CYP6s significantly increased for at least one time point within 48 h after the induction of imidacloprid at an LC50 concentration. The genes CYP6CY7, CYP6CY8, CYP6CY9, CYP6CY14 and CYP6CZ1 responded quickly at 3 h, while the genes CYP6CY12, CYP6CY16, CYP6CY18, CYP6CY20, CYP6CY48, CYP6DB1 and CYP6DD1 exhibited a relatively slower response. Among them, the relative expression levels of CYP6CY7 have the highest upregulation, which is more than 3-fold compared to those observed for the non-induction of treatment. The relative expression level of CYP6CY7 was highest at 12 h after induction, which was 3.55-fold higher than that of non-induction, followed by CYP6CY48, which was 3.36-fold higher than that of non-induction at 48 h after induction (Figure 2). Thus, CYP6CY7 might be the main gene that provides resistance to imidacloprid stress.

3.3. Relative Expressions of CYP6CY7 after RNAi

The inhibition efficiencies of CYP6CY7 were determined using qRT-PCR after RNAi. The results demonstrated that the relative expression levels of CYP6CY7 decreased by 58.49%, 73.80% and 69.30% at 3 h, 6 h and 12 h after treatment (Figure 3), which indicated that mRNA expression level of CYP6CY7 was inhibited at different time points after treatment.

3.4. P450 Enzyme Activities after RNA Inhibition

The enzyme activities of P450 were examined, as shown in Figure 4; after the inhibition of CYP6CY7, the enzyme activity of P450 was significantly decreased at 6 h, and it was decreased by 19.38% and 21.96% at 12 h and 24 h. The RNA inferences took effect after 6 h, and the effects were relieved after 48 h.

3.5. Sensitivity of A. glycines to Imidacloprid Treatment after RNA Inhibition

We measured the sensitivity of A. glycines to imidacloprid treatment after CYP6CY7 inhibition through leaf dipping methods. When A. glycines were exposed to imidacloprid-containing leaves for 12 h, the mortality of A. glycines inhibited with siRNA of CYP6CY7 increased by 13.13%. After exposure to imidacloprid-containing leaves for 24 h, the proportion of aphids inhibited with siRNA of CYP6CY7 also significantly increased (Figure 5).

4. Discussion

Regarding insect detoxification, metabolic systems are the key mechanisms for developing insecticide resistance [37]. Insect cytochrome P450 enzymes are present in organisms and are widely implicated in the detoxification of xenobiotics. The soybean aphid is a global agricultural pest that causes huge economic losses, and its increasing resistance to pesticides in recent years has made the effective control of this pest a problem that urgently needs to be solved. In this study, 12 cDNA sequences from the CYP6 family of the soybean aphid were obtained from the transcriptome database, named by the International P450 Nomenclature Committee and submitted to the NCBI database. The 12 amino acid sequences contain conserved regions common to insect P450s, thus being consistent with the characteristic sequences of this family. Their sequence identity was 58.97%, indicating that the subfamily sequence in the P450 family was relatively conserved.
The cytochrome P450 is inducible after exposing insects to a certain degree of heterologous stress, and a quick and significant increase in the levels of cytochrome P450 genes results in enhanced detoxification [38].
Multiple families and subfamilies of insect P450 genes exist, of which the CYP3 and CYP4 families are considered to be most closely related ones to drug resistance [39,40]. Indeed, the CYP4G8 gene of Helicoverpa armigera was overexpressed twice as much in the pyrethroid insecticide-resistant strain [41], and five P450 genes of the CYP4 family that may be associated with deltamethrin resistance were also previously identified in Culex pipiens [42]. As research in this field progressed, the functions of the CYP6 genes were discovered, with the genes CYP6CY7 and CYP6CY21 being related to α-cypermethrin cross-resistance in Aphis gossypii Glover [43], while the knockdown of one CYP4 gene and five CYP6 genes dramatically increased the sensitivity of A. gossypii to thiamethoxam or imidacloprid treatment [44]. The CYP6 family may be induced at different concentrations and different pesticide treatment times; in this study, the 12 CYP6 genes exhibited different upregulated expression patterns after imidacloprid treatment at different processing times.
Upon comparing multiple experiments at home and abroad, the relative expression levels of CYP6AB56 and CYP4M51 in Mamestra brassicae induced with deltamethrin at LD50 for 3 h, 6 h, 12 h, 24 h and 48 h showed a trend of an initial increase (reaching a maximum at 12 h) before gradually decreasing [27]. In imidacloprid-resistant strains of B and Q biotypes of Bemisia tabaci, the expression levels of CYP6CM1 showed a certain temporal effect [24], and this phenomenon was also observed in the CYP6AY1 and CYP6ER1 from the brown planthopper [26,45]. With the overexpression of CYP6 genes, induction caused activity enhancement in the detoxification system, increasing metabolic defenses.
RNAi technology is a universal method employed to identify resistance genes, and Bautista reported that after specific interference with the CYP6BG1 gene in the diamondback moth, the P450 enzyme activity decreased by 25.5%, thus reducing the larvae’s resistance to cypermethrin [46]. RNAi knockdown of CYP6BG1 in diamondback moth larvae resulted in a 45.5% decrease in P450 enzyme activity, thus increasing sensitivity to chlorantraniliprole by 26.4% [47]. The results showed that using RNAi with CYP6CY7 significantly decreased P450 activity by up to 21.96%. A further bioassay showed that the sensitivity of soybean aphids to imidacloprid significantly increased after the interference. This result indicated that the metabolic activity of the proteins encoded by the CYP6CY7 gene plays an important role in P450 metabolic activity, while the metabolic activity plays an important role in this insect’s insecticide detoxification mechanism.
After the RNAi of CYP6CY7, the interference effect on the soybean aphid did not appear until 3 h, and it remained at 12 h. The activity of cytochrome P450 significantly decreased after 12 h and lasted until 24 h, and the sensitivity of aphids to imidacloprid increased significantly after 24 h. The efficiency of RNAi depends on various factors, such as the mechanisms used to inhibit or interfere with RNAi in cells as well as degradation by nucleases or immune responses used to recognize and eliminate foreign dsRNA or siRNA molecules, and some cells may also have low expression or activity of RNA-induced silencing complex (RISC) components or other factors crucial for RNAi [48]. Due to differences in the onset times of RNAi, the changes in the corresponding enzyme also exhibit time specificity. At the same time, the sensitivity of soybean aphids to imidacloprid treatment caused by the detoxification function and mediated by different genes also exhibits time specificity. However, further research is needed to understand how the expression of genes mediates the time-specific changes in enzymes used for field application guidance.
RNAi technology can selectively interfere with target genes in target insects, effectively alleviating long-term concerns about pesticide residues and food safety; therefore, it shows high potential in pest control. The optimal effect of RNAi-mediated gene silencing is difficult to achieve, especially in aphids. Due to the fact that the abdominal cavity is filled with bodily fluids, the direct injection of double-stranded RNA is not only complicated but can also lead to the death of experimental insects due to mechanical injury, post-injection infection and other issues, thereby resulting in inaccurate test results.
It is difficult to eliminate malnutrition-related errors by feeding soybean aphids artificial diets for dsRNA interference. Additionally, the efficiency of the dsRNA molecules penetrating the membrane barrier of insect intestine to reach the target genes is reduced. Directly soaking soybean aphids in a dsRNA aqueous solution can reduce mechanical damage and achieve effective interference, but soybean aphids often escape from the soaking solution, and the soaking time required to achieve effective delivery cannot be guaranteed [34]. In addition, complete immersion in a high amount of dsRNA solution causes significant economic loss. Thus, we chose to use the body wall permeation method by introducing the nano-carrier into the insect body to not only ensure the survival rate of the experimental soybean aphid but also to create the potential for practical production.
dsRNA degrades its mRNA homologous, a process known as RNA interference. dsRNA was cleaved into short interfering RNAs (siRNAs) and hybridized with homologous mRNA, which induced its degradation. In general, both dsRNA and its metabolite, siRNA, can induce RNAi effects in different species. Compared with dsRNAs, siRNAs are easy to synthesize and to mass-produce at a low cost. siRNAs can also be designed to prevent their degradation and preserve stability [49,50]. siRNAs may also be developed as a specific nucleic acid insecticide [51]. Researchers have previously reported many successful RNAi experiments that used siRNA [52,53]. This experiment introduced siRNAs into soybean aphids and suppressed the relative expression level of the CYP6CY7, as well as the enzyme activities of P450s, thereby improving the sensitivity of soybean aphids to imidacloprid. Therefore, it provides a theoretical basis for RNAi experiments and the development of new insecticides.

5. Conclusions

In this study, we identified and characterized 12 P450 CYP6 genes from A. glycines. The overexpression of these genes after the induction of imidacloprid treatment at an LC50 concentration provided new information about the role of CYP6s in determining resistance to insecticides in A. glycines. Through siRNA interference by nanocarriers, we interfered with the expression of CYP6CY7, highly increasing the susceptibility of aphids to imidacloprid; the P450 enzyme activity was inhibited by 21.96%, the insecticidal efficacy of imidacloprid treatment on A. glycines increased and the mortality of the interfered A. glycines in response to imidacloprid treatment increased by 14.71%. Compared to dsRNA, siRNAs are easy to synthesize and to mass-produce and are decreasing in cost. Small-molecule siRNAs can also be designed to prevent degradation and improve stability, and they also have the potential for development as a specific nucleic acid insecticide. Therefore, this study provides a theoretical basis for further research into the potential metabolic detoxification mechanisms of CYP6s and the development of new synergists to increase insecticidal efficacy.

Supplementary Materials

The following supporting information can be downloaded via this link: https://www.mdpi.com/article/10.3390/insects15030188/s1; Table S1: Primers for cloning, qRT-PCR and RNA interference.

Author Contributions

Conceptualization, D.F.; methodology, X.Z. and D.F.; investigation, S.L., H.Y., X.Z., Z.S. and Y.W.; data curation, S.L., L.W., X.Z. and D.F.; writing—original draft preparation, S.L., J.L. and X.Z.; writing—review and editing, X.Z., H.Y., Y.W. and D.F.; visualization—X.Z. and Y.W. supervision, X.Z. and D.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Key Research and Development Program of China (2023YFD1401000).

Data Availability Statement

The data presented in this study are available upon reasonable request from the corresponding author.

Acknowledgments

We are grateful to Shen Jie from China Agricultural University for providing the nanocarrier of SPc as well as for useful advice to support our study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tilmon, K.J.; Hodgson, E.W.; O’Neal, M.E.; Ragsdale, D.W. Biology of the soybean aphid, Aphis glycines (Hemiptera: Aphididae) in the United States. J. Integr. Agric. 2011, 2, A1–A7. [Google Scholar] [CrossRef]
  2. Ragsdale, D.W.; Landis, D.A.; Brodeur, J.; Heimpel, G.E.; Desneux, N. Ecology and management of the soybean aphid in North America. Annu. Rev. Entomol. 2011, 56, 375–399. [Google Scholar] [CrossRef] [PubMed]
  3. Ragsdale, D.W.; McCornack, B.P.; Venette, R.C.; Potter, B.D.; MacRae, I.V.; Hodgson, E.W.; O’Neal, M.E.; Johnson, K.D.; O’Neil, R.J.; DiFonzo, C.D.; et al. Economic threshold for soybean aphid (Hemiptera: Aphididae). J. Econ. Entomol. 2007, 100, 1258–1267. [Google Scholar] [CrossRef] [PubMed]
  4. Hunt, D.; Foottit, R.; Gagnier, D.; Baute, T. First Canadian records of Aphis glycines (Hemiptera: Aphididae). Can. Entomol. 2003, 135, 879–881. [Google Scholar] [CrossRef]
  5. Hill, J.H.; Alleman, R.; Hogg, D.B.; Grau, C.R. First Report of Transmission of Soybean mosaic virus and Alfalfa mosaic virus by Aphis glycines in the New World. Plant Dis. 2001, 85, 561. [Google Scholar] [CrossRef] [PubMed]
  6. Clark, A.J.; Perry, K.L. Transmissibility of Field Isolates of Soybean Viruses by Aphis glycines. Plant Dis. 2002, 86, 1219–1222. [Google Scholar] [CrossRef] [PubMed]
  7. Venette, R.C.; Ragsdale, D.W. Assessing the Invasion by Soybean Aphid (Homoptera: Aphididae): Where Will It End? Ann. Entomol. Soc. Am. 2004, 97, 219–226. [Google Scholar] [CrossRef]
  8. Liu, J.; Zhao, K.J. Biology and control techniques of soybean aphid, Aphis glycines. Chin. Bull. Entomol. 2007, 44, 179–185. [Google Scholar]
  9. Nauen, R.; Denholm, I. Resistance of insect pests to neonicotinoid insecticides: Current status and future prospects. Arch. Insect Biochem. Physiol. 2005, 58, 200–215. [Google Scholar] [CrossRef]
  10. Tang, J.J.; Chen, X.; Zhang, C.J.; Wang, X. Advances of Imidacloprid Research and Its Applicationin in Crop Pest Control. Chin. Agr. Sci. Bull. 1999, 15, 38–40. [Google Scholar]
  11. Le Questel, J.Y.; Graton, J.; Cerón-Carrasco, J.P.; Jacquemin, D.; Planchat, A.; Thany, S.H. New insights on the molecular features and electrophysiological properties of dinotefuran, imidacloprid and acetamiprid neonicotinoid insecticides. Bioorg. Med. Chem. 2011, 19, 7623–7634. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, S.H.; Tang, M.; Wang, J.J.; Du, B.; Lu, D.M.; Wang, Z.Y. Research Summary in Neonicotinoids Insecticides. Shandong Chem. Ind. 2015, 44, 66–68. [Google Scholar]
  13. Zhang, A.N.; Han, L.L.; Zhao, K.J.; Zhang, W.L.; Xiao, J.F.; Chen, J.; Gao, L.T. Effects of semilethal and sublethal doses of imidacloprid on Aphis glycines (Hemiptera: Aphididae). Chin. J. Appl. Entomol. 2020, 57, 676–681. [Google Scholar]
  14. Xu, L.; Wang, J.H.; Mei, Y.; Li, D.Z. Research progress on the molecular mechanisms of insecticides resistance mediated by detoxification enzymes and transporters. Chin. J. Pestic. Sci. 2020, 22, 1–10. [Google Scholar]
  15. Chen, M.H.; Han, Z.J.; Qiao, X.F.; Qu, M.J. Mutations in acetylcholinesterase genes of Rhopalosiphum padi resistant to organophosphate and carbamate insecticides. Genome 2007, 50, 172–179. [Google Scholar] [CrossRef]
  16. Lu, Y.; Gao, X. Multiple mechanisms responsible for differential susceptibilities of Sitobion avenae (Fabricius) and Rhopalosiphum padi (Linnaeus) to pirimicarb. Bull. Entomol. Res. 2009, 99, 611–617. [Google Scholar] [CrossRef]
  17. Bass, C.; Field, L.M. Gene amplification and insecticide resistance. Pest. Manag. Sci. 2011, 67, 886–890. [Google Scholar] [CrossRef]
  18. Kang, X.L.; LI, Y.T.; Wang, K.; Zhang, M.; Duan, X.L.; Peng, X.; Chen, M.H. Molecular cloning, characterization and expression analysis of ATPbinding cassette transporter genes in the bird cherry-oat aphid, Rhopalosiphum padi(Hemiptera: Aphididae). Acta Entomol. Sin. 2015, 58, 593–602. [Google Scholar]
  19. Yu, S.J. Detoxification Mechanisms in Insects. In Encyclopedia of Entomology; Capinera, J.L., Ed.; Springer: Dordrecht, The Netherlands, 2008. [Google Scholar]
  20. Omura, T.; Ryo Sato, R. The Carbon Monoxide-binding Pigment of Liver Microsomes: I. Evidence for its hemoprotein nature. J. Biol. Chem. 1964, 239, 2370–2378. [Google Scholar] [CrossRef]
  21. Scott, J.G.; Liu, N.; Wen, Z. Insect cytochromes P450: Diversity, insecticide resistance and tolerance to plant toxins. Comp. Biochem. Physiol. C Pharmacol. Toxicol. Endocrinol. 1998, 121, 147–155. [Google Scholar] [CrossRef]
  22. Feyereisen, R. Insect P450 enzymes. Annu. Rev. Entomol. 1999, 44, 507–533. [Google Scholar] [CrossRef]
  23. Nauen, R.; Bass, C.; Feyereisen, R.; Vontas, J. The Role of Cytochrome P450s in Insect Toxicology and Resistance. Annu. Rev. Entomol. 2022, 67, 105–124. [Google Scholar] [CrossRef]
  24. 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, 634–644. [Google Scholar] [CrossRef]
  25. Karunker, I.; Morou, E.; Nikou, D.; Nauen, R.; Sertchook, R.; Stevenson, B.J.; Paine, M.J.; Morin, S.; Vontas, J. Structural model and functional characterization of the Bemisia tabaci CYP6CM1vQ, a cytochrome P450 associated with high levels of imidacloprid resistance. Insect Biochem. Mol. Biol. 2009, 39, 697–706. [Google Scholar] [CrossRef] [PubMed]
  26. Bao, H.; Gao, H.; Zhang, Y.; Fan, D.; Fang, J.; Liu, Z. The roles of CYP6AY1 and CYP6ER1 in imidacloprid resistance in the brown planthopper: Expression levels and detoxification efficiency. Pestic. Biochem. Physiol. 2016, 129, 70–74. [Google Scholar] [CrossRef] [PubMed]
  27. Zhou, X.; Fan, X.; Gao, Y.; Yang, J.; Qian, J.; Fan, D. Identification of two novel P450 genes and their responses to deltamethrin in the cabbage moth, Mamestra brassicae Linnaeus. Pestic. Biochem. Physiol. 2017, 141, 76–83. [Google Scholar] [CrossRef] [PubMed]
  28. Baum, J.A.; Bogaert, T.; Clinton, W.; Heck, G.R.; Feldmann, P.; Ilagan, O.; Johnson, S.; Plaetinck, G.; Munyikwa, T.; Pleau, M.; et al. Control of coleopteran insect pests through RNA interference. Nat. Biotechnol. 2007, 25, 1322–1326. [Google Scholar] [CrossRef] [PubMed]
  29. Liu, J.S.; Zhu, W.H.; Liao, W.L.; BU, X.L.; Jiang, W.Y. Ds rna delivery methods for rna interference in insects. Acta Entomol. Sin. 2016, 59, 685–691. [Google Scholar]
  30. He, B.; Chu, Y.; Yin, M.; Müllen, K.; An, C.; Shen, J. Fluorescent nanoparticle delivered dsRNA toward genetic control of insect pests. Adv. Mater. 2013, 25, 4580–4584. [Google Scholar] [CrossRef] [PubMed]
  31. 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. [Google Scholar] [CrossRef]
  32. 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]
  33. Bansal, R.; Mamidala, P.; Mian, M.A.; Mittapalli, O.; Michel, A.P. Validation of reference genes for gene expression studies in Aphis glycines (Hemiptera: Aphididae). J. Econ. Entomol. 2012, 105, 1432–1438. [Google Scholar] [CrossRef]
  34. Zheng, Y.; Hu, Y.; Yan, S.; Zhou, H.; Song, D.; Yin, M.; Shen, J. A polymer/detergent formulation improves dsRNA penetration through the body wall and RNAi-induced mortality in the soybean aphid Aphis glycines. Pest. Manag. Sci. 2019, 75, 1993–1999. [Google Scholar] [CrossRef] [PubMed]
  35. Feyereisen, R. Insect CYP genes and P450 enzymes. In Insect Molecular Biology and Biochemistry; E-Publishing Inc.: London, UK, 2012; pp. 236–316. [Google Scholar]
  36. Feyereisen, R. Evolution of insect P450. Biochem. Soc. Trans. 2006, 34, 1252–1255. [Google Scholar] [CrossRef] [PubMed]
  37. Ebadollahi, A.; Naseri, B.; Abedi, Z.; Setzer, W.N.; Changbunjong, T. Promising Insecticidal Efficiency of Essential Oils Isolated from Four Cultivated Eucalyptus Species in Iran against the Lesser Grain Borer, Rhyzopertha dominica (F.). Insects 2022, 13, 517. [Google Scholar] [CrossRef]
  38. Khan, S.; Uddin, M.N.; Rizwan, M.; Khan, W.; Farooq, M.; Shah, A.S.; Muhammad, M. Mechanism of insecticide resistance in insects/pests. Pol. J. Environ. Stud. 2020, 29, 2023–2030. [Google Scholar] [CrossRef]
  39. Strode, C.; Wondji, C.S.; David, J.P.; Hawkes, N.J.; Lumjuan, N.; Nelson, D.R.; Drane, D.R.; Karunaratne, S.H.; Hemingway, J.; Black, W.C., 4th; et al. Genomic analysis of detoxification genes in the mosquito Aedes aegypti. Insect Biochem. Mol. Biol. 2008, 38, 113–123. [Google Scholar] [CrossRef]
  40. Zhou, D.; Liu, X.; Sun, Y.; Ma, L.; Shen, B.; Zhu, C. Genomic Analysis of Detoxification Supergene Families in the Mosquito Anopheles sinensis. PLoS ONE 2015, 10, e0143387. [Google Scholar] [CrossRef]
  41. Pittendrigh, B.; Aronstein, K.; Zinkovsky, E.; Andreev, O.; Campbell, B.; Daly, J.; Trowell, S.; Ffrench-Constant, R.H. Cytochrome P450 genes from Helicoverpa armigera: Expression in a pyrethroid-susceptible and -resistant strain. Insect Biochem. Mol. Biol. 1997, 27, 507–512. [Google Scholar] [CrossRef] [PubMed]
  42. Shen, B.; Dong, H.Q.; Tian, H.S.; Ma, L.; Li, X.L.; Wu, G.L.; Zhu, C.L. Cytochrome P450 genes expressed in the deltamethrin-susceptible and resistant strains of Culex pipiens pallens. Pestic. Biochem. Physiol. 2003, 75, 19–26. [Google Scholar] [CrossRef]
  43. Zeng, X.; Pan, Y.; Song, J.; Li, J.; Lv, Y.; Gao, X.; Tian, F.; Peng, T.; Xu, H.; Shang, Q. Resistance Risk Assessment of the Ryanoid Anthranilic Diamide Insecticide Cyantraniliprole in Aphis gossypii Glover. J. Agric. Food Chem. 2021, 69, 5849–5857. [Google Scholar] [CrossRef]
  44. Lv, Y.; Wen, S.; Ding, Y.; Gao, X.; Chen, X.; Yan, K.; Yang, F.; Pan, Y.; Shang, Q. Functional Validation of the Roles of Cytochrome P450s in Tolerance to Thiamethoxam and Imidacloprid in a Field Population of Aphis gossypii. J. Agric. Food Chem. 2022, 16, 14339–14351. [Google Scholar] [CrossRef] [PubMed]
  45. Jin, R.H.; Mao, K.K.; Liao, X.; Xu, P.F.; Li, Z.; Ali, E.; Wan, H. Overexpression of CYP6ER1 associated with clothianidin resistance in Nilaparvata lugens (Stål). Pestic. Biochem. Physiol. 2019, 154, 39–45. [Google Scholar] [CrossRef] [PubMed]
  46. Bautista, M.A.; Miyata, T.; Miura, K.; Tanaka, T. RNA interference-mediated knockdown of a cytochrome P450, CYP6BG1, from the diamondback moth, Plutella xylostella, reduces larval resistance to permethrin. Insect Biochem. Mol. Biol. 2009, 39, 38–46. [Google Scholar] [CrossRef] [PubMed]
  47. Li, X.; Li, R.; Zhu, B.; Gao, X.; Liang, P. Overexpression of cytochrome P450 CYP6BG1 may contribute to chlorantraniliprole resistance in Plutella xylostella (L.). Pest. Manag. Sci. 2018, 74, 1386–1393. [Google Scholar] [CrossRef] [PubMed]
  48. Agrawal, N.; Dasaradhi, P.V.; Mohmmed, A.; Malhotra, P.; Bhatnagar, R.K.; Mukherjee, S.K. RNA interference: Biology, mechanism, and applications. Microbiol. Mol. Biol. Rev. 2003, 67, 657–685. [Google Scholar] [CrossRef]
  49. Yamaguchi, J.; Mizoguchi, T.; Fujiwara, H. siRNAs induce efficient RNAi response in Bombyx mori embryos. PLoS ONE 2011, 6, e25469. [Google Scholar] [CrossRef]
  50. Lares, M.R.; Rossi, J.J.; Ouellet, D.L. RNAi and small interfering RNAs in human disease therapeutic applications. Trends Biotechnol. 2010, 28, 570–579. [Google Scholar] [CrossRef]
  51. Gong, L.; Yang, X.; Zhang, B.; Zhong, G.; Hu, M. Silencing of Rieske iron-sulfur protein using chemically synthesised siRNA as a potential biopesticide against Plutella xylostella. Pest. Manag. Sci. 2011, 67, 514–520. [Google Scholar] [CrossRef] [PubMed]
  52. Kumar, M.; Gupta, G.P.; Rajam, M.V. Silencing of acetylcholinesterase gene of Helicoverpa armigera by siRNA affects larval growth and its life cycle. J. Insect Physiol. 2009, 55, 273–278. [Google Scholar] [CrossRef] [PubMed]
  53. Clemons, A.; Haugen, M.; Le, C.; Mori, A.; Tomchaney, M.; Severson, D.W.; Duman-Scheel, M. siRNA-mediated gene targeting in Aedes aegypti embryos reveals that frazzled regulates vector mosquito CNS development. PLoS ONE 2011, 6, e16730. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The alignment of the amino acid sequences of 12 CYP6 genes from A. glycines. The conserved motifs are shown in red squares. The black area indicates an identity of 100%, the pink area indicates an identity of 75% or more and the blue area indicates an identity of 50% or more. * indicates the position of the number next to it ±10.
Figure 1. The alignment of the amino acid sequences of 12 CYP6 genes from A. glycines. The conserved motifs are shown in red squares. The black area indicates an identity of 100%, the pink area indicates an identity of 75% or more and the blue area indicates an identity of 50% or more. * indicates the position of the number next to it ±10.
Insects 15 00188 g001
Figure 2. Effects of imidacloprid induction on the expression of CYP6s by A. glycines. The relative expression levels of CYPs were examined after imidacloprid induction at 3, 6, 12, 24 and 48 h. TBP and RPS9 were used as the internal reference genes. Expression levels of CYPs were analyzed using the 2−ΔΔCT method and by calculating the means ± SEs. Statistical differences at p < 0.05 are indicated by *, p < 0.01 are indicated by ** and the ns represents no significant differences.
Figure 2. Effects of imidacloprid induction on the expression of CYP6s by A. glycines. The relative expression levels of CYPs were examined after imidacloprid induction at 3, 6, 12, 24 and 48 h. TBP and RPS9 were used as the internal reference genes. Expression levels of CYPs were analyzed using the 2−ΔΔCT method and by calculating the means ± SEs. Statistical differences at p < 0.05 are indicated by *, p < 0.01 are indicated by ** and the ns represents no significant differences.
Insects 15 00188 g002
Figure 3. Expression levels of CYP6CY7 after the siRNA treatment of A. glycines at different times. The relative expression levels of the gene were compared with the NC after RNA interfered for 1, 2, 3, 6, 12, 24 and 48 h. The negative control was a kind of siRNA that showed no RNAi effects in A. glycines. The results are the mean ± SE of three independent replications. ** represents significant differences between different treatments (p < 0.01) and the ns represents no significant differences.
Figure 3. Expression levels of CYP6CY7 after the siRNA treatment of A. glycines at different times. The relative expression levels of the gene were compared with the NC after RNA interfered for 1, 2, 3, 6, 12, 24 and 48 h. The negative control was a kind of siRNA that showed no RNAi effects in A. glycines. The results are the mean ± SE of three independent replications. ** represents significant differences between different treatments (p < 0.01) and the ns represents no significant differences.
Insects 15 00188 g003
Figure 4. Enzyme activities of P450 in A. glycines after CYP6CY7 inhibition. RNA interferes with CYP6CY7 and NC for 1, 2, 3, 6, 12, 24 and 48 h. The results are the mean ± SE of three independent replications. ** represents significant differences between different treatments (p < 0.01) and the ns represents no significant differences.
Figure 4. Enzyme activities of P450 in A. glycines after CYP6CY7 inhibition. RNA interferes with CYP6CY7 and NC for 1, 2, 3, 6, 12, 24 and 48 h. The results are the mean ± SE of three independent replications. ** represents significant differences between different treatments (p < 0.01) and the ns represents no significant differences.
Insects 15 00188 g004
Figure 5. Mortality of A. glycines treated with imidacloprid LC50 after CYP6CY7 inhibition. The results are the mean ± SE of five independent replications. ** represents significant differences between different treatments (p < 0.01).
Figure 5. Mortality of A. glycines treated with imidacloprid LC50 after CYP6CY7 inhibition. The results are the mean ± SE of five independent replications. ** represents significant differences between different treatments (p < 0.01).
Insects 15 00188 g005
Table 1. The characteristics of CYP6s identified from A. glycines.
Table 1. The characteristics of CYP6s identified from A. glycines.
Gene NameAccession No.Length (bp)Length (aa)pIMw (kDa)
CYP6CY7MN05599717355178.4059.20
CYP6CY8MN0560001913 5148.6059.50
CYP6CY9MN05599920715228.3459.44
CYP6CY12MN05599816865058.6858.21
CYP6CY14MN05600324685137.5658.96
CYP6CY16MN05599517695138.3859.27
CYP6CY18MN05600215565088.7659.38
CYP6CY20MG82985620405099.10 59.20
CYP6CY48MN05599616395118.87 59.48
CYP6CZ1MN05599318375118.86 59.37
CYP6DB1MN05600122875139.11 59.58
CYP6DD1MN05599421085128.27 58.10
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, S.; Yang, H.; Wang, Y.; Wei, L.; Lyu, J.; Shan, Z.; Zhang, X.; Fan, D. RNA Interference Reveals the Impacts of CYP6CY7 on Imidacloprid Resistance in Aphis glycines. Insects 2024, 15, 188. https://doi.org/10.3390/insects15030188

AMA Style

Li S, Yang H, Wang Y, Wei L, Lyu J, Shan Z, Zhang X, Fan D. RNA Interference Reveals the Impacts of CYP6CY7 on Imidacloprid Resistance in Aphis glycines. Insects. 2024; 15(3):188. https://doi.org/10.3390/insects15030188

Chicago/Turabian Style

Li, Shuangyu, Hongjia Yang, Yixiao Wang, Lisi Wei, Jiawei Lyu, Zhimeng Shan, Xinxin Zhang, and Dong Fan. 2024. "RNA Interference Reveals the Impacts of CYP6CY7 on Imidacloprid Resistance in Aphis glycines" Insects 15, no. 3: 188. https://doi.org/10.3390/insects15030188

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