Next Article in Journal
Minding the Gap: Exploring Neuroinflammatory and Microglial Sex Differences in Alzheimer’s Disease
Next Article in Special Issue
Developmental and Nutritional Dynamics of Malpighian Tubule Autofluorescence in the Asian Tiger Mosquito Aedes albopictus
Previous Article in Journal
In Search of More Leaps to Realize the Precision Medicine of Migraine
Previous Article in Special Issue
Untargeted Lipidomics Analysis Unravels the Different Metabolites in the Fat Body of Mated Bumblebee (Bombus terrestris) Queens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification and Functional Analysis of an Epsilon Class Glutathione S-Transferase Gene Associated with α-Pinene Adaptation in Monochamus alternatus

1
Co-Innovation Center for Sustainable Forestry in Southern China, College of Forestry, Nanjing Forestry University, Nanjing 210037, China
2
Soochow College, Soochow University, Suzhou 215006, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(24), 17376; https://doi.org/10.3390/ijms242417376
Submission received: 24 October 2023 / Revised: 29 November 2023 / Accepted: 10 December 2023 / Published: 12 December 2023

Abstract

:
Alpha-pinene is one of the main defensive components in conifers. Monochamus alternatus (Coleoptera: Cerambycidae), a wood borer feeding on Pinaceae plants, relies on its detoxifying enzymes to resist the defensive terpenoids. Here, we assayed the peroxide level and GST activity of M. alternatus larvae treated with different concentrations of α-pinene. Meanwhile, a gst gene (MaGSTe3) was isolated and analyzed. We determined its expression level and verified its function. The results showed that α-pinene treatment led to membrane lipid peroxidation and thus increased the GST activity. Expression of MaGSTe3 was significantly upregulated in guts following exposure to α-pinene, which has a similar pattern with the malonaldehyde level. In vitro expression and disk diffusion assay showed that the MaGSTe3 protein had high antioxidant capacity. However, RNAi treatment of MaGSTe3 did not reduce the hydrogen peroxide and malonaldehyde levels, while GST activity was significantly reduced. These results suggested MaGSTe3 takes part in α-pinene adaptation, but it does not play a great role in the resistance of M. alternatus larvae to α-pinene.

1. Introduction

In the long evolutionary process, conifers have developed efficient methods to resist the invasion of herbivorous insects and pathogens [1]. The defensive systems of conifers include constitutive and induced defenses [2]. Constitutive defenses are comprehensive defenses based on the corkification of bark, resin secretion, and phenolic synthesis, while induced defenses are responses to specific attacks through the synthesis of secondary metabolites [3,4,5]. Meanwhile, insects have also developed several mechanisms to help them break through the defenses. Insects can reduce the toxicity of defensive compounds through metabolism induced by detoxification enzymes or decrease oxidative damage caused by the compounds through antioxidant enzymes [6,7].
Masson pine (Pinus massoniana Lamb) is an important species for afforestation and timber in China. It has a special terpenoid defense system that consists chiefly of α-pinene and β-pinene. Of the two, α-pinene exhibits stronger insecticidal activity than β-pinene [8]. A previous study has shown that α-pinene can disrupt energy metabolism by the uncoupling of oxidative phosphorylation and the inhibition of electron transfer [9]. Also, it can enhance the generation of reactive oxygen species (ROS), leading to oxidative stress in tissues [10]. As a result, it is believed that the defensive mechanism of α-pinene is associated with oxidative damage to insects [11,12].
In the past decades, P. massoniana forests in China have been devastated by the pine wood nematode (PWN, Bursaphelenchus xylophilus), which is the pathogen of pine wilt disease [13]. Usually, the PWN is carried by wood borers and infects host plants through wounds caused by their feeding behavior [14]. In China, its primary vector is Monochamus alternatus, whose preferred host is P. massoniana [15]. Therefore, prevention and control of M. alternatus have attracted increasing attention. Usually, M. altrenatus larvae feed on the xylem of the host, and their feeding behavior can significantly induce α-pinene synthesis in the phloem and xylem of P. massoniana [16]. As result, α-pinene stress is inevitable for them, and they must have methods to overcome the stress. A previous study showed that exposure to α-pinene caused differential expression of cytochrome P450 mono-oxygenases (P450s), uridine diphosphate glycosyltransferases (UGTs), glutathione S-transferases (GSTs), and ATP-binding cassettes (ABCs) in M. altrenatus larvae [17]. As GSTs have an important effect on oxidative stress resistance, they may also be involved in the resistance of M. altrenatus to α-pinene stress [18].
In this study, we measured the level of membrane lipid peroxidation of larvae under α-pinene stress. Meanwhile, we isolated and characterized a gst gene (MaGSTe3), which was the only gst gene significantly upregulated in the transcriptome data of our previous study [17], and studied its role in the α-pinene resistance of M. altrenatus larvae. Our results will help clarify the role of GSTs in the resistance of M. altrenatus to α-pinene and understand the adaptive mechanisms of M. altrenatus to toxic terpenes, thus laying a foundation for the development of its control strategy.

2. Results

2.1. The Effects of α-Pinene Treatment on M. alternatus Larvae

As is shown in Figure 1, the exposure to α-pinene caused hydrogen peroxide and MDA levels to increase in the medium- and high-dose groups after treatment for 24 and 48 h. In particular, the MDA levels rapidly reached 2.68- and 1.61-fold of the control, respectively, after treatment for 24 h and decreased to 1.54- and 1.48-fold of the control, respectively, 48 h after treatment (Figure 1B).
Similarly, the GST activity was also increased in the high- and medium-dose groups, which reached 5.59- and 3.60-fold of the control at 24 h after treatment, respectively, and decreased to 3.29- and 1.78-fold of the control at 48 h, respectively, after treatment (Figure 1C). Interestingly, though there was no significant difference with the control in hydrogen peroxide and MDA content of the low-dose group, the group still had significantly higher GST activities at 24 and 48 h after treatment, which were 2.01- and 1.74-fold of the control, respectively. These results indicated that GSTs were involved in the adaptability of α-pinene stress.

2.2. Physicochemical Properties and Bioinformatics Analysis of MaGSTe3

The length of the MaGSTe3 sequence is 648 bp, which encodes a protein with 215 amino acids with a predicted molecular weight of 23.58 kDa and a theoretical pI of 5.72. Subcellular localization shows that MaGSTe3 is in cytoplasm and has no transmembrane regions (Table A4). The 3D structure of MaGSTe3 shows that it contains 12 α-helices and 6 β-sheets (Figure A2), including a N-terminal domain with βαβαββα topology and a C-terminal domain. The conserved domain prediction shows that MaGSTe3 has multiple glutathione and substrate binding sites (Figure A3).
The results of BLASTx comparison showed MaGSTe3 had high homology with Anoplophora glabripennis (Table A4). The phylogenetic tree constructed with the maximum likelihood method presented the evolutionary relationship of MaGSTe3 (Figure 2). MaGSTe3 was clustered in the epsilon class with a Leptinotarsa decemlineata GST (LdGSTe10) in the same clade.

2.3. Response of MaGSTe3 to α-Pinene Stress

To further analyze the role of MaGSTe3 in α-pinene adaptability, qRT-PCR was performed to estimate the response of MaGSTe3 to α-pinene stress. The results showed that the relative expression of MaGSTe3 was, overall, not changed at 12 h after treatment but remarkably upregulated at 24 h and 48 h after treatment (Figure 3). In particular, MaGSTe3 was most highly upregulated in guts (Figure 3D). The transcriptional levels of low-, medium-, and high-dose groups was 5.76-, 6.51-, and 8.34-fold of control after treatment for 24 h, and 6.17-, 7.21-, and 6.41-fold of control after treatment for 48 h, respectively. Meanwhile, MaGSTe3 showed significant upregulation in the high-dose group at 24 and 48 h after treatment in hemolymph and at all three time points in epidermis (Figure 3A,C).

2.4. Enzymatic Properties of Recombinant MaGSTe3

The MaGSTe3 was amplified and inserted into pET-28a (+) and expressed in E. coli with IPTG induction. SDS-PAGE analysis (Figure A1) showed that protein bands with an approximate molecular weight of 28 kDa appeared in both precipitate and supernatant. The concentration of the MaGSTe3 solution obtained from purification was approximately 1 mg/mL.
The recombinant MaGSTe3 was incubated at a temperature range of 10 to 70 °C for 30 min, and the maximum activity was detected at 25 °C (Figure 4A). Also, the recombinant MaGSTe3 exhibited optimum catalytic activity at pH 7.7 (Figure 4B). The kinetic parameters determination showed that the Vmax and Km of recombinant MaGSTe3 were 2985.78 ± 347.15 μmol·min−1·mg−1 and 1.24 ± 0.21 mM, respectively. These results suggested that MaGSTe3 is catalytically active, displaying glutathione transferase activity.

2.5. Disk Diffusion Assay of MaGSTe3

Antioxidant activity of MaGSTe3 was detected by disk diffusion assay (Figure 5). E. coli cells carrying overexpressed MaGSTe3 were exposed to cumene hydroperoxide (CHP) to induce oxidative stress. After incubated for 24 h, the halo diameter of inhibition zones had no significant difference between treatment group and control when CHP level was 2–10 mM. Nevertheless, when the CHP level was increased to more than 20 mM, the halo diameter of inhibition zones could reduce more than 51% in the treatment group (Figure 5B).

2.6. RNAi of MaGSTe3

RNAi of MaGSTe3 was performed to further verify the function of MaGSTe3. The dsRNA was injected into the intersegment membrane of M. alternatus and caused significant reduction in the GST activity and expression level of MaGSTe3 (Figure 6A,B). However, the hydrogen peroxide and MDA levels did not show any differences (Figure 6C,D). These results indicated that there might be other enzymes involved in the antioxidation and detoxification of α-pinene stress.

3. Discussion

As the major defensive component in P. massoniana [19], α-pinene was believed to have strong biological activities against herbivorous insects [20]. A-pinene was observed to induce the antioxidant system in the Mediterranean flour moth, Ephestia kuehniella Zeller, and can inhibit feeding activity of the red turpentine beetle, Dendroctonus valens [7,21]. These phenomena suggest that oxidative damage may have occurred in the insect gut under the influence of α-pinene. As the main product of membrane peroxidation, the content of MDA can directly reflect the strength of oxidative damage [22]. In our study, the MDA level in the guts of M. alternatus larvae was rapidly increased after α-pinene treatment, and the content seems to be related to the concentration of α-pinene, suggesting that α-pinene might be the direct cause of the disorder in guts. Nevertheless, the MDA levels were decreased at 48 h after treatment and returned to normal at 72 h, which was similar with the trend of GST activity. Many studies have shown that oxidative stress induced by different factors all led to an increase in GST activity in insects [7,23,24]. Moreover, it was reported that GSTs participate in the reduction in lipid peroxides in Drosophila melanogaster [25]. As a result, GSTs may play roles in the adaptability of the α-pinene of M. alternatus by relieving the oxidative stress.
GSTs in insects are divided into six families [26]. It is believed that GSTs from delta and epsilon families are crucial for insecticide resistance and the detoxification of plant xenobiotics [27,28,29,30]. In Coleoptera, epsilon GSTs might play an important role in detoxification. There are at least 12 out of 25 DaGSTs found in the Chinese white pine beetle, Dendroctonus armandi, and 7 out of 14 PtsGSTs in Pagiophloeus tsushimanus, which are reported to respond to the defensive components of host trees that are classified as being in the epsilon family [31,32,33]. In our study, MaGSTe3 was significantly upregulated in guts when responding to α-pinene stress. Many studies have reported the abundant expression of gst genes in insect guts, especially in the midgut [29,34,35]. It is generally believed that the midgut is one of the tissues where insect GSTs mainly function, as it is the largest part of the digestive tract and the main target of plant defensive substances [36]. The role of MaGSTe3 in guts might be mediating oxidative stress responses. We also found that MaGSTe3 in the high-dose treatment group was consistently upregulated in the epidermis. Several studies have also reported that volatile pesticide substances could lead to overexpression of gst in the epidermis [37,38]. However, the function of GSTs in other tissues is still unknown. Considering that plant terpenoids like α-pinene are volatiles that can be ingested by contact with the cuticle or fumigation, MaGSTe3 may play a part in cuticular resistance with similar functions in guts [37,39].
Disk diffusion assays and RNAi treatment were performed for function verification of MaGSTe3. Smaller inhibition zones in overexpressed groups than in control demonstrated the antioxidant capacity of MaGSTe3. Meanwhile, RNAi treatment of MaGSTe3 significantly reduced the GST activity of M. alternatus larvae. However, RNAi treatment of MaGSTe3 did not reduce the level of hydrogen peroxide and MDA. Many studies have shown that resistance to oxidative stress based on GSTs usually involves multiple enzymes from the same or different GST classes [24,31,32]. Though MaGSTe3 is the only gst gene detected in the transcriptome of our study, we cannot deny the possibility of other GSTs participating. For example, GST genes from the sigma and omega classes, though their functions are not yet clear, were also observed to be upregulated when responding to oxidative stress, but their expression level is relatively low [31,40,41,42]. Moreover, epsilon GSTs were reported differentially regulated in the different feeding and terpenoids treatments [31,32]. There is a possibility that other gst genes, which were not originally upregulated, may have a substitute effect in M. alternatus larvae when MaGSTe3 was knocked down. Thus, further study on the expression profiling of MaGSTs is needed to clarify the role of GSTs in the α-pinene adaptation of M. alternatus. Another possibility is that MaGSTs do not play a major role in α-pinene resistance. Our previous study found that P450s and UGTs were the two genes with the highest number of upregulated expressions in detoxification enzymes, and knockdown of P450s increased mortality in the larvae under α-pinene stress [17]. Some studies reported that the P450 gene has specific expansion in conifer pests such as Dendrolimus punctatus, Thaumetopoea pityocampa, and D. ponderosae, indicating that P450s may be the key factor in resistance to host defensive substances [43,44]. It was also observed that UGTs are associated with eliminating oxidative stress [45]. Therefore, MaGSTe3 may play a supporting role in the detoxification process of P450, and knockdown of MaGSTe3 will not affect the detoxification of α-pinene in this case. However, the relationship between the P450s and GSTs of M. alternatus still needs further research.

4. Materials and Methods

4.1. Insect Collection and Treatment

The 2nd- and 3rd-instar M. alternatus larvae were collected from diseased and dead pines in Jiujiang city (115°56′35″ E, 29°34′1″ N), Jiangxi province, China, in September 2020. The larvae were disinfected with 75% ethanol and then reared in an incubator in the dark (26 ± 2 °C, RH = 70 ± 5%) with an artificial diet [46].
The recipe of the artificial diet was according to Chen et al. [46], which consisted of 100 g pine sawdust, 40 g agar, 20 g sucrose, 12.5 g yeast extract, 2 g sodium benzoate, 1 g potassium sorbate, 10 mL 0.5 M sulfuric acid, 25 g wheat germ powder, 1.5 g cholesterol, 4 g ascorbic acid, 20 g casein, 1 g choline chloride, and 800 mL water.
The 4th-instar larvae, which have relatively higher tolerance to α-pinene [17], were used in our study. After starvation for 24 h, the larvae were divided into four groups and fed an artificial diet containing α-pinene (Aladdin, Shanghai, China) of different concentrations (2.8, 5.6, and 11.2 mg/g), which were named the low-, medium- and high-dose groups, respectively. Larvae fed a diet containing no α-pinene were used as control. All treatment groups were used for experiments unless otherwise stated.

4.2. Determination of Hydrogen Peroxide, MDA, and Activity of GSTs

The larvae were dissected to take intestinal samples after treatment for different durations (24, 48, and 72 h). The samples were ground in an ice bath after accurately weighing them. Subsequently, the homogenate was centrifuged at 600× g for 10 min, and the supernatant was used for assays.
Determination of hydrogen peroxide was performed with a hydrogen peroxide assay kit (Nanjing Jiancheng Bioengineering Institute, Nanjing, China) according to the manufacturer’s instructions. In brief, the supernatant was mixed with molybdic acid solution. The complex generated by hydrogen peroxide and molybdic acid has its maximum absorption peak at 405 nm. Thus, the hydrogen peroxide content was calculated by the absorbance at 405 nm and defined as the amount of produced complex per gram of tissue homogenate protein.
MDA was determined using a malondialdehyde (MDA) assay kit (TBA method) (Nanjing Jiancheng Bioengineering Institute, Nanjing, China) according to the manufacturer’s instructions. Briefly, the supernatant was mixed with thiobarbituric acid (TBA) and trichloroacetic acid (TCA) and then heated in boiling water for 20 min at 95 °C. The sample solution was cooled and centrifuged at 1500× g for 10 min. Condensation products of MDA and TBA have their maximum absorption peak at 532 nm. Therefore, the MDA content was calculated using the absorbance at 532 nm and defined as the amount of condensation products per gram of tissue homogenate protein.
A GST activity assay was performed using a glutathione S-transferase (GST) activity assay kit (Sangon Biotech, Shanghai, China) according to the manufacturer’s instructions. To be brief, the supernatant was mixed with 1-chloro-2,4-dinitrobenzene (CDNB) and reducing glutathione solution (GSH). Absorbance at 340 nm was recorded at 30 s intervals to calculate the GST activity. Each milligram of tissue homogenate protein that catalyzes 1 µmol of CDNB binding to GSH per minute was defined as an enzyme activity unit.

4.3. RNA Isolation and cDNA Synthesis

Larvae of the low-dose group were used for total RNA isolation with a TRNzol Universal kit (Tiangen®, Beijing, China) under the manufacturer’s instructions. RNA integrity was evaluated on 1% agarose gels and quantified with NanoDrop 2000 (Thermo Scientific, Pittsburg, PA, USA). Samples with high-quality bands were used for cDNA synthesis with an Evo M-MLV RT Mix Tracking Kit (Accurate Biology, Changsha, China). Concentration and quality of cDNA samples were determined with NanoDrop 2000.

4.4. Amplification and Cloning of MaGSTe3

Based on the transcriptome database constructed in our previous study [17], the unique genes encoding the GSTs of the functional annotations were collected. The full-length open reading frame (ORF) of MaGSTe3 was confirmed using SeqBuilder of DNAStar (DNASTAR, Madison, WI, USA).
The synthesized cDNA from total RNA was used as the template for amplification with ApexHF HS DNA Polymerase FS (Accurate Biology, Changsha, China). The primers used for PCR reaction with complete ORF (Table A1) were designed using Primer 5.0. The reaction was performed in a 50 µL volume containing 3 µL cDNA, 1 µL of each primer (10 mM), 25 µL 2× ApexHF FS PCR Master Mix, and 20 µL nuclease-free water. The processes of amplification were 30 s of initial denaturation at 94 °C, a cycling protocol consisting of 35 cycles of denaturation at 98 °C for 10 s, annealing at Tm for 15 s, and extension at 72 °C for 5 s.
The PCR products were recovered using a SteadyPure Agarose Gel DNA Purification Kit (Accurate Biology, Changsha, China) and subsequently ligated to circular plasmids using a 5 min TA/Blunt-Zero Cloning Kit (Vazyme, Nanjing, China). The plasmids were transformed into Escherichia coli DHα5 (Vazyme, Nanjing, China). Single colonies were shaking cultured on Luria–Bertani (LB) medium for 1 h at 37 °C at 200 rpm. Positive bacterial fluid was selected and sent to Sangon Biotech (Shanghai) Co., Ltd. (Shanghai, China) for sequencing.

4.5. Sequence and Phylogenetic Analysis of MaGSTe3

The theoretical isoelectric points and the molecular weight of MaGSTe3 were predicted using ExPASy (https://web.expasy.org/, accessed on 4 March 2022). The subcellular localization was predicted with the website (http://cello.life.nctu.edu.tw/, accessed on 4 March 2022). The conserved domain of MaGSTe3 was predicted with the Conserved Domain Database of NCBI (https://www.ncbi.nlm.nih.gov/cdd/, accessed on 4 March 2022). The three-dimensional structure of MaGSTe3 was predicted through SWISS-MODEL (https://swissmodel.expasy.org/, accessed on 4 March 2022) [47].
The obtained sequences were aligned with 64 insect GST sequences collected from Uniprot (https://www.uniprot.org/, accessed on 4 March 2022) and NCBI (https://www.ncbi.nlm.nih.gov/, accessed on 4 March 2022) using DNAMAN 8.0 (Lynnon Biosoft, San Ramon, CA, USA). Then, a phylogenetic tree was constructed by the maximum likelihood method using Mega 7.0 (https://www.megasoftware.net/download_form, accessed on 1 May 2021), and the bootstrap value was set to 1000.

4.6. qRT-PCR

The cDNA used for qRT-PCR was obtained with the methods mentioned above. Primers (Table A2) were designed with Primer-BLAST (https://www.ncbi.nlm.nih. gov/tools/primer-blast/, accessed on 4 March 2022). To estimate the qPCR efficiency and validation, the cDNA was diluted to 200 ng/μL and then diluted to five gradients (1.0, 10−1, 10−2, 10−3, and 10−4). The standard curve was established after the amplification of cDNA of different dilutions and their initial concentrations. Primers with R2 > 0.98 and 100 ± 10% efficiency were used for subsequent experiments.
The qRT-PCR reactions were carried with the ABI 7900HT Real-Time PCR System using a Hieff qPCR SYBE Green Master Mix (Yeasen, Shanghai, China). The reaction was performed in a 20 μL volume containing 2 μL cDNA templates, 0.4 μL of each primer (10 μM), 10 μL of Hieff qPCR SYBE Green Master Mix, and 7.2 μL of nuclease-free water. The thermal cycle conditions were 5 min of initial denaturation at 95 °C, a cycling protocol consisting of 40 cycles of denaturation at 95 °C for 10 s, and annealing at 60 °C for 40 s. The ribosomal protein L10 (RPL10) gene of M. alternatus was used as the reference gene. The assay was performed in triplicate and repeated three times. The expression level was quantified using the 2−ΔΔCt method.

4.7. Prokaryotic Expression and Purification of Recombinant MaGSTe3 Protein

The entire coding sequence of MaGSTe3 was amplified with primers containing the specific sequences of the vector. The amplified fragments were then inserted into pET-28a (+) stored in our laboratory and then transformed into E. coli BL21 (DE3) (Vazyme, Nanjing, China). The cells were shaking-cultured under the condition of 37 °C and 200 rpm in LB liquid medium with kanamycin. When the optical density at 600 nm (OD600) reached the range of 0.4 to 0.6, 600 μL 500mM isopropyl β-d-1-thiogalactopyranoside (IPTG) was added to induce the expression of recombinant proteins at 20 °C at 200 rpm for 18 h. The harvested cells were sonicated and then centrifuged at 13,500× g and 4 °C for 25 min. The supernatant was purified using nickel–nitrilotriacetic acid (Ni–NTA) resin. The recombinant GST protein was evaluated by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS–PAGE). Purified protein was dialyzed with three dialysates (Table A3) in a dialysis bag (25 × 18 cm, 15 kDa retention molecular weight), frozen for 6 h, and then diluted to 1 mg/mL. The protein concentrations were determined using a total protein quantitative assay kit (Jiancheng, Nanjing, China).

4.8. Enzyme Activity Assays of the Recombinant MaGSTe3

To determine the activity of recombinant MaGSTe3, 1-chloro-2,4-dinitrobenzene (CDNB) was used as the substrate. All reactions were performed with a total reaction volume of 200 μL containing 2 μL protein, 2 μL GSH, 2 μL CNDB, and 194 μL 100mM PBS.
The kinetic parameters of recombinant MaGSTe3 were determined by various concentrations (0–400 mM) of CDNB substrate. The boiled inactivated recombinant MaGSTe3 was used as control, and each test was repeated three times independently.
Special activity (SA) was measured to construct an enzyme kinetic curve. OD340 was measured at 30 s intervals for 5 min for calculation of SA. The SA of GST was expressed as per mole of CDNB conjugated with GSH per minute per milligram of protein using the extinction coefficient of the resulting 2,4-dinitrophenyl-glutathione, which is about 9.6 mM/cm. The enzyme kinetic curve was fitted using the Michaelis–Menten equation by Origin 2021 (OriginLab, Northampton, MA, USA).
The optimum temperature of MaGSTe3 was assayed by incubating the recombinant protein in PBS (pH 7.4) at a 10–60 °C temperature gradient for 30 min. The optimum pH for MaGSTe3 activity was investigated with different PBS with a pH range of 4.0 to 11.0. The SA was measured, and the maximum SA of each experiment was considered as 100% of the relative enzyme activity.

4.9. Disk Diffusion Assay

A disk diffusion assay was performed following Burmeister et al. [48] with some modification. The E. coli BL21 (DE3) culture containing overexpressed MaGSTe3 was plated onto LB agar plates with kanamycin. Each plate contained about 5 × 108 cells, which were measured by an Auto Colony Counter MF2 (SHINESO, Hangzhou, China). After incubation at 37 °C for 1 h, filter discs 6 mm in diameter soaked with various concentrations of cumene hydroperoxide (CHP, 0, 25, 50, 100, and 200 mM) were placed on the plates and then incubated at 37 °C for 24 h. The bacteriostatic rings were analyzed using Auto Colony Counter MF2.

4.10. RNA Interference

To obtain dsRNA fragments, pairs of primers (Table A5) were designed based on the ORF of MaGSTe3 with Primer 5.0, and the T7 promoter (TTAATACGACTCACTATAGGG) was added to both ends of primers using the T7 RiboMAXTM Express RNAi System (Promega, Madison, WI, USA). A PCR reaction was performed for DNA used as a dsRNA template with the method described above. After recovery and purification, the products were transcribed into dsMaGSTe3 and subsequently injected into the 4th-instar larvae. The RNAi-treated larvae were treated with a diet containing 5.6 mg/g α-pinene, and their guts were taken after treatment for 24 h for hydrogen peroxide MDA and GST activity determination. Meanwhile, the RNAs were extracted for qRT-PCR using the above-mentioned methods.

4.11. Statistical Analysis

The data were analyzed using SPSS 25.0 (IBM, Armonk, NY, USA). The differences between treatment and control groups were analyzed by one-way ANOVA with a significance level of p < 0.05. The values were presented as the mean value ± standard deviation (SD) of three independent experiments.

5. Conclusions

Our results suggested that MaGSTe3 takes part in α-pinene adaptation in M. alternatus larvae through protecting tissues from oxidative stress but does not play a major role.

Author Contributions

Conceptualization, D.H. and W.-Y.C.; methodology, H.L, W.-Y.C. and X.X.; investigation: X.X., M.X., Y.D., F.T. and S.Z.; formal analysis, X.X., S.Z. and M.X.; writing—original draft preparation, X.X., M.X., Y.D. and F.T.; writing—review and editing, H.L. and D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Foundation of China (No. 31470650), the National Science Foundation of China (No. 32001322), and the Basic Research Program Natural Science Foundation of Jiangsu Province (No. BK20220412).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. SDS-PAGE analysis of MaGSTe3 after recombinant expression and purification. L1 and L6 are protein molecular standards. L2 and L4 are the supernatant and precipitate, respectively, of the empty pET-28a (+) vector after 1 mM IPTG induction. L3 and L5 are the supernatant and precipitate, respectively, of pET-28a (+)-inserted MaGSTe3 after 1 mM IPTG induction. L7 and L8 are purified MaGSTe3.
Figure A1. SDS-PAGE analysis of MaGSTe3 after recombinant expression and purification. L1 and L6 are protein molecular standards. L2 and L4 are the supernatant and precipitate, respectively, of the empty pET-28a (+) vector after 1 mM IPTG induction. L3 and L5 are the supernatant and precipitate, respectively, of pET-28a (+)-inserted MaGSTe3 after 1 mM IPTG induction. L7 and L8 are purified MaGSTe3.
Ijms 24 17376 g0a1
Figure A2. Three-dimensional structure of MaGSTe3 predicted by SWISS-MODEL. The red parts are α-helices and the yellow parts are β-sheets.
Figure A2. Three-dimensional structure of MaGSTe3 predicted by SWISS-MODEL. The red parts are α-helices and the yellow parts are β-sheets.
Ijms 24 17376 g0a2
Figure A3. The conserved domain of MaGSTe3.
Figure A3. The conserved domain of MaGSTe3.
Ijms 24 17376 g0a3

Appendix B

Table A1. Primer used in the GST genes cloning.
Table A1. Primer used in the GST genes cloning.
Prime NamePrimer Sequence (5′→3′)
MaGSTe3-FATGACGGTTATAGTGTATGGGAC
MaGSTe3-R TTAAGCCAGTTTGCTCTTAAACGCTG
Table A2. Primer used in the GSTs qRT-PCR.
Table A2. Primer used in the GSTs qRT-PCR.
Prime NamePrimer Sequence (5′→3′)
RPL-10-qFGACAAACGTTTCAGCGGAAC
RPL-10-qRTGCTGTTGATCGCCCAAAAC
MaGSTe3-qFCAGAAGAGGGCCGTGGTTA
MaGSTe3-qRCAGCAGCGTAAGTGCTTCTT
Table A3. Dialysates used for dialysis of purified protein MaGSTe3 protein.
Table A3. Dialysates used for dialysis of purified protein MaGSTe3 protein.
Dialysates NameComponents
Dialysates I40 mM PBS (pH = 7.4), 200 mM NaCl, 200 mM imidazole
Dialysates II20 mM PBS (pH = 7.4), 100 mM NaCl, 10 mM imidazole
Dialysates IIIsterile water
Note: PBS refers to phosphate buffer solution.
Table A4. The basic characteristics of MaGSTe3.
Table A4. The basic characteristics of MaGSTe3.
Gene NameLengthMwpIBLASTX Best HitSubcellular Location
(bp)(kDa) SpeciesAcc. No.E-ValueIdentity
(%)
MaGSTe3648 23.58 5.72 Anoplophora glabripennisXP_018564047.12 × 10−999Cytoplasmic
Table A5. Primers for RNA interference of MaGSTe3.
Table A5. Primers for RNA interference of MaGSTe3.
Prime NamePrimer Sequence (5′→3′)
ds-MaGSTe3-T7-FtaatacgactcactatagggAGTGAACACTATGGCTGGGG
ds-MaGSTe3-T7-RtaatacgactcactatagggGACGCGTAGTAAGGCAAAGC
ds-GFP-T7-FggatcctaatacgactcactataggATGAGTAAAGGAGAAGAACTTTTC
ds-GFP-T7-RggatcctaatacgactcactataggTTTGTATAGTTCATCCATGCCAT
ds-MaGSTe3-q-FACCAGTGCCACTACAACAATG
ds-MaGSTe3-q-RGTTCCTCTGACGACTGGACT

References

  1. Wilf, P.; Labandeira, C.C.; Johnson, K.R.; Coley, P.D.; Cutter, A.D. Insect herbivory, plant defense, and early Cenozoic climate change. Proc. Natl. Acad. Sci. USA 2001, 98, 6221–6226. [Google Scholar] [CrossRef]
  2. Keeling, C.I.; Bohlmann, J. Genes, enzymes and chemicals of terpenoid diversity in the constitutive and induced defence of conifers against insects and pathogens. New Phytol. 2006, 170, 657–675. [Google Scholar] [CrossRef]
  3. Trapp, S.; Croteau, R. Defensive resin biosynthesis in conifers. Annu. Rev. Plant Biol. 2001, 52, 689–724. [Google Scholar] [CrossRef]
  4. Franceschi, V.R.; Krokene, P.; Christiansen, E.; Krekling, T. Anatomical and chemical defenses of conifer bark against bark beetles and other pests. New Phytol. 2005, 167, 353–376. [Google Scholar] [CrossRef]
  5. Phillips, M.A.; Croteau, R.B. Resin-based defenses in conifers. Trends Plant Sci. 1999, 4, 184–190. [Google Scholar] [CrossRef]
  6. Liu, B.; Chen, H. Disruption of CYP6DF1 and CYP6DJ2 increases the susceptibility of Dendroctonus armandi to (+)-α-pinene. Pestic. Biochem. Physiol. 2022, 188, 105270. [Google Scholar] [CrossRef]
  7. Shahriari, M.; Zibaee, A.; Sahebzadeh, N.; Shamakhi, L. Effects of α-pinene, trans-anethole, and thymol as the essential oil constituents on antioxidant system and acetylcholine esterase of Ephestia kuehniella Zeller (Lepidoptera: Pyralidae). Pestic. Biochem. Physiol. 2018, 150, 40–47. [Google Scholar] [CrossRef]
  8. Chiu, C.C.; Keeling, C.I.; Bohlmann, J. Toxicity of pine monoterpenes to mountain pine beetle. Sci. Rep. 2017, 7, 8858. [Google Scholar] [CrossRef]
  9. Abrahim, D.; Francischini, A.C.; Pergo, E.M.; Kelmer-Bracht, A.M.; Ishii-Iwamoto, E.L. Effects of α-pinene on the mitochondrial respiration of maize seedlings. Plant Physiol. Biochem. 2003, 41, 985–991. [Google Scholar] [CrossRef]
  10. Singh, H.P.; Batish, D.R.; Kaur, S.; Arora, K.; Kohli, R.K. α-Pinene inhibits growth and induces oxidative stress in roots. Ann. Bot. 2006, 98, 1261–1269. [Google Scholar] [CrossRef]
  11. Fernanda López, M.; Cano-Ramírez, C.; Shibayama, M.; Zúñiga, G. α-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]
  12. Rufino, A.T.; Ribeiro, M.; Judas, F.; Salgueiro, L.; Lopes, M.C.; Cavaleiro, C.; Mendes, A.F. Anti-inflammatory and chondroprotective activity of (+)-α-pinene: Structural and enantiomeric selectivity. J. Nat. Prod. 2014, 77, 264–269. [Google Scholar] [CrossRef]
  13. Ye, J. Epidemic Status of Pine Wilt Disease in China and Its Prevention and Control Techniques and Counter Measures. Sci. Silvae Sin. 2019, 55, 1–10. [Google Scholar]
  14. Li, M.; Li, H.; Sheng, R.-C.; Sun, H.; Sun, S.-H.; Chen, F.-M. The first record of Monochamus saltuarius (Coleoptera; Cerambycidae) as vector of Bursaphelenchus xylophilus and its new potential hosts in China. Insects 2020, 11, 636. [Google Scholar] [CrossRef]
  15. Li, H.; Zhao, X.; Qiao, H.; He, X.; Tan, J.; Hao, D. Comparative transcriptome analysis of the heat stress response in Monochamus alternatus Hope (Coleoptera: Cerambycidae). Front. Physiol. 2020, 10, 1568. [Google Scholar] [CrossRef]
  16. Chen, R.; He, X.; Chen, J.; Gu, T.; Liu, P.; Xu, T.; Teale, S.A.; Hao, D. Traumatic resin duct development, terpenoid formation, and related synthase gene expression in Pinus massoniana under feeding pressure of Monochamus alternatus. J. Plant Growth Regul. 2019, 38, 897–908. [Google Scholar] [CrossRef]
  17. He, X.; Hao, Z.; Li, H.; Xu, T.; Chen, R.; Xia, X.; Li, S.; Hao, D. Transcriptome profiling and RNA interference reveals relevant detoxification genes in Monochamus alternatus response to (+)-α-pinene. J. Appl. Entomol. 2022, 146, 823–837. [Google Scholar] [CrossRef]
  18. Krishnan, N.; Kodrík, D. Antioxidant enzymes in Spodoptera littoralis (Boisduval): Are they enhanced to protect gut tissues during oxidative stress? J. Insect Physiol. 2006, 52, 11–20. [Google Scholar] [CrossRef]
  19. Wu, D.; Yang, Z.; Huang, Y. Analysis and evaluation of resin productivity and resin component among different half sibling families of Pinus massoniana. J. Beijing For. Univ. 2019, 41, 53–61. [Google Scholar]
  20. Mercier, B.; Prost, J.; Prost, M. The essential oil of turpentine and its major volatile fraction (α-and β-pinenes): A review. Int. J. Occup. Med. Environ. Health 2009, 22, 331–342. [Google Scholar] [CrossRef]
  21. Xu, B.B.; Liu, Z.D.; Sun, J.H. The effects of α-pinene on the feeding performance and pheromone production of Dendroctonus valens. Entomol. Exp. Et Appl. 2014, 150, 269–278. [Google Scholar] [CrossRef]
  22. Tsikas, D. Assessment of lipid peroxidation by measuring malondialdehyde (MDA) and relatives in biological samples: Analytical and biological challenges. Anal. Biochem. 2017, 524, 13–30. [Google Scholar] [CrossRef]
  23. Meng, J.-Y.; Zhang, C.-Y.; Zhu, F.; Wang, X.-P.; Lei, C.-L. Ultraviolet light-induced oxidative stress: Effects on antioxidant response of Helicoverpa armigera adults. J. Insect Physiol. 2009, 55, 588–592. [Google Scholar] [CrossRef]
  24. Sun, L.; Yin, J.; Du, H.; Liu, P.; Cao, C. Characterisation of GST genes from the Hyphantria cunea and their response to the oxidative stress caused by the infection of Hyphantria cunea nucleopolyhedrovirus (HcNPV). Pestic. Biochem. Physiol. 2020, 163, 254–262. [Google Scholar] [CrossRef]
  25. Singh, S.P.; Coronella, J.A.; Beneš, H.; Cochrane, B.J.; Zimniak, P. Catalytic function of Drosophila melanogaster glutathione S-transferase DmGSTS1-1 (GST-2) in conjugation of lipid peroxidation end products. Eur. J. Biochem. 2001, 268, 2912–2923. [Google Scholar] [CrossRef]
  26. Ding, Y.; Ortelli, F.; Rossiter, L.C.; Hemingway, J.; Ranson, H. The Anopheles gambiae glutathione transferase supergene family: Annotation, phylogeny and expression profiles. BMC Genom. 2003, 4, 35. [Google Scholar] [CrossRef]
  27. Jing, T.-X.; Wu, Y.-X.; Li, T.; Wei, D.-D.; Smagghe, G.; Wang, J.-J. Identification and expression profiles of fifteen delta-class glutathione S-transferase genes from a stored-product pest, Liposcelis entomophila (Enderlein)(Psocoptera: Liposcelididae). Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2017, 206, 35–41. [Google Scholar] [CrossRef]
  28. Meng, L.W.; Yuan, G.R.; Lu, X.P.; Jing, T.X.; Zheng, L.S.; Yong, H.X.; Wang, J.J. Two delta class glutathione S-transferases involved in the detoxification of malathion in Bactrocera dorsalis (Hendel). Pest Manag. Sci. 2019, 75, 1527–1538. [Google Scholar] [CrossRef]
  29. Tao, F.; Si, F.L.; Hong, R.; He, X.; Li, X.Y.; Qiao, L.; He, Z.B.; Yan, Z.T.; He, S.L.; Chen, B. Glutathione S-transferase (GST) genes and their function associated with pyrethroid resistance in the malaria vector Anopheles sinensis. Pest Manag. Sci. 2022, 78, 4127–4139. [Google Scholar] [CrossRef]
  30. Gonzalez, D.; Fraichard, S.; Grassein, P.; Delarue, P.; Senet, P.; Nicolaï, A.; Chavanne, E.; Mucher, E.; Artur, Y.; Ferveur, J.-F. Characterization of a Drosophila glutathione transferase involved in isothiocyanate detoxification. Insect Biochem. Mol. Biol. 2018, 95, 33–43. [Google Scholar] [CrossRef]
  31. Dai, L.; Ma, J.; Ma, M.; Zhang, H.; Shi, Q.; Zhang, R.; Chen, H. Characterisation of GST genes from the Chinese white pine beetle Dendroctonus armandi (Curculionidae: Scolytinae) and their response to host chemical defence. Pest Manag. Sci. 2016, 72, 816–827. [Google Scholar] [CrossRef]
  32. Gao, H.; Dai, L.; Fu, D.; Sun, Y.; Chen, H. Isolation, expression profiling, and regulation via host allelochemicals of 16 Glutathione S-Transferases in the Chinese White Pine Beetle, Dendroctonus armandi. Front. Physiol. 2020, 11, 546592. [Google Scholar] [CrossRef]
  33. Li, S.; Wang, J.; Chen, C.; Li, H.; Hao, D. Tolerance, biochemistry and related gene expression in Pagiophloeus tsushimanus (Coleoptera: Curculionidae) exposed to chemical stress from headspace host-plant volatiles. Agric. For. Entomol. 2022, 24, 189–203. [Google Scholar] [CrossRef]
  34. Gao, H.; Lin, X.; Yang, B.; Liu, Z. The roles of GSTs in fipronil resistance in Nilaparvata lugens: Over-expression and expression induction. Pestic. Biochem. Physiol. 2021, 177, 104880. [Google Scholar] [CrossRef]
  35. Li, H.M.; Buczkowski, G.; Mittapalli, O.; Xie, J.; Wu, J.; Westerman, R.; Schemerhorn, B.; Murdock, L.; Pittendrigh, B. Transcriptomic profiles of Drosophila melanogaster third instar larval midgut and responses to oxidative stress. Insect Mol. Biol. 2008, 17, 325–339. [Google Scholar] [CrossRef]
  36. Enayati, A.A.; Ranson, H.; Hemingway, J. Insect glutathione transferases and insecticide resistance. Insect Mol. Biol. 2005, 14, 3–8. [Google Scholar] [CrossRef]
  37. Huang, Y.; Xu, Z.; Lin, X.; Feng, Q.; Zheng, S. Structure and expression of glutathione S-transferase genes from the midgut of the Common cutworm, Spodoptera litura (Noctuidae) and their response to xenobiotic compounds and bacteria. J. Insect Physiol. 2011, 57, 1033–1044. [Google Scholar] [CrossRef]
  38. Chen, X.; Liu, J.; Zhang, C.; Li, Y.; Liu, T.; Wang, L.; Yu, Q.; Zhang, Y.; Lu, C.; Pan, M. The silkworm GSTe4 is sensitive to phoxim and protects HEK293 cells against UV-induced cell apoptosis. Bull. Entomol. Res. 2015, 105, 399–407. [Google Scholar] [CrossRef]
  39. Li, S.; Li, H.; Wang, J.; Chen, C.; Hao, D. Hormetic response and co-expression of cytochrome P450 and cuticular protein reveal the tolerance to host-specific terpenoid defences in an emerging insect pest, Pagiophloeus tsushimanus (Coleoptera: Curculionidae). J. Pest Sci. 2023, 96, 141–160. [Google Scholar] [CrossRef]
  40. Balakrishnan, B.; Su, S.; Wang, K.; Tian, R.; Chen, M. Identification, expression, and regulation of an omega class glutathione S-transferase in Rhopalosiphum padi (L.)(Hemiptera: Aphididae) under insecticide stress. Front. Physiol. 2018, 9, 427. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Yang, B.; Yu, N.; Luo, G.; Gao, H.; Lin, X.; Liu, Z. Insecticide resistance associated overexpression of two sigma GST genes assists Nilaparvata lugens to remedy oxidative stress from feeding on resistant rice variety. Pestic. Biochem. Physiol. 2022, 188, 105230. [Google Scholar] [CrossRef]
  42. Hassan, F.; Singh, K.P.; Ali, V.; Behera, S.; Shivam, P.; Das, P.; Dinesh, D.S. Detection and functional characterization of sigma class GST in Phlebotomus argentipes and its role in stress tolerance and DDT resistance. Sci. Rep. 2019, 9, 19636. [Google Scholar] [CrossRef]
  43. Zhang, S.; Shen, S.; Peng, J.; Zhou, X.; Kong, X.; Ren, P.; Liu, F.; Han, L.; Zhan, S.; Huang, Y. Chromosome-level genome assembly of an important pine defoliator, Dendrolimus punctatus (Lepidoptera; Lasiocampidae). Mol. Ecol. Resour. 2020, 20, 1023–1037. [Google Scholar] [CrossRef]
  44. Keeling, C.I.; Henderson, H.; Li, M.; Yuen, M.; Clark, E.L.; Fraser, J.D.; Huber, D.P.; Liao, N.Y.; Docking, T.R.; Birol, I. Transcriptome and full-length cDNA resources for the mountain pine beetle, Dendroctonus ponderosae Hopkins, a major insect pest of pine forests. Insect Biochem. Mol. Biol. 2012, 42, 525–536. [Google Scholar] [CrossRef]
  45. Cui, X.; Wang, C.; Wang, X.; Li, G.; Liu, Z.; Wang, H.; Guo, X.; Xu, B. Molecular Mechanism of the UDP-Glucuronosyltransferase 2B20-like Gene (AccUGT2B20-like) in Pesticide Resistance of Apis cerana cerana. Front. Genet. 2020, 11, 592595. [Google Scholar] [CrossRef]
  46. Chen, R.; Wang, L.; Lin, T.; Wei, Z.; Wang, Y.; Hao, D. Rearing techniques of Monochamus alternatus Hope (Coleoptera: Cerambycidae) on artificial diets. J. Nanjing For. Univ. 2017, 60, 199. [Google Scholar]
  47. Waterhouse, A.; Bertoni, M.; Bienert, S.; Studer, G.; Tauriello, G.; Gumienny, R.; Heer, F.T.; de Beer, T.A.P.; Rempfer, C.; Bordoli, L. SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Res. 2018, 46, W296–W303. [Google Scholar] [CrossRef]
  48. Burmeister, C.; Lüersen, K.; Heinick, A.; Hussein, A.; Domagalski, M.; Walter, R.D.; Liebau, E. Oxidative stress in Caenorhabditis elegans: Protective effects of the Omega class glutathione transferase (GSTO-1). FASEB J. 2008, 22, 343–354. [Google Scholar] [CrossRef]
Figure 1. The contents of (A) hydrogen peroxide and (B) MDA, and (C) the GST activity of M. alternatus larvae at 24, 48, and 72 h after treatment with three different concentrations (2.8, 5.6, and 11.2 mg/g) of α-pinene. The contents of hydrogen peroxide and MDA were represented as μmol/mg protein. Data are presented as mean values ± standard deviation (SD). The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant difference.
Figure 1. The contents of (A) hydrogen peroxide and (B) MDA, and (C) the GST activity of M. alternatus larvae at 24, 48, and 72 h after treatment with three different concentrations (2.8, 5.6, and 11.2 mg/g) of α-pinene. The contents of hydrogen peroxide and MDA were represented as μmol/mg protein. Data are presented as mean values ± standard deviation (SD). The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant difference.
Ijms 24 17376 g001
Figure 2. The phylogenetic tree of MaGSTe3. A total of 64 GST insect sequences from 5 species were used to construct the tree by the maximum likelihood (ML) method, with a total of 1000 bootstrap replications. The abbreviations of species are as follows: Dm—Drosophila melanogaster, Bm—Bombyx mori, Tc—Tribolium castaneum, Ld—Leptinotarsa decemlineata, and So—Sitophilus oryzae.
Figure 2. The phylogenetic tree of MaGSTe3. A total of 64 GST insect sequences from 5 species were used to construct the tree by the maximum likelihood (ML) method, with a total of 1000 bootstrap replications. The abbreviations of species are as follows: Dm—Drosophila melanogaster, Bm—Bombyx mori, Tc—Tribolium castaneum, Ld—Leptinotarsa decemlineata, and So—Sitophilus oryzae.
Ijms 24 17376 g002
Figure 3. Relative expression level of MaGSTe3 in (A) hemolymph, (B) fat body, (C) epidermis, and (D) guts at 12, 24, and 48 h after treatment with three different concentrations (2.8, 5.6, and 11.2 mg/g) of α-pinene. Data are presented as mean values ± SD. The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant differences.
Figure 3. Relative expression level of MaGSTe3 in (A) hemolymph, (B) fat body, (C) epidermis, and (D) guts at 12, 24, and 48 h after treatment with three different concentrations (2.8, 5.6, and 11.2 mg/g) of α-pinene. Data are presented as mean values ± SD. The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant differences.
Ijms 24 17376 g003
Figure 4. Enzymatic properties of recombinant MaGSTe3. (A) Relative activity of purified recombinant MaGSTe3 at different temperatures. (B) Relative activity of purified recombinant MaGSTe3 at various pH values.
Figure 4. Enzymatic properties of recombinant MaGSTe3. (A) Relative activity of purified recombinant MaGSTe3 at different temperatures. (B) Relative activity of purified recombinant MaGSTe3 at various pH values.
Ijms 24 17376 g004
Figure 5. Disc diffusion assays using E. coli overexpressing MaGSTe3. (A) Resistance of E. coli cells overexpressing MaGSTe3 to cumene hydroperoxide. The labels 1~5 on filter disks represent 25, 20, 10, 5, and 2 mM cumene hydroperoxide, respectively. (B) Histograms comparing halo diameters of inhibition zones. Data are presented as mean values ± SD. The differences between treatment groups and control were analyzed by Student’s t test, and asterisks mean significant differences: (**) p < 0.01.
Figure 5. Disc diffusion assays using E. coli overexpressing MaGSTe3. (A) Resistance of E. coli cells overexpressing MaGSTe3 to cumene hydroperoxide. The labels 1~5 on filter disks represent 25, 20, 10, 5, and 2 mM cumene hydroperoxide, respectively. (B) Histograms comparing halo diameters of inhibition zones. Data are presented as mean values ± SD. The differences between treatment groups and control were analyzed by Student’s t test, and asterisks mean significant differences: (**) p < 0.01.
Ijms 24 17376 g005
Figure 6. Effect of RNAi treatment on (A) GST activities, (B) relative expression level of MaGSTe3, (C) hydrogen peroxide, and (D) MDA levels. All larvae were treated with 5.6 mg·g−1 α-pinene. The contents of hydrogen peroxide and MDA were represented as μmol/mg protein. Data are presented as mean values ± SD. The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant difference.
Figure 6. Effect of RNAi treatment on (A) GST activities, (B) relative expression level of MaGSTe3, (C) hydrogen peroxide, and (D) MDA levels. All larvae were treated with 5.6 mg·g−1 α-pinene. The contents of hydrogen peroxide and MDA were represented as μmol/mg protein. Data are presented as mean values ± SD. The differences between different groups were analyzed by one-way ANOVA with a significance level of p < 0.05, and different letters showed significant difference.
Ijms 24 17376 g006
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

Xue, M.; Xia, X.; Deng, Y.; Teng, F.; Zhao, S.; Li, H.; Hao, D.; Chen, W.-Y. Identification and Functional Analysis of an Epsilon Class Glutathione S-Transferase Gene Associated with α-Pinene Adaptation in Monochamus alternatus. Int. J. Mol. Sci. 2023, 24, 17376. https://doi.org/10.3390/ijms242417376

AMA Style

Xue M, Xia X, Deng Y, Teng F, Zhao S, Li H, Hao D, Chen W-Y. Identification and Functional Analysis of an Epsilon Class Glutathione S-Transferase Gene Associated with α-Pinene Adaptation in Monochamus alternatus. International Journal of Molecular Sciences. 2023; 24(24):17376. https://doi.org/10.3390/ijms242417376

Chicago/Turabian Style

Xue, Mingyu, Xiaohong Xia, Yadi Deng, Fei Teng, Shiyue Zhao, Hui Li, Dejun Hao, and Wei-Yi Chen. 2023. "Identification and Functional Analysis of an Epsilon Class Glutathione S-Transferase Gene Associated with α-Pinene Adaptation in Monochamus alternatus" International Journal of Molecular Sciences 24, no. 24: 17376. https://doi.org/10.3390/ijms242417376

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