Identification and characterisation of seventeen glutathione S-transferase genes from the cabbage white butterfly Pieris rapae

Identification and characterisation of seventeen glutathione S-transferase genes from the cabbage white butterfly Pieris rapae

Accepted Manuscript Identification and characterisation of seventeen glutathione Stransferase genes from the cabbage white butterfly Pieris rapae Su ...

2MB Sizes 0 Downloads 65 Views

Accepted Manuscript Identification and characterisation of seventeen glutathione Stransferase genes from the cabbage white butterfly Pieris rapae

Su Liu, Yu-Xing Zhang, Wen-Long Wang, Bang-Xian Zhang, ShiGuang Li PII: DOI: Reference:

S0048-3575(17)30104-9 doi: 10.1016/j.pestbp.2017.09.001 YPEST 4111

To appear in:

Pesticide Biochemistry and Physiology

Received date: Revised date: Accepted date:

8 March 2017 30 August 2017 2 September 2017

Please cite this article as: Su Liu, Yu-Xing Zhang, Wen-Long Wang, Bang-Xian Zhang, Shi-Guang Li , Identification and characterisation of seventeen glutathione S-transferase genes from the cabbage white butterfly Pieris rapae. The address for the corresponding author was captured as affiliation for all authors. Please check if appropriate. Ypest(2017), doi: 10.1016/j.pestbp.2017.09.001

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT Identification and characterisation of seventeen glutathione S-transferase genes from the cabbage white butterfly Pieris rapae Su Liu# , Yu-Xing Zhang# , Wen-Long Wang, Bang-Xian Zhang, Shi-Guang Li*

These authors contributed equally to this work.

SC

RI

#

PT

College of Plant Protection, Anhui Agricultural University, Hefei, Anhui 230036, China

*

MA

NU

This manuscript includes 27 pages, 6 figures, 3 tables, 5 supplementary files

Corresponding author:

ED

Shi-Guang Li, College of Plant Protection, Anhui Agricultural University, 130 West

EP T

Changjiang Road, Hefei, Anhui 230036, China.

AC C

E-mail: [email protected]

ACCEPTED MANUSCRIPT

Abstract Insect glutathione S-transferases (GSTs) play essential roles in the detoxification of insecticides and other xenobiotic compounds. The cabbage white butterfly, Pieris rapae, is an economically important agricultural pest. In this study, 17 cDNA sequences encoding

PT

putative GSTs were identified in P. rapae. All cDNAs include a complete open reading

RI

frame and were designated PrGSTd1–PrGSTz2. Based on phylogenetic analysis, PrGSTs

SC

were divided into six classes (delta, epsilon, omega, sigma, theta and zeta). The exon- intron organizations of these PrGSTs were also analysed. Recombinant proteins of eight PrGSTs

NU

(PrGSTD1, PrGSTD2, PrGSTE1, PrGSTE2, PrGSTO1, PrGSTS1, PrGSTT1 and PrGSTZ1)

MA

were heterologously expressed in Escherichia coli, and all of these proteins displayed glutathione-conjugating activity towards 1-chloro-2,4-dinitrobenzene (CDNB). Expression

ED

patterns in various larval tissues, at different life stages, and following exposure to sublethal

EP T

doses of abamectin, chlorantraniliprole or lambda-cyhalothrin were determined by reverse transcription-quantitative PCR. The results showed that PrGSTe3, PrGSTs1, PrGSTs2, and

AC C

PrGSTs4 were mainly transcribed in the fat body, while PrGSTe2 was expressed predominantly in the Malpighian tubules. Four genes (PrGSTe2, PrGSTo4, PrGSTs4 and PrGSTt1) were mainly expressed in fourth- instar larvae, while others were ubiquitously expressed in egg, larval, pupa and/or adult stages. Abamectin treatment significantly upregulated ten genes (PrGSTd1, PrGSTd3, PrGSTe1, PrGSTe2, PrGSTo1, PrGSTo3, PrGSTs1, PrGSTs3, PrGSTs4 and PrGSTt1). Chlorantraniliprole and lambda-cyhalothrin treatment significantly upregulated nine genes (PrGSTd1, PrGSTd2, PrGSTe1, PrGSTe2, PrGSTe3, PrGSTs1, PrGSTs3, PrGSTs4 and PrGSTz1) and ten genes (PrGSTd1, PrGSTd3,

ACCEPTED MANUSCRIPT

PrGSTe1, PrGSTe2, PrGSTo1, PrGSTo2, PrGSTs1, PrGSTs2, PrGSTs3 and PrGSTz2), respectively. These GSTs are potentially involved in the detoxification of insecticides.

Keywords

AC C

EP T

ED

MA

NU

SC

RI

PT

Pieris rapae, GST, transcriptome analysis, insecticide treatment, detoxification

ACCEPTED MANUSCRIPT

1. Introduction Glutathione S-transferases (GSTs) are a superfamily of detoxifying enzymes present in both vertebrates and invertebrates [1]. In insects, GSTs play important roles in the detoxification of various harmful xenobiotic and endobiotic compounds, such as synthetic

PT

insecticides, plant allelochemicals, and lipid peroxides [2]. GSTs are able to catalyse the

RI

conjugation of reduced glutathione (GSH) with xenobiotics and endobiotics, making them

SC

less toxic and easier to excrete from cells [3]. Some other GSTs noncatalytically bind endogenous and exogenous compounds rather than metabolising them [4]. Apart from the

NU

detoxification function, insect GSTs are also required for odorant inactivation, ecdysteroid

MA

biosynthesis and larval development [5-7]. Furthermore, a number of GSTs have the peroxidase activity that protect against oxidative stress [8-9].

ED

There are two types of GSTs present in insects: cytosolic and microsomal [1]. Most

EP T

insect GSTs belong to the cytosolic group, and they have been divided into six classes (delta, epsilon, omega, sigma, theta and zeta) according to their sequence identity, genomic

AC C

structure, and biochemical properties [10]. GSTs that cannot be classified using this nomenclatural system are assigned to an 'unclassified' subgroup [10]. GSTs have two domains that are responsible for the detoxification function: a GSH binding domain (G-site), which is conserved in the N-terminal region of different classes of GSTs, and a hydrophobic substrate binding domain (H-site), which is more variable and located in the C-terminal region [3]. Upregulation of GST genes and enhanced activity of GST proteins has been associated with insecticide detoxification in many insect species [4]. Also, heterologously expressed

ACCEPTED MANUSCRIPT

insect GST proteins are capable of metabolising a variety of insecticides including organochlorines, organophosphates, and pyrethroids [3, 11-12]. Recently, RNA interference (RNAi) has been used to investigate the function of insect GSTs, and in many species including Aedes aegypti, Locusta migratoria, Nilaparvata lugens and Bemisia tabaci, an

PT

involvement in detoxifying insecticides has been demonstrated [13-16].

RI

To date, 37, 28, 8, 36, 28, 29, 24, 9, 28 and 14 cytosolic GST genes have been identified

SC

in the model insect species Drosophila melanogaster, Anopheles gambiae, Apis mellifera, Tribolium castaneum, Dendroctonus ponderosae, Leptinotarsa decemlineata, Acyrthosiphon

NU

pisum, N. lugens, L. migratoria and Rhodnius prolixus, respectively [15, 17-20]. GST genes

MA

have also been discovered in the model species Bombyx mori (23 genes) and Plutella xylostella (22 genes) of the Order Lepidoptera, and in the agricultural pests Spodoptera

ED

litura (37 genes) and Cnaphalocrocis medinalis (25 genes) [21-24]. However, these species

EP T

are all moths, and GST genes in butterflies remain poorly characterised. The small white butterfly, Pieris rapae, is a severe agricultural pest distributed

AC C

worldwide that causes serious yield losses in Brassicaceae crops [25]. In the past decades, management of P. rapae in China mainly relies on the spraying of insecticides. Abamectin, chlorantraniliprole and lambda-cyhalothrin are three insecticides widely used for controlling the pest insect [26]. Recently, however, these insecticides became less efficient to control the pest even at relatively high doses (S. Liu and S.-G. Li, personal observation). It is possible that GSTs in P. rapae contribute to the detoxification of insecticides. In this study, we identified 17 GST genes (PrGSTs) from P. rapae using previously released transcriptome datasets [27]. We also analysed the exon- intron organizations and

ACCEPTED MANUSCRIPT

phylogenetic relationships of these PrGSTs. Eight PrGST proteins were heterologously expressed in Escherichia coli, and their catalytic activities were investigated. Transcription of some genes varied in different larval tissues, at different developmental stages, and following exposure to various concentrations of abamectin, chlorantraniliprole and

PT

lambda-cyhalothrin. To our knowledge, this is the first report o n the large-scale identification

SC

RI

and characterisation of GST genes in this economically important insect pest.

2. Materials and methods

NU

2.1. Insects

MA

P. rapae individuals used in this study originated from a colony collected from cabbage fields in an experimental farmland of Anhui Agricultural University, Hefei, Anhui, China.

ED

Larvae were reared on the leaves of Chinese cabbage (Brassica pekinensis) and adults were

EP T

fed on a 10% (v/v) honey solution. Animals were reared at 25 ± 1°C with 65% relative humidity and a 16:8 h light:dark photoperiod.

AC C

2.2. RNA extraction and cDNA synthesis Total RNA was isolated using RNAiso Plus reagent (Takara, Dalian, China) and treated with RNase-free DNase I (Takara, Dalian, China) to remove potential contaminants from genomic DNA. The quality and concentration of RNA were determined by agarose gel electrophoresis and NanoDrop 2000 spectrophotometer (Thermo Scientific, Wilmington, DE). First-strand cDNA was reverse-transcribed using the TransScript First-Strand cDNA Synthesis SuperMix (Transgen, Beijing, China).

ACCEPTED MANUSCRIPT

2.3. Homology searching and sequence verification cDNA sequences encoding GSTs were retrieved from previously released P. rapae transcriptome datasets [27] using the TBLASTN algorithm in the Basic Local Alignment Search Tool (BLAST) program [28]. Annotated GST protein sequences from model insect

PT

species (including D. melanogaster, A. gambiae, B. mori and P. xylostella) were used as

RI

queries with a cut-off E-value of 1 × 10-5 .

SC

To confirm that the identified GST sequences were not chimeric, gene-specific primers (Table S1) were designed to amplify complete or partial open reading frames (ORFs).

NU

First-strand cDNA from fourth- instar larvae was used as a template. PCR products were

Searching

for

EP T

2.4. Bioinformatic analyses

ED

sequenced in both 5' and 3' directions.

MA

analysed by agarose gel, and DNA bands of the expected size were excised, purified, and

orthologs

was

performed

using

the

BLAST

program

residues

AC C

(http://blast.ncbi.nlm.nih.gov/blast.cgi) with a cut-off E-value of 1 × 10-5 [28]. Catalytic were

predicted

by

searching

the

Conserved

Domain

database

(http://www.ncbi.nlm.nih.gov/structure/cdd/cdd.shtml) [29]. Multiple sequence alignment was performed using the Clustal Omega program (http://www.ebi.ac.uk/tools/msa/clustalo/) [30]. Phylogenetic tree construction was performed by MEGA6.0 software using the neighbour-joining method with the pairwise deletion option [31]. To evaluate the branch strength of the tree, a 1000 bootstrap replication analysis was performed [31]. The GenBank accession numbers of sequences used are listed in Table S2. The genomic DNA of P. rapae

ACCEPTED MANUSCRIPT

was download from a lepidopteran genome database (http://prodata.swmed.edu/LepDB/) [32], and the exon- intron structure of PrGSTs was determined by aligning cDNA sequences with

genomic

DNA

sequences

using

Splign

program

PT

(https://www.ncbi.nlm.nih.gov/sutils/splign/splign.cgi) [33].

RI

2.5. Protein expression and purification

SC

The complete ORFs of eight PrGSTs (PrGSTd1, PrGSTd2, PrGSTe1, PrGSTe2, PrGSTo1, PrGSTs1, PrGSTt1, and PrGSTz1) were cloned using gene-specific primers (Table

NU

S1) and ligated into the pEASY-Blunt E1 expression vector (Transgen, Beijing, China). Each

MA

recombinant protein will be expressed as a fusion protein with a N-terminal 6×His·tag. Each expression construct was verified by DNA sequencing, and the correct constructs were

ED

transformed into the E. coli OrigamiB (DE3) cells. Bacteria were cultured in Luria-Bertani

tetracycline)

and

EP T

(LB) medium (containing 50 μg/ml of ampicillin, 30 μg/ml of kanamycin, and 12.5 μg/ml of grown

at

37

°C

until

OD600

was

0.6,

then

isopropyl

AC C

β-D-1-thiogalactopyranoside (IPTG) was added to 0.5 mM final concentration. After growing at 30 °C for 6 h, cells were harvested by centrifugation at 3000 × g for 5 min, resuspended in lysis buffer [20 mM Tris·HCl (pH 7.4), 500 mM NaCl, 15% glycerol, and 1 mM phenylmethanesulfonyl fluoride (PMSF)], and lysed by sonication on ice using a probe sonicator (Xinzhi Biotech., Ningbo, China). The recombinant proteins were all in soluble form. Proteins were purified by using a HisTrap column (GE Healthcare, Uppsala, Sweden) with a linear gradient of 0–300 mM imidazole, and desalted by using a Centricon filter device (10 kD cut-off, Millipore, Ireland). The purity of recombinant protein was analyzed

ACCEPTED MANUSCRIPT

by 12% (w/v) sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). The concentration of protein was measured using a Pierce BCA protein assay kit (Thermo Scientific, Wilmington, DE).

PT

2.6. Activity assay

RI

The 1-chloro-2,4-dinitrobenzene (CDNB) and GSH were purchased from Sigma-Aldrich

SC

(St Louis, MO) and dissolved in ethanol and H2 O, respectively. The activity assay was performed according to a protocol described by Samra et al [11]. Briefly, a 200 μl reaction

NU

mixture contained 200 ng protein, 1 mM CDNB, 5 mM GSH, and 1% (v : v) ethanol in 0.1

MA

M sodium phosphate buffer (pH 6.5). The increases of absorbance were monitored at 15 s intervals at 340 nm for 3 min. The CDNB-conjugating activity (μmol/min/mg) was

ED

calculated using the molar extinction coefficient (ε340 = 9600 M−1 cm−1 ) of the resultant

EP T

2,4-dinitrophenyl- glutathione. The assays were biologically repeated three times, the absorbance was recorded on a Multiskan Go microplate reader (Thermo Scientific,

AC C

Wilmington, DE).

2.7. Reverse transcription-quantitative PCR (RT-qPCR) RT-qPCR was used to investigate the expression profiles of PrGSTs in various larval tissues, including integument, fat body, midgut, and Malpighian tubules. Tissues were dissected from more than 100 fourth- instar larvae and stored at −80°C until RNA extractions were carried out. Transcription profiles were also investigated in different developmental stages, including 300 eggs, 60 fourth- instar larvae, 60 pupae, and a mixture of 30 adult males

ACCEPTED MANUSCRIPT

and 30 adult females. RNA isolation was performed as described in Section 2.2, and first-strand cDNA was synthesized using the ReverTra Ace qPCR RT kit (Toyobo, Osaka, Japan). Each cDNA sample was diluted to 10 ng/μl with nuclease-free water. Primers for RT-qPCR are listed in Table S1. Two housekeeping genes (18S rRNA and

PT

β-actin) were used as internal references to normalise target gene expression. In a pre-test,

RI

the amplification efficiency (listed in Table S1) of each primer pair was calculated from the

SC

slope of the log- linear portion of the curve generated by amplification from serially diluted cDNA samples. RT-qPCR was performed in triplicate in 96-well reaction plates (Bio-Rad,

NU

Hercules, CA). Each reaction (20 μl volume) contained 10 μl SYBR Green Real-time PCR

MA

Master Mix (Toyobo, Osaka, Japan), 0.4 μl (0.2 μM) of each primer, 1 μl (10 ng) cDNA template, and 8.2 μl nuclease-free water. RT-qPCR was performed on a CFX96 Real-Time

ED

System (Bio-Rad, Hercules, CA) with the following parameters: one cycle at 95°C for 30 s,

EP T

and 40 cycles at 95°C for 5 s, and 60°C for 25 s. To confirm that only one single gene was detected by the fluorescence dye, a heat-dissociation protocol was added at the end of the

AC C

thermal cycle. A no-template control and a no-reverse transcriptase control were included on each reaction plate to detect potential contamination. Reactions for all samples were independently repeated three times, and quantitative variation in gene expression among different samples was calculated by a modified version of the Pfaffl method [34].

2.8. Insecticide treatment Abamectin, chlorantraniliprole and lambda-cyhalothrin were purchased from Dr Ehrenstorfer GmbH (Augsburg, Germany), the purity of these insecticides are all ≥ 94%.

ACCEPTED MANUSCRIPT

The three chemicals were diluted with analytical- grade acetone to produce a stock solution from which serial decimal dilutions were carried out. In previous study, three sublethal doses, LD5 , LD20 and LD50 (5%, 20% and 50% of the test animals are killed at 24 h, respectively), for each insecticide were determined and the values are listed in Table S3.

PT

The freshly molted (<24 h) fourth- instar larvae were used for the bioassay. Larvae were

RI

placed into a clean glass petri dish (9-cm diameter) containing a piece (5 × 5 cm) of fresh

SC

cabbage (B. pekinensis) leaves. A droplet of 1 μl of insecticide solution at each dose was applied topically on the dorsal part of larval middle abdomen with a 10 μl microsyringe

NU

(Gaoge Industry and Trade Co. Ltd., Shanghai, China). Control insects were treated with 1 μl

MA

of acetone only. The rearing conditions for treated larvae were controlled at 25 ± 1°C and 65% relative humidity. At 6 h after the treatment, surviving insects were collected for

ED

RT-qPCR analysis (as described in Section 2.7). Each treatment was biologically repeated

EP T

three times, and in each repeat 30 larvae were used.

AC C

2.9. Statistical analysis

Data were analysed using Data Processing System (DPS) software v9.5 [35]. Student's t-test and one-way analysis of variance (ANOVA) with Tukey's post-hoc test were performed, respectively, for comparing difference between two samples and differences among multiple samples. The level of significance was set at p <0.05.

3. Results 3.1. Identification and classification of P. rapae GSTs

ACCEPTED MANUSCRIPT

A total of 17 cDNA sequences encoding putative GSTs were identified from P. rapae transcriptome data and designated PrGSTd1–PrGSTz2 (Table 1). In order to confirm that the assembled PrGSTs were not chimeric, PCR analysis was performed and the results showed that all PrGSTs were amplified from larval cDNA with gene-specific primers (data not

PT

shown). DNA sequencing subsequently confirmed that the sequences of amplified PrGSTs

RI

were identical to those retrieved from the transcriptome data. All PrGST cDNAs included a

SC

complete ORF, and the size of deduced PrGST proteins ranged from 203 to 287 amino acid residues (Table 1), sharing between 12% and 76% amino acid sequence identity (Table S4).

NU

A BLASTX search of the best hits showed that PrGSTs share relatively high sequence

MA

identity (56–95%) with their respective orthologs from other lepidopteran species (Table 1). The results of Conserved Domain revealed G-sites in the N-terminal regions in 10 of the

ED

PrGSTs but not in PrGSTE1, PrGSTE2, PrGSTO1, PrGSTO2, PrGSTO3, PrGSTO4 and

EP T

PrGSTZ2, and H-sites in the C-terminal region of all PrGSTs (Fig. 1). Phylogenetic analysis was performed to better understand the classification of PrGSTs.

AC C

The 17 PrGSTs clearly segregated into six classes (delta, epsilon, omega, sigma, theta and zeta) (Fig. 2). Based on this classification, P. rapae possesses three delta-class GSTs, three epsilon GSTs, four omega GSTs, four sigma GSTs, one theta GST, and two zeta GSTs (Fig. 2; Table 1).

3.2. Exon-intron organization of PrGSTs The position of exons and introns in each PrGST gene was analysed by aligning a cDNA sequence with a genomic DNA sequence (Fig. 3). The size of one intron in PrGSTe1 gene

ACCEPTED MANUSCRIPT

was not identified due to gaps in the genomic DNA (Fig. 3). A total of 64 introns were found in the 17 PrGST genes. No intronless gene was observed. The donor- and acceptor-sites of these introns all conformed to the classical GT–AG splice junction rules (data not shown). Most PrGST genes have three or four introns, whereas PrGSTd1 and PrGSTo3 both have

PT

five introns (Fig. 3). The PrGSTd3 and PrGSTe2 possessed the smallest (two) and the largest

SC

RI

(six) numbers of introns, respectively (Fig. 3).

3.3. Comparison of GSTs in P. rapae and other insects

NU

The number of annotated GSTs in P. rapae and other representative insect species

MA

belonging to different Orders (Lepidoptera, Diptera, Coleoptera, Hymenoptera, Hemiptera and Orthoptera) are listed in Table 2. This number clearly varies greatly between species,

ED

with a particularly large number (27 to 39 genes) in Diptera, Coleoptera and Orthoptera, but

EP T

far fewer in the hymenopteran A. mellifera (10 genes) and the hemipteran N. lugens (11 genes) (Table 2). Within the Order Lepidoptera, P. rapae (17 genes) possesses fewer GSTs

AC C

than P. xylostella (22 genes), B. mori (24 genes), C. medinalis (25 genes) and S. litura (37 genes) (Table 2). There are fewer delta (three) and epsilon (three) class GSTs in P. rapae than in other lepidopteran species (Table 2).

3.4. Functional characterisation of recombinant PrGST proteins Recombinant proteins of eight PrGSTs (PrGSTD1, PrGSTD2, PrGSTE1, PrGSTE2, PrGSTO1, PrGSTS1, PrGSTT1, and PrGSTZ1) were heterologously expressed in E. coli as histidine- fusion proteins (Fig. S1). The CDNB-conjugating activities of these proteins were

ACCEPTED MANUSCRIPT

characterised in vitro (Table 3). The results showed that one delta-class GST, PrGSTD2, had the highest activity (4.61 μmol/min/mg) towards CDNB than other PrGSTs; whereas a theta-class GST, PrGSTT1, exhibited the lowest activity (0.028 μmol/min/mg) to catalyse the conjugation of CDNB. The activities of other PrGSTs ranged from 0.49 to 1.57

RI

PT

μmol/min/mg (Table 3).

SC

3.5. Expression of PrGSTs in different larval tissues and at different developmental stages Using RT-qPCR, expression profiles of PrGSTs were investigated in four different larval

NU

tissues (integument, fat body, midgut and Malpighian tubules). The results showed that four

MA

genes (PrGSTe3, PrGSTs1, PrGSTs2 and PrGSTs4) were mainly transcribed in the fat body; only one gene, PrGSTe2, was mainly expressed in Malpighian tubules (Fig. 4). None of the

ED

genes were specifically distributed in the integument or midgut (Fig. 4). The remaining

EP T

genes were ubiquitously expressed in all the tested tissues or in at least two tissues. For example, PrGSTd1 was expressed predominantly in both the fat body and midgut, whilst

AC C

PrGSTo2 was omnipresently transcribed in all the four tissues examined (Fig. 4). Expression patterns of PrGSTs at different developmental stages were also determined, and four genes (PrGSTe2, PrGSTo4, PrGSTs4 and PrGSTt1) were mainly expressed during the fourth- instar larval stage (Fig. 5). However, no genes were predominantly expressed in egg, pupal or adult stages. Some genes were mainly expressed in two or three developmental stages, such as PrGSTd1, which was predominantly expressed in larval, pupal and adult stages, while PrGSTe1 was transcribed mainly in larval and pupal stages (Fig. 5).

ACCEPTED MANUSCRIPT

3.6. Expression of PrGSTs in larvae exposed to insecticides To investigate whether the expression of PrGSTs respond to the synthetic insecticides abamectin, chlorantraniliprole and lambda-cyhalothrin, PrGST transcript levels were determined in larvae following exposure to sublethal doses (LD5 , LD20 , and LD50 ) of these

PT

insecticides. The result revealed that the expression of eight genes (PrGSTd1, PrGSTd3,

RI

PrGSTe1, PrGSTe2, PrGSTo1, PrGSTo3, PrGSTs1, PrGSTs3 and PrGSTs4) was

SC

significantly upregulated by LD50 dose of abamectin, compared with control insects (Fig. 6A). The LD5 and LD20 doses of abamectin also significantly elevated five (PrGSTe1,

NU

PrGSTe2, PrGSTs1, PrGSTs3 and PrGSTs4) and eight (PrGSTd3, PrGSTe1, PrGSTe2,

MA

PrGSTo3, PrGSTs1, PrGSTs3, PrGSTs4 and PrGSTt1) genes, respectively (Fig. 6A). Similarly, PrGST genes were significantly upregulated by chlorantraniliprole and

ED

lambda-cyhalothrin; specifically, chlorantraniliprole significantly stimulated PrGSTd1,

EP T

PrGSTd2, PrGSTe1, PrGSTe2, PrGSTe3, PrGSTs1, PrGSTs3, PrGSTs4 and PrGSTz1 (Fig. 6B), whilst lambda-cyhalothrin significantly upregulated PrGSTd1, PrGSTd3, PrGSTe1,

AC C

PrGSTe2, PrGSTo1, PrGSTo2, PrGSTs1, PrGSTs2, PrGSTs3 and PrGSTz2 (Fig. 6C). Conversely, three genes (PrGSTd2, PrGSTo4 and PrGSTz1) were significantly downregulated after exposure to different concentrations of abamectin (Fig. 6A). Similarly, chlorantraniliprole downregulated PrGSTd3, PrGSTo1, PrGSTs2, PrGSTt1 and PrGSTz2, while lambda-cyhalothrin decreased the mRNA levels of PrGSTe3, PrGSTs4, PrGSTt1 and PrGSTz1 (Fig. 6B, C).

ACCEPTED MANUSCRIPT

4. Discussion Since whole-genome information for P. rapae is still unavailable, searching of transcriptome datasets can be used to identify specific genes, including GST genes. This approach has been used successfully for other insects lacking genomic data, such as S. litura

PT

[21], Laodelphax striatellus [36], C. medinalis [22], and Bactrocera dorsalis [37]. In the

RI

present study, we identified 17 PrGST genes, which is greater than the number present in A.

SC

mellifera [38], L. striatellus [36], and N. lugens [15], but less than the number in some other insects, especially lepidopteran species. For example, there are 22, 24, 25 and 37 GST genes

NU

in P. xylostella, B. mori, C. medinalis and S. litura, respectively [21-24]. Although we

MA

performed exhaustive searches of the available transcriptome data, it is of course possible that additional GST genes may be present in P. rapae that were not detected in this study.

ED

Based on phylogenetic analysis, the 17 PrGSTs were divided into six classes (delta,

EP T

epsilon, omega, sigma, theta and zeta) (Fig. 2). Delta and epsilon are insect-specific classes, and GSTs from these two classes often have detoxification functions and have been related to

AC C

resistance to insecticides [3-4]. For instance, delta GSTs from Culex pipiens and Cydia pomonella are able to metabolise 1,1,1-trichloro-2,2-bis(4-chlorophenyl)ethane (DDT) and lambda-cyhalothrin, respectively [11-12]. Furthermore, silencing of epsilon GSTs in Aedes aegypti and B. dorsalis by RNAi increases susceptibility to pyrethroid and malathion, respectively [13, 39]. P. rapae has fewer delta and epsilon GSTs compared with other lepidopteran species such as P. xylostella, B. mori, C. medinalis and S. litura. This is probably due to within-species variation, and the deficit of delta and epsilon GSTs may be functionally compensated by GSTs from other classes.

ACCEPTED MANUSCRIPT

To date, GSTs from many different insect species have been heterologously expressed and purified [8-9, 11-12]. Furthermore, the biochemical characteristics of these proteins have been analysed in vitro, and the results showed that the catalytic properties of insect GSTs varied greatly. For instance, an omega GST (AccGSTO1) from Apis cerana cerana

PT

displayed low activity for CDNB (0.015 μmol/min/mg) [40], whilst a delta GST (CpGSTD1)

RI

from Culex pipiens had strong ability to conjugate CDNB (> 40 μmol/min/mg) [11]. In the

SC

current study, recombinant proteins of eight PrGSTs were produced in the E. coli expression system and all of them displayed catalytic activities towards CDNB in the presence of GSH

NU

(Table 3). The result indicated that these PrGSTs were functional proteins and might play a

MA

significant role in xenobiotic detoxification. Although biochemical study was not performed for other PrGSTs, they were expected to have catalytic functions and might be involved in

ED

important physiological pathways.

EP T

The insect fat body, midgut, and Malpighian tubules are well-known for their critical roles in the detoxification of xenobiotic compounds [41-43]. In many lepidopteran species

AC C

such as B. mori, P. xylostella and C. medinalis, several GST genes are highly expressed in these tissues [22-24]. In this study, we found four genes (PrGSTe3, PrGSTs1, PrGSTs2, and PrGSTs4) that were expressed predominantly in the fat body, and one gene (PrGSTe2) that was mainly transcribed in Malpighian tubules (Fig. 4). These PrGSTs are likely associated with the detoxification of xenobiotics. Furthermore, we found several genes that were at a moderate expression level in integument (Fig. 4). Transcription of GST genes in integument has been reported in many insect species including A. cerana cerana, S. litura, and Bombus ignitus [8, 44-45], implying a possible role of GSTs in the initial detoxification of xenobiotic

ACCEPTED MANUSCRIPT

compounds. Other PrGST genes were ubiquitously expressed in various tissues, indicating a basic but presumably important physiological function in P. rapae. In many insect species, GST mRNA levels vary greatly during different life stages. For instance, GSTs1 in P. xylostella is highly expressed in egg and larval stages, but not in pupal

PT

and adult stages [24]. By contrast, a sigma-class GST gene in Chilo suppressalis is expressed

RI

most highly in adults [46]. In the present study, four PrGSTs (PrGSTe2, PrGSTo4, PrGSTs4

SC

and PrGSTt1) were predominantly expressed during the fourth- instar larval stage (Fig. 5), suggesting that they might be related to the detoxification of secondary metabolites from

NU

host plants and insecticides. Several GST genes in N. lugens, L. striatellus, S. furcifera, P.

MA

xylostella and B. dorsalis are enriched during the egg stage and might regulate hormones throughout embryonic development [15, 24, 36-37]. However, no egg stage-specific GSTs

ED

were identified in P. rapae.

EP T

Abamectin, chlorantraniliprole and lambda-cyhalothrin have been widely used to control a variety of insect pests, including P. rapae [26]. Recently, variation in susceptibility to these

AC C

insecticides was observed in N. lugens [9], P. xylostella [47-48], S. exigua [49], C. medinalis [50], C. suppressalis [51-52] and Apolygus lucorum [53]. Although mutations in some key proteins such as the glutamate-gated chloride channel, ryanodine receptor, and voltage- gated sodium channel can contribute to reduced susceptibility [53-55], detoxifying enzymes such as GSTs, esterases and cytochrome P450s may also play important roles in the detoxification insecticides [12, 56-59]. Additionally, insect GSTs may also be involved in protecting against the oxidative stress induced by insecticides [8-9]. In this study, ten, eight, and ten PrGSTs were upregulated following exposure to chlorantraniliprole, abamectin, and

ACCEPTED MANUSCRIPT

lambda-cyhalothrin, respectively (Fig. 6). These genes are therefore potential candidates involved in the detoxification of the tested insecticides. Downregulation of PrGSTs was also observed in larvae treated with insecticides (Fig. 6). Decreased GST mRNA levels following insecticide exposure has also been reported for

PT

many other insect species such as N. lugens, L. striatellus, S. furcifera, B. dorsalis, C.

RI

medinalis and L. decemlineata [15, 20, 22, 36-37]. Han et al [20] suggested that

SC

downregulation of a subset of GSTs may be an adaptive mechanism in insects that reduces the total activity of GST enzymes to prevent excessive GST activity from exhausting the

NU

supply of GSH. However, functional studies are needed to test this hypothesis.

MA

In conclusion, we identified 17 GST genes in P. rapae by analysing previously published transcriptome data. The exon-intron organizations and phylogenetic relationships of PrGSTs

ED

were investigated, and the catalytic activities of eight PrGST proteins were also examined.

EP T

PrGST genes displayed distinct expression profiles in various larval tissues and during different life stages. The transcription of several genes was induced following exposure to

AC C

abamectin, chlorantraniliprole and lambda-cyhalothrin, consistent with an involvement in the detoxification of insecticides. This is the first report on the large-scale identification and characterisation of GST genes in P. rapae. The results will pave the way for a better understanding of detoxification enzymes in this insect species.

Conflict of interest The authors declare that there is no conflict of interest in this paper.

ACCEPTED MANUSCRIPT

Acknowledgements This work was supported by the National Key Research and Development Program of China (grant number 2016YFD0200205-7), the National Natural Science Foundation of China (grant number 31401734), and the Anhui Provincial Natural Science Foundation

RI

PT

(grant number 1708085QC50).

References

SC

[1] J.D. Hayes, J.U. Flanagan, I.R. Jowsey, Glutathione transferases, Annu. Rev. Pharmacol. Toxicol. 45

NU

(2005) 51–88.

[2] A.J. Ketterman, C. Saisawang, J. Wongsantichon, Insect glutathione transferases, Drug Metab. Rev. 43

MA

(2011) 253–265.

[3] A.A. Enayati, H. Ranson, J. Hemingway, Insect glutathione transferases and insecticide resistance, Insect Mol. Biol. 14 (2005) 3–8.

ED

[4] X. Li, M.A. Schuler, M.R. Berenbaum, Molecular mechanisms of metabolic resistance to synthetic and natural xenobiotics, Annu. Rev. Entomol. 52 (2007) 231–253.

EP T

[5] M.E. Rogers, M.K. Jani, R.G. Vogt, An olfactory-specific glutathione-S-transferase in the sphinx moth Manduca sexta, J. Exp. Biol. 202 (1999) 1625–1637. [6] S. Enya, T. Daimon, F. Igarashi, H. Kataoka, M. Uchibori, H. Sezutsu, et al., The silkworm glutathione

AC C

S-transferase gene noppera-bo is required for ecdysteroid biosynthesis and larval development, Insect Biochem. Mol. Biol. 61 (2015) 1–7. [7] S. Enya, T. Ameku, F. Igarashi, M. Iga, H. Kataoka, T. Shinoda, et al., A Halloween gene noppera-bo encodes a glutathione S-transferase essential for ecdysteroid biosynthesis via regulating the behaviour of cholesterol in Drosophila, Sci. Rep. 4 (2014) 6586. [8] H. Yan, F. Meng, H. Jia, X. Guo, B. Xu, The identification and oxidative stress response of a zeta class glutathione S-transferase (GSTZ1) gene from Apis cerana cerana, J. Insect Physiol. 58 (2012) 782– 791. [9] J.G. Vontas, G.J. Small, J. Hemingway, Glutathione S-transferases as antioxidant defence agents

ACCEPTED MANUSCRIPT confer pyrethroid resistance in Nilaparvata lugens, Biochem. J. 357 (2001) 65–72. [10] Y. Ding, F. Ortelli, L. Rossiter, J. Hemingway, H. Ranson, The Anopheles gambiae glutathione transferase supergene family: annotation, phylogeny and expression profiles, BMC Genomics 4 (2003) 35. [11] A.I. Samra, S.G. Kamita, H.-W. Yao, A.J. Cornel, B.D. Hammock, Cloning and characterization of

PT

two glutathione S-transferases from pyrethroid-resistant Culex pipiens, Pest Manag. Sci. 68 (2012) 764–772.

RI

[12] J. Liu, X. Yang, Y. Zhang, Characterization of a lambda-cyhalothrin metabolizing glutathione S-transferase CpGSTd1 from Cydia pomonella (L.), Appl. Microbiol. Biotechnol. 98 (2014) 8947–

SC

8962.

NU

[13] N. Lumjuan, S. Rajatileka, D. Changsom, J. Wicheer, P. Leelapat, L.-a. Prapanthadara, et al., The role of the Aedes aegypti epsilon glutathione transferases in conferring resistance to DDT and pyrethroid

MA

insecticides, Insect Biochem. Mol. Biol. 41 (2011) 203–209.

[14] G. Qin, M. Jia, T. Liu, X. Zhang, Y. Guo, K.Y. Zhu, et al., Characterization and functional analys is of four glutathione S-transferases from the migratory locust, Locusta migratoria, PLoS ONE 8 (2013)

ED

e58410.

[15] W.-W. Zhou, Q.-M. Liang, Y. Xu, G.M. Gurr, Y.-Y. Bao, X.-P. Zhou, et al., Genomic insights into the

EP T

glutathione S-transferase gene family of two rice planthoppers, Nilaparvata lugens (Stål) and Sogatella furcifera (Horváth) (Hemiptera: Delphacidae), PLoS ONE 8 (2013) e56604. [16] X. Yang, C. He, W. Xie, Y. Liu, J. Xia, Z. Yang, et al., Glutathione S-transferases are involved in

AC C

thiamethoxam resistance in the field whitefly Bemisia tabaci Q (Hemiptera: Aleyrodidae), Pestic. Biochem. Physiol. 134 (2016) 73–78. [17] C. Keeling, M. Yuen, N. Liao, T.R. Docking, S. Chan, G. Taylor, et al., Draft genome of the mountain pine beetle, Dendroctonus ponderosae Hopkins, a major forest pest, Genome Biology. 14 (2013) R27. [18] X. Wang, X. Fang, P. Yang, X. Jiang, F. Jiang, D. Zhao, et al., The locust genome provides insight into swarm formation and long-distance flight, Nat. Commun. 5 (2014) 2957. [19] R. Schama, N. Pedrini, M.P. Juárez, D.R. Nelson, A.Q. Torres, D. Valle, et al., Rhodnius prolixus supergene families of enzymes potentially associated with insecticide resistance, Insect. Biochem.

ACCEPTED MANUSCRIPT Mol. Biol. 69 (2016) 91–104. [20] J.-B. Han, G.-Q. Li, P.-J. Wan, T.-T. Zhu, Q.-W. Meng, Identification of glutathione S-transferase genes in Leptinotarsa decemlineata and their expression patterns under stress of three insecticides, Pestic. Biochem. Physiol. 133 (2016) 26–34. [21] N. Zhang, J. Liu, S.-N. Chen, L.-H. Huang, Q.-L. Feng, S.-C. Zheng, Expression profiles of

PT

glutathione S-transferase superfamily in Spodoptera litura tolerated to sublethal doses of chlorpyrifos, Insect Sci. 23 (2016) 675–687.

RI

[22] S. Liu, X.-J. Rao, M.-Y. Li, M.-F. Feng, M.-Z. He, S.-G. Li, Glutathione S-transferase genes in the rice leaffolder, Cnaphalocrocis medinalis (Lepidoptera: Pyralidae): identification and expression

SC

profiles, Arch. Insect Biochem. 90 (2015) 1–13.

NU

[23] Q. Yu, C. Lu, B. Li, S. Fang, W. Zuo, F. Dai, et al., Identification, genomic organization and expression pattern of glutathione S-transferase in the silkworm, Bombyx mori, Insect Biochem. Mol.

MA

Biol. 38 (2008) 1158–1164.

[24] Y. You, M. Xie, N. Ren, X. Cheng, J. Li, X. Ma, et al., Characterization and expression profiling of glutathione S-transferases in the diamondback moth, Plutella xylostella (L.), BMC Genomics 16

ED

(2015) 152.

[25] D. Grzywacz, A. Rossbach, A. Rauf, D.A. Russell, R. Srinivasan, A.M. Shelton, Current control

EP T

methods for diamondback moth and other brassica insect pests and the prospects for improved management with lepidopteran-resistant Bt vegetable brassicas in Asia and Africa, Crop Prot. 29 (2010) 68–79.

AC C

[26] P.-s. Wang, Y. Lyu, J. Dong, L. Jing, Z.-q. Yuan, J.-g. Yang, et al., Control effect of 13 pesticides on Pieris rapae in the cauliflower field, Agrochemicals 56 (2017) 300–302. [27] L. Qi, Q. Fang, L. Zhao, H. Xia, Y. Zhou, J. Xiao, et al., De novo assembly and developmental transcriptome analysis of the small white butterfly Pieris rapae, PLoS ONE 11 (2016) e0159258. [28] S.F. Altschul, T.L. Madden, A.A. Schäffer, J. Zhang, Z. Zhang, W. Miller, et al., Gapped BLAST and PSI-BLAST: a new generation of protein database search programs, Nucleic Acids Res. 25 (1997) 3389–3402. [29] A. Marchler-Bauer, M. K. Derbyshire, N.R. Gonzales, S. Lu, F. Chitsaz, L.Y. Geer, et al., CDD: NCBI's conserved domain database, Nucleic Acids Res. 43 (2015) D222–D226.

ACCEPTED MANUSCRIPT [30] F. Sievers, A. Wilm, D. Dineen, T.J. Gibson, K. Karplus, W. Li, et al., Fast, scalable generation of high‐ quality protein multiple sequence alignments using Clustal Omega, Mol. Syst. Biol. 7 (2011) 539. [31] K. Tamura, G. Stecher, D. Peterson, A. Filipski, S. Kumar, MEGA6: Molecular Evolutionary Genetics Analysis version 6.0, Mol. Biol. Evol. 30 (2013) 2725–2729.

PT

[32] J. Shen, Q. Cong, L.N. Kinch, D. Borek, Z. Otwinowski, N.V. Grishin, Complete genome of Pieris rapae, a resilient alien, a cabbage pest, and a source of anti-cancer proteins, F1000Res. 5 (2016)

RI

2631.

[33] Y. Kapustin, A. Souvorov, T. Tatusova, D. Lipman, Splign: algorithms for computing spliced

SC

alignments with identification of paralogs, Biol. Direct 3 (2008) 20.

NU

[34] S. Liu, Z.-J. Gong, X.-J. Rao, M.-Y. Li, S.-G. Li, Identification of putative carboxylesterase and glutathione S-transferase genes from the antennae of the Chilo suppressalis (Lepidoptera: Pyralidae),

MA

J. Insect Sci. 15 (2015) 103.

[35] Q.-Y. Tang, C.-X. Zhang, Data Processing System (DPS) software with experimenta l design, statistical analysis and data mining developed for use in entomological research, Insect Sci. 20 (2013)

ED

254–260.

[36] W.-W. Zhou, X.-W. Li, Y.-H. Quan, J. Cheng, C.-X. Zhang, G. Gurr, et al., Identification and

EP T

expression profiles of nine glutathione S-transferase genes from the important rice phloem sap-sucker and virus vector Laodelphax striatellus (Fallén) (Hemiptera: Delphacidae), Pest Manag. Sci. 68 (2012) 1296–1305.

AC C

[37] F. Hu, W. Dou, J.-J. Wang, F.-X. Jia, J.-J. Wang, Multiple glutathione S-transferase genes: identification and expression in oriental fruit fly, Bactrocera dorsalis, Pest Manag. Sci. 70 (2014) 295–303.

[38] C. Claudianos, H. Ranson, R.M. Johnson, S. Biswas, M.A. Schuler, M.R. Berenbaum, et al., A deficit of detoxification enzymes: pesticide sensitivity and environmental response in the honeybee, Insect Mol. Biol. 15 (2006) 615–636. [39] X.-P. Lu, L.-L. Wang, Y. Huang, W. Dou, C.-T. Chen, D. Wei, et al., The epsilon glutathione S-transferases contribute to the malathion resistance in the oriental fruit fly, Bactrocera dorsalis (Hendel), Comp. Biochem. Physiol. C 180 (2016) 40–48.

ACCEPTED MANUSCRIPT [40] F. Meng, Y. Zhang, F. Liu, X. Guo, B. Xu, Characterization and mutational analysis of omega-class GST (GSTO1) from Apis cerana cerana, a gene involved in response to oxidative stress, PLoS ONE 9 (2014) e93100. [41] L. Després, J.-P. David, C. Gallet, The evolutionary ecology of insect resistance to plant chemicals, Trends Ecol. Evol. 22 (2007) 298–307.

PT

[42] J.A.T. Dow, S.A. Davies, The Malpighian tubule: rapid insights from post-genomic biology, J. Insect Physiol. 52 (2006) 365–378.

RI

[43] E.L. Arrese, J.L. Soulages, Insect fat body: energy, metabolism, and regulation, Annu. Rev. Entomol. 55 (2010) 207–225.

SC

[44] Y. Huang, Z. Xu, X. Lin, Q. Feng, S. Zheng, Structure and expression of glutathione S-transferase

NU

genes from the midgut of the common cutworm, Spodoptera litura (Noctuidae) and their response to xenobiotic compounds and bacteria, J. Insect Physiol. 57 (2011) 1033–1044.

MA

[45] B. Y. Kim, W.L. Hui, K.S. Lee, H. Wan, H.J. Yoon, Z.Z. Gui, et al., Molecular cloning and oxidative stress response of a sigma-class glutathione S-transferase of the bumblebee Bombus ignitus, Comp. Biochem. Physiol. B 158 (2011) 83–89.

ED

[46] J. Huang, S. Wu, G. Ye, Molecular characterization of the sigma class glutathione S-transferase from Chilo suppressalis and expression analysis upon bacterial and insecticidal challenge, J. Econ.

EP T

Entomol. 104 (2011) 2046–2053.

[47] X. Pu, Y. Yang, S. Wu, Y. Wu, Characterisation of abamectin resistance in a field-evolved multiresistant population of Plutella xylostella, Pest Manag. Sci. 66 (2010) 371–378.

AC C

[48] X. Wang, X. Li, A. Shen, Y. Wu, Baseline susceptibility of the diamondback moth (Lepidoptera: Plutellidae) to chlorantraniliprole in China, J. Econ. Entomol. 103 (2010) 843–848. [49] T. Lai, J. Li, J. Su, Monitoring of beet armyworm Spodoptera exigua (Lepidoptera: Noctuidae) resistance to chlorantraniliprole in China, Pestic. Biochem. Physiol. 101 (2011) 198–205. [50] S.-K. Zhang, X.-B. Ren, Y.-C. Wang, J. Su, Resistance in Cnaphalocrocis medinalis (Lepidoptera: Pyralidae) to new chemistry insecticides, J. Econ. Entomol. 107 (2014) 815–820. [51] Y.P. He, C.F. Gao, W.M. Chen, L.Q. Huang, W.J. Zhou, X.G. Liu, et al., Comparison of dose responses and resistance ratios in four populations of the rice stem borer, Chilo suppressalis (Lepidoptera: Pyralidae), to 20 insecticides, Pest Manag. Sci. 64 (2008) 308–315.

ACCEPTED MANUSCRIPT [52] C. Gao, R. Yao, Z. Zhang, M. Wu, S. Zhang, J. Su, Susceptibility baseline and chlorantraniliprole resistance monitoring in Chilo suppressalis (Lepidoptera: Pyralidae), J. Econ. Entomol. 106 (2013) 2190–2194. [53] C. Zhen, X. Gao, A point mutation (L1015F) of the voltage-sensitive sodium channel gene associated with lambda-cyhalothrin resistance in Apolygus lucorum (Meyer–Dür) population from the transgenic

PT

Bt cotton field of China, Pestic. Biochem. Physiol. 127 (2016) 82–89. [54] B. Troczka, C.T. Zimmer, J. Elias, C. Schorn, C. Bass, T.G.E. Davies , et al., Resistance to diamide

RI

insecticides in diamondback moth, Plutella xylostella (Lepidoptera: Plutellidae) is associated with a mutation in the membrane-spanning domain of the ryanodine receptor, Insect Biochem. Mol. Biol. 42

SC

(2012) 873–880.

NU

[55] X. Wang, R. Wang, Y. Yang, S. Wu, A.O. O'Reilly, Y. Wu, A point mutation in the glutamate-gated chloride channel of Plutella xylostella is associated with resistance to abamectin, Insect Mol. Biol. 25

MA

(2016) 116–125.

[56] W.-H. Jiang, W.-P. Lu, W.-C. Guo, Z.-H. Xia, W.-J. Fu, G.-Q. Li, Chlorantraniliprole susceptibility in Leptinotarsa decemlineata in the north Xinjiang Uygur Autonomous Region in China, J. Econ.

ED

Entomol. 105 (2012) 549–554.

[57] M. Riga, D. Tsakireli, A. Ilias, E. Morou, A. Myridakis, E.G. Stephanou, et al., Abamectin is

EP T

metabolized by CYP392A16, a cytochrome P450 associated with high levels of acaricide resistance in Tetranychus urticae, Insect Biochem. Mol. Biol. 46 (2014) 43–53. [58] S.H.P.P. Karunaratne, N.J. Hawkes, M.D.B. Perera, H. Ranson, J. Hemingway, Mutated sodium

AC C

channel genes and elevated monooxygenases are found in pyrethroid resistant populations of Sri Lankan malaria vectors, Pestic. Biochem. Physiol. 88 (2007) 108–113. [59] X. Chen, L. Yuan, Y. Du, Y. Zhang, J. Wang, Cross-resistance and biochemical mechanisms of abamectin resistance in the western flower thrips, Frankliniella occidentalis, Pestic. Biochem. Physiol. 101 (2011) 34–38.

ACCEPTED MANUSCRIPT

Figure captions Fig. 1. Predicted GSH binding sites (G-sites) and substrate binding sites (H-sites) of deduced PrapGST proteins.

PT

Fig. 2. Phylogenetic analysis of GSTs from Acyrthosiphon pisum (Ap-prefix), Anopheles

RI

gambiae (Ag), Apis mellifera (Am), Bactrocera dorsalis (Bd), Bombyx mori (Bm),

SC

Cnaphalocrocis medinalis (Cm), Drosophila melanogaster (Dm), Leptinotarsa decemlineata (Ld), Locusta migratoria (Lm), Nilaparvata lugens (Nl), Plutella xylostella (Px), Pieris

NU

rapae (Pr), Spodoptera litura (Sl), and Tribolium castaneum (Tc). Bootstrap support values

MA

based on 1000 replicates are indicated by colouring from green (0) to red (100) on each node. Insect GSTs are classified into six classes (delta, epsilon, omega, sigma, theta and

ED

zeta) and an 'unclassified' subgroup. The seventeen P. rapae GSTs (PrGSTs) are coloured

EP T

red. GenBank accession numbers of sequences used are listed in Table S2.

AC C

Fig. 3. Schematic diagram of the exon-intron structure of each PrGST gene.

Fig. 4. Relative expression levels of PrGSTs in various larval tissues. IG, integument; FB, fat body; MG, midgut; MT, Malpighian tubes. Expression levels in different tissues were normalized relative to that in the integument. Different lowercase letters indicate significant variation in transcription among different samples (one-way ANOVA with Tukey's post-hoc test, p < 0.05).

ACCEPTED MANUSCRIPT

Fig. 5. Relative expression levels of PrGSTs during different life stages. E, egg stage; L, fourth- instar larval stage; P, pupal stage; A, adult stage. Expression levels at different life stages were normalized relative to that in the egg stage. Different lowercase letters indicate significant variation in transcription among different samples (one-way ANOVA with

RI

PT

Tukey's post-hoc test, p < 0.05).

SC

Fig. 6. Relative expression levels of PrGSTs in larvae exposed to three different concentrations (LD5 , LD20 , and LD50 ) of (A) abamectin, (B) chlorantraniliprole, and (C)

NU

lambda-cyhalothrin. The transcriptional level of each gene in insecticide-treated individuals

MA

was normalized relative to that in acetone-treated (control) individuals. * and ** indicate

AC C

EP T

ED

significant upregulation and downregulation, respectively (Student's t-test, p < 0.05).

SC

RI

PT

ACCEPTED MANUSCRIPT

AC C

EP T

ED

MA

NU

Figure 1

AC C

Figure 2

EP T

ED

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC C

EP T

ED

MA

Figure 3

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC C

EP T

ED

MA

Figure 4

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC C

EP T

ED

MA

Figure 5

AC C

Figure 6

EP T

ED

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Table 1 Details of the 17 GSTs identified in P. rapae. Group Gene name

GenBank ORF

Mw

acc. no.

(kDa)

(aa)

pI

BLASTX top hit Identity Reference organism

Acc. no.

E-value

PrGSTd1KX229707 244

6.2

27.4 Amyelois transitella

XP_0131923911E-128 71

PrGSTd2KX229708 216

6.2

24.4 Danaus plexippus

EHJ63706

PrGSTd3KX229723 214

6.6

24.3 Papilio polytes

NP_0012983812E-95

Epsilon PrGSTe1 KX229709 225

8.7

26.3 Papilio xuthus

KPJ03136

2E-109 68

PrGSTe2 KX229710 253

6.3

28.7

PT

(%) Delta

AIZ46901

6E-109 61

P46430

2E-93

63

6.1

24.6 Manduca sexta

Omega PrGSTo1 KX229712 254

6.1

28.9 Danaus plexippus

EHJ65985

3E-137 72

PrGSTo2 KX229713 287

8.8

33.3 Bombyx mori

ABD36306

2E-110 56

PrGSTo3 KX229714 241

7.1

28.8 Chilo suppressalis

AKS40347

1E-145 81

PrGSTo4 KX229715 275

8.2

31.7 Danaus plexippus

EHJ65985

6E-105 57

Sigma PrGSTs1 KX229716 204

5.8

23.2 Amyelois transitella

XP_0131955713E-101 68

PrGSTs2 KX229717 205

6.8

23.4 Bombyx mori

NP_0010369943E-86

PrGSTs3 KX229718 203

9.0

23.6

AAF23078

NU

SC

PrGSTe3 KX229711 215

MA

medinalis

RI

Cnaphalocrocis

9E-131 89

60

61

PrGSTs4 KX229719 206

ED

Choristoneura 4E-103 70

fumiferana

8.5

23.6 Papilio machaon

XP_0143644454E-81

59

7.0

26.5 Danaus plexippus

EHJ70012

5E-106 65

PrGSTt1 KX229720 224

Zeta

PrGSTz1 KX229721 219

7.1

25.1 Chilo suppressalis

AKS40351

2E-148 95

PrGSTz2 KX229722 219

9.4

25.3 Spodoptera litura

AIH07599

4E-99

AC C

EP T

Theta

65

ACCEPTED MANUSCRIPT

Table 2 Comparison of GST genes in various insect species. Species

Delta Epsilon Omega Sigma Theta Zeta Unclassified Microsomal Total

Lepidoptera Pieris rapae 3 Plutella xylostella 5

2 2

0 2

– –

17 22

8

4

2

1

2

2

1

24

Cnaphalocrocis medinalis

4

9

3

5

0

2

2



25

Spodoptera litura 4 Drosophila 11

15 14

3 5

6 1

1 4

2 2

1 0

5 1

37 38

melanogaster Anopheles

12

8

1

1

2

1

3

3

31

gambiae Aedes aegypti

8

8

1

1

4

1

3

1

27

Culex

17

10

1

2

6

0

3

0

39

quinquefasciatus Tribolium

3

19

4

7

1

1

0

1

36

castaneum Dendroctonus ponderosae

6

12

2

2

1

0



28

Leptinotarsa

3

10

1

0

5

SC

RI

PT

4

5

4

4

1

2

1

30

1

4

1

1

0

2

10

5

0

2

8

3

1

0

0

19

vitripennis Acyrthosiphon

10

0

0

6

2

0

0

2

20

1

1

3

1

1

0

2

11

0

3

12

2

1

0

0

28

EP T

Nasonia

pisum Nilaparvata lugens 2 Locusta migratoria

AC C

Orthoptera

1 1

Bombyx mori

decemlineata Hymenoptera Apis mellifera

Hemiptera

4 2

NU

Coleoptera

4 5

ED

Diptera

3 5

MA

Order

10

Data are collated from [15, 17, 18, 20-24]. "–" denotes data are not shown in the literature (P. xylostella, C. medinalis and D. ponderosae), or that transcriptome searching was not performed (P. rapae).

ACCEPTED MANUSCRIPT

Table 3 The CDNB-conjugating activities of eight recombinant PrGST proteins. Activity (μmol/min/mg)

PrGSTD1 PrGSTD2

0.76 ± 0.08 4.61 ± 0.35

PrGSTE1 PrGSTE2

0.92 ± 0.05 1.57 ± 0.2

PrGSTO1 PrGSTS1

0.49 ± 0.07 0.93 ± 0.09

PrGSTT1

0.028 ± 0.006

PrGSTZ1

1.2 ± 0.15

AC C

EP T

ED

MA

NU

SC

RI

PT

Protein

ACCEPTED MANUSCRIPT

AC C

EP T

ED

MA

NU

SC

RI

PT

Graphical abstract

ACCEPTED MANUSCRIPT

Highlights A total of 17 GST genes (PrGSTs) were identified in Pieris rapae



Eight PrGST proteins were functionally characterised



Some PrGSTs display tissue- and developmental stage-specific expression



Several PrGST genes are significantly upregulated by insecticides



PrGSTs are potentially involved in insecticide detoxification

AC C

EP T

ED

MA

NU

SC

RI

PT