Skip to main content
Advertisement
  • Loading metrics

Resistance to Bacillus thuringiensis Cry1Ac toxin requires mutations in two Plutella xylostella ATP-binding cassette transporter paralogs

  • Zhaoxia Liu,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Resources, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Shu Fu,

    Roles Formal analysis, Investigation, Methodology, Resources, Software

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Xiaoli Ma,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Resources

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China, Xinjiang Key Laboratory of Biological Resources and Genetic Engineering, College of Life Science and Technology, Xinjiang University, Urumqi, China

  • Simon W. Baxter,

    Roles Writing – review & editing

    Affiliations Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China, School of BioSciences, The University of Melbourne, Melbourne, Victoria, Australia

  • Liette Vasseur,

    Roles Writing – review & editing

    Affiliations Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China, Department of Biological Sciences, Brock University, St. Catharines, Ontario, Canada

  • Lei Xiong,

    Roles Formal analysis, Investigation, Methodology, Resources

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Yuping Huang,

    Roles Formal analysis, Investigation, Methodology, Resources

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Guang Yang,

    Roles Conceptualization, Funding acquisition, Writing – review & editing

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Shijun You ,

    Roles Conceptualization, Funding acquisition, Project administration, Supervision, Writing – review & editing

    sjyou@fafu.edu.cn (SY); msyou@fafu.edu.cn (MY)

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

  • Minsheng You

    Roles Conceptualization, Funding acquisition, Project administration, Supervision, Writing – review & editing

    sjyou@fafu.edu.cn (SY); msyou@fafu.edu.cn (MY)

    Affiliations State Key Laboratory of Ecological Pest Control for Fujian-Taiwan Crops and College of Life Science, Fujian Agriculture and Forestry University, Fuzhou, China, Institute of Applied Ecology, Fujian Agriculture and Forestry University, Fuzhou, China, Joint International Research Laboratory of Ecological Pest Control, Ministry of Education, Fuzhou, China

Abstract

The diamondback moth, Plutella xylostella, is a cosmopolitan pest and the first species to develop field resistance to toxins from the gram-positive bacterium Bacillus thuringiensis (Bt). Although previous work has suggested that mutations of ATP-binding cassette transporter subfamily C2 (ABCC2) or C3 (ABCC3) genes can confer Cry1Ac resistance, here we reveal that P. xylostella requires combined mutations in both PxABCC2 and PxABCC3 to achieve high-level Cry1Ac resistance, rather than simply a mutation of either gene. We identified natural mutations of PxABCC2 and PxABCC3 that concurrently occurred in a Cry1Ac-resistant strain (Cry1S1000) of P. xylostella, with a mutation (RA2) causing the mis-splicing of PxABCC2 and another mutation (RA3) leading to the premature termination of PxABCC3. Genetic linkage analysis showed that RA2 and RA3 were tightly linked to Cry1Ac resistance. Introgression of RA2 and RA3 enabled a susceptible strain (G88) of P. xylostella to obtain high resistance to Cry1Ac, confirming that these genes confer resistance. To further support the role of PxABCC2 and PxABCC3 in Cry1Ac resistance, frameshift mutations were introduced into PxABCC2 and PxABCC3 singly and in combination in the G88 strain with CRISPR/Cas9 mediated mutagenesis. Bioassays of CRISPR-based mutant strains, plus genetic complementation tests, demonstrated that the deletion of PxABCC2 or PxABCC3 alone provided < 4-fold tolerance to Cry1Ac, while disruption of both genes together conferred >8,000-fold resistance to Cry1Ac, suggesting the redundant/complementary roles of PxABCC2 and PxABCC3. This work advances our understanding of Bt resistance in P. xylostella by demonstrating mutations within both PxABCC2 and PxABCC3 genes are required for high-level Cry1Ac resistance.

Author summary

Bacillus thuringiensis (Bt) foliar sprays and transgenic crops expressing Bt toxins are used extensively to control insect pests, but the evolution of resistance limits their efficacy. Multiple studies have reported that ATP-binding cassette (ABC) transporters are important Bt receptors, and mutations in either ABCC2 or ABCC3 can lead to Cry1Ac-toxin resistance, although this process is not fully understood. In this study, we applied both forward and reverse genetic analyses to demonstrate that high-level Bt-Cry1Ac resistance in Plutella xylostella requires concurrent mutations in both PxABCC2 and PxABCC3. We identified inactivating mutations in these two genes from a Cry1Ac-resistant strain (Cry1S1000) of P. xylostella and conducted genetic linkage analysis, which supported the role that PxABCC2 and PxABCC3 were the causal genes of Cry1Ac resistance. We then knocked out PxABCC2 and PxABCC3 in a P. xylostella susceptible reference strain (G88) to confirm that high-level Cry1Ac resistance requires mutation of PxABCC2 and PxABCC3, rather than a mutation of either one gene. This finding expands our understanding of complex Bt resistance processes and may be relevant to Bt-Cry1Ac resistance in other lepidopteran insects.

Introduction

Bacillus thuringiensis (Bt) is a ubiquitous gram-positive bacterium that can produce insecticidal protein toxins used for the control of many lepidopteran, coleopteran and dipteran pests [1,2]. Bt toxins often show high insect specificity and are safe for non-target organisms [1], however, like many synthetic agrochemicals, evolution of resistance among pests is threatening their efficacy [3]. By 2019, practical field-evolved resistance to Bt has been documented in at least 22 cases and nine insect species [37]. Complete understanding of the mechanisms causing Bt resistance remains limited, although mutations or expression changes in toxin-binding proteins, such as cadherin-like protein, alkaline phosphatase (ALP), aminopeptidase N (APN), and ABC transporters, appear to play important roles in mediating the resistance level [8,9].

ATP-binding cassette transporters, or ABC transporters, consist of two cytosolic nucleotide-binding domains (NBDs) and two integral transmembrane domains (TMDs), each with six membrane-spanning helices [10]. Mutations in ABC transporter subfamily C2 (ABCC2) are associated with Bt (Cry1A toxins) resistance in a number of lepidopteran species [1117]. The receptor functionality of ABCC2 for Cry1A toxins has been well characterized using heterologous expression in cultured cell lines, manipulating mRNA abundance with RNA interference and with transgenic Drosophila [14,1822]. ABC transporter subfamily C3 (ABCC3), a paralog of ABCC2, has also been reported to mediate Cry1A toxicity [14,2224]. Recently, Wang et al. [25] reported mutations in ABCC2 plus ABCC3 were required to achieve high-level Cry1Ac resistance in Helicoverpa armigera, suggesting functional redundancy as toxin receptors.

The diamondback moth, Plutella xylostella, is a notorious pest of cruciferous crops worldwide. Field resistance to Bt spray formulas first arose in this species within a population from Hawaii, USA [5,26]. Baxter et al. [17] have identified a 30-bp deletion in ABCC2 within a Cry1Ac-resistant strain (NO-QAGE) derived from this Hawaiian population. Reduced expressions in ABCC2 and ABCC3 have also been observed in resistant P. xylostella strains from other countries, including the DBM1Ac-R strain collected from Loxahatchee (Florida, USA) [22]. A recent gene-editing-based investigation provides further evidence of the roles of ABCC2 and ABCC3 in Bt resistance: knockout of either of the two P. xylostella genes, PxABCC2 and PxABCC3, confers high levels of resistance to Bt Cry1Ac (724 and 413 folds compared to the reference strain DBM1Ac-S) [16]. However, these resistance levels to Cry1Ac were more than 4,000-fold lower than a selected resistant strain [16].

In this study, we tested the hypothesis that mutations in both P. xylostella ABCC2 and ABCC3 are required to achieve a high-level Cry1Ac resistance in natural populations, while mutations in either gene confer low-level resistance. To do so, the sequences of PxABCC2 and PxABCC3 were compared between a Cry1Ac-resistance strain (Cry1S1000) and a susceptible strain (G88) of P. xylostella to identify the potential mutations. We then tested the genetic linkage of these two genes with Cry1Ac resistance and analyzed their contribution to Bt resistance by introgressing PxABCC2 and PxABCC3 from Cry1S1000 into G88. Finally, we introduced mutations into these ABCC genes separately and together in a reference strain with the CRISPR/Cas9 system, and then assessed their capacity to confer resistance to Cry1Ac using bioassays and genetic complementation tests.

Results

Resistance of the Cry1S1000 strain to Cry1Ac protoxin

Bioassays performed with artificial diet overlay assays indicated that the concentration of Cry1Ac protoxin capable of killing 50% of resistant Cry1S1000 larvae (LC50) was higher than 1,000 μg/ml (S1 Table). The Cry1S1000 strain was > 8,000 fold more resistant to Cry1Ac than the susceptible reference, G88 (S1 Table). In subsequent generations without Bt selection, Cry1S1000 maintained stable resistance to Cry1Ac (LC50 > 1,000 μg/ml) and showed 100% survival at a diagnostic concentration of 0.5 μg/ml of Cry1Ac protoxin (S2 Table).

ABCC2 and ABCC3 sequence comparison between P. xylostella strains

Full-length PxABCC2 transcripts were sequenced using midgut cDNA template generated from Bt-susceptible G88 larvae. The coding sequence was 4,044 bp and encoded a predicted 1,347 aa protein (S1 and S2 Figs) that displayed a typical structure among known lepidopteran ABCCs [11], including two transmembrane domains (TMD1 and TMD2), each with six transmembrane helices, and two nucleotide-binding domains (NBD1 and NBD2) (Fig 1A) containing conserved Walker A, B and C motifs (S2 Fig).

thumbnail
Fig 1. ABCC2 in the Cry1Ac-susceptible and -resistant strains of P. xylostella.

(A) Schematic structure of PxABCC2 protein showing the position of deletions (path lines in purple) and insertions (orange triangles) in the Cry1S1000 strain. The predicted protein contains two transmembrane domains (TMD1 and TMD2) each with 6 transmembrane helices (dark green cylinders) and two nucleotide-binding domains (NBD1 and NBD2, yellow ovals). ECL: extracellular loop; ICL: intracellular loop. Transmembrane helices were predicted by Phobius (http://phobius.sbc.su.se/). (B) Schematic representation of PxABCC2 exonic sequence of the susceptible G88 strain (ABCC2_S) and five transcript variants (ABCC2_R1-5) of the resistant Cry1S1000 strain. Exons are numbered (1–26). (C) Genomic structure of PxABCC2 illustrating sequences of SA2 and RA2 allele. Underlined letters GT/AG indicate the splice sites.

https://doi.org/10.1371/journal.ppat.1008697.g001

Five independent PxABCC2 transcript variants from the resistant Cry1S1000 strain were found to contain a range of insertions or deletions, however, all lacked exon 7 (ABCC2_R1-5 in Fig 1B, Table 1, S1 Fig). Two isoforms (ABCC2_R1 and ABCC2_R5) had in-frame deletions in exon 5 (Fig 1B), disrupting the portion of PxABCC2 protein from intracellular loop 2 to NBD1 (Fig 1A, Table 1, S2 Fig). The remaining three isoforms (ABCC2_R2-4) had frameshift mutations that introduced premature stop codons at extracellular loop 3 (Fig 1A, Table 1, S2 Fig).

thumbnail
Table 1. PxABCC2 isoforms of P. xylostella strain Cry1S1000.

https://doi.org/10.1371/journal.ppat.1008697.t001

All Cry1S1000 PxABCC2 isoforms lacked exon 7, however, genomic DNA amplicon sequencing revealed the coding sequence was present in the genome and not simply a chromosomal deletion. A point mutation (G>T) in the 3′ splice junction of intron 6 was identified, which would disrupt pre-mRNA processing and prevent the incorporation of exon 7 into the mature mRNA transcript. Hereafter, we refer to this genomic DNA mutation as “RA2” (Fig 1C and S3 Fig).

Alignments between gDNA and cDNA sequences also found an 8-bp deletion in ABCC2_R1 was caused by an alternative 5′ splice site in intron 5, and a 29-bp deletion in ABCC2_R2 was caused by an alternative 3′ splice site in intron 14 (Table 1, S3 and S4 Figs). The insertions were found in two other isoforms, which were caused by the retention of introns in the mature mRNA (intron 12 in ABCC2_R3 and introns 11, 15 and 16 in ABCC2_R4, S4S6 Figs). Finally, an 804-bp deletion in ABCC2_R5 was caused through skipping exons 6–10 (Fig 1B).

Intron 6 varied in size between G88 (581 bp, SA2) and Cry1S1000 (429 bp, RA2), which enabled development of an allele-specific PCR assay (AS-PCR) capable of quickly differentiating between susceptible and resistant alleles (S7 Fig).

PxABCC3 cDNA from the G88 (ABCC3_S) contained an open reading frame (ORF) of 4,047 bp and encoded a predicted protein of 1,348 aa (S8 and S9 Figs). Similar to PxABCC2, the PxABCC3 protein also consisted of two transmembrane domains (TMD1 and TMD2) and two nucleotide-binding domains (NBD1 and NBD2) (Fig 2A and S9 Fig). A point mutation (C>T) identified at nucleotide 2131 in Cry1S1000 PxABCC3 cDNA (ABCC3_R) (Fig 2B and S8 Fig) introduced a premature stop codon just prior to transmembrane helix 7, truncating the protein (Fig 2A and S9 Fig). This allele was referred as “RA3”, and the susceptible PxABCC3 allele “SA3”. Size variation was again used to develop an AS-PCR in PxABCC3 using primers spanning intron 13 (480 bp in G88 and 546 bp in Cry1S1000) (S10 Fig).

thumbnail
Fig 2. ABCC3 in the Cry1Ac-susceptible and -resistant strains of P. xylostella.

(A) Schematic illustration of PxABCC3 protein showing the position of the premature stop codon (asterisk in red) from the resistant Cry1S1000 strain. The PxABCC3 protein is predicted to contain two transmembrane domains with 6 transmembrane helices (dark green cylinders) for TMD1 and 6 transmembrane helices (light blue cylinders) for TMD2, and two nucleotide-binding domains (NBD1 and NBD2). Transmembrane helices, except for TM1 and TM2, were predicted by Phobius. The dotted lines, TMD2 in light blue and NBD2 in yellow represent the predicted region missing in the PxABCC3 from Cry1S1000. (B) The partial cDNA sequences and the deduced protein sequences from the G88 and Cry1S1000 strains show the point mutation of C2131T (the letter in red) in the Cry1S1000 strain that leads to the premature termination of PxABCC3. Amino acid residues highlighted in light blue are the portion of transmembrane helix 7.

https://doi.org/10.1371/journal.ppat.1008697.g002

Genetic linkage between RA2/RA3 and Cry1Ac resistance

We tested for genetic linkage between RA2/RA3 alleles and Cry1Ac resistance using 10 backcross families, which were obtained from single-pair crosses between Cry1S1000 (RR) and an F1 (RS, from Cry1S1000♀ × G88♂) (Fig 3A). Bioassays were performed on 50 larvae from each of 10 backcross families (n = 500) and 224 survived after exposed to 0.5 μg/ml Cry1Ac protoxin. A total of 447 untreated control larvae were also collected (S3 Table).

thumbnail
Fig 3. Genetic linkage analysis of RA2 and RA3 with Cry1Ac resistance.

(A) Diagram showing the strategy for backcrossing and bioassays. A resistant female (Cry1S1000) was crossed with a susceptible male (G88) to produce F1 progeny. Two types of backcrosses were made between F1 and Cry1S1000: BCa1-5 families were produced by single-pair crosses between an F1 male and Cry1S1000 female, and BCb1-5 families were produced by single-pair cross between an F1 female and Cry1S1000 male. Only the resistant backcross progeny (RR) survived the discriminating 0.5 μg/ml Cry1Ac bioassay. (B,C) Genotyping of PxABCC2 (B) and PxABCC3 (C) in Cry1S1000 (RA2RA2 or RA3RA3) or G88 (SA2SA2 or SA3SA3) individuals, and their F1 progeny (RA2SA2 or RA3SA3) using AS-PCR.

https://doi.org/10.1371/journal.ppat.1008697.g003

We first genotyped PxABCC2 and PxABCC3 in all cross parents using AS-PCRs (Fig 3B and 3C) and confirmed Cry1S1000 individuals were homozygous for the RA2 and RA3 alleles and only detected SA2 and SA3 alleles among G88 individuals. All F1 progeny from Cry1S1000 × G88 mating were heterozygous for the RA2/SA2 and RA3/SA3 alleles (Fig 3B and 3C). Among untreated control backcross progeny, we expected half the individuals to be heterozygous (RA2/SA2, RA3/SA3) and half to be homozygous for ABCC mutations (RA2/RA2, RA3/RA3) and observed numbers did not significantly differ (216 heterozygous, 231 homozygous, Fisher’s exact test P > 0.41 for each control family) (S3 Table). Among insecticide treated larvae, no heterozygous RA2/SA2, RA3/SA3 individuals were detected, indicating the Cry1Ac discriminating dose was sufficient to kill carriers of ABCC2 and ABCC3 G88 alleles (Fisher’s exact test for each treated family, P < 0.0005) (S3 Table). Therefore, Cry1Ac resistance was genetically linked to RA2 and RA3 alleles in the Cry1S1000 strain.

Introgression of RA2/RA3 alleles into the genetic background of G88

To eliminate potential differences in fitness between strains, and assess whether loci that were not linked to ABCC mutations contributed to Cry1Ac resistance, RA2 and RA3 alleles were introgressed into G88. Cry1S1000 males were crossed to G88 females, then F1 progeny backcrossed to G88 for a further six generations. Selection of the RA2 allele was performed in each generation backcrossing using AS-PCR. Random mating between progeny of backcross six with heterozygous genotypes (RA2SA2) were used to establish a near-isogenic line to the G88 strain (G88-RA2) that was homozygous for both the RA2 and RA3 alleles (S4 Table). Diet bioassays confirmed the G88-RA2 strain was equally resistant to Cry1S1000 (>1,000 μg/ml Cry1Ac protoxin, S1 Table).

Reciprocal crosses between resistant strains (Cry1S1000 and G88-RA2) and G88 were performed and susceptibility of F1 progeny to the diagnostic concentration (0.5 μg/ml) of Cry1Ac protoxin was assessed. The mass crosses between each of the two resistant strains with G88 resulted in no survival of F1 progeny (i.e., Cry1S1000♀ × G88♂, G88♀ × Cry1S1000♂, G88-RA2♀ × G88♂, and G88♀ × G88-RA2♂) (S5 Table). This indicated that the Cry1S1000 and G88-RA2 strains exhibited autosomal and completely recessive (h = 0) resistance to Cry1Ac at the diagnostic concentration. Third-instar larval progeny from ten single-pair crosses between Cry1S1000 and G88-RA2 were assayed and showed 100% survival after a 5-day bioassay (n = 385), confirming the strains shared a common, recessive resistance locus (Index of commonality = 1) (S5 Table).

Mutagenesis of PxABCC2 and PxABCC3 mediated by CRISPR/Cas9

To knock out PxABCC2 and PxABCC3, we injected mixtures of sgRNA and Cas9 protein (or mRNA) into fresh eggs from the wild type G88 strain. Successful site-specific mutations within PxABCC2 and PxABCC3 were identified within mosaic G0 moths (S6 Table).

Four homozygous mutant PxABCC2 strains were generated: G88-ABCC2--1 had a 43-bp insertion in exon 1, G88-ABCC2--2 a 28-bp deletion in exon 3, G88-ABCC2--3 a 16-bp deletion in exon 3, and G88-ABCC2--4 a 40-bp deletion in exon 20 (Fig 4A and S11 Fig). Embryos microinjected with PxABCC2-exon 3 sgRNA that failed to generate targeted indel mutation were used to generate the line G88-No-Indel and acted as a negative control to analyze potential off-target effect (Fig 4A and S11B Fig). The four ABCC2 CRISPR/Cas9 strains were all predicted to carry PxABCC2 frameshift mutations producing truncated proteins (S12A Fig).

thumbnail
Fig 4. Mutagenesis of PxABCC2 and PxABCC3 induced by CRISPR/Cas9.

(A) Representative sequences from the wild type (G88) and four homozygous PxABCC2 mutant strains, showing the deletions and insertions (indels) in exon 1, 3 and 20 of PxABCC2. (B) Partial sequences from the G88 and two homozygous PxABCC3 mutants showing the indels at the target sequence (ABCC3-sg1) in exon 1 of PxABCC3. (C) Representative sequences from two PxABCC3 mutants and three PxABCC2/PxABCC3 double mutant strains showing the indels at the target sequence (ABCC2-sg2) in exon 3 of PxABCC2.

https://doi.org/10.1371/journal.ppat.1008697.g004

Nano-LC-MS/MS sequencing of brush border membrane vesicles (BBMV) proteins ~ 120–200 kDa in the G88 strain midgut identified 1,685 peptides, 34 of which were specific to PxABCC2 (S12B–S12D Fig). Although 1,179 peptides were detected from the G88-ABCC2--1 mutant, none of them were PxABCC2 sequences (S7 Table), supporting total disruption of PxABCC2 protein in this PxABCC2-knockout strain.

For PxABCC3, two homozygous mutant strains with either a 1-bp insertion or a 11-bp deletion in exon 1 were generated (G88-ABCC3--1 and G88-ABCC3--2) (Fig 4B and S13 Fig). Nano-LC-MS/MS analysis for ~ 90–200 kDa midgut BBMV proteins detected 2,239 tryptic peptides from the G88 strain, 33 of which were specific to the PxABCC3 protein (S14 Fig). A total of 2,553 and 1,968 peptides were detected from G88-ABCC3--1 (S8 Table) and G88-ABCC3--2 (S9 Table), respectively, and none were PxABCC3 peptides. This supported a complete loss of PxABCC3 protein in both PxABCC3-knockout strains.

To produce PxABCC2/PxABCC3 double mutant strains, frameshift mutations in PxABCC2 exon 3 were then introduced through microinjecting embryos from the G88-ABCC3--1 and G88-ABCC3--2 CRISPR/Cas9 strains. Two double mutant strains were established from G88-ABCC3--1 strain, including G88-A2-A3--1 (8-bp PxABCC2 exon 3 insertion) and G88-A2-A3--2 (35-bp exon 3 deletion) (Fig 4C), and one double mutant strain was established from G88-ABCC3--2 that we name G88-A2-A3--3 (49-bp PxABCC2 exon 3 deletion) (Fig 4C). Nano-LC-MS/MS analysis of ~ 120–250 kDa midgut BBMV proteins from G88 larvae yielded 2,048 peptides, 27 of which were PxABCC2 and 40 PxABCC3 (S15 Fig). However, we found no evidence of PxABCC2 and PxABCC3 proteins in the double mutant G88-A2-A3--1 when the 3,159 peptides sequenced were examined (S10 Table), indicating complete knockout of both proteins in this strain.

Effect of PxABCC2 and PxABCC3 mutations on P. xylostella susceptibility to Cry1Ac protoxin

The bioassays indicated that knockout of PxABCC2 in four mutant strains (i.e., G88-ABCC2--1, G88-ABCC2--2, G88-ABCC2--3 and G88-ABCC2--4) resulted in a decline in susceptibility to Cry1Ac protoxin between 2.45–3.78 folds relative to the G88 reference strain. The negative control of G88-No-Indel strain showed similar Cry1Ac susceptibility as the G88 strain. The two PxABCC3 mutant strains, G88-ABCC3--1 and G88-ABCC3--2, surprisingly showed a slight increase in susceptibility in comparison to G88. The three PxABCC2/PxABCC3 double mutant strains showed > 8,000-fold resistance to Cry1Ac compared with the G88 strain (Fig 5A, S1 Table).

thumbnail
Fig 5. Response of different P. xylostella strains to Cry1Ac protoxin.

(A) Resistance level (fold) of two resistant strains (Cry1S1000 and G88-RA2), a negative control (G88-No-Indel), four CRISPR-based PxABCC2 mutants, two CRISPR-based PxABCC3 mutants, three CRISPR-based PxABCC2/PxABCC3 double mutants and the F1 progeny from mass crosses between each of three CRISPR-based mutants and G88-RA2 compared to the G88 strain. The resistance ratio (fold) for G88 is 1. (B) Survivals at the diagnostic concentration (0.5 μg/ml) of Cry1Ac protoxin of P. xylostella larvae from two resistant strains (Cry1S1000 and G88-RA2), a susceptible strain (G88), a CRISPR-based PxABCC2 mutant (G88-ABCC2--1), a CRISPR-based PxABCC3 mutant (G88-ABCC3--1), a CRISPR-based PxABCC2/PxABCC3 double mutant (G88-A2-A3--1), and the F1 progeny from single-pair crosses between each of three CRISPR-based mutants and each of two resistant strains or G88. Asterisks indicate 0% survival. Index of commonality (C) indicates the proportion of the two strains sharing the common resistance locus or loci.

https://doi.org/10.1371/journal.ppat.1008697.g005

To further determine the contribution of recessive mutations in two ABCC loci to Cry1Ac resistance, we performed complementation tests for PxABCC2 and PxABCC3 between resistant strains and CRISPR-based mutants. The average survival rates when exposed to 0.5 μg/ml Cry1Ac protoxin were 11.7% (7/60) for G88-ABCC2--1, 11.7% (0.0% - 20.0%) for the F1 progeny from single-pair crosses between G88-ABCC2--1 and Cry1S1000, and 2.6% (0.0% - 21.7%) for the F1 progeny from single-pair crosses between G88-ABCC2--1 and G88-RA2 (Fig 5B). The results suggested that two mutations within PxABCC2 conferred a low level of resistance to Cry1Ac. Similarly, no survival of F1 progeny from single-pair crosses between G88-ABCC3--1 and each of two resistant strains was observed, indicating that two mutations at the PxABCC3 locus failed to confer Cry1Ac resistance. However, the F1 progeny produced from crossing the double mutant strain G88-A2-A3--1 and Cry1S1000 or G88-RA2 were highly resistant to Cry1Ac like their parent strains (Fig 5B). These results showed that two mutations at each of PxABCC2 and PxABCC3 loci were required for high levels of resistance.

Consistent with the results from single-pair crosses, the F1 progeny from mass crosses between G88-ABCC2--1 and G88-RA2, and between G88-ABCC3--1 and G88-RA2 retained comparable susceptibility to Cry1Ac as G88 (Fig 5A, S1 Table). In addition, the bioassay test showed that F1 progeny from a mass cross between the G88-A2-A3--1 CRISPR generated strain and G88-RA2 introgression strain exhibited > 8,000-fold resistance to Cry1Ac compared with G88 (Fig 5A, S1 Table).

Discussion

The complex mode of action of Bt insecticides requires activated toxins to interact with receptors on the brush border membrane of the insect midgut [27]. In this study, we provide rigorous support for the duel involvement of two ABC transporter subfamily genes, PxABCC2 and PxABCC3, in mediating Bt Cry1Ac protoxin susceptibility in P. xylostella. CRISPR/Cas9-mediated mutagenesis verified that high-level resistance to Cry1Ac only occurred after introducing mutations into both PxABCC2 and PxABCC3 alleles, while mutations in one gene alone had little effect. Therefore, PxABCC2 and PxABCC3 appear to function as redundant Cry1Ac receptors for determining susceptibility of Cry1Ac in P. xylostella. This finding supports recent research by Wang et al. [25] demonstrating that the knockout of ABCC2 and ABCC3 are required for Cry1Ac resistance in H. armigera.

Here we identified five mutated transcripts of PxABCC2 (ABCC2_R1-5) in the resistant Cry1S1000 strain, and although variation in transcripts sequence was observed, all appeared to be directly caused by the RA2 splice-site mutation of intron 6. One of these transcripts, ABCC2_R1 was previously identified by Guo et al. and reported as isoform VIIa from a resistant strain (DBM1Ac-R) of P. xylostella [22], and the loss of exon 7 from sequenced transcripts was a common feature of both these studies [22].

The PxABCC2 RA2 splice-site mutation described here in Cry1S1000 is independent from another 30-bp deletion in exon 20 of another Bt resistant P. xylostella NO-QAGE strain [17]. This suggests that multiple mutations can occur in Bt-selected populations. Expressing PxABCC2 in Drosophila midgut confirmed the transmembrane protein can act as a Cry1Ac toxin receptor [21], however, mutations in PxABCC3 have never been reported in NO-QAGE nor any P. xylostella Bt resistant strain to the best of our knowledge. A premature stop codon predicted in PxABCC3 (RA3) of Cry1S1000 here appeared to prevent translation.

Reducing expression levels of either PxABCC2 or PxABCC3 mRNA through RNAi knockdown has previously been shown to reduce mortality in wild type P. xylostella larvae when challenged with discriminating doses of Cry1Ac protoxin [22]. Simultaneous reduction of PxABCC2, PxABCC3 and membrane-bound ALP (PxmALP) transcripts using RNAi reduces mortality more than ABCC2 or ABCC3 alone [22], providing some support for a combined role of receptors in Cry1Ac mode of action. Reducing or eliminating PxABCC2 and PxABCC3 through transcriptional regulators has also been proposed as a mechanism for Cry1A resistance development without requiring genetic mutations within the genes themselves [22].

Although many lepidopteran insects are susceptible to Cry1Ac toxins, it is currently unclear whether mutations in ABCC2 and ABCC3 are required for high-level resistance in all species. Some of the first examples with Heliothis virescens and Bombyx mori [11,12] provide functional support that ABCC2 is required for Cry1A susceptibility. The B. mori Cry1Ab/c resistant strain, “C2”, was rescued to a susceptible phenotype through piggyback transformation that introduced a wild-type copy of BmABCC2 [11]. Mutations in BmABCC3 have not been reported, however, only rescue of one receptor may be required for function. CRISPR/Cas9-mediated double knockout of H. armigera ABCC2 and ABCC3 confers a >15,000-fold resistance to Cry1Ac [25]. Knockout of ABCC2 in Ostrinia furnacalis causes fairly high level of resistance to Cry1F, but not Cry1A toxins [28]. In Spodoptera exigua, however, ABCC2 mutation appears to be sufficient to confer resistance to Cry1Ac [15].

Expressing ABCC2 in cell lines of both H. virescens and B. mori cause cells to swell or lyse and the effect is synergized when co-expressed with another well-characterized Bt toxin receptor, a cadherin-like protein (CaLP) [19,20]. However, the involvement of P. xylostella CaLP in Cry1Ac toxicity is still controversial [2931]. Synergism between ABCC2 and CaLP has also been demonstrated in vivo. Mutations in either the Trichoplusia ni cadherin or ABCC2 only slightly reduce larval susceptibility to a variant of Cry1Ac, while deletion of both proteins confers much higher levels of resistance [32]. Deficiencies of H. virescens 12-cadherin-domain protein (HevCaLP) and ABCC2 (HevABCC2) offer higher resistance to Cry1Ac than a defect in either one [12].

The results here support redundant roles between PxABCC2 and PxABCC3 in conferring Cry1Ac susceptibility, and differ from two previous studies with diamondback moth [16,22]. In their cases, reduced expression of either PxABCC2 or PxABCC3 using RNAi decreased Cry1Ac susceptibility of P. xylostella larvae [22] and knockout of either PxABCC2 or PxABCC3 using CRISPR/Cas9 caused high levels of resistance to Cry1Ac [16]. Although we used artificial-diet bioassay and the previous studies used leaf-dip bioassay, this seems unlikely to be responsible for such differences in these results. Transcripts of ABCC2 and ABCC3 sequenced from our susceptible reference strain (G88) all contained full-length open reading frames, however, Guo et al. [22] reported a low frequency of alternatively spliced isoforms from the control strain in their study. Control strains that express different ABCC2 and ABCC3 alleles, or control strains that produce different mRNA isoforms could potentially explain the discrepancies between our work and the earlier works of Guo et al. [16,22].

Generating PxABCC2 mutation in the G88 wild-type strain led to a slightly higher resistance to Cry1Ac protoxin, yet PxABCC3 mutations caused higher susceptibility. Similarly, cultured cells of B. mori and S. exigua exhibit higher Cry1Aa susceptibility when expressing ABCC2 than ABCC3 [24,33]. Contribution of ABCC2 in determining toxicity of Cry1Ac may therefore be more significant than that of ABCC3. In B. mori, the extracellular loop 1 (ECL1) in BmABCC2/3 transporters determines receptor specificity for Cry1A toxins [33]. The loop 1 Q125 residue of SfABCC2 is also the key factor for the high toxicity of Cry1Ac to Hi5 cells expressing this protein [34]. Structure-based evidence therefore is necessary to further verify the roles of ECL1 in determining the specificity of PxABCC2 and PxABCC3 for Cry1Ac toxicity, and to better address their potential synergism in terms of functional complementation or redundancy. In addition, PxABCC2 has been associated with Cry1Ac oligomerization and oligomer membrane insertion [35]. Our knockout mutants of PxABCC2, PxABCC3 or both genes provide new opportunities to test this hypothesis and investigate whether PxABCC3 plays a similar role in Cry1Ac oligomerization.

Our work indicates that P. xylostella has evolved resistance to Cry1Ac protoxin through mutations in both PxABCC2 and PxABCC3. DNA-based screening may be a useful tool for monitoring the frequency of P. xylostella field resistance to Cry1A toxins, caused by mutations in ABCC genes. However, additional allelic variants causing mutations that affect toxin binding are likely to exist in field populations, and this needs to be considered. The present finding provides strong evidence for understanding the roles of ABC transporters in Bt mode of action and the genetic mechanisms that lead to Bt resistance in P. xylostella.

Materials and methods

Insect sample

The Plutella xylostella strains were established by Dr. Anthony M. Shelton (Cornell University, USA) and provided to the Institute of Applied Ecology, Fujian Agriculture and Forestry University, in 2016. The insecticide susceptible reference strain, Geneva 88 (G88), was collected from the New York State Agricultural Experiment Station in 1988 and maintained on artificial diet without exposure to insecticides [36]. A population collected from Loxahatchee, Florida, in 1992 was resistant to B. thuringiensis subsp. kurstaki spray formula [37]. It was subsequently reselected with transgenic broccoli that only expressed the Cry1Ac toxin [38] and referred to as Cry1Ac-R [39]. The resistant strain Cry1Ac-R was then selected using 1000 μg/ml Cry1Ac protoxin in 2016 and reared on the artificial diet for over 50 generations without further insecticide exposure. We refer to this resistant strain as Cry1S1000. Both of the P. xylostella strains were kept at 26±1°C, 65% RH, and photoperiod of 16: 8 h (L: D). Adults were provisioned with 10% honey solution.

Bioassays

Cry1Ac protoxin was prepared from B. thuringiensis subsp. kurstaki strain HD-73 (provided by Key Laboratory of Biopesticide and Chemical Biology, Ministry of Education, Fujian Agriculture and Forestry University) as previously described by Kain et al. [40] and serially diluted in ultrapure water. An artificial diet overlay assay, similar to that described by Zhao et al. [39] and Kain et al. [40], was used to determine the susceptibility of P. xylostella larvae to Cry1Ac. Bioassays were performed on five replicates for five different concentrations of Cry1Ac protoxin, plus an insecticide free control. For each replicate, pre-warmed liquid artificial diet (5 mL) was poured into a 30 mL plastic cup, and 0.3 mL aliquot of Cry1Ac protoxin was then evenly distributed over the surface of the diet (surface area ≈ 7 cm2). The assay was maintained at room temperature and after 1.5–2 hours of air-drying, ten early-molted 3rd-instar larvae were placed in the cup. Larval mortality was recorded at day 5, and mortality rate was calibrated according to Abbott’s formula [41].

Cloning and sequencing of cDNA and genomic DNA

Midguts were dissected from G88 (n = 5) and Cry1S1000 (n = 5) 4th-instar larvae and total RNA was extracted from each individual using the TransZol Up Plus RNA Kit (TranGen, Beijing, China). cDNA synthesis was performed with the TransScript One-Step gDNA Removal and cDNA Synthesis Super Mix (TranGen, Beijing, China) according to the manufacturer’s instructions. PCR amplification of full length PxABCC2 and PxABCC3 transcripts were performed in a 25-μl reaction system containing 0.5 μl of cDNA template, 0.4 μM of primer (S11 Table), 12.5 μl of 2 × Phanta Max Buffer, and 0.5 μl of Phanta Max Super-Fidelity DNA Polymerase (Vazyme, Nanjing, China). Reactions were carried out with an initial denaturation at 95°C for 3 min and followed by 35 cycles of denaturation at 95°C for 15 s, annealing at 55°C for 15 s and extension at 72°C for 5 min. After examination by agarose gel electrophoresis, the amplified products were purified using the Gel Extraction Kit (Omega, Morgan Hill, GA, USA), cloned into the pJET1.2/blunt Cloning Vector using the CloneJET PCR Cloning Kit (Thermo Scientific, Waltham, MA, USA), and sequenced by Biosune Biotech Company (Fuzhou, China).

Genomic DNA (gDNA) was isolated from each of the individual carcasses remaining after midgut dissection using the Tissue DNA Kit (Omega, Morgan Hill, GA, USA). The gDNA was used as template for PCR amplification of four PxABCC2 fragments (gDNA_PF1—PF4). Amplification reactions (25 μl) contained 50 ng of gDNA template, 0.4 μM of primer (S11 Table), 12.5 μl of 2 × Phanta Max Buffer, and 0.5 μl of Phanta Max Super-Fidelity DNA Polymerase (Vazyme, Nanjing, China). All fragments were amplified with the extension time adjusted according to the fragment size, then cloned and sequenced as previously described.

Allele-specific PCR assay

Based on the size variation of PxABCC2 intron 6 and PxABCC3 intron 13 between Cry1S1000 and G88, to detect resistant (RA2 or RA3) alleles, we used allele-specific PCR and two pairs of primers (S11 Table) that flank either RA2 or RA3 mutations. All adults that needed to be genotyped for RA2 or RA3 allele were sacrificed and gDNA was isolated individually using the Tissue DNA Kit (Omega, Morgan Hill, GA, USA). PCR amplification was performed in a 25-μl reaction system under the conditions described in section [Cloning and sequencing of cDNA and genomic DNA]. PCR products and DL1000 DNA marker (TAKARA, Dalian, China) were separated using 3% agarose gel electrophoresis to differentiate the size of DNA bands of resistant (RA2 and RA3) and susceptible (SA2 and SA3) alleles. To validate allele-specific PCR, amplified products were cloned and sequenced from Cry1S1000 (n = 5) and G88 (n = 5) as described in section [Cloning and sequencing of cDNA and genomic DNA].

Genetic linkage analysis of RA2 and RA3 with Cry1Ac resistance

To test for genetic linkage between Cry1Ac resistance in the Cry1S1000 strain and RA2 and RA3, a single-pair cross was performed between a Cry1S1000 female and G88 male to generate F1 progeny. Ten single pair backcrosses (BC) were performed by crossing an F1 male and Cry1S1000 female (BCa1—BCa5) or an F1 female and Cry1S1000 male (BCb1—BCb5). For each backcross family, 36–48 3rd-instar larvae were reared on the artificial diet (control), while 50 larvae were reared on artificial diet supplemented with 0.5 μg/ml of Cry1Ac protoxin for five days. Mortality was recorded and all surviving larvae were allowed to complete development before gDNA was isolated.

Introgression of the RA2 and RA3 alleles from Cry1S1000 into G88

The Cry1S1000 Cry1Ac resistance locus was introgressed into the genetic background of G88. First, males from Cry1S1000 (RA2RA3/RA2RA3) were mass crossed with G88 females (SA2SA3/SA2SA3) to produce heterozygous F1’s (RA2RA3/SA2SA3) and the male progeny were then mass crossed with G88 females to generate BC1 progeny. BC1 males were individually crossed with G88 females (single-pair mating) to generate BC2 progeny and sacrificed after mating for genotyping using allele-specific PCR assays. Only lines containing an RA2 allele (RA3 was not genotyped during introgression crosses as the two genes are tightly linked) were retained. Additional four single-pair backcrosses (BC3—BC6) and corresponding molecular assays were performed using the same method of backcrossing between BC1 males and G88 females. Subsequently, BC6 siblings were crossed in single pairs to generate BC6F1 progeny and lines containing heterozygous parents (RA2SA2) were retained after genotyping. Similarly, BC6F1 siblings were crossed (single-pair mating) to generate BC6F2 progeny and genotyped. Only lines containing homozygous parents (RA2RA2) were retained and constituted a homozygous strain of G88-RA2 (S16 Fig). Finally, the G88-RA2 strain was genotyped to confirm it was fixed for the RA3 allele.

Determination of inheritance mode of Cry1Ac resistance

Reciprocal crosses were made between each of two resistant strains (Cry1S1000 and G88-RA2) and G88 to generate F1 heterozygotes using 15 virgin females from one strain and 15 males from the other. Bioassays were performed using 0.5 μg/ml Cry1Ac protoxin on Cry1S1000, G88-RA2, G88 and their F1 progeny from reciprocal crosses. To evaluate maternal effects and sex linkage of resistance in Cry1S1000 and G88-RA2, we compared the survival of F1 progeny from reciprocal crosses between Cry1S1000 and G88, and between G88-RA2 and G88. Dominance (h), ranging from 0 (completely recessive resistance) to 1 (completely dominant resistance), was estimated based on the survival of each resistant strain (Cry1S1000 and G88-RA2), G88 and their F1 progeny using the single-concentration method [42].

Preparation of sgRNA and Cas9

sgRNA target sites containing 5'-N20NGG-3' (with the PAM sequence underlined) were selected in PxABCC2 exons 1, 3, and 20 and in PxABCC3 exon 1. A template for in vitro transcription of sgRNA’s was obtained using PCR that included two oligonucleotides. The first contained the T7 polymerase binding site and sgRNA target sequence indicated by twenty underlined N’s (5′-TAATACGACTCACTATAGGNNNNNNNNNNNNNNNNNNNNGTTTTAGAGCTAGAAATAGCAAGTTAA-3′), and the second contained the remaining sequence common to all sgRNAs (5′-AAAAGCACCGACTCGGTGCCACTTTTTCAAGTTGATAACGGAC TAGCCTTATTTTAACTTGCTATTTCTAGCTCTAAAA-3′). All sgRNA target sequences for PxABCC2 and PxABCC3 are listed in S12 Table. PCR amplification was performed in a 50-μl reaction system containing 0.4 μM of primer, 5 μl of 10× PCR buffer for KOD-Plus-Neo, 5 μl of 2 mM dNTPs, 3 μl of 25 mM MgSO4, and 1 μl of KOD-Plus-Neo (TOYOBO, Osaka, Japan) with the following program: 98°C for 2 min, 35 cycles at 98°C for 10 s, 55°C for 30 s, and 68°C for 30 s. The PCR products were examined by agarose gel electrophoresis and purified using the Gel Extraction Kit (Omega, Morgan Hill, GA, USA). sgRNAs were generated via in vitro transcription using the HiScribe T7 Quick High Yield RNA Synthesis Kit (New England Biolabs, Ipswich, MA, USA) and purified through phenol-chloroform extraction and ethanol precipitation. sgRNAs were then stored at -80°C until usage.

Cas9 mRNA was generated by in vitro transcription using the HiScribe T7 ARCA mRNA Kit (with tailing) (New England Biolabs, Ipswich, MA, USA) with the template of a linearized PTD1-T7-Cas9 vector [43], and purified as described for sgRNA. Cas9 protein (GenCrispr Cas9-N-NLS Nuclease) was purchased from the GeneScript Corporation (Nanjing, China).

Microinjection of P. xylostella embryos

A parafilm sheet (5 cm × 15 cm) precoated with dry radish seedling powder was provided to G88 females for 15–20 minutes to facilitate egg collection. The parafilm sheet was retrieved and fixed to a glass slide to inject the embryos with a mixture containing Cas9 (300 ng/μl of Cas9 mRNA or 600 ng/μl of Cas9 protein) and 100 ng/μl of sgRNA. Injections were performed within 1–2 hours of lay egg collection using the IM 300 Microinjector (Narishige, Tokyo, Japan) mounted on the SZX16 Stereo Microscope (Olympus, Tokyo, Japan). Injected embryos were incubated without light at 26±1°C in a Petri dish containing a moist tissue paper until hatching. Hatched 1st-instar larvae were transferred to the artificial diet (representing the G0).

Screening and development of homozygous mutant strains

Virgin G0 adults that survived microinjection were individually sexed and coupled with virgin G88 adults for single-pair mating. After successful oviposition, G0 individuals were sacrificed and gDNA isolated. PCR amplicons spanning sgRNA target sites were generated then sequenced from the G0’s, and only lines containing mutations were retained. G1 siblings were crossed in single pairs to produce G2 progeny and only lines that contain heterozygous parents sharing the same allelic mutation based on molecular identification were retained. Finally, G2 siblings were crossed (single-pair mating) again to generate G3 progeny and genotyped. Only lines containing homozygous mutant parents were kept to generate the homozygous mutant strains. If homozygous individuals could not be obtained from the G2 progeny, additional sibling crosses were performed (S17 Fig).

Identification of CRISPR/Cas9 mediated mutations first required isolating gDNA from whole adults after mating using the Tissue DNA Kit (Omega, Morgan Hill, GA, USA). Amplicons spanning the sgRNA target sites for PxABCC2 and PxABCC3 were generated using primers listed in S11 Table. PCR amplification was performed as described in section [Cloning and sequencing of cDNA and genomic DNA] and products were Sanger sequenced (Biosune Biotech Company, Fuzhou, China).

Preparation of midgut brush border membrane vesicles (BBMV)

BBMV were prepared using the differential magnesium precipitation method [44]. The 4th-instar larvae were dissected longitudinally in cold buffer A (300 mM mannitol, 5 mM EGTA, 17 mM Tris-HCl, pH 7.5) on ice to isolate the midgut epithelium. For each strain analyzed, 10 dissected midguts were pooled and homogenized in 750 μl of buffer A containing 1 mM PMSF using the TissueLyser II mixer (Qiagen, Dusseldorf, Germany) and a 3 mm stainless-steel bead at 22 Hz for two 1-min periods, separated by a 1-min cooling interval on ice. Subsequently, an equal volume of 24 mM MgCl2 solution was added to the midgut homogenate and mixed. The mixture was incubated on ice for 15 min, and then centrifuged at 2,500 g for 15 min at 4°C. The supernatant was collected and stored on ice. A second extraction was performed by resuspending the pellet using half the volumes of cold buffer A solution, then the same process as the first extraction with equal volume of 24 mM MgCl2. The second 2,500 g supernatant was combined with the first one and centrifuged at 30,000 g for 30 min at 4°C. The final pellet was resuspended in 50 μl of cold half-strength buffer A. Protein concentrations were measured by the BCA assay kit (Solarbio, Beijing, China) with bovine serum albumin (BSA) as a standard.

Protein identification

PxABCC2 protein from the midgut BBMV was determined by western blotting. BBMV proteins (20 μg) were separated using 7.5% SDS-PAGE and transferred onto polyvinylidene difluoride (PVDF) membranes (Millipore, Bedford, MA, USA) using transfer buffer (192 mM Glycine, 24.8 mM Tris, 20% methanol). Subsequently, the membranes were blocked in PBS (137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 1.76 mM KH2PO4, pH 7.4) containing 4% BSA at 4°C overnight. Membranes were probed separately with antibodies in PBS containing 1% BSA for 1 hour. All incubation steps were performed on a TS-100 shaker (Qilinbeier, Haimen, China) at room temperature and PVDF membranes were washed four times (5 min each time) with PBST (PBS containing 0.25% Tween-20) between steps. The antibodies were rabbit polyclonal anti-PxABCC2 antibody (1: 1,000 dilution, ABclonal, Wuhan, China) and anti-rabbit IgG (1: 10,000 dilution, Affinity, Cincinnati, OH, USA), or mouse monoclonal anti-alpha-tubulin (control, 1: 5000 dilution, Sigma-Aldrich, Saint Louis, MO, USA) and anti-mouse IgG (1: 50,000 dilution, Sigma, Saint Louis, MO, USA). Membranes were visualized with the Clarity Western ECL Substrate (Bio-Rad, Hercules, CA, USA).

PxABCC2 protein was further confirmed by nano-liquid chromatography-tandem mass spectrometry (nano-LC-MS/MS). Since PxABCC3 lacked primary antibody for the western blotting, its protein from the midgut BBMV was identified by nano-LC-MS/MS. Total BBMV proteins (30 μg) were separated on 7.5% SDS-PAGE. The regions expected to contain PxABCC2 and PxABCC3 proteins (~150 kDa) were excised from Coomassie blue stained gels, preserved in PBS buffer and shipped with ice packs to Huada Protein Research Center (Shenzhen, China) for tryptic digestion and analysis by nano-LC-MS/MS.

Allelic complementation test

Allelic complementation tests were performed between Cry1S1000 and introgression strain G88-RA2 using 10 single-pair reciprocal crosses and challenged their F1 progeny (25 to 50 larvae per family) with 0.5 μg/ml Cry1Ac protoxin. To determine whether the homozygous mutations in PxABCC2 or PxABCC3 alone or combined together confer Cry1Ac resistance, we performed 8 single-pair reciprocal crosses between three CRISPR/Cas9 lines and Cry1S1000, G88-RA2 or G88. Subsequently, the F1 progeny from each of these 8 single-pair crosses (20 to 45 larvae per family), along with six parental strains (three CRISPR/Cas9 lines, Cry1S1000, G88-RA2 and G88; 50 to 60 larvae per strain), were assayed with 0.5 μg/ml Cry1Ac protoxin. We then used the surviving individuals to calculate the index of commonality (C), which varies from close to or < 0 (resistance conferred by alleles at different loci) to close to or > 1 (resistance conferred by alleles at a shared locus or loci) [45].

In another additional experiment, each of three CRISPR/Cas9 mutant lines were mass crossed with G88-RA2 (mass mating, 15 pairs) to generate F1 progeny. These F1 progeny individuals were tested with at least five concentrations of Cry1Ac protoxin to get the LC50 using artificial diet overlay assay as described in section [Bioassays].

Data analyses

All statistical analyses were performed using IBM SPSS Statistics 22. LC50 values for Cry1Ac larval bioassays were calculated using probit analysis. Significant differences in the LC50 values were defined by non-overlapping 95% confidence limits [46]. Fisher’s exact test was used to determine significant differences between observed and expected genotypes among genetic crosses.

Supporting information

S1 Table. Susceptibility of different P. xylostella strains to Cry1Ac protoxin.

https://doi.org/10.1371/journal.ppat.1008697.s001

(DOC)

S2 Table. Resistant response of the Cry1S1000 strain to Cry1Ac protoxin at different generations.

https://doi.org/10.1371/journal.ppat.1008697.s002

(DOC)

S3 Table. Genetic linkage of the RA2 and RA3 alleles with Cry1Ac resistance.

https://doi.org/10.1371/journal.ppat.1008697.s003

(DOC)

S4 Table. Genotype of the RA2 and RA3 alleles in P. xylostella during introgression.

https://doi.org/10.1371/journal.ppat.1008697.s004

(DOC)

S5 Table. Inheritance of Cry1Ac resistance in the Cry1S1000 and G88-RA2 strains of P. xylostella.

https://doi.org/10.1371/journal.ppat.1008697.s005

(DOC)

S6 Table. Targeted mutagenesis of PxABCC2 and PxABCC3 induced by CRISPR/Cas9.

https://doi.org/10.1371/journal.ppat.1008697.s006

(DOC)

S7 Table. A total of 1,179 peptides identified from the G88-ABCC2--1 mutant strain.

https://doi.org/10.1371/journal.ppat.1008697.s007

(XLS)

S8 Table. A total of 2,553 peptides identified from the G88-ABCC3--1 mutant strain.

https://doi.org/10.1371/journal.ppat.1008697.s008

(XLS)

S9 Table. A total of 1,968 peptides identified from the G88-ABCC3--2 mutant strain.

https://doi.org/10.1371/journal.ppat.1008697.s009

(XLS)

S10 Table. A total of 3,159 peptides identified from the G88-A2-A3--2 mutant strain.

https://doi.org/10.1371/journal.ppat.1008697.s010

(XLS)

S11 Table. Sequences of primers used to amplify PxABCC2 and PxABCC3 in this study.

https://doi.org/10.1371/journal.ppat.1008697.s011

(DOC)

S12 Table. Target sequences of sgRNAs for PxABCC2 and PxABCC3.

https://doi.org/10.1371/journal.ppat.1008697.s012

(DOC)

S1 Fig. Alignment of PxABCC2 cDNA sequences from the G88 and Cry1S1000 strains.

Sequence alignment of PxABCC2 cDNA from G88 (ABCC2_S, GenBank accession no. MN652064) and five isoforms from Cry1S1000 (ABCC2_R1-5, GenBank accession no. MN652066-MN652070). Asterisks denote consensus sequences. Exon regions (yellow and green) are shown in the ABCC2_S. Deletions and insertions are respectively highlighted in gray and blue. Codons highlighted in red indicate the premature stop codons.

https://doi.org/10.1371/journal.ppat.1008697.s013

(DOC)

S2 Fig. Alignment of predicted amino acid sequences of PxABCC2 from G88 (ABCC2_S) and Cry1S1000 (ABCC2_R1-5).

Asterisks denote consensus sequences. Transmembrane domains (TM) (blue), Walker A and B sequences (yellow), C motif (green) and region for antibody recognition (underlined) are shown in the sequence of PxABCC2 protein encoded by ABCC2_S. Deletions are highlighted in gray.

https://doi.org/10.1371/journal.ppat.1008697.s014

(DOC)

S3 Fig. Alignment of partial gDNA sequence (gDNA_PF1, GenBank accession no. MN660237) of PxABCC2 and partial sequence of ABCC2_R1 cDNA from Cry1S1000.

Asterisks denote consensus sequences. Exons (yellow) and introns (gray) are shown in the sequence of gDNA_PF1. Deletions in cDNA are highlighted in green. Letter highlighted in red indicates the point mutation in gDNA. The primer sequences are underlined in red.

https://doi.org/10.1371/journal.ppat.1008697.s015

(DOC)

S4 Fig. Alignment of partial gDNA sequence (gDNA_PF2, GenBank accession no. MN660238) of PxABCC2 and partial sequences of ABCC2_R2/4 cDNA from Cry1S1000.

Asterisks denote consensus sequences. Exons (yellow) and introns (gray) are shown in the fragment sequence of gDNA_PF2. Deletions and insertions in cDNA are respectively highlighted in green and blue. The primer sequences are underlined in red.

https://doi.org/10.1371/journal.ppat.1008697.s016

(DOC)

S5 Fig. Alignment of partial gDNA sequence (gDNA_PF3, GenBank accession no. MN660239) of PxABCC2 and partial sequences of ABCC2_R3/4 cDNA from Cry1S1000.

Exons (yellow) and introns (gray) are shown in the sequence of gDNA_PF3. Insertions in cDNA are highlighted in blue. The primer sequences are underlined in red.

https://doi.org/10.1371/journal.ppat.1008697.s017

(DOC)

S6 Fig. Alignment of partial gDNA sequence (gDNA_PF4, GenBank accession no. MN660240) of PxABCC2 and partial sequence of ABCC2_R4 cDNA from Cry1S1000.

Asterisks denote consensus sequences. Exons (yellow) and introns (gray) are shown in the fragment sequence of gDNA_PF4. Insertions in cDNA are highlighted in blue. The primer sequences are underlined in red.

https://doi.org/10.1371/journal.ppat.1008697.s018

(DOC)

S7 Fig. RA2 mutation in intron 6 of Cry1S1000 PxABCC2.

(A) Genotyping of PxABCC2 in G88 and Cry1S1000 individuals using AS-PCR. (B) Alignment of gDNA sequences of the SA2 allele (GenBank accession no. MN660243) and the RA2 allele (GenBank accession no. MN660241). Letter in red indicates the point mutation in 3′ splice site of intron 6. Two sequences highlighted in orange are the primer of 33fPxABCC2, and a sequence matched by the primer of 55PxABCC2.

https://doi.org/10.1371/journal.ppat.1008697.s019

(TIF)

S8 Fig. Alignment of PxABCC3 cDNA sequences from the strains of G88 (ABCC3_S, GenBank accession no. MN652065) and Cry1S1000 (ABCC3_R, GenBank accession no. MN652071).

Asterisks denote consensus sequences. Codons highlighted in red indicate premature stop codon.

https://doi.org/10.1371/journal.ppat.1008697.s020

(DOC)

S9 Fig. Alignment of predicted amino acid sequences of PxABCC3 from the G88 and Cry1S1000 strains.

Asterisks denote consensus sequences. Transmembrane domains (TM) (blue), Walker A and B sequences (yellow), and C motif (green) are shown in the sequence of PxABCC3 protein encoded by the susceptible allele. Transmembrane helices, except for TM1 and TM2, were predicted by Phobius.

https://doi.org/10.1371/journal.ppat.1008697.s021

(DOC)

S10 Fig. RA3 mutation in exon 14 of Cry1S1000 PxABCC3.

(A) Genotyping of PxABCC3 in G88 and Cry1S1000 individuals using AS-PCR. (B) Alignment of gDNA sequences of the SA3 allele (GenBank accession no. MN660244) and the RA2 allele (GenBank accession no. MN660242). Letter in red indicates the point mutation in exon 14. Two sequences highlighted in orange are the primer of 58PxABCC3, and a sequence matched by the primer of 37rPxABCC3.

https://doi.org/10.1371/journal.ppat.1008697.s022

(TIF)

S11 Fig. Representative sequencing trace of the PCR fragments from wild type (G88), heterozygotes and homozygous mutant strains containing sgRNA target sites 1 (A), 2 (B) and 3 (C) for PxABCC2, respectively.

https://doi.org/10.1371/journal.ppat.1008697.s023

(TIF)

S12 Fig. Protein identification of PxABCC2 in different P. xylostella strains.

(A) Analysis of PxABCC2 in the wild type G88 strain, four PxABCC2 knockout strains and G88-No-Indel using the western blotting. PxABCC2 antibody recognizes a region from extracellular loop 3 to partial nucleotide-binding domain 1 as indicated in the amino acid sequence of full-length PxABCC2 protein (S2 Fig). (B) SDS-PAGE profile of midgut BBMV protein from G88 and one PxABCC2 knockout strain, G88-ABCC2--1. The dashed rectangles indicate the region of SDS-PAGE gel excised and analyzed by nano-LC-MS/MS. (C) Details of 34 peptides specific to PxABCC2 identified from the G88 strain. None of PxABCC2 peptides were detected from the G88-ABCC2--1 strain. (D) Map of the full-length PxABCC2 protein showing the position of 31 peptides (red arrows) specific to PxABCC2 identified from the G88 strain. Black arrow indicates where the protein of G88-ABCC2--1 is expected to be truncated. The predicted tryptic fragments (no detection from the mutant strain using nano-LC-MS/MS analysis) for expected truncated PxABCC2 from G88-ABCC2--1 are indicated with blue arrows. Numbers indicate the position of amino acid residues.

https://doi.org/10.1371/journal.ppat.1008697.s024

(TIF)

S13 Fig. Representative sequencing trace of the PCR fragments containing sgRNA target sites for PxABCC3 in PxABCC3 mutant strains.

https://doi.org/10.1371/journal.ppat.1008697.s025

(TIF)

S14 Fig. Protein identification of PxABCC3 in different P. xylostella strains.

(A) SDS-PAGE profile of BBMV protein from the strain of G88 and two PxABCC3 knockout strains of G88-ABCC3--1 and G88-ABCC3--2. The dashed rectangles indicate the region of SDS-PAGE gel excised and analyzed by nano-LC-MS/MS. (B) Details of the 33 peptides specific to PxABCC3 identified from the G88 strain. None of PxABCC3 peptides were detected from two PxABCC3 knockout strains. (C) Map of the full-length PxABCC3 protein showing the position of 33 peptides (red arrows) specific to PxABCC3 identified from the G88 strain. Black arrows indicate where the proteins of G88-ABCC3--1 and G88-ABCC3--2 are expected to be truncated. The predicted tryptic fragments (no detection from two mutant strains using nano-LC-MS/MS analysis) for expected truncated PxABCC3 proteins from G88-ABCC3--1 and G88-ABCC3--2 are indicated with blue arrows. Numbers indicate the position of amino acid residues.

https://doi.org/10.1371/journal.ppat.1008697.s026

(TIF)

S15 Fig. Protein identification of PxABCC2 and PxABCC3 in different P. xylostella strains.

(A) SDS-PAGE profile of BBMV protein from the G88 strain and one PxABCC2/PxABCC3 double mutant strain, G88-A2-A3--1. The dashed rectangles indicate the region of SDS-PAGE gel excised and analyzed by nano-LC-MS/MS. (B) Details of the 27 peptides specific to PxABCC2 identified from the G88 strain. None of PxABCC2 peptides were detected from the G88-A2-A3--1 strain. (C) Map of the full-length PxABCC2 protein showing the position of 28 peptides (red arrows) specific to PxABCC2 identified from the G88 strain. Black arrow indicates where the protein of G88-A2-A3--1 is expected to be truncated. The predicted tryptic fragments (no detection from the mutant strain using nano-LC-MS/MS analysis) for expected truncated PxABCC2 proteins from G88-A2-A3--1 are indicated with blue arrows. Numbers indicate the position of amino acid residues. (D) Details of the 40 peptides specific to PxABCC3 identified from the G88 strain. None of PxABCC3 peptides were detected from the G88-A2-A3--1 strain. (E) Map of the full-length PxABCC3 protein showing the position of 40 peptides (red arrows) specific to PxABCC3 identified from the G88 strain. Black arrow indicates where the protein of G88-A2-A3--1 is expected to be truncated. The predicted tryptic fragments (no detection from the mutant strain using nano-LC-MS/MS analysis) for expected truncated PxABCC3 proteins from G88-A2-A3--1 are indicated with blue arrows. Numbers indicate the position of amino acid residues.

https://doi.org/10.1371/journal.ppat.1008697.s027

(TIF)

S16 Fig. Schematic strategy of continuous backcrossing for introgression of RA2 and RA3 alleles from the resistant Cry1S1000 strain into the genetic background of the susceptible G88 strain of P. xylostella.

Square and circle represent a male adult and a female adult, respectively. BC1-6: 1–6 generations of backcrossing. BC6F1 and BC6F2 are two additional generations for sibling crosses in single pair to generate a homozygous strain with RA2RA2 (G88-RA2). SA2SA2/RA2SA2 underlined with the red line represent the filtered individuals of each generation after molecular identification using AS-PCR.

https://doi.org/10.1371/journal.ppat.1008697.s028

(TIF)

S17 Fig. Schematic strategy for screening of homozygous mutant strains.

Each rectangle represents a paired female and male. Red rectangles represent the single-pair families kept for the following generations whose parent(s) harbored the desire frameshift mutations as determined by PCR and direct sequencing. “A” shown in the rectangles is the wild type allele; “a1”, “a2” and “a3” are mutant alleles with different type of indel mutations.

https://doi.org/10.1371/journal.ppat.1008697.s029

(TIF)

Acknowledgments

We are very grateful to Prof Anthony M. Shelton and his team at Cornell University for their kind supply of a susceptible strain (G88) and a Cry1Ac-resistant strain (Cry1Ac-R) of the diamondback moth.

References

  1. 1. Usta C. Microorganisms in biological pest control-a review (bacterial toxin application and effect of environmental factors). In: Silva-Opps M, editor. Current progress in biological research. Rijeka: InTech; 2013. p. 133–147.
  2. 2. Palma L, Munoz D, Berry C, Murillo J, Caballero P. Bacillus thuringiensis toxins: an overview of their biocidal activity. Toxins. 2014; 6:3296–3325. pmid:25514092.
  3. 3. Tabashnik BE, Carriere Y. Surge in insect resistance to transgenic crops and prospects for sustainability. Nat Biotechnol. 2017; 35:926–935. pmid:29020006.
  4. 4. Tabashnik BE, Carrière Y, Gassmann A. Global patterns of resistance to Bt crops highlighting pink bollworm in the United States, China, and India. J Econ Entomol. 2019; 112:2513–2523. pmid:31254345.
  5. 5. Tabashnik BE, Cushing NL, Finson N, Johnson MW. Field development of resistance to Bacillus thuringiensis in diamondback moth (Lepidoptera: Plutellidae). J Econ Entomol. 1990; 83:1671–1676.
  6. 6. Janmaat AF, Myers J. Rapid evolution and the cost of resistance to Bacillus thuringiensis in greenhouse populations of cabbage loopers, Trichoplusia ni. Proc R Soc Lond B. 2003; 270:2263–2270. pmid:14613613.
  7. 7. Calles-Torrez V, Knodel JJ, Boetel MA, French BW, Fuller BW, Ransom JK. Field-evolved resistance of Northern and Western corn rootworm (Coleoptera: Chrysomelidae) populations to corn hybrids expressing single and pyramided Cry3Bb1 and Cry34/35Ab1 Bt proteins in North Dakota. J Econ Entomol. 2019; 112:1875–1886. pmid:31114868.
  8. 8. Peterson B, Bezuidenhout CC, Van den Berg J. An overview of mechanisms of Cry toxin resistance in lepidopteran insects. J Econ Entomol. 2017; 110:362–377. pmid:28334065.
  9. 9. Pardo-López L, Soberón M, Bravo A. Bacillus thuringiensis insecticidal three-domain Cry toxins: mode of action, insect resistance and consequences for crop protection. FEMS Microbiol Rev. 2013; 37:3–22. pmid:22540421.
  10. 10. Heckel DG. Learning the ABCs of Bt: ABC transporters and insect resistance to Bacillus thuringiensis provide clues to a crucial step in toxin mode of action. Pestic Biochem Physiol. 2012; 104:103–110.
  11. 11. Atsumi S, Miyamoto K, Yamamoto K, Narukawa J, Kawai S, Sezutsu H, et al. Single amino acid mutation in an ATP-binding cassette transporter gene causes resistance to Bt toxin Cry1Ab in the silkworm, Bombyx mori. Proc Natl Acad Sci USA. 2012; 109:e1591–e1598. pmid:22635270.
  12. 12. Gahan LJ, Pauchet Y, Vogel H, Heckel DG. An ABC transporter mutation is correlated with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS Genet. 2010; 6:e1001248. pmid:21187898.
  13. 13. Xiao Y, Zhang T, Liu C, Heckel DG, Li X, Tabashnik BE, et al. Mis-splicing of the ABCC2 gene linked with Bt toxin resistance in Helicoverpa armigera. Sci Rep. 2014; 4:6184. pmid:25154974.
  14. 14. Park Y, Gonzalez-Martinez RM, Navarro-Cerrillo G, Chakroun M, Kim Y, Ziarsolo P, et al. ABCC transporters mediate insect resistance to multiple Bt toxins revealed by bulk segregant analysis. BMC Biol. 2014; 12:46. pmid:24912445.
  15. 15. Huang J, Xu Y, Zuo Y, Yang Y, Tabashnik BE, Wu Y. Evaluation of five candidate receptors for three Bt toxins in the beet armyworm using CRISPR-mediated gene knockouts. Insect Biochem Mol Biol. 2020; 121:103361. pmid:32199887.
  16. 16. Guo Z, Sun D, Kang S, Zhou J, Gong L, Qin J, et al. CRISPR/Cas9-mediated knockout of both the PxABCC2 and PxABCC3 genes confers high-level resistance to Bacillus thuringiensis Cry1Ac toxin in the diamondback moth, Plutella xylostella (L.). Insect Biochem Mol Biol. 2019; 107:31–38. pmid:30710623.
  17. 17. Baxter SW, Badenes-Perez FR, Morrison A, Vogel H, Crickmore N, Kain W, et al. Parallel evolution of Bacillus thuringiensis toxin resistance in lepidoptera. Genetics. 2011; 189:675–679. pmid:21840855.
  18. 18. Martínez-Solís M, Pinos D, Endo H, Portugal L, Sato R, Ferré J, et al. Role of Bacillus thuringiensis Cry1A toxins domains in the binding to the ABCC2 receptor from Spodoptera exigua. Insect Biochem Mol Biol. 2018; 101:47–56. pmid:30077769.
  19. 19. Bretschneider A, Heckel DG, Pauchet Y. Three toxins, two receptors, one mechanism: mode of action of Cry1A toxins from Bacillus thuringiensis in Heliothis virescens. Insect Biochem Mol Biol. 2016; 76:109–117. pmid:27456115.
  20. 20. Tanaka S, Miyamoto K, Noda H, Jurat-Fuentes JL, Yoshizawa Y, Endo H, et al. The ATP-binding cassette transporter subfamily C member 2 in Bombyx mori larvae is a functional receptor for Cry toxins from Bacillus thuringiensis. FEBS J. 2013; 280:1782–1794. pmid:23432933.
  21. 21. Stevens T, Song S, Bruning JB, Choo A, Baxter SW. Expressing a moth abcc2 gene in transgenic Drosophila causes susceptibility to Bt Cry1Ac without requiring a cadherin-like protein receptor. Insect Biochem Mol Biol. 2017; 80:61–70. pmid:27914919
  22. 22. Guo Z, Kang S, Chen D, Wu Q, Wang S, Xie W, et al. MAPK signaling pathway alters expression of midgut ALP and ABCC genes and causes resistance to Bacillus thuringiensis Cry1Ac toxin in diamondback moth. PLoS Genet. 2015; 11:e1005124. pmid:25875245.
  23. 23. Chen Z, He F, Xiao Y, Liu C, Li J, Yang Y, et al. Endogenous expression of a Bt toxin receptor in the Cry1Ac-susceptible insect cell line and its synergistic effect with cadherin on cytotoxicity of activated Cry1Ac. Insect Biochem Mol Biol. 2015; 59:1–17. pmid:25662100.
  24. 24. Endo H, Tanaka S, Imamura K, Adegawa S, Kikuta S, Sato R. Cry toxin specificities of insect ABCC transporters closely related to lepidopteran ABCC2 transporters. Peptides. 2017; 98:86–92. pmid:28416297
  25. 25. Wang J, Ma H, Zhao S, Huang J, Yang Y, Tabashnik BE, et al. Functional redundancy of two ABC transporter proteins in mediating toxicity of Bacillus thuringiensis to cotton bollworm. PLoS Pathog. 2020; 16:e1008427. pmid:32191775.
  26. 26. Furlong MJ, Wright DJ, Dosdall LM. Diamondback moth ecology and management: problems, progress, and prospects. Annu Rev Entomol. 2013; 58:517–541. pmid:23020617.
  27. 27. Bravo A, Gill SS, Soberon M. Mode of action of Bacillus thuringiensis Cry and Cyt toxins and their potential for insect control. Toxicon. 2007; 49:423–435. pmid:17198720.
  28. 28. Wang X, Xu Y, Huang J, Jin W, Yang Y, Wu Y. CRISPR-mediated knockout of the ABCC2 gene in Ostrinia furnacalis confers high-level resistance to the Bacillus thuringiensis Cry1Fa toxin. Toxins. 2020; 12:246. pmid:32290427
  29. 29. Park Y, Herrero S, Kim Y. A single type of cadherin is involved in Bacillus thuringiensis toxicity in Plutella xylostella. Insect Mol Biol. 2015; 24:624–633. pmid:26331576.
  30. 30. Guo Z, Kang S, Zhu X, Wu Q, Wang S, Xie W, et al. The midgut cadherin-like gene is not associated with resistance to Bacillus thuringiensis toxin Cry1Ac in Plutella xylostella (L.). J Invertebr Pathol. 2015; 126:21–30. pmid:25595643.
  31. 31. Baxter SW, Zhao JZ, Gahan LJ, Shelton AM, Tabashnik BE, Heckel DG. Novel genetic basis of field-evolved resistance to Bt toxins in Plutella xylostella. Insect Mol Biol. 2005; 14:327–334. pmid:15926902.
  32. 32. Wang S, Kain W, Wang P. Bacillus thuringiensis Cry1A toxins exert toxicity by multiple pathways in insects. Insect Biochem Mol Biol. 2018; 102:59–66. pmid:30278206.
  33. 33. Endo H, Tanaka S, Adegawa S, Ichino F, Tabunoki H, Kikuta S, et al. Extracellular loop structures in silkworm ABCC transporters determine their specificities for Bacillus thuringiensis Cry Toxins. J Biol Chem. 2018; 293:8569–8577. pmid:29666188.
  34. 34. Liu L, Chen Z, Yang Y, Xiao Y, Liu C, Ma Y, et al. A single amino acid polymorphism in ABCC2 loop 1 is responsible for differential toxicity of Bacillus thuringiensis Cry1Ac toxin in different Spodoptera (Noctuidae) species. Insect Biochem Mol Biol. 2018; 100:59–65. pmid:29964167.
  35. 35. Ocelotl J, Sanchez J, Gómez I, Tabashnik BE, Bravo A, Soberón M. ABCC2 is associated with Bacillus thuringiensis Cry1Ac toxin oligomerization and membrane insertion in diamondback moth. Sci Rep. 2017; 7:2386. pmid:28539590.
  36. 36. Shelton AM, Cooley RJ, Kroening MK, Wilsey WT, Eigenbrode SD. Comparative analysis of two rearing procedures for diamondback moth (Lepidoptera: Plutellidae). J Entomol Sci. 1991; 26:17–26.
  37. 37. Tang JD, Gilboa S, Roush RT, Shelton AM. Inheritance, stability, and lack-of-fitness costs of field-selected resistance to Bacillus thuringiensis in diamondback moth (Lepidoptera: Plutellidae) from Florida. J Econ Entomol. 1997; 90:732–741.
  38. 38. Metz TD, Roush RT, Tang JD, Shelton AM, Earle ED. Transgenic broccoli expressing a Bacillus thuringiensis insecticidal crystal protein: implications for pest resistance management strategies. Mol Breed. 1995; 1:309–317.
  39. 39. Zhao JZ, Li YX, Collins HL, Shelton AM. Examination of the F2 screen for rare resistance alleles to Bacillus thuringiensis toxins in the diamondback moth (Lepidoptera: Plutellidae). J Econ Entomol. 2002; 95:14–21. pmid:11942749
  40. 40. Kain W, Zhao J, Janmaat AF, Myers JH, Shelton AM, Wang P. Inheritance of resistance to Bacillus thuringiensis Cry1Ac toxin in a greenhouse-derived strain of cabbage looper (Lepidoptera: Noctuidae). J Econ Entomol. 2004; 97:2073–2078. pmid:15666767.
  41. 41. Abbott WS. A method of computing the effectiveness of an insecticide. J Econ Entomol. 1925; 18:265–267.
  42. 42. Liu Y, Tabashnik BE. Inheritance of resistance to the Bacillus thuringiensis toxin Cry1C in the diamondback moth. Appl Environ Microbiol. 1997; 63:2218–2223. pmid:16535623.
  43. 43. Huang Y, Chen Y, Zeng B, Wang Y, James AA, Gurr GM, et al. CRISPR/Cas9 mediated knockout of the abdominal-A homeotic gene in the global pest, diamondback moth (Plutella xylostella). Insect Biochem Mol Biol. 2016; 75:98–106. pmid:27318252.
  44. 44. Wolfersberger M, Luethy P, Maurer A, Parenti P, Sacchi FV, Giordana B, et al. Preparation and partial characterization of amino acid transporting brush border membrane vesicles from the larval midgut of the cabbage butterfly (Pieris brassicae). Comp Biochem Physiol Part A: Physiol. 1987; 86:301–308.
  45. 45. Zhang H, Tian W, Zhao J, Jin L, Yang J, Liu C, et al. Diverse genetic basis of field-evolved resistance to Bt cotton in cotton bollworm from China. Proc Natl Acad Sci USA. 2012; 109:10275–10280. pmid:22689968.
  46. 46. Tabashnik BE, Cushing NL, Johnson MW. Diamondback moth (Lepidoptera: Plutellidae) resistance to insecticides in Hawaii: Intra-Island variation and cross-resistance. J Econ Entomol. 1987; 80:1091–1099.