Skip to main content

Greenhouse test of spraying dsRNA to control the western flower thrips, Frankliniella occidentalis, infesting hot peppers

Abstract

Background

The western flower thrips Frankliniella occidentalis is an insect pest that damages various crops, including hot peppers. It is a vector of a plant pathogen, tomato spotted wilt virus. To control this pest, chemical insecticides have been used in the past, but the control efficacy is unsatisfactory owing to rapid resistance development by F. occidentalis.

Methodology

: This study reports a novel control technology against this insect pest using RNA interference (RNAi) of the vacuolar-type ATPase (vATPase) expression. Eight subunit genes (vATPase-A ∼ vATPase-H) of vATPase were obtained from the F. occidentalis genome and confirmed for their expressions at all developmental stages.

Results

Double-stranded RNAs (dsRNAs) specific to the eight subunit genes were fed to larvae and adults, which significantly suppressed the corresponding gene expressions after 24-h feeding treatment. These RNAi treatments resulted in significant mortalities, in which the dsRNA treatments at ∼2,000 ppm specific to vATPase-A or vATPase-B allowed complete control efficacy near 100% mortality in 7 days after treatment. To prevent dsRNA degradation by the digestive proteases during oral feeding, dsRNAs were formulated in a liposome and led to an enhanced mortality of the larvae and adults of F. occidentalis. The dsRNAs were then sprayed at 2,000 ppm on F. occidentalis infesting hot peppers in a greenhouse, which resulted in 53.5–55.9% control efficacy in 7 days after treatment. Even though the vATPases are conserved in different organisms, the dsRNA treatment was relatively safe for non-target insects owing to the presence of mismatch sequences compared to the dsRNA region of F. occidentalis.

Conclusion

These results demonstrate the practical feasibility of spraying dsRNA to control F. occidentalis infesting crops.

Peer Review reports

Background

The western flower thrips, Frankliniella occidentalis, is classified as Thripidae among the different families of the Tenebrantia suborder. It has a wide host spectrum and causes massive economic damage to various ornamental and other horticultural crops [1]. In particular, it causes both direct and indirect damage to crops by feeding or ovipositing on the host plants and by transmitting a plant pathogen called tomato spotted wilt virus (TSWV) [2]. The high potential of this insect to invade and colonize new habitats has been well documented in many countries [3]. Indeed, it was found in most of the counties in Korea within a decade since its first detection in 1993 [4].

Relatively small body size, cryptic behavior, short generation time, and high reproduction rate allow the thrips to frequently attack crops [5]. Chemical insecticides have been used to control the thrips but are not always satisfactory in their control efficacies because of rapid development of insecticide resistance [3]. Usually, plants develop a series of constitutive and inducible defense mechanisms against insect-feeding stress. Breeding insect-resistant varieties has been considered effective for controlling thrips. However, insects also evolve antidefense mechanisms, including behavioral, physiological, and biochemical adaptations, to increase survival and reproduction against the host plants [6]. Indeed, F. occidentalis induced detoxifying potential by elevating its glutathione-S-transferase activity against secondary metabolites, such as tannins, alkaloids, total phenols, flavonoids, and lignin [7].

RNA interference (RNAi) is defined as sequence-specific silencing of a target gene expression by shortening the mRNA lifetime [8]. In the RNAi pathway, the target mRNA is degraded by hybridization with its specific antisense RNA derived from a double-stranded RNA (dsRNA) after a catalytic activity of an RNase (= Dicer) [9]. RNAi has provided a novel and powerful reverse genetic tool for identifying gene functions and shows great potential in pest management [10]. By targeting genes essential for the survival of specific pest insects, RNAi can be used to selectively kill such pest insects without adversely affecting the non-target species [11].

The present study was aimed at developing novel control tactics against F. occidentalis using RNAi against the vacuolar-type ATPase (vATPase) gene. vATPase is highly conserved among eukaryotic organisms via its catalytic activity to acidify various intracellular compartments by pumping protons across the plasma membranes with the help of ATP hydrolysis [12]. It has two domains, namely a membrane V0 domain performing H+ channel hydrolysis and an external V1 domain performing ATP hydrolysis [13]. The V1 domain is composed of eight subunits (A-H), in which three copies of the A and B subunits are catalytic [14]. Baum et al. [15] demonstrated the control efficacy of dsRNA specific to vATPase-A against a corn rootworm. In addition, various other insect pests, such as the fruit fly (Drosophila melanogaster), beetle (Tribolium castaneum), aphid (Acyrthosiphon pisum), and moth (Manduca sexta), were selectively exterminated when fed species-specific dsRNA-targeting vATPase transcripts [11]. This motivated us to re-evaluate the control efficacy of dsRNA specific to vATPase gene of F. occidentalis.

Methods

Insect rearing

The larvae and adults of F. occidentalis were collected from the National Academy of Agricultural Sciences (Jeonju, Korea). The rearing conditions were maintained at a temperature of 25 ± 2 °C, with photoperiod of 14:10 h (L:D) and 65 ± 5% relative humidity. A circular breeding container (100 ⋅ 40 mm, SPL, Seoul, Korea) was used to rear the thrips via continuous supply of soybean, Phaseolus coccineus, as a diet. The beans were germinated, and the sprouted seed kernels were provided as a diet for the larvae and adults.

Bioinformatics

Eight subunit sequences of the F. occidentalis vATPase V1 domain (A-H) were obtained with GenBank accession numbers of XM_026418961.1, KP234253.1, XM_026421419.1, XM_026424871.1, XM_026438253.1, XM_026421632.1, XM_026416289.1, and XM_026438737.1. The resulting sequences were subjected to open reading frame (ORF) analysis using ORFfinder (https://www.ncbi.nlm.nih.gov/orffinder/).

RNA extraction, cDNA synthesis, RT-PCR, and RT-qPCR

RNA samples were extracted from all developmental stages with around 50 individuals per stage using Trizol reagent (Invitrogen, Carlsbad, CA, USA) under manufacturer’s instructions. The RNA concentrations were measured using a spectrophotometer (NanoDrop, Thermo Scientific, Wilmington, DE, USA) after resuspending in nuclease-free water. The extracted RNAs (60 ng per reaction) were used for cDNA synthesis using RT-Premix (Intron Biotechnology, Seoul, Korea) containing the oligo dT primer based on manufacturer’s instructions.

The synthesized cDNAs were used for PCR amplification using DNA Taq polymerase (GeneALL, Seoul, Korea) with gene-specific forward and reverse primers under the following conditions: 94 Â°C for 5 min for the initial heat treatment, followed by 35 cycles of denaturation at 94 Â°C for 1 min, different annealing temperatures (52–55 Â°C) for 1 min, and extension at 72 Â°C for 1 min. The PCR was terminated with a final chain extension at 72 Â°C for 10 min. Each PCR mixture (25 µL) consisted of the DNA template, dNTP (2.5 mM each), 10 pmol of each forward and reverse primer, and Taq polymerase (2.5 unit/µL).

Quantitative PCR (qPCR) was conducted using a Real-time PCR machine (Step One Plus Real-Time PCR System, Applied Biosystems, Singapore) and the Power SYBR Green PCR Master Mix (Life Technologies, Carlsbad, CA, USA) according to the procedures of Bustin et al. [16]. The reaction mixture (20 µL) contained 10 µL of the Power SYBR Green PCR Mix, 2 µL of the cDNA template (60 ng/µL), and 1 µL each of the forward and reverse primers. An elongation factor (EF1) was used as the reference gene. The quality of the PCR products was assessed by melting curve analysis. Quantitative analysis was performed using the comparative CT (2−∆∆CT) method [17], where each experiment was replicated three times with individually prepared samples.

dsRNA preparation

Template DNAs were amplified with gene-specific primers containing T7 promoter sequence at 5’ end. These were used for in vitro transcription using the MEGAscript RNAi Kit (Ambion, Austin, TX, USA) according to manufacturer’s instructions. The resulting dsRNAs were mixed with Metafectene PRO (Biontex, Plannegg, Germany), a transfection reagent, at 1:1 ratio and incubated at room temperature for 30 min to form liposomes. For field assays, two dsRNAs specific to vATPase-A and vATPase-B were produced at a large scale by Genolution Inc. (Seoul, Korea). In the production, three grades of dsRNA quality were produced: ‘Grade 1’ represented the dsRNA mixture without any purification, ‘Grade 2’ was a dsRNA product purified by alcohol precipitation, and ‘Grade 3’ was a dsRNA product purified by 0.22 Î¼m membrane filtration.

Insecticidal efficacy test of dsRNA against F. occidentalis in the laboratory

Before laboratory bioassays, RNAi efficacy was assessed. Different dsRNAs (dsATPaseA∼dsATPaseH) specific to the eight vATPase subunits were prepared using the MEGAscript kit described above. These dsRNAs were fed to the larvae and adults through the bean diet. The beans were soaked in the dsRNA suspension for 20 min. After removing the excess moisture, the treated beans were supplied to 10 individuals of F. occidentalis in a circular breeding container (100 â‹… 40 mm) (= an experimental unit). After 24 h of dsRNA-treated diet feeding, untreated fresh beans were supplied to the test thrips. After the 24-h dsRNA treatment, the RNAi efficiencies were measured over 0–24 h by RT-qPCR. Each treatment was replicated three times. As a control, the viral gene CpBV302 [18] was used.

For bioassay after the 24-h of dsRNA-treated diet feeding, untreated fresh beans were supplied to the test thrips. Mortality was monitored for 7 days after treatment (DAT). Each treatment used 10 individuals and was replicated three times.

Control efficacy of dsRNA against F. occidentalis in the greenhouse

The study greenhouse contained five rows of cultivated young (about 30 cm height) hot peppers, where the average distance between adjacent rows was around 50 cm, and the average distance between adjacent plants was 40 cm. Each plant was considered as one experimental unit and infested with 80–90 thrips before application of the dsRNA mixture. These experimental units were deployed under a randomized complete block design. Each treatment was replicated with three blocks. To cover the entire hot pepper plant including both sides of leaves, each spray volume per plant was ~ 6.5 mL, which contained 1,994.2 Âµg/mL for dsATPaseA or 2,025.1 Âµg/mL for dsATPaseB. For the control, the same volume of water was sprayed.

Influences of dsATPaseA and dsATPaseB on non-target insects

The non-target insects included Thrips tabaci, Plutella xylostella, Tenebrio molitor, and Tribolium castanum. These test insects were obtained from laboratory cultures of T. tabaci [19], P. xylostella [18], T. molitor [20], and T. castaneum [21]. The larval stages of these species were used for the bioassays. dsATPase (2,000 ppm) was fed to the larvae with the treated diets via welsh onion for T. tabaci, cabbage for P. xylostella and T. molitor, and nuts for T. castaneum. Each treatment used 10 individuals and was replicated three times. The mortality was assessed at 7 DAT under the rearing conditions.

Statistical analysis

All studies were analyzed with one-way ANOVA using PROC GLM of SAS program [22]. The mortality data were subjected to arcsine transformation and used for ANOVA. The means were compared with the least-squares difference (LSD) test. The study experiments involved three biologically independent replicates that were plotted as mean ± standard error using Sigma plot. The BioRender program was used for the schematic diagram (https://biorender.com/).

Results

Expression profiles of eight subunit genes of vATPase in F. occidentalis

Eight subunits (‘A-H’) of vATPase V1 domain (Fig. 1A) were obtained from F. occidentalis genome. These genes were highly similar (73–99%) to other insect orthologs in the predicted amino acid sequences (Table S2). All eight genes were expressed in the larvae, pupae, and adults (Fig. 1B). However, the expression levels of the different subunits varied at each stage. Moreover, similar expression patterns were exhibited among the different stages, with relatively high expressions of vATPase-E and vATPase-H.

Fig. 1
figure 1

Redrawn from Chen et al. [23]. B Expression analysis of vATPase subunits at different developmental stages. An elongation factor EF1 was used to normalize the expression levels at different stages. Each measurement was replicated with three independent samplings. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test)

Identification and expression profiles of vATPase subunits of F. occidentalis. A Molecular structures of the vATPase domains (subunits): V0 (a-e) and V1 (A-H).

Individual RNAi of the eight vATPase subunits and subsequent insecticidal activities

RNAi specific to each of the eight vATPase subunits was performed by feeding a diet soaked in dsRNA suspension at 500 ppm (Fig. 2). After 24-h of feeding dsRNA, the expression level of each subunit was monitored over 24 h by RT-qPCR (Fig. 2A). All RNAi treatments reduced the expression levels of the target vATPase subunits by more than 50%. Except for vATPase-G, the other vATPase subunits exhibited the highest reductions in expression levels 12 h after dsRNA feeding. Under these RNAi conditions, the treated thrips suffered from significant (P < 0.05) mortalities. In both the larval and adult stages, more than 80% mortalities were recorded for treatments with dsRNA specific to vATPase-A or vATPase-B, which were used for the subsequent bioassays (Fig. 2B).

Fig. 2
figure 2

RNAi and bioassays by feeding dsRNA (500 ppm) specific to vATPase subunit (A-H) genes against F. occidentalis. A RNAi efficiencies of eight individual dsRNAs in adults. A viral gene CpBV302 was used as the control dsRNA (dsCON). An elongation factor EF1 was used to normalize the expression levels. Each measurement was replicated with three independent samplings. B Bioassays of the individual RNAi against larvae and adults. Mortality was assessed at 7 days after treatment (DAT) and corrected by the mortality of control treatment obtained by feeding dsCON. Each treatment used 10 individuals and was replicated three times. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test)

Production of effective dsRNAs at a large scale and their control efficacies against F. occidentalis

For the field assays, the two most effective dsRNAs specific to vATPase-A and vATPase-B were produced at a large scale in three grades, which differed in purity levels from Grade 1 to Grade 3 (Fig. 3A). All three grades were prepared at 500 ppm of active gradient and applied to the thrips. These dsRNAs also showed high control efficacies against both larvae and adults, which were observed in treatments using dsRNAs prepared with the MEGAscript kit. However, the control efficacies were not different among the different purity levels of the dsRNAs produced at a large scale.

For the Grade-1 dsRNAs produced at a large scale, the dose-mortality was assessed against the adults of F. occidentalis (Fig. 3B). With increase in the dsRNA amount, the thrips suffered high mortalities. At around 2,000 ppm, both dsRNAs resulted in 100% mortality. With respect to the median lethal dose (LC50), however, the dsRNA specific to vATPase-B was more potent than that specific to vATPase-A (Fig. 3C).

Fig. 3
figure 3

Evaluation of the company-made dsRNAs for insecticidal activities against F. occidentalis. Two dsRNAs specific to vATPase-A and vATPase-B were produced at a large scale. A Comparative analysis of the insecticidal activities of three different grades of dsRNAs based on purity. A viral gene CpBV302 was used as the control dsRNA. Mortality was assessed at 7 days after treatment (DAT) and corrected by the mortality of control treatment obtained by feeding the control dsRNA. B Dose-mortality curves of the two Grade-1 dsRNAs against adult thrips. Test dsRNA suspensions were applied to the thrips by the feeding method. Each treatment used 10 individuals and was replicated three times. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test) for each developmental stage or target gene. C Median lethal concentrations (LC50) of the two dsRNAs against adult thrips

Comparative toxicity analyses of potent dsRNAs against F. occidentalis

A previous screening [24] among 57 gene-specific dsRNAs proposed TLR6 as a potent target to effectively exterminate F. occidentalis. The insecticidal activity of the dsRNA specific to TLR6 was compared with those of dsRNAs specific to vATPase-A and vATPase-B by the feeding bioassay used in our current study at 500 ppm of dsRNA concentration (Fig. 4A). In both larvae and adults, the dsRNA specific to vATPase-B or TLR6 was more potent than the treatment with the dsRNA specific to vATPase-A. The two potent dsRNAs were further assessed at different doses (Fig. 4B). There were no significant differences for different doses on the larvae (F = 5.82; df = 1, 28; P = 0.0524) and adults (F = 0.08; df = 1, 6; P = 0.7868). In addition, these two dsRNAs were not very different at the median lethal doses (Fig. 4C).

Fig. 4
figure 4

Insecticidal activities among the three dsRNAs (dsATPase A, dsATPase B, and dsTLR6) specific to vATPase-A, vATPase-B, and TLR6, respectively, against F. occidentalis. A Comparative analysis of the dsRNAs by feeding a dose of 500 ppm to both larvae and adults. A viral gene CpBV302 was used as the control dsRNA. Mortality was assessed at 7 days after treatment (DAT) and corrected by the mortality of control treatment obtained by feeding the control dsRNA. B Dose-mortality curves of the two dsRNAs against the larvae and adults. Test insecticide suspensions were applied to the thrips by the feeding method. Each treatment used 10 individuals and was replicated three times. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test). C Median lethal time (LT50) of the two dsRNAs against larvae and adults

Enhanced insecticidal activity by dsRNA formulation with liposome against F. occidentalis

To prevent degradation of the dsRNA during feeding, EDTA was added to inactivate the degradation enzyme(s) or liposome formulation was devised (Fig. 5). The addition of EDTA to the dsRNAs specific to vATPase-A and vATPase-B did not significantly increase the control efficacy against larvae or adults. However, the liposome formulation using metafectene significantly enhanced the insecticidal activity of the dsRNA specific to vATPase-B against both developmental stages of F. occidentalis. Enhanced control efficacy of liposome formation was detected in the treatment with the dsRNA specific to vATPase-A, but not significantly different compared to the control dsRNA treatment without any addition.

Fig. 5
figure 5

Enhancement of insecticidal activity by liposome formation of dsRNA against F. occidentalis. The ‘Control’ treatment represents plain dsRNA without any formulation. Liposome formation (‘+ Metafectene’) used a lipofectin reagent, metafectene, in a 1:1 volume ratio mixture with the dsRNA suspension (Grade 1). ‘+EDTA’ represents the addition of 3% EDTA to the dsRNA suspension. All dsRNA treatments used a feeding assay at the dose of 500 ppm. A viral gene CpBV302 was used as the control dsRNA. Mortality was assessed at 7 days after treatment (DAT) and corrected by the mortality of control treatment obtained by feeding the control dsRNA. Each treatment used 10 individuals and was replicated three times. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test)

Field assays of vATPase-A and vATPase-B against F. occidentalis infesting hot pepper

Liposome formulation of dsRNA (∼2,000 ppm) specific to vATPase-A or vATPase-B was sprayed on F. occidentalis infesting hot peppers (Fig. 6). The relatively young host plants were evenly sprayed with ∼6.5 mL of the dsRNA suspension (Fig. 6A). Before spraying, each plant had 88–98 thrips, including both larvae and adults (Fig. 6B). Control efficacies were observed at 3 DAT: dsRNA spray targeting vATPase-A showed 17.27% control efficacy while that with vATPase-B showed 13.53% control efficacy. The control efficacies of the dsRNA treatments increased at 7 DAT, recording 53.53% and 55.88% for dsRNAs specific to vATPase-A and vATPase-B, respectively.

Fig. 6
figure 6

Field test of dsRNAs specific to vATPase for controlling F. occidentalis infesting hot peppers in a greenhouse. A Schematic diagram of spraying dsRNAs in the greenhouse, where the row-to-row distance was around 50 cm, and the plant-to-plant distance was 40 cm. Each treatment was replicated under a randomized complete block design, in which each plant represented an experimental unit. Control treatment represents water spraying without any dsRNAs. B Control efficacies of two dsRNAs specific to vATPase-A and vATPase-B. Mortalities were assessed at 3 and 7 days after treatment (DAT).

Target-specific control efficacy of dsRNA specific to vATPase-B

Considering the importance and conserved nature of vATPase in all organisms, any non-target damage owing to dsRNA application must be assessed. For this assay, five other insect species were compared with F. occidentalis for control efficacies after treatment with dsRNA specific to vATPase-B (Fig. 7). As expected, > 80% control efficacy of the dsRNA treatment was detected against F. occidentalis while the dsRNA treatment showed less than 40% control efficacies for the other non-target insects. When the DNA sequences corresponding to dsRNA were compared among these insects, the sequence identities appeared to be positively correlated (r = 0.6582) with the control efficacies, although the correlation coefficient was not significant (P = 0.1552). Interestingly, the identical 20-nucleotide (∼ short interfering RNA (siRNA) size after dsRNA process) site was detected between vATPase genes of F. occidentalis and T. tabaci (Fig. S2). T. tabaci was highly influenced by dsRNA specific to vATPase-B of F. occidentalis among the selected non-target species.

Fig. 7
figure 7

Influence of dsRNA specific to vATPase-B against F. occidentalis (Fo) and non-target insects, namely Thrips tabaci (Tt), Plutella xylostella (Px), Maruca vitrata (Mv), Tenebrio molitor (Tm), and Tribolium castanum (Tc). Test dsRNAs at 2,000 ppm were applied to the test insects by the feeding method. Mortality was assessed at 7 days after treatment (DAT). Each treatment used 10 individuals and was replicated three times. The letters above the standard deviation bars indicate significant differences among the means at Type I error = 0.05 (LSD test). A viral gene CpBV302 was used as the control dsRNA. Mortality was assessed at 7 days after treatment (DAT) and corrected by the mortality of control treatment obtained by feeding the control dsRNA. DNA sequences corresponding to the dsRNAs of F. occidentalis were compared among these insects (Fig. S2)

Discussion

This study tested the insecticidal activities of dsRNAs specific to vATPase against F. occidentalis. In the RNAi application, dsRNAs were delivered by oral administration via a diet coated with the dsRNA solutions. The typical dsRNA delivery methods include microinjection, feeding, and soaking [25]. Among these delivery methods, the current study used feeding of dsRNA against F. occidentalis and showed effective RNAi efficacy resulting in significant insecticidal activities. Relatively high RNAi efficiencies via feeding dsRNA are usually observed in F. occidentalis unlike other sucking insects such as aphids because of the unique feeding behavior of the thrips exhibiting a pierce-sucking type followed by feeding the exuding plant juice from the broken plant tissues probably containing the treated dsRNA on the plant surface [24]. In addition, Whitten et al. [26] successfully fed F. occidentalis media containing dsRNA-expressing bacteria, which resulted in significant mortality of the insects, especially those at the larval stage, by targeting the alpha-tubulin gene. RNAi has been used to control gene expression during the development of eukaryotes, including insects [27]. In response to dsRNA, cellular enzymes Dicer or Dicer-like proteins are activated to cleave it into double-stranded siRNAs (20–24 nucleotides). One of the siRNA strands is loaded onto the RNA-induced silencing complex (RISC) to surveil the complementary sequences, which then undergo translational arrest or are degraded by the action of the argonaute (AGO) core protein in the RISC complex [28, 29]. In insects, RNAi has been considered for pest control by oral administration of dsRNA [10]. The uptake of dsRNA from the insect gut is mediated either by specific transmembrane proteins or by endocytosis [30]. These suggest the practical application of dsRNA for con-trolling F. occidentalis under field conditions. The present study tested the control efficacies of dsRNAs targeting the vATPase of F. occidentalis.

A genome of F. occidentalis encodes all eight subunits of the V1 domain. Individual dsRNAs specific to each of these subunits were tested to compare their insecticidal activities, which showed that dsRNAs specific to vATPase-A and vATPase-B were relatively the most potent. Three copies of these two subunits form a catalytic V1 domain of the enzyme [14]. This vacuolar-type ATP synthase is involved in solute transport across the plasma membranes as well as lumen pH regulation [31]. Thus, the RNAi of the gene transcripts must have been detrimental to the thrips’ survival. Han et al. [24] screened the RNAi efficiencies of 57 genes of F. occidentalis, including vATPase, and proposed four dsRNAs specific to the Toll-like receptor 6 (TLR6), apolipophorin (apoLp), coatomer protein subunit epsilon (COPE), and sorting/assembly machinery component (SAM) genes to be effective for exterminating thrips by oral administration. However, dsRNAs specific to vATPase subunits were highly potent in several insects, including F. occidentalis. Indeed, a comparative analysis in the current study showed that the dsRNA specific to vATPase-B was equivalent to that of TLR6 in control efficacy against both larvae and adults of F. occidentalis. Badillo-Vargas et al. [32] reported a significant decrease in survival and fertility by introducing dsRNA into F. occidentalis via microinjection targeting vATPase-B. The oral administration of dsRNA specific to vATPase was effective by direct feeding in solution rather than feeding via diet in F. occidentalis [33]. vATPase is well known for mediating saliva secretion in insects. In the blowfly Calliphora vicina, serotonin induces saliva secretion by elevating cAMP levels, which activate protein kinase A to assemble and activate vATPase in the apical membrane [34]. Saliva plays a crucial role in the feeding of F. occidentalis because it contains various functional factors for digestion and detoxification against toxic secondary metabolites from the host plants [35]. Thus, its suppression by RNAi would lead to significant lethality against F. occidentalis.

Liposome formulation of dsRNA specific to vATPase-B slightly enhanced the control efficacy against F. occidentalis. This formulation might be effective to minimize dsRNA degradation in the gut lumen from enzymatic attacks [36]. Specifically, dsRNases are suspected to be potent enzymes for degrading dsRNAs and leading to major obstacles influencing successful RNAi [37]. Since dsRNases were first identified in silkworms, several dsRNases have been reported in different insects [38]. These dsRNases participate in dsRNA degradation and reduce RNAi efficiency because the RNAi treatments against these dsRNase genes effectively protect dsRNAs [39]. Liposome formulation is likely to protect the dsRNA from enzymatic attacks by dsRNases, as demonstrated in Spodoptera litura [40]. This explains the enhancement of the control efficacy by liposome formation of the dsRNA specific to vATPase-B against F. occidentalis in our current study.

Spraying dsRNAs specific to vATPase-A and vATPase-B resulted in significant control efficacies in F. occidentalis infesting hot peppers in the greenhouse. However, the control efficacies under field conditions were much lower (∼55%) than those in the laboratory assays at the same treatment concentrations. The use of sprayable dsRNA-based pest control technology was proven to be effective in the Colorado potato beetle Leptinotarsa decemlineata [41]. However, this dsRNA control technique was applicable owing to the differential sensitivities among the different insects [42]. In particular, the highly susceptible insects appeared to exhibit systemic RNAi, in which the dsRNA was not only capable of entering the gut cells but also spread to other tissues [43]. The relatively efficient RNAi in F. occidentalis via feeding delivery suggests that systemic RNAi may operate in this insect species. This needs to be clarified in future studies. In addition, subsequent RNAi studies must assess several other factors that contribute to efficient RNAi, such as core RNAi machinery components and dsRNA metabolism, in F. occidentalis. Moreover, the low control efficiency under field conditions suggests environmental factors contributing to RNAi efficiency, which should be protected by development of an optimal formation such as chitosan nanoparticle [44].

The dsRNA specific to vATPase-B was relatively safe on non-target insects. The hazard to these non-target organisms appeared to be positively correlated with the sequence similarity among species in the dsRNA region of vATPase-B. A recent environmental risk assessment of dsRNAs expressed in a genetically modified maize variety against different non-target organisms, including pollinators and pollen feeders, soil dwelling detritivores, predators and parasitoids, aquatic detritivores, insectivorous birds, and wild mammals, showed that they are not harmful at environmentally relevant concentrations [45]. This supports spraying of dsRNAs to control F. occidentalis in agricultural fields.

Altogether, feeding of dsRNA was effective for suppressing the target gene expressions in F. occidentalis. The suppression of vATPase-B expression by RNAi was lethal to F. occidentalis. Thus, RNAi via dsRNA feeding allows a novel strategy against F. occidentalis for sustainable agriculture because it allows greater specificity than chemical insecticides [32]. Indeed, a commercial product of the dsRNA via mass production was shown to be toxic to the thrips. This paves the path for practical application of dsRNAs specific to vATPase-B to control field populations of F. occidentalis.

Data Availability

All data generated or analyzed during this study are included in this published article and its supplementary information.

Abbreviations

TSWV:

Tomato spotted wilt virus

RNAi:

RNA interference

dsRNA:

double-stranded RNA

vATPase:

Vaculolar-type ATPase

DAT:

Days after treatment

LSD:

Least-squares difference

RISC:

RNA-induced silencing complex

AGO:

Argonaute

TLR:

Toll-like receptor

References

  1. Clarke GM, Gross S, Matthews M, Catling PC, Baker B, Hewitt CL, Crowther D, Saddler SR. Environmental pest species in Australia, Australia: state of the Environment. Second Technical Paper Series (Biodiversity. Australia: Department of the Environment and Heritage. Canberra; 2000.

    Google Scholar 

  2. Nachappa P, Challacombe J, Margolies DC, Nechols JR, Whitfield AE, Rotenberg D. Tomato spotted wilt virus benefits its thrips vector by modulating metabolic and plant defense pathways in tomato. Front Plant Sci. 2020;11:575564.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Reitz SR, Gao Y, Kirk WDJ, Hoddle MS, Leiss KA, Funderburk JE. Invasion biology, ecology, and management of western flower thrips. Annu Rev Entomol. 2020;65:17–37.

    Article  CAS  PubMed  Google Scholar 

  4. Lee GS, Lee JH, Kang SH, Woo KS. Thrips species (Thysanoptera: Thripidae) in winter season and their vernal activities on Jeju island. Korea. J Asia Pac Entomol. 2001;4:115–22.

    Article  Google Scholar 

  5. He Z, Guo JF, Reitz SR, Lei ZR, Wu SY. A global invasion by the thrips, Frankliniella occidentalis: current virus vector status and its management. Insect Sci. 2020;27:626–45.

    Article  PubMed  Google Scholar 

  6. Beran F, Petschenka G. Sequestration of plant defense compounds by insects: from mechanisms to insect-plant co-evolution. Annu Rev Entomol. 2022;67:163–80.

    Article  PubMed  Google Scholar 

  7. Zhang T, Liu L, Jia Y, Zhi J, Yue W, Li D, Zeng G. Induced resistance combined with RNA interference attenuates the counter adaptation of the western flower thrips. Int J Mol Sci. 2022;23:10886.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391:806–11.

    Article  CAS  PubMed  Google Scholar 

  9. Hannon GJ. RNA interference. Nature 2002;418:244 – 51.

  10. Huvenne H, Smagghe G. Mechanisms of dsRNA uptake in insects and potential of RNAi for pest control: a review. J Insect Physiol. 2010;56:227–35.

    Article  CAS  PubMed  Google Scholar 

  11. Whyard S, Singh AD, Wong S. Ingested double stranded RNAs can act as species-specific insecticides. Insect Biochem Mol Biol. 2009;39:824–32.

    Article  CAS  PubMed  Google Scholar 

  12. Wang S, Han Y, Nabben M, Neumann D, Luiken JJFP, Glatz JFC. Endosomal v-ATPase as a sensor determining myocardial substrate preference. Metabolites. 2022;12:579.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Song Q, Meng B, Xu H, Mao Z. The emerging roles of vacuolar-type ATPase-dependent lysosomal acidification in neurodegenerative diseases. Transl Neurodegener. 2020;9:17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Cipriano DJ, Wang Y, Bond S, Hinton A, Jefferies KC, Qi J, Forgac M. Structure, and regulation of the vacuolar ATPases. Biochim Biophys Acta. 2008;1777:599–604.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Baum JA, Bogaert T, Clinton W, Heck GR, Feldmann P, Ilagan O, Johnson S, Plaetinck G, Munyikwa T, Pleau M, Vaughn T, Roberts J. Control of coleopteran insect pests through RNA interference. Nat Biotechnol. 2007;25:1322–6.

    Article  CAS  PubMed  Google Scholar 

  16. Bustin SA, Benes V, Garson JA, Hellemans J, Huggett J, Kubista M, Mueller R, Nolan T, Pfaffl WM, Shipley LG, Vandesompele J, Wittwer TC. The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem. 2009;55:611–22.

    Article  CAS  PubMed  Google Scholar 

  17. Livak KJ, Schimittgen TD. Analysis of relative gene expression data using real time quantitative PCR and the 2–∆∆CT method. Methods. 2001;25:402–8.

    Article  CAS  PubMed  Google Scholar 

  18. Park B Kim Y. Transient transcription of a putative RNase containing BEN domain encoded in Cotesia plutellae bracovirus induces an immunosuppression of the diamondback moth, Plutella xylostella. J Invertebr Pathol. 2010;105:156–63.

    Article  Google Scholar 

  19. Khan F, Roy MC, Kim Y. Thelytokous reproduction of onion thrips, Thrips tabaci Lindeman 1889, infesting welsh onion and genetic variation among their subpopulations. Insects. 2022;13:78.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Roy MC, Kim Y. Tolerance of the mealworm beetle, Tenebrio molitor, to an entomopathogenic nematode, Steinernema feltiae, at two infection foci, the intestine and the hemocoel. J Invertebr Pathol. 2020;174:107428.

    Article  CAS  PubMed  Google Scholar 

  21. Monika B, Bhadriraju S, Paul WF, Michael RL, Michelle H, Rajshekhar H. Susceptibility of Tribolium castaneum life stages exposed to elevated temperatures during heat treatments of a pilot flour mill: influence of sanitation, temperatures attained among mills floors, and costs. J Econ Entomol. 2012;105:709–17.

    Article  Google Scholar 

  22. SAS Institute Inc. SAS/STAT user’s guide, Release 6.03, Ed. Cary, NC. 1989.

  23. Chen F, Kang R, Liu J, Tang D. The V-ATPases in cancer and cell death. Cancer Gene Ther. 2022;3:1–13.

    Google Scholar 

  24. Han SH, Kim JH, Kim K, Lee SH. Selection of lethal genes for ingestion RNA interference against western flower thrips, Frankliniella occidentalis, via leaf disc-mediated dsRNA delivery. Pestic Biochem Physiol. 2019;161:47–53.

    Article  CAS  PubMed  Google Scholar 

  25. Zhu KY, Palli SR. Mechanisms, applications, and challenges of insect RNA interference. Annu Rev Entomol. 2020;65:293–311.

    Article  CAS  PubMed  Google Scholar 

  26. Whitten MMA, Facey PD, Sol RD, Fernández-Martínez LT, Evans MC, Mitchell JJ, Bodger OG, Dyson PJ. Symbiont mediated RNA interference in insects. Proc R Soc. 2016;283:20160042.

    Google Scholar 

  27. Swevers L, Liu J, Smagghe G. Defense mechanisms against viral infection in Drosophila: RNAi and non-RNAi. Viruses. 2018;10:230.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Hammond SM. Dicing and slicing: the core machinery of the RNA interference pathway. FEBS Lett. 2005;579:5822–9.

    Article  CAS  PubMed  Google Scholar 

  29. Ding SW, Voinnet O. Antiviral immunity directed by small RNAs. Cell. 2007;130:413–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Cappelle K, de Oliveira CF, Van Eynde B, Christiaens O, Smagghe G. The involvement of clathrin-mediated endocytosis and two Sid-1-like transmembrane proteins in double-stranded RNA uptake in the Colorado potato beetle midgut. Insect Mol Biol. 2016;25:315–23.

    Article  CAS  PubMed  Google Scholar 

  31. Freeman SA, Grinstein S, Orlowski J. Determinants, maintenance, and function of organellar pH. Physiol Rev. 2023;103:515–606.

    Article  CAS  PubMed  Google Scholar 

  32. Badillo-Vargas IE, Rotenberg D, Schneweis BA, Whitfield AE. RNA interference tools for the western flower thrips, Frankliniella occidentalis. J Insect Physiol. 2007;76:36–46.

    Article  Google Scholar 

  33. Andongma AA, Greig C, Dyson PJ, Flynn N, Whitten MMA. Optimization of dietary RNA interference delivery to western flower thrips Frankliniella occidentalis and onion thrips Thrips tabaci. Arch Insect Biochem Physiol. 2020;103:e21645.

    Article  CAS  PubMed  Google Scholar 

  34. Rein J, Voss M, Blenau W, Walz B, Baumann O. Hormone-induced assembly and activation of V-ATPase in blowfly salivary glands is mediated by protein kinase A. Am J Physiol Cell Physiol. 2008;294:C56–65.

    Article  CAS  PubMed  Google Scholar 

  35. Stafford-Banks CA, Rotenberg D, Johnson BR, Whitfield AE, Ullman DE. Analysis of the salivary gland transcriptome of Frankliniella occidentalis. PLoS ONE. 2014;9:e94447.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Castellanos NL, Smagghe G, Sharma R, Oliveira EE, Christiaens O. Liposome encapsulation and EDTA formula-tion of dsRNA targeting essential genes increase oral RNAi-caused mortality in the neotropical stink bug Euschistus heros. Pest Manag Sci. 2019;75:537–48.

    Article  CAS  PubMed  Google Scholar 

  37. Wang K, Peng Y, Pu J, Fu W, Wang J, Han Z. Variation in RNAi efficacy among insect species is attributable to dsRNA degradation in vivo. Insect Biochem Mol Biol. 2016;77:1–9.

    Article  PubMed  Google Scholar 

  38. Arimatsu Y, Kotani E, Sugimura Y, Furusawa T. Molecular characterization of a cDNA encoding extracellular dsRNase and its expression in the silkworm, Bombyx mori. Insect Biochem Mol Biol. 2007;37:176–83.

    Article  CAS  PubMed  Google Scholar 

  39. Giesbrecht D, Heschuk D, Wiens I, Boguski D, LaChance P, Whyard S. RNA interference is enhanced by knock-down of double-stranded RNases in the yellow fever mosquito aedes aegypti. Insects. 2020;11:327.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Yao Y, Lin DJ, Cai XY, Wang R, Hou YM, Hu CH, Gao SJ, Wang JD. Multiple dsRNases involved in exogenous dsRNA degradation of fall armyworm Spodoptera frugiperda. Front Physiol. 2022;13:850022.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Rodrigues TB, Mishra SK, Sridharan K, Barnes ER, Alyokhin A, Tuttle R, Kokulapalan W, Garby D, Skizim NJ, Tang YW, Manley B, Aulisa L, Flannagan RD, Cobb C, Narva KE. First sprayable double-stranded RNA-based biopesticide product targets proteasome subunit beta type-5 in Colorado potato beetle (Leptinotarsa decemlineata). Front Plant Sci. 2021;12:728652.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Terenius O, Papanicolaou A, Garbutt JS, Eleftherianos I, Huvenne H, Kanginakudru S, Albrechtsen M, An C, Aymeric JL, Barthel A, Bebas P, Bitra K, Bravo A, Chevalier F, Collinge DP, Crava CM, de Maagd RA, Duvic B, Erlandson M, Faye I, Felföldi G, Fujiwara H, Futahashi R, Gandhe AS, Gatehouse HS, Gatehouse LN, Giebultowicz JM, Gómez I, Grimmelikhuijzen CJ, Groot AT, Hauser F, Heckel DG, Hegedus DD, Hrycaj S, Huang L, Hull JJ, Iatrou K, Iga M, Kanost MR, Kotwica J, Li C, Li J, Liu J, Lundmark M, Matsumoto S, Meyering-Vos M, Millichap PJ, Monteiro A, Mrinal N, Niimi T, Nowara D, Ohnishi A, Oostra V, Ozaki K, Papakonstantinou M, Popadic A, Rajam MV, Saenko S, Simpson RM, Soberón M, Strand MR, Tomita S, Toprak U, Wang P, Wee CW, Whyard S, Zhang W, Nagaraju J, ffrench-Constant RH, Herrero S, Gordon K, Swevers L, Smagghe G. RNA interference in Lepidoptera: an overview of successful and unsuccessful studies and implications for experimental design. J Insect Physiol. 2011;57:231–45.

    Article  CAS  PubMed  Google Scholar 

  43. Joga MR, Zotti MJ, Smagghe G, Christiaens O. RNAi efficiency, systemic properties, and novel delivery methods for pest insect control: what we know so far. Front Physiol. 2016;7:553.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Zhang X, Mysore K, Flannery E, Michel K, Severson DW, Zhu KY, Duman-Scheel M. Chitosan/interfering RNA nanoparticle mediated gene silencing in disease vector mosquito larvae. J Vis Exp. 2015;97:e52523.

    Google Scholar 

  45. Boeckman CJ, Anderson JA, Linderblood C, Olson T, Roper J, Sturtz K, Walker C, Woods R. Environmental risk assessment of the DvSSJ1 dsRNA and the IPD072Aa protein to non-target organisms. GM Crops Food. 2021;12:459–78.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not Applicable.

Funding

This work was carried out under the support of the Cooperative Research Program for Agriculture Science & Technology Development (Project No. PJ01578901) funded by the Rural Development Administration (RDA), Republic of Korea. RDA played no role in the design of the study and collection, analysis, interpretation of data, and in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

YK conceived and designed the experiments. FK performed the molecular work. YK and FK analyzed and interpreted all data. FK, MK and YK wrote manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yonggyun Kim.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Conflicting interests

The authors declare that they have no conflicts of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khan, F., Kim, M. & Kim, Y. Greenhouse test of spraying dsRNA to control the western flower thrips, Frankliniella occidentalis, infesting hot peppers. BMC Biotechnol 23, 10 (2023). https://doi.org/10.1186/s12896-023-00780-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12896-023-00780-y

Keywords