Skip to main content
PLOS Pathogens logoLink to PLOS Pathogens
. 2021 Mar 10;17(3):e1009424. doi: 10.1371/journal.ppat.1009424

Coordination between terminal variation of the viral genome and insect microRNAs regulates rice stripe virus replication in insect vectors

Wan Zhao 1,2, Jinting Yu 1,2, Feng Jiang 2,3, Wei Wang 1, Le Kang 1,2,3, Feng Cui 1,2,*
Editor: Aiming Wang4
PMCID: PMC7984632  PMID: 33690727

Abstract

Maintenance of a balance between the levels of viral replication and selective pressure from the immune systems of insect vectors is one of the prerequisites for efficient transmission of insect-borne propagative phytoviruses. The mechanism regulating the adaptation of RNA viruses to insect vectors by genomic variation remains unknown. Our previous study demonstrated an extension of the 3’-untranslated terminal region (UTR) of two genomic segments of rice stripe virus (RSV). In the present study, a reverse genetic system for RSV in human cells and an insect vector, the small brown planthopper Laodelphax striatellus, was used to demonstrate that the 3’-terminal extensions suppressed viral replication in vector insects by inhibiting promoter activity due to structural interference with the panhandle structure formed by viral 3’- and 5’-UTRs. The extension sequence in the viral RNA1 segment was targeted by an endogenous insect microRNA, miR-263a, which decreased the inhibitory effect of the extension sequence on viral promoter activity. Surprisingly, the expression of miR-263a was negatively regulated by RSV infection. This elaborate coordination between terminal variation of the viral genome and endogenous insect microRNAs controls RSV replication in planthopper, thus reflecting a distinct strategy of adaptation of phytoviruses to insect vectors.

Author summary

Mutations frequently happen when insect-transmitted RNA viruses circulate between insect vectors and plant or mammalian hosts. However, the significance of these mutations for viral fitness in the two distinct organisms is poorly understood. We discovered that a high proportion of rice stripe virus (RSV) had terminally extended genomes when the virus infected insect vectors. In the present study, we found that the extension sequence suppressed viral replication in insect vectors by impairing a special structure formed by the two ends of the viral genomes. An endogenous insect small RNA was able to bind the extension sequence to relieve the inhibitory effect. However, the expression of this small RNA was reduced in the presence of RSV to ultimately maintain the inhibitory effect of the extension sequence. This elaborate coordination between virus and vector enables a limited level of RSV replication that does not produce serious damage to vectors, thus reflecting a distinct strategy of adaptation of insect-transmitted plant viruses.

Introduction

Insect-borne propagative phytoviruses and arboviruses usually induce serious or fatal diseases in plant and mammalian hosts in contrast to the asymptomatic infection in insect vectors [1,2]. Maintenance of a balance between the levels of viral replication and selective pressure from the immune systems of a host or a vector is one of the major factors that determine the distinct outcome of the infection. Genetic variation is frequently observed in insect-borne RNA viruses during viral circulation in two distinct organisms. In addition to nucleotide deletion, insertion, and substitution in the coding regions, various mutations appear in the untranslated terminal regions (UTRs) of the viral genomes [3,4]. The mechanism of the regulation of adaptation of a virus to insect vectors by variations in the UTRs is currently unknown.

The UTRs of single-stranded RNA viruses play the key roles in viral replication and gene transcription. The 3’- and 5’-UTRs of negative-strand RNA viruses form a panhandle structure via distal complementary 15–19 nucleotides (nt) to provide a binding site for viral RNA-dependent RNA polymerase (RdRp) [5,6]. The 3’-UTRs of certain negative-strand RNA viruses, such as influenza A virus, contain the core promoter elements [5]. Mutations in the 3’-UTRs usually negatively impact viral replication, transcription, or translation [7]. Endogenous microRNAs (miRNAs) of the host cells have been frequently reported to target the 3’-UTRs of RNA viruses to regulate viral replication [8,9].

Our previous study demonstrated 16 and 15 nt extensions in the 3’-UTRs of two genomic segments, negative-sense RNA1 and ambisense RNA2, of rice stripe virus (RSV). RSV is a single-stranded RNA virus of the Tenuivirus genus [10] efficiently transmitted by small brown planthopper, Laodelphax striatellus, in a persistent-propagative manner. RSV causes serious destructive rice diseases in Asian countries [11]. The genome of RSV consists of four RNA segments encoding a nucleocapsid protein (NP), an RdRp, and five nonstructural proteins [12,13]. Similar to other negative-strand RNA viruses, the 5’- and 3’-terminal sequences of each genomic RNA of RSV have sufficient complementarity to form a panhandle structure [14,15]. We have demonstrated that the relative ratios of 3’-terminally extended types of genomic RNA1 and RNA2 increased from around 7% to 35% or 49% with viral replication in planthopper and decreased to less than 3% in the late-stage of infection in the plant hosts, and the extension sequences were deleterious for viral replication in the plants [10]. However, the mechanism and biological significance of the maintenance of the 3’-terminally extended RSV genomes in insect vectors are unknown.

This study investigated the influence of 3’-terminal extensions of the viral genome on the viral promoter activity using a reverse genetic system for RSV in human cells and in small brown planthopper. The adverse effect of the 3’-extension sequences regulates viral replication level in planthopper. However, the extension sequence is responsible for targeting of the viral genome by a conserved insect miRNA.

Results

The 3’ extensions suppress viral replication in planthopper

We have previously demonstrated that insect-derived RSV have considerably higher ratios of RNA1 and RNA2 with the extended 3’-UTRs (3’-EUTRs) than those in plant-derived RSV, and the 3’ extensions inhibit viral replication in the plant hosts [10]. To determine whether the 3’ extensions play a similar function in insect vectors, nonviruliferous planthoppers were fed an artificial diet containing equal amounts of viral RNAs derived from viruliferous planthoppers or from rice. Higher levels of viral RNAs were detected in the original crude extracts from insects than in those from rice (Fig 1A), and significantly higher ratios of RNA1 or RNA2 with the 3’-EUTRs were detected in insect-derived RSV than those in plant-derived RSV (Fig 1B). After dilution at various dilution factors and mixing with the artificial diet, the diluted viral crude extracts contained equal amounts of insect- and plant-derived viral RNAs (Fig 1C); however, the ratios of RNA1 or RNA2 with the 3’-EUTRs were no longer different between RSVs derived from two sources (Fig 1D) due to low virus titer. The accumulation of RNA1 and RNA2 in planthopper was assayed by quantitative real-time PCR (qPCR) at 2 h, 1 d, 2 d, 5 d, and 8 d after the insects were fed the artificial diet containing equal amounts of insect- or plant-derived viral RNAs. Less than 2 d after administration, similar RNA levels of RNA1 or RNA2 were detected in the two groups of planthoppers (Fig 1E and 1F). Starting from 2 d after inoculation (DAI), the amounts of RNA1 and RNA2 in planthoppers infected with insect-derived RSV were always lower than those in planthoppers infected with plant-derived RSV corresponding to the higher ratios of RNA1 or RNA2 with the 3’-EUTRs in planthoppers infected with insect-derived RSV (Fig 1E and 1F). This result demonstrated that the 3’ extensions are detrimental to viral replication in insect vectors.

Fig 1. The 3’ extensions suppress viral replication in planthopper.

Fig 1

(A) Relative RNA levels of the viral NP gene, genomic RNA1 and RNA2, and 3’ extensions of RNA1 (RNA1-3’-EUTR) and RNA2 (RNA2-3’-EUTR) in crude RSV extract from RSV-infected small brown planthopper (insect-derived RSV) and rice (plant-derived RSV). Viruliferous fourth-instar nymphs and rice leaves with obvious stripe disease symptoms were sampled for virus extraction. (B) Relative ratio of the 3’ extension (3’-EUTR) to RNA1 or RNA2 in insect- and plant-derived RSVs. (C) Relative RNA levels of viral NP, RNA1, RNA2, RNA1-3’-EUTR, and RNA2-3’-EUTR in the diluted crude viral extracts containing equal amounts of insect- or plant-derived RSVs. (D) Relative ratio of the 3’-EUTR to RNA1 or RNA2 in the diluted crude viral extracts containing equal amounts of insect- and plant-derived RSVs. (E) Relative RNA levels of viral RNA1 and relative ratio of the 3’-EUTR to RNA1 in planthopper at various time points after feeding the artificial diet containing equal amounts of insect- or plant-derived RSVs. (F) Relative RNA levels of viral RNA2 and relative ratio of the 3’-EUTR to RNA2 in planthopper at various time points after feeding the artificial diet containing equal amounts of insect- or plant-derived RSVs. NS, no significant differences. The values are presented as the mean ± SE. *, P < 0.05. **, P < 0.01. ***, P < 0.001.

The 3’ extensions inhibit viral promoter activity in vitro

Prediction of the secondary RNA structure suggested that the 3’ extensions modify a panhandle structure formed by the 3’- and 5’-UTRs of viral RNA1 or RNA2 (S1 Fig) and may thus influence the promoter activity. To verify this hypothesis, we used the pPol I plasmid to generate a reverse genetic system for RSV in human embryonic kidney 293T cells due to lack of planthopper cell lines. The pPol I plasmid is driven by the human RNA polymerase I (Pol I), which normally directs the synthesis of ribosomal RNAs (rRNAs) [16]. The complementary sequence of Renilla luciferase (Rluc) flanked by the 3’- and 5’-UTRs of RSV RNA1 or RNA2 were inserted downstream of the Pol I promoter to construct the wild-type plasmids named P1-WT for RNA1 and P2-WT for RNA2. The diagrams of the recombinant pPol I plasmids are presented in Fig 2A. RSV NP and RdRp were incorporated into the pcDNA plasmid to generate the recombinant pcDNA-NP and pcDNA-RdRp plasmids, which express NP and RdRp, respectively. When the P1-WT or P2-WT plasmids were cotransfected with pcDNA-RdRp or pcDNA-RdRp and pcDNA-NP into 293T cells, the Rluc reporter RNA containing viral 3’- and 5’-UTRs was transcribed by Pol I. Rluc reporter RNA served as a template to generate Rluc mRNA under the promoter contained in the viral 3’-UTR in the presence of viral RdRp and NP. The pGL4.13 plasmid that constitutively expresses firefly luciferase was used as a reference to normalize the Rluc activity. The results showed that relative Rluc activity was low in 293T cells cotransfected with P1-WT or P2-WT and pcDNA-RdRp (Fig 2B). The relative Rluc activity was considerably increased by the addition of NP expression (Fig 2B) demonstrating that viral RdRp can recognize the panhandle structure formed by the 3’- and 5’-UTRs of viral RNA1 or RNA2 with the assistance of NP, and the viral 3’-UTRs had promoter activities.

Fig 2. The 3’ extensions inhibit viral promoter activity.

Fig 2

(A) The diagrams of the recombinant pPol I plasmids with human Pol I promoter. The antisense Renilla luciferase coding sequence (Rluc) was flanked by the 5’-UTR and variable length of 3’-UTRs of RNA1 or RNA2 of RSV. Three types of the pPol I plasmids were constructed based on variable length of the 3’-UTR: P1/P2-WT with the wild type 3’-UTRs of RNA1 or RNA2; P1/P2-EUTR with the 16 or 15 nt extended 3’-UTRs of RNA1 or RNA2; and P1/P2-TUTR with the truncated 3’-UTRs lacking the terminal 16 nt regions of RNA1 or RNA2. R, gene encoding hepatitis delta virus ribozyme. (B) The promoter activity of various recombinant pPol I plasmids in human 293T cells with the coexpression of RSV NP and RdRp according to the data of the dual luciferase reporter assay. The pGL4.13 plasmid constitutively expressing firefly luciferase was transfected as a reference. The relative activity of Renilla luciferase (Rluc) to firefly luciferase (Fluc) is presented as the mean ± SE. ***, P < 0.001. The coexpression of RSV NP and RdRp encoded by the pcDNA3.1 plasmid was assayed by western blot using an in-house generated anti-NP monoclonal antibody and anti-RdRp polyclonal antibodies. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control. (C) Alignment of the sequences surrounding the Pol I transcription initiation sites of Laodelphax striatellus and other eukaryotes (Homo sapiens, Rattus norvegicus, Mus musculus, Gallus gallus, Xenopus laevis, Aedes albopictus, Glossina morsitans, and Drosophila melanogaster). The transcription initiation site is labeled as +1. (D) The diagrams of the recombinant pPol I plasmids with Pol I promoter of small brown planthopper (LsPol I). The antisense GFP coding sequence was flanked by the 5’-UTR and 3’-UTR (LsP1/P2-WT) or the 5’-UTR and 3’-EUTR (LsP1/P2-EUTR) of RSV RNA1 or RNA2. (E) Detection of the GFP signals in the gut and salivary glands of viruliferous planthopper 6 d after the injection of LsP1-WT or LsP2-WT. No signals were observed in the two organs of nonviruliferous planthoppers administered the same injection. (F) Comparison of the GFP signals in the gut of viruliferous planthopper 6 d after the injection of equal amounts of LsP1-WT and LsP1-EUTR or equal amounts of LsP2-WT and LsP2-EUTR. For (E) and (F), the nuclei (Nc) were stained with Hoechst. Scale bars: 100 μm. (G) Western blot of the GFP protein levels in the gut of (F) using an anti-GFP polyclonal antibody. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control.

Then, we replaced the 3’-UTRs with the extended 3’-UTRs (3’-EUTRs) of viral RNA1 and RNA2 to construct the mutant plasmids named P1-EUTR for RNA1 and P2-EUTR for RNA2. Another type of mutant plasmids with truncated 3’-UTRs (3’-TUTRs) lacking the terminal 16 nt of viral RNA1 and RNA2 was constructed and named P1-TUTR and P2-TUTR, respectively (Fig 2A). When equal amounts of P1/P2-WT, P1/P2-EUTR, and P1/P2-TUTR were transfected into 293T cells in combination with the vectors for the expression of NP and RdRp, the cells transfected with P1-EUTR or P2-EUTR had considerably lower relative Rluc activities than those in the cells transfected with P1-WT or P2-WT and similar activities to those in the cells transfected with P1-TUTR or P2-TUTR (Fig 2B). The results of these in vitro experiments showed that the 3’ extensions of RNA1 and RNA2 inhibit the promoter activity of viral 3’-UTRs in human cells.

The 3’ extensions inhibit viral promoter activity in planthopper

The in vitro results indicated that viral 3’ extensions influence the promoter activity in human cells. Considering the species specificity of RNA Pol I promoters, we cloned the Pol I promoter of small brown planthopper to replace the original human Pol I promoter by querying the L. striatellus genome [17] based on the high conservation of Pol I promoter sequences [18,19]. The nucleotides from 15,877 to 15,906 in contig36944 have high identity with the sequences adjacent to the known Pol I transcription initiation sites of other eukaryotes, and T at position -1 and A or G at positions +1 and -2, respectively, are conserved (Fig 2C) [19]. Thus, the 370 bp sequence upstream of the planthopper Pol I transcription initiation site (from -1 to -370) (GenBank registration number MW353861) was cloned as a putative Pol I promoter.

The human Pol I promoters in the P1/P2-WT and P1/P2-EUTR plasmids were replaced by the 370 bp planthopper Pol I promoter, and Rluc was replaced by GFP for convenient observation in vivo. The new recombinant plasmids were named LsP1/P2-WT and LsP1/P2-EUTR (Fig 2D). When equivalent amounts of LsP1-WT or LsP2-WT were injected into viruliferous and nonviruliferous planthopper nymphs, GFP signals were clearly detected in the gut and salivary glands of viruliferous planthoppers at 6 DAI, and no signal was observed in these tissues in nonviruliferous planthoppers. Furthermore, the GFP signals were stronger in the gut of viruliferous insects than that in the salivary glands (Fig 2E). This result confirmed that the planthopper Pol I promoter was cloned and demonstrated that the promoters of the 3’-UTRs of viral RNA1 and RNA2 are active in vivo. Planthoppers injected with LsP1-EUTR or LsP2-EUTR had considerably weaker GFP signals in the gut at 6 DAI compared with that in viruliferous planthoppers injected with LsP1-WT or LsP2-WT (Fig 2F). The data of western blot also demonstrated lower protein expression of GFP in the gut after the injection with LsP1-EUTR or LsP2-EUTR (Fig 2G) demonstrating that the 3’ extensions inhibit the promoter activity of viral RNA1 or RNA2 in planthopper. Therefore, these in vivo experiments supported the hypothesis that the 3’ extensions of RNA1 and RNA2 influence the promoter activity of viral 3’-UTRs to control viral replication in planthopper.

The extended 3’-UTR of viral RNA1 is targeted by planthopper miR-263a

To determine whether the 3’ extensions of viral RNA1 and RNA2 are regulated by endogenous miRNAs in planthopper, we predicted candidate miRNAs that may target the 3’ termini of viral RNA1 or RNA2 using the small RNA (sRNA) sequencing data from our previous study [20]. Two out of 213 putative miRNAs from small brown planthopper, dme-miR-263a-5p (miR-263a) and ame-miR-190 (miR-190), were predicted by both the miRanda and RNAhybrid algorithm to specifically target the 3’-EUTRs of RNA1 and RNA2, respectively. The seed sequences from 2 to 8 nt of miR-263a were a nearly perfect match to the 3’-EUTR (with a G:U wobble base pair), but not to the 3’-UTR of RNA1, and the seed sequences of miR-190 had two G:U wobble base pairs with the 3’-EUTR and did not match the 3’-UTR of RNA2 (S2 Fig). Direct interactions of these two miRNAs with the target viral RNA sequences were initially verified by dual luciferase assay in 293T cells. Relative Rluc activities in the cells transfected with the construct containing the 3’-EUTR of RNA1 were significantly decreased in the presence of miR-263a at the concentrations from 10 nM to 100 nM in a dose-dependent manner (Fig 3A). In contrast, transfection of the cells with the construct containing the 3’-UTR of RNA1 did not result in suppressed Rluc activities in the presence of miR-263a (Fig 3A). This result showed that miR-263a targeted the 3’-EUTR, but not the 3’-UTR, of RSV RNA1. Additionally, the relative Rluc activities in the cells transfected with the constructs containing 3’-EUTR or 3’-UTR of viral RNA2 were not influenced by miR-190 at the concentrations from 50 pM to 100 nM (Fig 3B) implying that these two types of 3’ termini are not the targets of miR-190.

Fig 3. The extended 3’ UTR of viral RNA1 is targeted by planthopper miR-263a.

Fig 3

(A) Dual luciferase reporter assay in 293T cells cotransfected with miR-263a mimics and the recombinant pLightSwitch plasmid containing the 3’-untranslated terminal region (3’-UTR) or the extended 3’-UTR (3’-EUTR) of RSV RNA1 segment. (B) Dual luciferase reporter assay in 293T cells cotransfected with miR-190 mimics and the recombinant pLightSwitch plasmid containing the 3’-UTR or 3’-EUTR of RSV RNA2 segment. The mimic control (NC) was used in the control group instead of miR-263a or miR-190. The pGL4.13 plasmid expressing firefly luciferase (Fluc) was used as a reference control. The relative activity of Renilla luciferase (Rluc) encoded by the recombinant pLightSwitch normalized to Fluc is presented as the mean ± SE. (C) RIP analysis combined with real-time quantitative PCR (qPCR). An anti-Ago1 monoclonal antibody was used to pull down RNAs from viruliferous planthoppers injected with agomir-263a, agomir-190, or agomir control (agomir-NC). Mouse IgG was used as a negative control. The amount of 3’-EUTRs of RSV RNA1 or RNA2 in the Ago1 immunoprecipitates relative to that in the IgG samples (fold enrichment) was quantified using qPCR and compared in the groups treated with agomir-263a or agomir-190 injection and agomir-NC injection. miR-263a and miR-190 in the precipitates were assayed using qPCR. The level of Ago1 in the precipitates is shown in the western blot. NS, no significant differences. (D) Biotin pulldown assay coupled with qPCR was used to quantify the amount of 3’-EUTR of RNA1 in viruliferous planthoppers injected with biotinylated miR-263a (bio-miR-263a) or with control biotinylated scrambled sequences (bio-Scr). The transcript levels of actin were quantified and used as a negative control. From (A) to (D), the values are presented as the mean ± SE. NS, no significant differences. *, P < 0.05. **, P < 0.01. ***, P < 0.001. (E) Double fluorescence in situ hybridization (FISH) for miR-263a and the 3’-EUTR of RNA1 in the gut cells of viruliferous planthopper. Probes for Caenorhabditis elegans cel-miR-67-3p and rice PsbP gene were used in the control group. The probes for microRNAs were labeled with biotin. The probes for the 3’-EUTR of RNA1 and PsbP were labeled with digoxigenin. The nuclei were stained with Hoechst. Scale bars: 50 μm.

To verify the interaction between miR-263a and the 3’-EUTR of viral RNA1 in vivo, we performed RNA immunoprecipitation (RIP) assay using an in-house generated planthopper anti-argonaute 1 (Ago1) monoclonal antibody [21] in viruliferous planthoppers after injected with agomir of miR-263a or miR-190, which made the endogenous levels of the two miRNAs increase by 4- or 1.3-fold (Fig 3C). The 3’-EUTR of viral RNA1 was significantly enriched in the Ago1-immunoprecipitated RNAs from viruliferous planthoppers injected with agomir-263a compared to that in the control group injected with agomir-NC (Fig 3C). On the other hand, no significant enrichment in 3’-EUTR of viral RNA2 was observed in Ago1-immunoprecipitated RNAs from insects injected with agomir-190 compared to that in the control group (Fig 3C). Biotinylated miRNA pulldown assay confirmed the interaction between miR-263a and the 3’-EUTR of RNA1. Higher amounts of the 3’-EUTR of RNA1 were pulled down by biotinylated-miR-263a from viruliferous planthoppers than that by actin RNA used as a negative control (Fig 3D). Furthermore, colocalization of miR-263a and the 3’-EUTR of RNA1 in the gut cells was observed in the cytoplasm and partial region of the nucleus according to the data of double fluorescence in situ hybridization (FISH) (Fig 3E).

miR-263a facilitates RSV replication by counteracting the inhibitory effect of RNA1 3’ extension

Since the 3’-EUTR of viral RNA1 is targeted by miR-263a, we aimed to determine whether miR-263a regulates the influence of the 3’ extension on the viral promoter activity. When P1-EUTR was cotransfected with miR-263a mimic or a control mimic (mimic-NC) into 293T cells in the presence of viral RdRp and NP, the transfection of miR-263a mimic significantly increased the normalized Rluc activity encoded by P1-EUTR compared to that in the mimic-NC-transfected group but did not influence the Rluc activity in the cells transfected with P1-WT (Fig 4A). Northern blot using a probe for Rluc mRNA detected the same size of the Rluc transcript driven by 3’-EUTR of RNA1 with or without miR-263a mimic indicating that miR-263a does not degrade the Rluc transcript (Fig 4B). Injection of the mixture of LsP1-EUTR and agomir-263a into viruliferous planthoppers increased the GFP expression levels in the gut compared to that in the control group injected with the mixture of LsP1-EUTR and agomir-NC according to the values of the fluorescence signals and western blot data (Fig 4C and 4D). In contrast, agomir-263a did not influence the GFP expression in the insects treated with LsP1-WT (Fig 4C and 4D). These results indicated that miR-263a counteracts the inhibitory effect of the 3’ extension on the promoter activity of RNA1 3’-UTR.

Fig 4. miR-263a facilitates RSV replication by counteracting the inhibitory effect of RNA1 3’ extension.

Fig 4

(A) The promoter activity of the recombinant pPol I plasmids P1-WT and P1-EUTR in the presence of miR-263a mimics or mimic control (NC) in human 293T cells was determined by dual luciferase reporter assay. P1-WT, P1-EUTR, and other experimental details are the same as described in the legend to Fig 2A and 2B. (B) Northern blot shows the size of Rluc transcript encoded by P1-EUTR in the presence of miR-263a mimics or NC in 293T cells; the signal was detected using a biotin-labeled LNA oligonucleotide probe for Rluc. 18S rRNA was used as an internal control. (C) GFP signal in the gut of viruliferous planthopper 6 d after coinjection of agomir-263a with LsP1-WT or LsP1-EUTR. The agomir control (agomir-NC) was injected as a negative control. LsP1-WT and LsP1-EUTR are described in Fig 2D. The nuclei (Nc) were stained with Hoechst. Scale bars: 100 μm. (D) The western blot results show the GFP protein levels in the gut of (C); the signal was detected using the anti-GFP polyclonal antibodies. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control. (E) The RNA levels of miR-263a, viral RNA1, RdRp, and NP in the whole body of viruliferous planthopper determined by real-time quantitative PCR (qPCR) 4 d after the injection of antagomir-263a or agomir-263a. The antagomir control (antagomir-NC) or agomir-NC were injected as a negative control. (F) Western blot analysis of RSV NP protein in the samples of (E) using an in-house generated anti-NP monoclonal antibody. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control. (G) The results of qPCR show the RNA levels of miR-263a, viral RNA1, RdRp, and NP in planthoppers infected by RSV for 5 d after the injection of agomir-263a or agomir-NC. Nonviruliferous planthoppers were initially fed RSV-infected rice for 1 d before the agomir injection. The values in (A), (E), and (G) represent the mean ± SE. NS, no significant differences. *, P < 0.05. **, P < 0.01. ***, P < 0.001.

To determine the role of miR-263a in RSV replication, we delivered antagomir or agomir of miR-263a into viruliferous planthoppers. The treatment with antagomir led to a 62% decrease in the expression of miR-263a, and the RNA levels of viral RNA1, RdRp, and NP in the whole body were declined by 65%, 49%, and 55%, respectively, at 4 DAI (Fig 4E). The protein level of NP was significantly decreased (Fig 4F). In contrast, the addition of miR-263a agomir did not upregulate the RNA levels of viral RNA1, RdRp, and NP or the protein level of NP in the whole body (Fig 4E and 4F) probably due to a high viral load in viruliferous planthoppers. The effects of antagomir or agomir of miR-263a on RSV load in the gut dissected from viruliferous planthoppers at 4 DAI were similar to those observed in the whole body assays (S3 Fig). To evaluate the role of miR-263a agomir in RSV replication, nonviruliferous planthoppers were initially fed RSV-infected rice for 1 d; then, the insects were injected with miR-263a agomir. The RNA levels of RNA1, RdRp, and NP were significantly upregulated by miR-263a agomir at 4 DAI (Fig 4G). Therefore, miR-263a facilitated RSV replication by eliminating the inhibitory effect of the 3’ extension on the promoter activity of RNA1 3’-UTR.

RSV suppresses miR-263a expression in planthopper

The effect of RSV infection on miR-263a expression was investigated using qPCR. No significant changes in miR-263a expression were detected in the whole body of viruliferous planthoppers compared to that in nonviruliferous planthoppers (Fig 5A). However, a significant downregulation of miR-263a expression was observed in the gut and salivary glands of viruliferous insects, and miR-263a expression was higher in the gut compared with that in the salivary glands, brain, fat body, ovary, and testis (Fig 5B). Considering that RSV proliferates mainly in the gut, the expression of miR-263a was further tested in the gut during RSV infection. After nonviruliferous planthoppers were infected with RSV for 2, 5, and 8 d by feeding RSV-containing crude extract, the expression of miR-263a in the gut was dramatically decreased concomitant to an increase in the virus load over time (Fig 5C).

Fig 5. RSV suppresses miR-263a expression in planthopper.

Fig 5

(A) Relative transcript levels of miR-263a in nonviruliferous (N) and viruliferous (V) fourth-instar nymphs of planthopper. (B) Relative transcript levels of miR-263a in six tissues of N and V planthoppers quantified by qPCR. (C) Expression patterns of miR-263a and viral NP in the gut of planthopper assayed using qPCR 2–8 d after the infection with RSV from RSV-infected rice. (D) Relative transcript levels of miR-263a in 293T cells assayed by qPCR after the transfection with various amounts of the recombinant pcDNA3.1 plasmid containing a partial sequence of pri-miR-263a centered pre-miR-263a (pc-pri-miR-263a). The empty pcDNA3.1 plasmid (pc-NC) was transfected and used as a negative control. (E) Relative transcript levels of pre-miR-263a and pri-miR-263a assayed by qPCR in the whole body of N and V planthoppers. (F) Relative transcript levels of pre-miR-263a and pri-miR-263a assayed by qPCR in the nuclei of N and V planthoppers. (G) Relative transcript levels of miR-263a and pri-miR-263a in the whole body of planthoppers 5 d after the injection of crude RSV preparations and double-stranded RNAs of RSV NP (dsNP) or double-stranded RNAs of GFP (dsGFP). The values from (A) to (G) were determined by real-time quantitative PCR and represent the mean ± SE. NS, no significant differences. *, P < 0.05. **, P < 0.01. ***, P < 0.001.

To determine the reason for a reduction in miR-263a in response to RSV infection, we used the genome sequence of small brown planthopper to predict primary miR-263a (pri-miR-263a) and stem-loop precursor miR-263a (pre-miR-263a) (S4 Fig) [17]. Partial pri-miR-263a sequence of 455 bp (GenBank registration number MW353862) was able to produce a mature miR-263a transcript in 293T cells treated with various template concentrations (Fig 5D) thus validating the identity of pri-miR-263a and pre-miR-263a. Then, the transcript levels of pri-miR-263a and pre-miR-263a were measured using qPCR and compared in nonviruliferous and viruliferous gut. The results showed that pri-miR-263a and pre-miR-263a were downregulated in the gut cells (Fig 5E) and even in the nuclei of the gut cells in viruliferous planthopper (Fig 5F) where primary miRNA transcription and precursor miRNA processing occur. When double-stranded RNA of RSV NP (dsNP-RNA) and RSV crude preparations were simultaneously injected into nonviruliferous planthoppers, viral replication in terms of NP RNA level was restrained (S5 Fig), and the transcript levels of miR-263a and pri-miR-263a were significantly upregulated compared to those in the dsGFP-RNA-injected group at 5 DAI (Fig 5G). Therefore, RSV suppresses miR-263a expression by inhibiting the transcription of pri-miR-263a.

Discussion

RSV is a typical insect-borne propagative phytovirus that circulates and replicates in insect vectors and host plants and maintains a limited replication level in insect vectors. In this study, we showed a molecular mechanism that controls the replication level of RSV in insect vectors. When planthoppers acquire virions from rice, most virions have normal genomes with complementary 3’- and 5’-UTRs. With viral replication, a 16 nt and a 15 nt sequence are added at the 3’-UTRs of RNA1 and RNA2. The 16 and 15 nt extensions in the 3’ termini of the RSV genomes impair viral promoter activity resulting in the inhibition of viral replication. This inhibitory effect of the 3’ extension of viral RNA1 was decreased by endogenous insect miR-263a, which targets the extended sequence. However, the expression of miR-263a was negatively regulated by RSV infection. This interaction between RSV and insect miRNAs secures the controlled replication of RSV in planthopper reflecting a distinct strategy of adaptation of phytoviruses to the insect vectors.

This study is the first to determine the promoter activity of the panhandle structure formed by RSV 3’- and 5’-UTRs in insects. Terminal complementary sequences are ubiquitous in RNA viruses, especially negative-strand RNA viruses, and play the key role in viral replication and expression [6,22]. The 3’- and 5’-termini of RSV genomic RNAs were initially detected in the RSV isolate T derived from RSV-infected maize host. The data of two-dimensional mobility shift analyses indicated that the 18–20 nt sequences at the 5’- and 3’-termini of each viral RNA are complementary to each other and generate a panhandle structure [14]. The promoter functions of the 3’- and 5’-termini of RSV were explored in vitro using a model RNA transcribed from a synthetic DNA template, T7 RNA polymerase, and viral 16–17 nt terminal conserved sequences of all four genomic RNA segments. Approximately 50 nt RNA products were successfully identified by incubation with solubilized viral proteins (containing at least RdRp and NP) [15]. However, the promoter activity was not verified in vivo. A reverse genetic system was constructed in the present study to test viral promoter activity in planthopper. This system is based on a modified pPol I plasmid with the Pol I promoter of small brown planthopper and was used to transcribe the RNA template containing viral 3’- and 5’-UTRs recognized by RSV RdRp to drive the expression of the GFP gene. The promoter activities corresponding to the GFP signals were detected in the gut and salivary glands of viruliferous planthopper indicating the validity of this reverse genetic system and confirming the promoter functions of the terminal panhandle structure of RSV in vivo.

The extensions at the 3’-UTRs of RSV inhibited viral replication. Variations in the viral UTRs generally inhibit viral replication and/or infection [7,23]. Mutagenic analysis revealed that the conserved terminal 11 nt in the terminal complementary 3’- and 5’-UTRs of Bunyamwera virus are critical for viral replication and transcription [24]. Separate mutations within the cyclization sequences of the 5’- or 3’-UTR of Kunjin virus significantly inhibited viral replication in BHK cells, and simultaneous compensatory mutations in both cyclization sequences restored viral replication ability, suggesting that base pairing is vital for Kunjin virus replication [22]. Furthermore, the UTRs of RNA viruses account for differences in viral fitness in the host and vector cells. For example, mutations in the 3’-UTR of Chikungunya virus reduced viral replication only in the mosquito vector cells, but not in the mammalian host cells, suggesting that adaptation to the vectors rather than to the hosts is the major evolutionary force acting on viral 3’-UTR [4]. The duplicated stem loop structure of the 3’-UTR was beneficial for dengue virus replication in mammalian cells, while deleterious in mosquito cells, suggesting that viral 3’-UTR is involved in insect-host adaptation cycle [25]. The enrichment of 3’-terminally extended RSV in planthopper is good for maintaining a limited replication level for virus to strike a balance with insect immune systems without causing severe diseases. It is possible that the extension sequences are generated by viral RdRp with the template of viral genomic RNAs because the 16 nt and 15 nt extension sequences at the 3’-UTRs of RNA1 and RNA2 are respectively identical to the 8874 nt to 8889 nt fragment in the upstream region of RNA1 and the 3241 nt to 3255 nt fragment in the upstream region of RNA2 [10].

miR-263a promoted RSV replication by targeting the extended 3’-UTR of the viral genome. In contrast to the positive regulation, cellular miRNAs usually suppress viral replication or infection by direct targeting of the viral genomes. For example, human miR-32 effectively inhibited the accumulation of the retrovirus primate foamy virus type 1 [26]. Human hsa-miR-296-5p restricted enterovirus A71 replication [27]. Human miR-125-5p and miR-122 inhibited hepatitis B virus replication by targeting the coding regions of the S protein and polymerase or the 3’-UTR of the core protein [28,29]. Positive regulation of viral replication by cellular miRNAs due to direct interaction with viral genomes is unusual. For example, human miR-122 benefits hepatitis C virus replication by binding to the 5’-UTR of the viral genome to enhance viral protein translation via recruitment of the 48S ribosomal subunit [30] and to increase viral RNA accumulation by protecting the 5′-terminus of genomic RNA [31,32]. We demonstrate that miR-263a can rescue the viral promoter activity of the extended UTR, which was not apparently degraded by miR-263a. The miR-263a, possibly with the aid of Ago1, covers the extension sequences, rendering the UTRs-formed panhandle structure exposed to viral RdRp for binding.

Thus, we discovered a strategy of RSV adaptation in vector insects mediated by the interaction between terminal variation of the viral genome and endogenous insect miRNAs. This elaborate coordination limits the replication of RSV to avoid serious damage to planthopper.

Materials and methods

Virus, insect vector, and host plant preparation

The viruliferous small brown planthopper populations were established from a rice field population collected in Nanjing, Jiangsu Province, China. Viruliferous and nonviruliferous planthoppers were reared separately on 2–3 cm seedlings of rice (Oryza sativa Huangjinqing) in glass incubators as described previously [33]. To maintain the RSV-carrying frequency of the viruliferous strain of at least 90%, nonviruliferous individuals were identified and eliminated via dot enzyme-linked immunosorbent assay (dot-ELISA) using an in-house generated monoclonal anti-NP antibody every three months [33].

Cells

Human embryonic kidney HEK293T (293T) cells were maintained in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (FBS) at 37°C in an atmosphere of 5% CO2.

Feeding RSV crude preparations to planthoppers

Insect- or plant-derived crude RSV preparations were extracted from the RSV-infected rice leaves or viruliferous planthoppers, respectively [33]. Nonviruliferous fourth-instar planthopper nymphs were fed an artificial diet containing equal amounts of insect- or plant-derived RSV virions as described previously [33]. After feeding the diet for 24 h, the RSV-infected nymphs were transferred to healthy rice seedlings and incubated for 2 h, 1 d, 2 d, 5 d, and 8 d. All planthoppers were collected for RNA extraction and quantitative real-time PCR (qPCR) analysis.

After feeding the artificial diet containing an appropriate concentration of plant-derived RSV for 2 d, 5 d, and 8 d, the gut from the nymphs was dissected for RNA extraction and qPCR analysis.

Double-stranded RNA synthesis and delivery

dsRNA was generated using a T7 RiboMAX express RNAi system (Promega, Madison, Wisconsin, USA) and purified using a Wizard SV gel and PCR clean-up system (Promega) following the manufacturers’ protocols. A 427-bp dsRNA for viral NP gene was generated using the dsNP-F and dsNP-R primers (S1 Table). A 420-bp dsRNA for the green fluorescent protein (GFP) gene was amplified as a negative control (S1 Table). dsRNAs (23 nL of 6 mg mL-1 solution) with crude preparations of insect-derived RSV were delivered into the hemolymph of nonviruliferous fourth-instar nymphs by microinjection through a glass needle using a Nanoliter 2000 system (World Precision Instruments, Sarasota, Florida, USA).

RNA preparation and quantitative real-time PCR for the assays of miRNAs, viral RNAs, and genes

Five to eight insects, or 20 salivary glands, 10 brains, 8 guts, and 8 fat bodies from nonviruliferous or viruliferous fourth-instar planthopper nymphs, or 8 samples of ovaries and testes from nonviruliferous or viruliferous planthopper adults, or RIP pulldown materials were used for RNA extraction by Trizol reagent (Invitrogen). DNase (Promega) was added to eliminate DNA contamination in the RNA samples. After quality control with a NanoDrop spectrophotometer (Thermo Scientific, Waltham, MA, USA) and gel electrophoresis, 2 μg of total RNA was reverse transcribed to cDNA in 20 μL by using a Superscript III first-strand synthesis system (Invitrogen) and random primers (Promega, Madison, WI, USA) according to the manufacturer’s instructions [33]. For reverse transcription of miRNAs, 1 μg of RNA was reverse transcribed by using miRNA RT enzyme mix of a miRcute plus miRNA first-strand cDNA kit (Tiangen, Beijing, China) [34].

The relative expression levels of miRNAs, planthopper genes, or viral RNAs were quantified by a miRcute miRNA qPCR detection kit (Tiangen) or LightCycler 480 SYBR Green I Master (Roche, Basel, Switzerland) using a LightCycler 480 instrument (Roche), respectively. Eight biological replicates were used for statistical analysis. U6 snRNA and planthopper EF2 gene [33] were used as endogenous controls for miRNAs, planthopper mRNAs, or viral RNAs, respectively. Viral RNA1 and RNA2 with the 3’-EUTR were identified by amplifying the 3’-UTR containing the extended terminal sequences. The fragments from the inner regions of RNA1 and RNA2 were amplified to measure the total RNA levels of RNA1 and RNA2, respectively. The ratios of RNA1 or RNA2 to the 3’ EUTRs were determined by associating the relative RNA levels of the 3’ EUTRs to those of the RNA levels of RNA1 and RNA2. Primers are listed in S1 Table.

To selectively quantify pri-miR-263a, pre-miR-263a, and miR-263a, specific primers were designed based on previous publications [34]. Briefly, the cDNA samples reverse transcribed by a Superscript III first-strand synthesis system (Invitrogen) were used to quantify pri-miR-263a. The cDNA samples reverse transcribed by a miRcute miRNA qPCR detection kit (Tiangen) were used to quantify pre-miR-263a and mature miR-263a. All products of PCR amplification were sequenced to verify the specificity of the primers. Differences were statistically evaluated using Student’s t-test to compare two means using SPSS 17.0. Primers are listed in S1 Table.

Determination of the viral promoter activity in vitro

To determine the promoter activities of viral 3’-UTRs, the pPol I plasmid (a gift from the laboratory of Prof. Yi Shi) containing the promoter sequence of human RNA polymerase I (Pol I) was used as a backbone to construct the recombinant plasmids. The reverse complementary sequence of Renilla luciferase gene flanked by the 5’-UTR and variable length of 3’-UTRs of RNA1 or RNA2 of RSV was inserted into the pPol I plasmid at the BglII site to generate the P1/P2-WT recombinant plasmids. The 3’-UTRs were replaced by 3’-EUTRs to generate the P1/P2-EUTR plasmids, respectively. Two deletion mutants P1/P2-TUTR lacking the terminal 16 nt of 3’-UTR of RNA1 or RNA2 were constructed based on the P1/P2-WT plasmids, respectively, using a KOD-Plus mutagenesis kit (Toyobo Bio-Technology, Osaka, Japan). The full-length ORFs of viral NP and RdRp were cloned from viruliferous planthopper cDNA by KOD FX DNA polymerase (Toyobo Bio-Technology) and were inserted into the pcDNA3.1 vector (Invitrogen) between the NotI and XbaI or BamHI and NotI restriction sites, respectively (S1 Table). The pGL4.13 plasmid (Promega) constitutively expressing firefly luciferase was used as a reference. A total of 0.5 μg of various recombinant pPol I plasmids (including P1/P2-WT, P1/P2-EUTR, and P1/P2-TUTR), 1 μg of pcDNA3.1-RdRp, 1 μg of pcDNA3.1-NP, and 0.1 μg of the pGL4.13 vector were cotransfected into 293T cells by Lipofectamine 3000 (Invitrogen). After transfection at 37°C for 45 h, the cells were harvested for dual luciferase reporter assay performed by using a Dual-Glo luciferase assay system (Promega) and a luminometer (Promega). Renilla luciferase expression was normalized to firefly luciferase expression. Six independent replicates were assayed for each treatment. Differences were statistically evaluated using Student’s t-test to compare two means using SPSS 17.0. Primers used in this experiment are listed in S1 Table.

Bioinformatics analysis of RNA polymerase I promoter of planthopper

In eukaryotes, the RNA Pol I promoter is generally located in the intergenic spacers between the 28S and 18S rRNA coding sequences [17], and the transcription initiation site is usually present within a region from 3,000 to 5,000 bp upstream of the 18S rRNA gene [18]. Thus, the 5,000 bp sequence upstream of the planthopper 18S rRNA gene was identified in contig36944 by querying the L. striatellus genome [17] with the 18S rRNA DNA sequence (GenBank accession no. AY591581.1). Possible RNA Pol I transcription initiation site was predicted by homology search with Bioedit software (version 7.2.5) based on highly conserved nucleotides around the transcription initiation site of 18S rRNA in eukaryotes [18]. The base frequency adjacent to the Pol I transcription initiation sites of L. striatellus and other eukaryotes (Homo sapiens, Rattus norvegicus, Mus musculus, Gallus gallus, Xenopus laevis, Aedes albopictus, Glossina morsitans, and Drosophila melanogaster) was identified using the WebLOGO tool.

Functional analysis of the effects of the 3’-extensions on viral promoter activities in vivo

A sequence ranging from -1 to -370 bp of the predicted planthopper RNA Pol I promoter was cloned using the LsPol I-F/L-Pol I-R primers (S1 Table) and was used to replace the original human Pol I promoter of the P1-EUTR or P2-EUTR vectors, respectively. Additionally, the RLuc gene was replaced by the GFP gene to generate the LsP1-EUTR or LsP2-EUTR plasmids (5 μg; Genewiz), respectively. Then, LsP1-WT or LsP2-WT with normal 3’-UTR of RNA1 or RNA2 were generated by using a KOD-Plus mutagenesis kit (Toyobo Bio-Technology), respectively (S1 Table). The LsP1/P2-WT plasmids (23 nL of 500 ng μL-1 solution) were injected into nonviruliferous or viruliferous fourth-instar nymphs. Then, equal amounts of the LsP1-WT and LsP1-EUTR or LsP2-WT and LsP2-EUTR plasmids (23 nL, 500 ng μL-1) were simultaneously injected into viruliferous fourth-instar nymphs. Six days after the injection, the nymphs were collected for protein extraction and tissue dissection. GFP protein level in the gut was evaluated by western blot using an anti-GFP monoclonal antibody (Abcam, Cambridge, UK). GADPH was detected by anti-GADPH polyclonal antibodies (Abcam) and was used as an internal control. The gut and salivary glands were dissected for GFP signal estimation performed using a Leica TCS SP5 confocal microscope (Leica Microsystems, Solms, Germany).

microRNA analysis and prediction of a potential target in planthopper

The small RNA sequencing data of viruliferous planthopper were obtained in our previous study [20]. To generate the planthopper miRNA datasets, clean reads mapped to rRNAs, tRNAs, snoRNAs, snRNAs, or virus-derived sRNAs identified in our previous studies [17,20] with perfect matching were discarded. The remaining sRNA reads were mapped to the planthopper genome with perfect matching allowing up to one mismatch using the mapper module of the miRDeep2 (version 2.0.0.5) software package. The statistics of the read positions (discrete manner) and frequency of reads within a potential hairpin are scored as measurement of the posterior probability that the hairpin is the stem-loop of planthopper miRNAs. The quantifier module was used to evaluate the expression levels based on read counts.

To determine whether 3’ extensions altered viral RNA-miRNA interactions, the miRanda [35] and RNAhybrid [36] algorithms were used to predict the potential binding sites of planthopper miRNAs in the 3’ extensions within approximately 200 bp upstream flanks of the RSV genome. The minimum free energy (MFE) of the RNA duplex was analyzed by miRanda using a cutoff of -18 kcal mol−1. The parameters of RNAhybrid were set as 1 hits per target, free energy threshold of –20 kcal mol−1, helix constraint from 2 to 8, max bulge loop length of 2, max internal loop length of 8, and two G:U included in the seed.

miRNA target validation

The putative targets of miR-263a (101 bp sequence containing the 3’extension of viral RNA1 and the upstream flanks) or the putative targets of miR-190 (200 bp sequence containing 15 bp of the 3’ extension of viral RNA2 and the upstream flanks) were cloned and inserted into the luciferase reporter vector (pLightSwitch UTR plasmid; Active Motif, Carlsbad, CA, USA) between the BglII and NcoI restriction sites to generate pLightSwitch-RNA1-3’EUTR and pLightSwitch-RNA2-3’EUTR using the pLS-R1-EUTR-F/R and pLS-R2-EUTR-F/R primers, respectively (S1 Table). Sequences containing the 3’-UTRs of RNA1 or RNA2 were also inserted to generate pLightSwitch-RNA1-3’UTR and pLightSwitch-RNA2-3’UTR, respectively, by using a KOD-Plus mutagenesis kit (Toyobo Bio-Technology) using the pLS-R1-UTR-F/R and pLS-R2-UTR-F/R primers (S1 Table). The pGL4.13 plasmid expressing firefly luciferase (Fluc) was used as a reference control. The sequence of a Caenorhabditis elegans miRNA, cel-miR-67-3p (5’-UCACAACCUCCUAGAAAGAGUAGA-3’), was used as a control mimic (NC). 293T cells were cotransfected with the recombinant pLightSwitch plasmids (1 μg), miRNA/control mimics (RiboBio, China) at various concentrations (50 pM, 10 nM, 50 nM, and 100 nM), and 1 μg of the pGL4.13 plasmid using Lipofectamine 3000 (Invitrogen). miRNA mimics corresponded to the sequences of miR-263a (5’-AATGGCACTGGAAGAATTCACGGG-3’) and miR-190 (5’-AGATATGTTTGATATTCTTGG-3’). The relative activity of Renilla luciferase (Rluc) encoded by the recombinant pLightSwitch plasmids normalized to Fluc activity is presented as the mean ± SE. The activities of firefly and Renilla luciferases were measured 45 h after the transfection using a Dual-Glo luciferase assay system (Promega, Madison, USA).

miRNA antagomir and agomir injection

miRNA inhibition or overexpression was performed by the injection of antagomir or agomir into the insects [34]. The sequence of cel-miR-67-3p was used as a negative control of antagomir or agomir (antagomir-NC or agomir-NC). A total volume of 13.8 nL of antagomir-263a/NC, agomir-263a/NC, antagomir-190/NC, or agomir-190/NC (500 μM; RiboBio) was delivered into the fourth-instar viruliferous nymphs by microinjection using a Nanoliter 2000 system (World Precision Instruments). Four days after the injection, planthoppers were collected for RNA or protein isolation.

Nonviruliferous fourth-instar nymphs were fed an artificial diet containing crude rice-derived RSV extract as described previously. After feeding RSV for 24 h, the RSV-infected nymphs were injected with agomir-263a or agomir-NC, transferred to healthy rice plants, and incubated for additional 1 d and 4 d for collection. Eight replicates and five insects per replicate at each time point were prepared. Samples from this experiment were used for RNA extraction and qPCR analysis.

RNA immunoprecipitation (RIP) assays

The RIP assay was performed using a Magna RIP RNA-binding protein immunoprecipitation kit (Merck Millpore, Billerica, MA, USA) [21]. Four days after the injection with agomir-263a/NC or agomir-190/NC, 20 fourth-instar viruliferous nymphs were homogenized in ice-cold RIP lysis buffer; the homogenate was incubated with 20 μL of magnetic beads and 2 μg of a monoclonal Ago1 antibody [21] or normal mouse IgG (Abcam) at 4°C overnight. A 1/10 fraction of the supernatant of the lysate was stored as an “input” sample. The immunoprecipitates were digested with proteinase K to remove proteins and release RNAs. Total RNA was extracted by TRIzol reagent and reverse transcribed into cDNA using random primers. qPCR was performed to detect the expression levels of the target fragments. The RNA level of each target segment relative to that in the IgG control sample is reported as the mean ± SE. A value of Ago/IgG above one indicates that the target segment may interact with the corresponding Ago protein. Student’s t-test was performed to evaluate the differences between the two means using SPSS 17.0.

Pulldown assay with biotinylated miRNA

Biotinylated miR-263a (bio-miR-263a; 23 nL) or control biotinylated scrambled sequences (bio-Scr) (50 nM; RiboBio) was injected into the fourth-instar viruliferous nymphs by microinjection using a Nanoliter 2000 system (World Precision Instruments). Two days after the injection, the nymphs were collected, homogenized, and lysed with ice-cold lysis buffer (0.02 M Tris-HCl, 0.1 M KCl, 5 mM MgCl2, and 0.5% NP-40, pH 7.5) supplemented with a proteinase inhibitor (Sigma, Santa Clara, CA, USA) and an RNase inhibitor (Promega) for 10 min. The lysates were centrifuged at 13,000 rpm for 10 min at 4°C, and 1/10 of the supernatant was saved as an “input”. To exclude RNA and protein complexes, M-280 streptavidin beads (Thermo Scientific) were blocked in lysis buffer containing RNase-free bovine serum albumin (BSA) and yeast tRNA (Sigma). The remaining supernatant of the lysate was incubated with the beads at 4°C for 4 h and sequentially washed with lysis buffer, low salt buffer, and high salt buffer. Bound RNAs were isolated using TRIzol reagent. RNA from biotin-labeled samples (20 μL) and 1 μg of RNA from the “input” samples were used for qPCR. Four replicates were performed. Student’s t-test was performed to evaluate the differences between the two means using SPSS 17.0, and P values < 0.05 were considered statistically significant. The miRNA enrichment was calculated as follows:

Enrichmentratio=biomiRpulldown/bioSrcpulldownbiomiRinput/bioSrcinput (1)

Fluorescence in situ hybridization

Double fluorescence in situ hybridization (dFISH) was performed in the gut of viruliferous planthopper. Prehybridization solution contained biotin-labeled miR-263a probe (1 μmol mL-1; RiboBio) and digoxigenin (DIG)-labeled RNA1 3’-EUTR probe (50 ng μL-1) for hybridization. Approximately 200 bp of the RNA probe for 3’-EUTR of RNA1 was synthesized by a T7/SP6 RNA transcription kit (Roche). The gut was dissected from viruliferous fourth-instar nymphs and fixed in 4% (wt vol-1) paraformaldehyde at 4°C overnight. After washing with PBST, the gut was digested with proteinase K (20 μg mL-1; Tiangen) at 37°C for 15 min and hybridized with the miRNA probe and viral RNA probe at 37°C overnight. The midgut was washed with 0.2× SSC at 37°C to remove nonspecific binding. An anti-DIG alkaline phosphatase-conjugated antibody (1:500) and an anti-biotin antibody (1:100) were used for signal detection. Probes for cel-miR-67-3p and rice PsbP gene (KF460579) were used in the control group. The fluorescent signal of DIG or biotin was generated by HNPP/Fast Red or fluorescein-tyramide (Perkin-Elmer, Waltham, MA, USA) and recorded using a Leica TCS SP5 confocal microscope (Leica Microsystems). Primers used for probe synthesis are listed in S1 Table.

Functional analysis of the effect of miR-263a on viral promoter activities in vitro and in vivo

To determine the effect of miR-263a on viral promoter activities in vitro, 80 ng of miR-263a/NC mimics were cotransfected with 0.5 μg of the P1-EUTR or P1-WT plasmids into 293T cells with coexpression of RSV RdRp and NP. Forty-five hours after the transfection at 37°C, the cells were harvested for dual luciferase reporter assay.

To determine the effect of miR-263a on viral promoter activities in planthopper, equal amounts of LsP1-WT and LsP1-EUTR (23 nL, 500 ng μL-1) were microinjected into viruliferous fourth-instar nymphs by microinjection using a Nanoliter 2000 system (World Precision Instruments). Two days after the injection, a total volume of 13.8 nL (500 μM) of agomir-263a/agomir-NC was delivered into the nymphs. Four days after the additional injection, the gut and salivary glands were dissected for GFP detection.

Western blot

Proteins from 0.5 mL of 293T cells (1 × 106 cells mL−1) or 20 guts or 10 bodies of the treated fourth-instar planthopper nymphs were extracted using 1×PBS (pH 7.2) at 4°C, separated by SDS-PAGE, and detected by using anti-GADPH polyclonal antibodies (Abcam), an in-house generated anti-NP monoclonal antibody, in-house generated anti-RdRp polyclonal antibodies, or anti-Renilla luciferase polyclonal antibodies (Abcam). Signal of the immunoblot was detected by an eECL western blot kit (CWBIO, Beijing, China).

Northern blot

Total RNA was extracted from 293T cells after cotransfection with the equivalent amounts of P1-EUTR, pcDNA-NP, pcDNA-RdRp, and miR-263a mimics/NC. RNA molecules were resolved by electrophoresis through a 0.9% agarose denaturing gel and transferred onto a positively charged nylon membrane (Invitrogen). Biotin-labeled LNA oligonucleotide probes (20 ng mL-1, RiboBio) for the antisense sequence of Rluc mRNA (5’-bio-CCTACGAGCACCAAGACAAGATCAAGGCCA-bio-3’) and 18S rRNA (5’-bio-CGGAACTACGACGGTATCTG-bio-3’) were hybridized to the membrane in prehybridization/hybridization buffer (Invitrogen) at 42°C for 16 h. The membranes were blocked with blocking buffer (Invitrogen) for 15 min at room temperature followed by incubation with hybridization buffer (Invitrogen) containing a streptavidin-HRP conjugate (1:500, Invitrogen) for 1 h at RT. Detection was carried out by an eECL western blot kit (CWBIO).

Overexpression of miR-263a in 293T cells

A miRNA-expressing vector based on the pcDNA 3.1 expression vector (Invitrogen) was constructed to overexpress miR-263a in 293T cells. Approximately 500 bp fragments centered around pre-miR-263a were amplified from the planthopper genome and ligated into the pcDNA 3.1 plasmid between the KpnI and XhoI sites. The empty pcDNA 3.1 plasmid was used as a negative control. miR-263a expression was performed by transfection of the recombinant plasmids into 293T cells using Lipofectamine 3000. miR-263a overexpression efficiency was verified by qPCR using the primers listed in S1 Table.

Supporting information

S1 Fig

The secondary structure of RSV genomic RNA1 (A) and RNA2 (B) with the regular 3’ untranslated terminal region (3’-UTR) or extended untranslated terminal region (3’-EUTR). The complementary 3’ and 5’ termini are marked with red boxes and enlarged on the right side.

(TIF)

S2 Fig

Sequence alignments of miR-263a (A) and miR-190 (B) with the predicted target site in the 3’ terminus of RSV RNA1 or RNA2. The 3’ extensions are shown in red. The G:U wobble pairs are labeled.

(TIF)

S3 Fig. The effects of antagomir and agomir of miR-263a on RSV replication in the gut of viruliferous planthopper.

(A) Relative transcript levels of miR-263a in the gut of viruliferous planthopper determined by real-time quantitative PCR (qPCR) 4 d after the injection of antagomir-263a or agomir-263a. Antagomir control (antagomir-NC) or agomir control (agomir-NC) were injected and used as the negative controls. (B) Relative RNA levels of RSV RNA1, RdRp, and NP in the gut of viruliferous planthopper measured by qPCR 4 d after the injection of antagomir-263a or antagomir-NC. (C) Relative RNA levels of RSV RNA1, RdRp, and NP in the gut of viruliferous planthopper measured by qPCR 4 d after the injection of agomir-263a or agomir-NC. The values in (A), (B), and (C) represent the mean ± SE. NS, no significant differences. *, P < 0.05. ***, P < 0.001. (D) Western blot analysis of RSV NP protein in the samples of (B) and (C) using an in-house generated anti-NP monoclonal antibody. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control.

(TIF)

S4 Fig. Predicted stem-loop structure of pre-miR-263a.

Mature miR-263a is marked with a line. Base-pair probabilities (from 0 to 1) are shown in the heatmap.

(TIF)

S5 Fig. Relative RNA levels of RSV NP in the whole body of planthopper 5 d after the injection of crude RSV preparations and double-stranded RNAs of RSV NP (dsNP) or double-stranded RNAs of GFP (dsGFP).

The values were determined by real-time quantitative PCR and are presented as the mean ± SE. *, P < 0.05.

(TIF)

S1 Table. Primers used in this study.

(DOCX)

Acknowledgments

We are grateful to Prof. Yi Shi from Institute of Microbiology, Chinese Academy of Sciences for providing the pPol I plasmid and guiding us to construct the reverse genetic system.

Data Availability

All relevant data are within the manuscript and its Supporting Information files.

Funding Statement

This work was supported by grants from the National Natural Science Foundation of China (No. 32090012) for W.Z., the State Key Research Development Program of China (No. 2019YFC1200504) for W.W., the Chinese Academy of Sciences (No. ZDBS-LY-SM027) for F.C., and the Youth Innovation Promotion Association, CAS (No. 2019086) for W.Z. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Cheng G, Liu Y, Wang P, Xiao X. Mosquito defense strategies against viral infection. Trends Parasitol. 2016;32(3):177–186. Epub 2015/12/03. 10.1016/j.pt.2015.09.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Blanc S, Gutierrez S. The specifics of vector transmission of arboviruses of vertebrates and plants. Curr Opin Virol. 2015;15:27–33. Epub 2015/08/04. 10.1016/j.coviro.2015.07.003 . [DOI] [PubMed] [Google Scholar]
  • 3.Wang L, Valderramos SG, Wu A, Ouyang S, Li C, Brasil P, et al. From mosquitos to humans: Genetic evolution of Zika virus. Cell Host & Microbe. 2016;19(5):561–565. WOS:000375595500001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Chen R, Wang E, Tsetsarkin KA, Weaver SC. Chikungunya virus 3’ untranslated region: adaptation to mosquitoes and a population bottleneck as major evolutionary forces. PLoS Pathog. 2013;9(8):e1003591–e. 10.1371/journal.ppat.1003591 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Pflug A, Guilligay D, Reich S, Cusack S. Structure of influenza A polymerase bound to the viral RNA promoter. Nature. 2014;516(7531):355–360. Epub 2014/11/20. 10.1038/nature14008 . [DOI] [PubMed] [Google Scholar]
  • 6.Barr JN, Elliott RM, Dunn EF, Wertz GW. Segment-specific terminal sequences of Bunyamwera bunyavirus regulate genome replication. Virology. 2003;311(2):326–338. 10.1016/s0042-6822(03)00130-2 . [DOI] [PubMed] [Google Scholar]
  • 7.Flick R, Elgh F, Pettersson RF. Mutational analysis of the uukuniemi virus (Bunyaviridae family) promoter reveals two elements of functional importance. J Virol. 2002;76(21):10849–10860. 10.1128/jvi.76.21.10849-10860.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Dubey SK, Shrinet J, Sunil S. Aedes aegypti microRNA, miR-2944b-5p interacts with 3’UTR of chikungunya virus and cellular target vps-13 to regulate viral replication. PLOS Negl Trop Dis. 2019;13(6):e0007429. 10.1371/journal.pntd.0007429 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Sedano CD, Sarnow P. Hepatitis C virus subverts liver-specific miR-122 to protect the viral genome from exoribonuclease Xrn2. Cell Host & Microbe. 2014;16(2):257–264. Epub 2014/08/15. 10.1016/j.chom.2014.07.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Zhao W, Xu Z, Zhang X, Yang M, Kang L, Liu R, et al. Genomic variations in the 3’-termini of Rice stripe virus in the rotation between vector insect and host plant. New Phytol. 2018;219(3):1085–1096. Epub 2018/06/09. 10.1111/nph.15246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Toriyama S. Rice stripe virus: prototype of a new group of viruses that replicate in plants and insects. Microbiol Sci. 1986;3(11):347–351. Epub 1986/11/01. . [PubMed] [Google Scholar]
  • 12.Toriyama S, Takahashi M, Sano Y, Shimizu T, Ishihama A. Nucleotide sequence of RNA 1, the largest genomic segment of rice stripe virus, the prototype of the tenuiviruses. J Gen Virol. 1994;75:3569–3579. 10.1099/0022-1317-75-12-3569 . [DOI] [PubMed] [Google Scholar]
  • 13.Zhu Y, Hayakawa T, Toriyama S, Takahashi M. Complete nucleotide sequence of RNA 3 of rice stripe virus: an ambisense coding strategy. J Gen Virol. 1991;72:763–767. Epub 1991/04/01. 10.1099/0022-1317-72-4-763 . [DOI] [PubMed] [Google Scholar]
  • 14.Takahashi M, Toriyama S, Kikuchi Y, Hayakawa T, Ishihama A. Complementarity between the 5′- and 3′-terminal sequences of rice stripe virus RNAs. J Gen Virol. 1990;71:2817–2821. 10.1099/0022-1317-71-12-2817 . [DOI] [PubMed] [Google Scholar]
  • 15.Barbier P, Takahashi M, Nakamura I, Toriyama S, Ishihama A. Solubilization and promoter analysis of RNA polymerase from rice stripe virus. J Gen Virol. 1992;66(10):6171–6174. 10.1128/JVI.66.10.6171-6174.1992 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Pleschka S, Jaskunas R, Engelhardt OG, Zürcher T, Palese P, García-Sastre A. A plasmid-based reverse genetics system for influenza A virus. J Virol. 1996;70(6): 4188–4192. Epub 1996/06/01. 10.1128/JVI.70.6.4188-4192.1996 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Zhu J, Jiang F, Wang X, Yang P, Bao Y, Zhao W, et al. Genome sequence of the small brown planthopper, Laodelphax striatellus. Gigascience. 2017;6(12):1–12. Epub 2017/11/15. 10.1093/gigascience/gix109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Grummt I. Life on a planet of its own: regulation of RNA polymerase I transcription in the nucleolus. Genes Dev. 2003;17(14):1691–1702. Epub 2003/07/17. 10.1101/gad.1098503R . [DOI] [PubMed] [Google Scholar]
  • 19.Wang Z, Duke GM. Cloning of the canine RNA polymerase I promoter and establishment of reverse genetics for influenza A and B in MDCK cells. Virol J. 2007;4:102. Epub 2007/10/25. 10.1186/1743-422X-4-102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Yang M, Xu Z, Zhao W, Liu Q, Li Q, Lu L, et al. Rice stripe virus-derived siRNAs play different regulatory roles in rice and in the insect vector Laodelphax striatellus. BMC Plant Biol. 2018;18(1):219. Epub 2018/10/06. 10.1186/s12870-018-1438-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Zhao W, Li Q, Cui F. Potential functional pathways of plant RNA virus-derived small RNAs in a vector insect. Methods. 2019;183:38–42. Epub 2019/10/28. 10.1016/j.ymeth.2019.10.006 . [DOI] [PubMed] [Google Scholar]
  • 22.Khromykh AA, Meka H, Guyatt KJ, Westaway EG. Essential role of cyclization sequences in flavivirus RNA replication. J Virol. 2001;75(14):6719–6728. 10.1128/JVI.75.14.6719-6728.2001 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Men R, Bray M, Clark D, Chanock RM, Lai CJ. Dengue type 4 virus mutants containing deletions in the 3’ noncoding region of the RNA genome: analysis of growth restriction in cell culture and altered viremia pattern and immunogenicity in rhesus monkeys. J Virol. 1996;70(6):3930–3937. Epub 1996/06/01. 10.1128/JVI.70.6.3930-3937.1996 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Kohl A, Dunn EF, Lowen AC, Elliott RM. Complementarity, sequence and structural elements within the 3’ and 5’ non-coding regions of the Bunyamwera orthobunyavirus S segment determine promoter strength. J Gen Virol. 2004;85: 3269–3278. Epub 2004/10/16. 10.1099/vir.0.80407-0 . [DOI] [PubMed] [Google Scholar]
  • 25.Villordo SM, Filomatori CV, Sanchez-Vargas I, Blair CD, Gamarnik AV. Dengue virus RNA structure specialization facilitates host adaptation. PLoS Pathog. 2015;11: e1004604. 10.1371/journal.ppat.1004604 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Lecellier CH, Dunoyer P, Arar K, Lehmann-Che J, Eyquem S, Himber C, et al. A cellular microRNA mediates antiviral defense in human cells. Science. 2005;308(5721):557–560. Epub 2005/04/23. 10.1126/science.1108784 . [DOI] [PubMed] [Google Scholar]
  • 27.Zheng Z, Ke X, Wang M, He S, Li Q, Zheng C, et al. Human microRNA hsa-miR-296-5p suppresses enterovirus 71 replication by targeting the viral genome. J Virol. 2013;87(10):5645–5656. Epub 2013/03/06. 10.1128/JVI.02655-12 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Potenza N, Papa U, Mosca N, Zerbini F, Nobile V, Russo A. Human microRNA hsa-miR-125a-5p interferes with expression of hepatitis B virus surface antigen. Nucleic Acids Res. 2011;39(12):5157–5163. Epub 2011/02/11. 10.1093/nar/gkr067 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Chen Y, Shen A, Rider PJ, Yu Y, Wu K, Mu Y, et al. A liver-specific microRNA binds to a highly conserved RNA sequence of hepatitis B virus and negatively regulates viral gene expression and replication. FASEB J. 2011;25(12):4511–4521. Epub 2011/09/08. 10.1096/fj.11-187781 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Henke JI, Goergen D, Zheng J, Song Y, Schüttler CG, Fehr C, et al. microRNA-122 stimulates translation of hepatitis C virus RNA. Embo J. 2008;27(24):3300–3310. Epub 2008/11/21. 10.1038/emboj.2008.244 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Machlin ES, Sarnow P, Sagan SM. Masking the 5’ terminal nucleotides of the hepatitis C virus genome by an unconventional microRNA-target RNA complex. Proc Natl Acad Sci U S A. 2011;108(8):3193–3198. Epub 2011/01/10. 10.1073/pnas.1012464108 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Shimakami T, Yamane D, Jangra RK, Kempf BJ, Spaniel C, Barton DJ, et al. Stabilization of hepatitis C virus RNA by an Ago2-miR-122 complex. Proc Natl Acad Sci U S A. 202;109(3):941–946. Epub 2012/01/03. 10.1073/pnas.1112263109 . [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Zhao W, Yang P, Kang L, Cui F. Different pathogenicities of Rice stripe virus from the insect vector and from viruliferous plants. New Phytol. 2016;210(1):196–207. Epub 2015/11/21. 10.1111/nph.13747 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Guo X, Ma Z, Du B, Li T, Li W, Xu L, et al. Dop1 enhances conspecific olfactory attraction by inhibiting miR-9a maturation in locusts. Nat Commun. 2018;9(1):1193. Epub 2018/03/24. 10.1038/s41467-018-03437-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Enright AJ, John B, Gaul U, Tuschi T, Sander C, Marks DS. MicroRNA targets in Drosophila. Genome Biol. 2003;5:R1. 10.1186/gb-2003-5-1-r1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Rehmsmeier M, Steffen P, Hochsmann M, Giegerich R. Fast and effective prediction of microRNA/target duplexes. RNA. 2004;10(10):1507–1517. Epub 2004/09/24. 10.1261/rna.5248604 [DOI] [PMC free article] [PubMed] [Google Scholar]

Decision Letter 0

Shou-Wei Ding, Aiming Wang

12 Feb 2021

Dear Dr. Cui,

Thank you very much for submitting your manuscript "Coordination between terminal variation of the viral genome and insect microRNAs regulates rice stripe virus replication in vector insects" for consideration at PLOS Pathogens. As with all papers reviewed by the journal, your manuscript was reviewed by members of the editorial board and by several independent reviewers. The reviewers appreciated the attention to an important topic. Based on the reviews, we are likely to accept this manuscript for publication, providing that you modify the manuscript according to the review recommendations.

Please prepare and submit your revised manuscript within 30 days. If you anticipate any delay, please let us know the expected resubmission date by replying to this email. 

When you are ready to resubmit, please upload the following:

[1] A letter containing a detailed list of your responses to all review comments, and a description of the changes you have made in the manuscript. 

Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out

[2] Two versions of the revised manuscript: one with either highlights or tracked changes denoting where the text has been changed; the other a clean version (uploaded as the manuscript file).

Important additional instructions are given below your reviewer comments.

Thank you again for your submission to our journal. We hope that our editorial process has been constructive so far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Aiming Wang, Ph.D

Associate Editor

PLOS Pathogens

Shou-Wei Ding

Section Editor

PLOS Pathogens

Kasturi Haldar

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0001-5065-158X

Michael Malim

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0002-7699-2064

***********************

Reviewer Comments (if any, and for reference):

Reviewer's Responses to Questions

Part I - Summary

Please use this section to discuss strengths/weaknesses of study, novelty/significance, general execution and scholarship.

Reviewer #1: This manuscript involves investigating the role of 3’ RNA extensions in the RNA1 and RNA2 genome segments of Rice stripe virus. The extensions were shown to inhibit virus levels and to act at the virus genome replication stage. The extensions were more prevalent in the planthopper vector, and a planthopper miRNA interacts with the RNA1 extension and reduces its inhibitory effect. In general, the experiments are well done and the data convincing. Suggestions for improvement are provided.

Reviewer #2: The manuscript submitted by Zhao et al. entitled “Coordination between terminal variation of the viral genome and 2 insect microRNAs regulates rice stripe virus replication in vector insects” describe an interesting story that the interaction of rice stripe virus (RSV, a member of tenuivirus) with it’s insect vector host small brown hopper was elaborately regulated. By using a reverse genetic system for RSV in human cells and a vector insect, they demonstrated that the 3'-terminal extensions of RSV RNA 1 & 2 suppress viral replication in vector insects by inhibiting promoter activity due to the formation of a panhandle structure by viral 3’- UTRs. They further showed that the extension sequence in the viral RNA1 segment was targeted by an endogenous insect microRNA, miR31 263a, which compromised the inhibitory effect of the extension sequence on viral promoter activity. However, the expression of miR-263a was down-regulated by RSV infection. The research is well designed and results are convincing. But before acceptance for publication, there are some issues needed to be addressed.

Reviewer #3: Zhao et al investigated the effect of 3’UTR extensions of RNA1 and RNA2 segments of rice stripe virus in viral replication in the planthopper Laodelphax striatellus. They first showed that the extensions are mainly produced when the virus replicates in planthoppers compared to when it replicates in the rice plant. The extensions were shown to inhibit replication of RSV. The authors went through extensive experimentation to show that the 3’UTRs of RNA1 and RNA2 serve as promoters and the extensions in the 3’UTRs interfere with their promoter activities by disrupting the panhandle structure normally formed by viral 3’ and 5’UTRs. Next, they found that a planthopper miRNA, miR-263a, specifically binds to the extended 3’UTR of RNA1 and decreases the inhibitory effect of the extension sequences. RSV infection itself reduced the transcription of the primary transcript of miR-263a and hence the abundance of the mature miRNA, a mechanism for regulating virus replication in planthoppers. This is a comprehensive and well executed work, with conclusions supported by the results. I only have minor comments.

**********

Part II – Major Issues: Key Experiments Required for Acceptance

Please use this section to detail the key new experiments or modifications of existing experiments that should be absolutely required to validate study conclusions.

Generally, there should be no more than 3 such required experiments or major modifications for a "Major Revision" recommendation. If more than 3 experiments are necessary to validate the study conclusions, then you are encouraged to recommend "Reject".

Reviewer #1: Line 121 – need to provide rationale/explanation here for why using Human cells to study plant/insect virus

Lines 156-169 – the details of acquiring the planthopper Pol I sequence is distracting and is not needed here – suggest moving it to M&M section

Line 223 and Fig.3E – in the FISH experiment, how would the miR-263a probe (which is complementary to miR-263a) hybridize with mirR-263a when it is base-paired to the 3’ extension?

The Discussion needs improvement:

Provide an integrate findings into working model – cycling from plant to insect.

Gamarnik lab work on Dengue’s 3’UTR in human vs mosquito cells should be discussed.

Do the extensions also inhibit replication in plant cells?

Possible mechanisms for how the 3’-extension is added should be mentioned.

Ideas on how the 3’UTR extension are preferentially added or maintained in insect cells.

Reviewer #2: Major:

1. Authors should provide detailed information of RSV population before and after given planthopper. The information includes sequences 3’ – terminal of RNA 1 & 2 and percentage of major groups such, no extension, with extensions.

2. In all of the results presented, RNA or 3’-EUTR level was only detected by Q-RT-PCR. But did not provide how the results were calculated. The authors should use other methods such sequencing or northern blots to confirm their results.

Reviewer #3: none

**********

Part III – Minor Issues: Editorial and Data Presentation Modifications

Please use this section for editorial suggestions as well as relatively minor modifications of existing data that would enhance clarity.

Reviewer #1: Abstract – line29 change to “…due to structural interference with the panhandle structure formed…”

Line 73 – change to “.., negative-sense RNA1 and ambisense RNA2, of

Line 73 – change semicolon to a period

Figure1C – red bar for NP should have white gap where the scale changes in the y-axis

Line 119 – change “damage” to “modify”

Line 122 – change to “the human RNA polymerase I (Pol I) normally directs the synthesis of ribosomal RNAs (rRNA) [16]. – delete “that function as an RNA template”.

Line 131 – change “, rRNA of…” to “, the Rluc reporter RNA……..was transcribed by Pol I.”

Line 132 – change “Rluc rRNA served…” to “Rluc reporter RNA served..”

Line 151 – change to “inhibit the promoter activity of viral 3’-UTRs in human cells.”

Line 155 – change to “activity in human cells”

Line 195 – please indicate the criteria used to select candidate miRNAs

Fig3 – please comment on the endogenous levels of miR-263a vs the concentrations being injected

For Figs3, 4 and 5 – might want to consider standardizing the color scheme of the graphs – Figs 3 and 5 are in black & white, whereas Fig 4 is in color.

Line 353 – please expand more specifically on how the miR interaction and Ago binding could “relieve” the inhibitory effect of the 3’-extension

Reviewer #2: Minor:

1. In Fig. 1 A to D, the sampling time should be provided either in results or fig legends.

2. In Fig. 2 B, how much of P1-WT and P2-WT were added in these two experiments?

3. Line 35, rice strip virus is member of Tenuivirus genus, not Phytoreovirus.

4. Line 41 “rice strip viruses” should be corrected to “rice strip virus”.

5. Line 77 “an RdRp” should be “a RdRp”.

Reviewer #3: Line 2 and the rest of the manuscript: I believe “insect vectors” is more commonly used than “vector insects”. The latter doesn’t read as well as the former. In a small number of places in the manuscript “insect vectors” has been used. It would be good to have consistency, either way.

Line 27: insert the scientific name of the planthopper as it is not mentioned in the title or the abstract.

Line 56: one of the major…

Line 100 and 109: technically, the amount of RNA is not equivalent to the number of virions. Virion should be replaced with viral RNAs

Line 139: with the assistance of….

Line 230: the authors should be consistent with the usage of mimic/agomir. There are other instances of interchangeable usage of the terms.

Line 285: injected into…

Line 456 and 583: Functional analysis of …..

Fig. 3E: The figure is not really clear in that it doesn’t have a defined nucleus and cytoplasmic region. Does the enlarged image in the lower row represent one cell?

**********

PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: No

Reviewer #2: No

Reviewer #3: No

Figure Files:

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email us at figures@plos.org.

Data Requirements:

Please note that, as a condition of publication, PLOS' data policy requires that you make available all data used to draw the conclusions outlined in your manuscript. Data must be deposited in an appropriate repository, included within the body of the manuscript, or uploaded as supporting information. This includes all numerical values that were used to generate graphs, histograms etc.. For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5.

Reproducibility:

To enhance the reproducibility of your results, PLOS recommends that you deposit laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. For instructions see http://journals.plos.org/plospathogens/s/submission-guidelines#loc-materials-and-methods

Decision Letter 1

Shou-Wei Ding, Aiming Wang

25 Feb 2021

Dear Dr. Cui,

We are pleased to inform you that your manuscript 'Coordination between terminal variation of the viral genome and insect microRNAs regulates rice stripe virus replication in insect vectors' has been provisionally accepted for publication in PLOS Pathogens.

Before your manuscript can be formally accepted you will need to complete some formatting changes, which you will receive in a follow up email. A member of our team will be in touch with a set of requests.

Please note that your manuscript will not be scheduled for publication until you have made the required changes, so a swift response is appreciated.

IMPORTANT: The editorial review process is now complete. PLOS will only permit corrections to spelling, formatting or significant scientific errors from this point onwards. Requests for major changes, or any which affect the scientific understanding of your work, will cause delays to the publication date of your manuscript.

Should you, your institution's press office or the journal office choose to press release your paper, you will automatically be opted out of early publication. We ask that you notify us now if you or your institution is planning to press release the article. All press must be co-ordinated with PLOS.

Thank you again for supporting Open Access publishing; we are looking forward to publishing your work in PLOS Pathogens.

Best regards,

Aiming Wang

Associate Editor

PLOS Pathogens

Shou-Wei Ding

Section Editor

PLOS Pathogens

Kasturi Haldar

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0001-5065-158X

Michael Malim

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0002-7699-2064

***********************************************************

Reviewer Comments (if any, and for reference):

Acceptance letter

Shou-Wei Ding, Aiming Wang

5 Mar 2021

Dear Dr. Cui,

We are delighted to inform you that your manuscript, "Coordination between terminal variation of the viral genome and insect microRNAs regulates rice stripe virus replication in insect vectors," has been formally accepted for publication in PLOS Pathogens.

We have now passed your article onto the PLOS Production Department who will complete the rest of the pre-publication process. All authors will receive a confirmation email upon publication.

The corresponding author will soon be receiving a typeset proof for review, to ensure errors have not been introduced during production. Please review the PDF proof of your manuscript carefully, as this is the last chance to correct any scientific or type-setting errors. Please note that major changes, or those which affect the scientific understanding of the work, will likely cause delays to the publication date of your manuscript. Note: Proofs for Front Matter articles (Pearls, Reviews, Opinions, etc...) are generated on a different schedule and may not be made available as quickly.

Soon after your final files are uploaded, the early version of your manuscript, if you opted to have an early version of your article, will be published online. The date of the early version will be your article's publication date. The final article will be published to the same URL, and all versions of the paper will be accessible to readers.

Thank you again for supporting open-access publishing; we are looking forward to publishing your work in PLOS Pathogens.

Best regards,

Kasturi Haldar

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0001-5065-158X

Michael Malim

Editor-in-Chief

PLOS Pathogens

orcid.org/0000-0002-7699-2064

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 Fig

    The secondary structure of RSV genomic RNA1 (A) and RNA2 (B) with the regular 3’ untranslated terminal region (3’-UTR) or extended untranslated terminal region (3’-EUTR). The complementary 3’ and 5’ termini are marked with red boxes and enlarged on the right side.

    (TIF)

    S2 Fig

    Sequence alignments of miR-263a (A) and miR-190 (B) with the predicted target site in the 3’ terminus of RSV RNA1 or RNA2. The 3’ extensions are shown in red. The G:U wobble pairs are labeled.

    (TIF)

    S3 Fig. The effects of antagomir and agomir of miR-263a on RSV replication in the gut of viruliferous planthopper.

    (A) Relative transcript levels of miR-263a in the gut of viruliferous planthopper determined by real-time quantitative PCR (qPCR) 4 d after the injection of antagomir-263a or agomir-263a. Antagomir control (antagomir-NC) or agomir control (agomir-NC) were injected and used as the negative controls. (B) Relative RNA levels of RSV RNA1, RdRp, and NP in the gut of viruliferous planthopper measured by qPCR 4 d after the injection of antagomir-263a or antagomir-NC. (C) Relative RNA levels of RSV RNA1, RdRp, and NP in the gut of viruliferous planthopper measured by qPCR 4 d after the injection of agomir-263a or agomir-NC. The values in (A), (B), and (C) represent the mean ± SE. NS, no significant differences. *, P < 0.05. ***, P < 0.001. (D) Western blot analysis of RSV NP protein in the samples of (B) and (C) using an in-house generated anti-NP monoclonal antibody. GADPH was detected using anti-GADPH polyclonal antibodies and was used as an internal control.

    (TIF)

    S4 Fig. Predicted stem-loop structure of pre-miR-263a.

    Mature miR-263a is marked with a line. Base-pair probabilities (from 0 to 1) are shown in the heatmap.

    (TIF)

    S5 Fig. Relative RNA levels of RSV NP in the whole body of planthopper 5 d after the injection of crude RSV preparations and double-stranded RNAs of RSV NP (dsNP) or double-stranded RNAs of GFP (dsGFP).

    The values were determined by real-time quantitative PCR and are presented as the mean ± SE. *, P < 0.05.

    (TIF)

    S1 Table. Primers used in this study.

    (DOCX)

    Attachment

    Submitted filename: Responses to comments-R.docx

    Data Availability Statement

    All relevant data are within the manuscript and its Supporting Information files.


    Articles from PLoS Pathogens are provided here courtesy of PLOS

    RESOURCES