Skip to main content
  • Research article
  • Open access
  • Published:

Salicylic acid treatment and expression of an RNA-dependent RNA polymerase 1 transgene inhibit lethal symptoms and meristem invasion during tobacco mosaic virus infection in Nicotiana benthamiana

Abstract

Background

Host RNA-dependent RNA polymerases (RDRs) 1 and 6 contribute to antiviral RNA silencing in plants. RDR6 is constitutively expressed and was previously shown to limit invasion of Nicotiana benthamiana meristem tissue by potato virus X and thereby inhibit disease development. RDR1 is inducible by salicylic acid (SA) and several other phytohormones. But although it contributes to basal resistance to tobacco mosaic virus (TMV) it is dispensable for SA-induced resistance in inoculated leaves. The laboratory accession of N. benthamiana is a natural rdr1 mutant and highly susceptible to TMV. However, TMV-induced symptoms are ameliorated in transgenic plants expressing Medicago truncatula RDR1.

Results

In MtRDR1-transgenic N. benthamiana plants the spread of TMV expressing the green fluorescent protein (TMV.GFP) into upper, non-inoculated, leaves was not inhibited. However, in these plants exclusion of TMV.GFP from the apical meristem and adjacent stem tissue was greater than in control plants and this exclusion effect was enhanced by SA. TMV normally kills N. benthamiana plants but although MtRDR1-transgenic plants initially displayed virus-induced necrosis they subsequently recovered. Recovery from disease was markedly enhanced by SA treatment in MtRDR1-transgenic plants whereas in control plants SA delayed but did not prevent systemic necrosis and death. Following SA treatment of MtRDR1-transgenic plants, extractable RDR enzyme activity was increased and Western blot analysis of RDR extracts revealed a band cross-reacting with an antibody raised against MtRDR1. Expression of MtRDR1 in the transgenic N. benthamiana plants was driven by a constitutive 35S promoter derived from cauliflower mosaic virus, confirmed to be non-responsive to SA. This suggests that the effects of SA on MtRDR1 are exerted at a post-transcriptional level.

Conclusions

MtRDR1 inhibits severe symptom development by limiting spread of virus into the growing tips of infected plants. Thus, RDR1 may act in a similar fashion to RDR6. MtRDR1 and SA acted additively to further promote recovery from disease symptoms in MtRDR1-transgenic plants. Thus it is possible that SA promotes MtRDR1 activity and/or stability through post-transcriptional effects.

Background

Salicylic acid (SA) is a vital signal molecule involved in maintenance and activation of plant defenses. SA is required for the limitation of pathogen spread during the hypersensitive response (HR), which is a genetically determined resistance mechanism whereby pathogens are restricted to the immediate vicinity of an infection site. Triggering the HR can induce an additional induced resistance mechanism called systemic acquired resistance (SAR) that is also SA-dependent. SAR is effective against a very broad spectrum of pathogens, including viruses, oomycetes, fungi and bacteria [1, 2].

SAR, induced either as the result of the HR or by application of resistance-inducing chemicals, is associated with dramatic changes in the transcriptome [3, 4]. These changes include increased transcription of genes encoding pathogenesis-related (PR) proteins, several of which contribute to defense against fungi, oomycetes and bacteria [5] but not against viruses [6–8]. Indeed, induced resistance to viruses remains poorly understood [9]. SA-induced resistance to viruses is not mediated by any of the known PR proteins and is not dependent on the transcriptional activator ‘Non-Expressor of PR proteins 1’ (NPR1), which is required for PR gene induction and effective SA-induced resistance and SAR against non-viral microbial pathogens [10, 11].

One of the mechanisms that could potentially underlie SA-induced resistance to viruses is RNA silencing. The importance of RNA silencing in anti-viral defense may be inferred from the fact that most plant viruses possess counter-defense proteins (viral suppressors of RNA silencing: VSRs) that inhibit the activity or stability of one or more components of the host silencing machinery. The demonstration that the 2b VSR of cucumber mosaic virus (CMV), which inhibits RNA silencing through binding of small RNAs [12–14], also inhibited SA-induced resistance to the replication and local movement of this virus [15], indicated a relationship between SA-induced resistance and silencing. This idea was reinforced by subsequent studies showing that the HC-Pro VSRs of potyviruses can have various effects on SA-mediated resistance to virus spread [16, 17].

An independent line of evidence implicating RNA silencing in SA-induced resistance to viruses was discovered by Xie et al. [18], who found that in tobacco SA increased the accumulation of the transcript encoding RNA-dependent RNA polymerase (RDR) 1 (NtRDR1). Arabidopsis thaliana was also found to possess an SA-inducible RDR1 gene [19]. RDRs are host enzymes that can initiate or amplify RNA silencing, including antiviral silencing, through synthesis of dsRNA molecules that serve as substrates for Dicer-like (DCL) nucleases [20–22]. The products of DCL-mediated cleavage are short-interfering dsRNAs that, after further processing to ssRNA, direct sequence-specific cleavage or translational arrest of homologous target RNA molecules by Argonaute (AGO) proteins [23, 24]. Plants possess a family of RDR paralogs, with six members occurring in Arabidopsis [25]. RDR1 and RDR6 (which is not affected by SA) have demonstrable roles in limiting virus infection or decreasing virus titre in Arabidopsis and Nicotiana species [18, 19, 26–30]. Interestingly, the gene encoding AGO2, one of the AGO proteins with a known antiviral role [31–34], was rendered more sensitive to induction by SA in the presence of a transgene expressing the 2b VSR [4]. This further supports a connection between antiviral silencing and SA-induced resistance.

Although ample evidence suggests a role for silencing in SA-induced resistance, RNA silencing is not absolutely required for successful SA-induced resistance to viruses. It was shown in Arabidopsis that DCLs 2, 3, and 4 are dispensable for SA-induced resistance to TMV and CMV [35]. Thus, it appears that SA induces multiple antiviral systems that include not only RNA silencing but also other mechanisms; for example, the SA-triggered inhibition of virus replication and movement induced via mitochondria-based signaling processes [36, 37].

Meanwhile, the role of RDR1 in virus resistance remains incompletely understood. Transgenic tobacco expressing an anti-sense construct for NtRDR1 were not compromised in their ability to exhibit SA-induced resistance to tobacco mosaic virus (TMV) and potato virus X (PVX) (reported in [18]). More recently, it was shown that RDR1 expression is inducible not only by SA but also by a wide range of other phytohormones, including jasmonic acid, and that RDR1 regulates insect resistance in Nicotiana attenuata [38–41]. The Nicotiana benthamiana accession commonly used for research is a natural rdr1 mutant that expresses a non-functional form of the enzyme (NbRDR1m), which explains why this plant is hypersusceptible to TMV [42, 43]. Consistent with an important role for RDR1 in virus resistance, TMV-induced disease was ameliorated in transgenic N. benthamiana plants constitutively expressing MtRDR1 from Medicago truncatula [37, 42]. However, Ying and colleagues [44] presented data showing that transgenic expression of tobacco NtRDR1 in N. benthamiana enhanced plant susceptibility to the potyvirus plum pox virus. This was surprising in the light of previous work showing that down-regulation of NtRDR1 expression in tobacco plants increased susceptibility to another potyvirus, potato virus Y [29]. To better understand the role of RDR1 in SA-mediated antiviral defense, we explored the effects of SA on virus accumulation and movement and the effects of SA on the gene expression and activity of RDR1 in N. benthamiana.

Results

Constitutive MtRDR1 expression and SA treatment inhibit spread of TMV in the vicinity of the meristem in N. benthamiana

Using TMV engineered to express the green fluorescent protein (TMV.GFP) we investigated the effect of SA on TMV movement in MtRDR1-transgenic N. benthamiana plants. Five-to-six week old MtRDR1-transgenic and empty vector (EV)- control plants (plants from a line transformed with an ‘empty’ transformation vector) were sprayed on the leaves with either 1 mM SA or a control solution once daily for four consecutive days prior to inoculation with TMV.GFP. Plants were monitored daily for the appearance of GFP fluorescence in the upper, non-inoculated leaves. In six independent experiments it was noted that SA treatment consistently resulted in a delay of 1 to 2 days in the first appearance of TMV.GFP in the upper, non-inoculated leaves. However, there was no apparent difference between control and MtRDR1-transgenic plants in the timing or patterning of TMV.GFP spread into the upper non-inoculated leaves (Fig. 1a), though MtRDR1-transgenic plants were less severely stunted than the control plants (Additional file 1). Western blot analysis of TMV coat protein accumulation in the systemically infected leaves at 14 days post-inoculation (dpi) showed that viral coat protein was present in the upper leaves of both groups of plants (Fig. 1b). When the plants were studied under UV illumination to assess TMV.GFP fluorescence in non-inoculated tissue, there was no marked difference in the timing of first appearance or the apparent extent of GFP fluorescence between control and MtRDR1-transgenic N. benthamiana plants in the upper leaves (Fig. 1c).

Fig. 1
figure 1

TMV.GFP accumulation in upper, non-inoculated leaves in N. benthamiana plants is not prevented by expression of an MtRDR1-transgene or by treatment with salicylic acid. Five-week-old MtRDR1-transgenic plants (MtRDR1), or plants transformed with an ‘empty vector’ transformation vector (EV-control), were pre-treated with a solution of 1 mM SA or a control solution [water amended with 0.05 % (v/v) ethanol] prior to inoculation with GFP-tagged TMV. a The movement of TMV.GFP from directly-inoculated leaves into non-inoculated leaves was monitored daily using a hand-held UV lamp and the first appearance of GFP fluorescence in non-inoculated leaves recorded. There was no apparent difference in the timing of appearance of TMV.GFP in non-inoculated leaves between the two groups of plants although systemic infection with TMV.GFP was delayed by SA treatment in both types of plant. There were 23 plants in each treatment group. b Western blot analysis of TMV.GFP accumulation in systemically-infected leaf tissue of water- (top) and SA- (bottom) treated plants using anti-TMV coat protein (CP). In each case leaf tissue samples were harvested at 14 days post-inoculation from the uppermost three leaves above the inoculated leaf. Three samples were taken from each treatment group (one sample = one plant). Equal loading of gel lanes with protein is shown by accumulation of ribulose-1, 5-bisphosphate carboxylase/oxygenase large subunit (LSU) revealed by Ponceau S staining of the western blot membrane. c The extent of GFP fluorescence in the upper leaves of water-treated TMV.GFP-infected empty vector control (control) and MtRDR1-transgenic plants (MtRDR1-transgenic) appeared similar when visualised using a hand-held UV lamp and photographed at 14 days post-inoculation. GFP fluorescence was less extensive in SA-treated plants of both groups. The data and photographs above are from one experiment, out of a total of four independent experiments. Scale bar = 8 cm

However, closer examination of the uppermost regions of the stems of SA-treated and untreated control and MtRDR1-expressing plants revealed some striking differences in the extent of TMV.GFP spread (Fig. 2). Microscopic observation of fluorescence indicated that SA treatment inhibited the extent of spread of TMV.GFP into the upper stem regions of non-transgenic and EV-control plants (Fig. 2a, Additional file 2). TMV.GFP was excluded entirely from the vicinity of the meristem and much of the upper stems of transgenic plants expressing MtRDR1 (Fig. 2a). Further examination showed that SA treatment enlarged the zone from which TMV.GFP was excluded in MtRDR1-transgenic N. benthamiana plants (Fig. 2b).

Fig. 2
figure 2

SA treatment enhances RDR1-mediated exclusion of TMV from tissue adjacent to the apical meristem. Empty vector control (plants transformed with an ‘empty’ transformation vector), non-transformed and MtRDR1-transgenic plants were treated with 1 mM SA in 0.05 % (v/v) ethanol (+) or a control solution of 0.05 % (v/v) ethanol (-) prior to inoculation with TMV.GFP. By 14 days post-inoculation the upper leaves of all plants with and without SA treatment showed TMV.GFP fluorescence in the upper leaves. However, when leaves were removed from the stem it was apparent under epifluorescence microscopy (a) that spread of TMV.GFP into tissue adjacent to meristems of non-transgenic and control transgenic plants was inhibited by SA treatment and that in MtRDR1-transgenic plants, the virus did not enter stem tissue adjacent to the meristem (scale bar, 2 mm). (b) Stems (with leaves removed) of TMV.GFP infected plants were observed under a UV lamp and photographed with a digital camera, revealing that in MtRDR1-transgenic plants the exclusion of TMV.GFP from stem tissue proximal to the meristem was enhanced in SA-treated plants. Photographs in (a) and (b) are from two independent experiments (scale bar = 1 cm)

Recovery of MtRDR1-transgenic N. benthamiana plants from severe TMV disease was enhanced in SA-treated plants

Neither TMV.GFP, nor the TMV.30B vector from which it was derived, induce strong disease symptoms in N. benthamiana [45, 46]. However, infection of this host by the U1 strain of TMV causes strong symptoms culminating in systemic necrosis and death of infected plants. Interestingly, TMV-infected MtRDR1-expressing N. benthamiana plants developed milder symptoms and the plants did not die [37, 42]. Here we found that SA treatment slowed, but did not prevent, TMV-induced death of EV-control transgenic plants and that all plants (SA-treated and untreated) had died by 35 dpi (Fig. 3a). By this time point no MtRDR1-transgenic plants had died although they did exhibit TMV U1-induced leaf necrosis, preceded by other symptoms (leaf curling and chlorosis) in many leaves, with the onset of necrosis being slower in SA-treated plants. Necrosis did not become sufficiently extensive to entirely kill MtRDR1-transgenic plants. The progress of disease development on MtRDR1-transgenic plants was monitored until 75 dpi and it was noted that on all of these plants the spread of necrosis abated and that newly emerging leaves were green and that growth of the plants resumed (Fig. 3b). This apparent recovery from disease was further enhanced on MtRDR1-transgenic plants that had been treated with SA prior to inoculation with TMV. SA-treated MtRDR1-transgenic plants had grown markedly taller than untreated plants by 75 dpi (Fig. 3b). Western blot analysis using anti-coat protein serum showed that TMV was present in the new leaf tissue produced by the plants (Fig. 3c) indicating that although the MtRDR1-expressing transgenic plants recovered from TMV-induced disease and that this recovery was enhanced in SA-treated plants, they did not develop true resistance to the virus.

Fig. 3
figure 3

The effects of MtRDR1 expression and SA treatment on the recovery of N. benthamiana plants from TMV-induced plant death. a Five-week-old transgenic control (plants transformed with an ‘empty’ transformation vector: EV-control) and MtRDR1-transgenic (MtRDR1) plants were pre-treated with a solution of 1 mM SA in 0.05 % (v/v) ethanol (+SA) or a 0.05 % (v/v) ethanol control solution prior to inoculation with TMV U1. Plants were monitored over a 5-week period and the number of plants killed by the virus recorded. With the transgenic control plant group, TMV U1 had killed all plants within 31 days of inoculation. Pre-treatment with 1 mM SA delayed plant death by a few days but did not prevent it. TMV U1-induced plant death was suppressed in all the MtRDR1-expressing N. benthamiana plants, regardless of whether they were treated with water or with SA. b Symptoms at 75 days post-inoculation with TMV U1 in transgenic control and MtRDR1-transgenic N. benthamiana plants. MtRDR1-transgenic plants pre-treated with SA (+) or treated with control solution (-) recovered but SA-treated plant growth was more vigorous (scale bar = 8 cm). Data and photographs are from one experiment, out of a total of four independent experiments. c Western blot analysis of TMV using anti-TMV coat protein (TMV CP) serum shows that TMV was present in the newly emerged, green leaf tissue in recovering MtRDR1-transgenic plants treated with SA (+) or control solution (-). Leaf samples taken at 75 days post-inoculation or mock-inoculation (each lane is loaded with protein extracted from one plant). Equal loading was confirmed by Ponceau S staining of the western blot membrane with the position of the ribulose 1,5-bisphosphate carboxylase/oxygenase large subunit (LSU) indicated

SA treatment increased extractable RDR activity in MtRDR1-transgenic N. benthamiana plants but did not alter MtRDR1 transcript accumulation

SA had a long-lasting effect on MtRDR1-transgenic N. benthamiana plants that resulted in an improvement in recovery from TMV-induced disease and resumption of growth (Fig. 3b). This suggested that SA enhances or primes RDR1-mediated defense against TMV. However, MtRDR1 transgene expression in these plants is under the control of the cauliflower mosaic virus 35S promoter [42] that is, despite containing an as-1 element shared with many SA-responsive plant promoters (for example, that of PR1a [47]), not activated by this phytohormone [48].

To confirm that the 35S promoter used in the construction of the MtRDR1-transgenic plants was behaving as expected, i.e. that it was not responding to SA, we extracted total RNA from leaves of control-transgenic and MtRDR1-transgenic N. benthamiana plants 72 h after they were infiltrated with either 1 mM SA or a control solution. Semi-quantitative reverse transcription-polymerase chain reaction (RT-PCR) assays showed that SA, but not the control solution, induced expression of the PR1a gene, indicating that this SA concentration was effective for the induction of SA-responsive gene expression (Fig. 4a, b). RT-PCR also indicated that whereas the transcript encoding the native, non-functional RDR1 of N. benthamiana (NbRDR1m) was induced by SA, the transgene-encoded MtRDR1 transcript remained at a similar level in MtRDR1-transgenic plants in the presence or absence of SA treatment (Fig. 4c). Analysis by RT-quantitative PCR (RT-qPCR) confirmed that steady state levels of MtRDR1 transcript were not affected by SA treatment in the MtRDR1-transgenic N. benthamiana plants (Fig. 4d, Additional file 3). Thus, the 35S promoter used in the construction of the MtRDR1-transgenic plants is unaffected by SA, meaning that the observed enhancement of RDR1-mediated antiviral action by SA (Figs. 2 and 3b) must be regulated at a post-transcriptional level.

Fig. 4
figure 4

RDR activity, but not MtRDR1 transcript accumulation, is induced by SA treatment in MtRDR1-transgenic N. benthamiana plants. Semi-quantitative RT-PCR analysis of PR1 transcript accumulation in leaves of (a) transgenic (empty vector) control and (b) MtRDR1-transgenic plants infiltrated with a control solution of 0.05 % (v/v) ethanol or a solution of 1 mM SA in 0.05 % (v/v) ethanol. Infiltrated leaf tissue samples were harvested for RNA extraction at 72 h post-infiltration. PR1 transcript accumulation levels were determined by RT-PCR after 40 cycles of PCR and compared relative to the accumulation levels of the elongation factor 1 alpha (EF1α) transcript. Increased PR1 accumulation confirmed that SA was taken up by the tissues and was effective in inducing transcriptional changes. c Semi-quantitative RT-PCR analysis showed that there was little difference in MtRDR1 transcript accumulation in MtRDR1-transgenic plants infiltrated with control solution or 1 mM SA. NbRDR1m transcript accumulation was up-regulated in both transgenic control and MtRDR1-transgenic N. benthamiana plants after SA treatment, although the NbRDR1m protein itself is non-functional. MtRDR1 and NbRDR1m transcript accumulation levels after 27 and 35 cycles, respectively, were compared relative to the accumulation levels of EF1α. Infiltrated tissue samples were harvested at 72 h post-infiltration. d RT-qPCR analysis of MtRDR1 transcript levels in leaves of empty vector control and MtRDR1-transgenic plants infiltrated with water control or 1 mM SA solution. MtRDR1 was not detected in empty vector control plants. Mean values for relative MtRDR1 levels (based on duplicate technical replicate values; 100 = mean value for the transcript level in untreated MtRDR1 plants) obtained from three plants (one plant = one independent sample) have been given for each treatment group. Error bars represent standard errors of the mean for the three samples. Relative transcript levels of MtRDR1 were calculated using the 2-ΔΔC(t) method [59] using EF1α as an internal reference. e Enhancement of RDR activity by SA in MtRDR1-transgenic plants. Leaves from tobacco (N. tabacum, included as a positive control) and transgenic empty vector control and MtRDR1-transgenic N. benthamiana plants were infiltrated with water (-) containing 0.05 % (v/v) ethanol or 2.5 mM SA in 0.05 % (v/v) ethanol (+) and harvested after 48 h for preparation of RDR1-enriched extracts. The proxy for RDR1 activity was the incorporation of α-[32P] CTP into nascent RNA analysed by liquid scintillation counting of radioactivity incorporated (counts per minute) into trichloroacetic acid-precipitable material. Error bars are standard errors for the mean for three technical replicates (RDR assays) per sample

In order to determine if SA treatment affected RDR activity levels, extracts enriched in RDR activity [18] were prepared from non-infected control transgenic and MtRDR1-transgenic N. benthamiana leaves infiltrated with SA or control solution 48 h prior to tissue harvesting (Fig. 4e). N. tabacum (tobacco) possesses the SA-inducible gene NtRDR1, which encodes a functional RDR1 [18], and so samples from non-transgenic N. tabacum were included in the assays as positive controls for SA-inducible RDR activity. Control-treated EV N. benthamiana plants contained active RDR not attributable to RDR1 and as expected, MtRDR1-transgenic N. benthamiana plants contained a higher level of extractable RDR activity than that detected in the EV control plants (Fig. 4e). Although treatment with SA did not increase the amount of active RDR extractable from EV-control plants (consistent with the known absence of SA-inducible RDR activity in N. benthamiana [42]), SA treatment did increase RDR activity in MtRDR1-transgenic plants. This effect was only observed when SA was applied to intact plant tissue, and not when RDR extracts from untreated N. tabacum or MtRDR1-expressing N. benthamiana plants were incubated in vitro with SA or with a biologically inactive isomer of SA (Additional file 4).

MtRDR1 protein is detectable in RDR extracts from SA-treated MtRDR1-transgenic N. benthamiana plants

Since the in vitro RDR activity of extracts from MtRDR1-transgenic plants was increased approximately 2.5 fold when the plants were pre-treated with SA, we investigated whether there was any change in the accumulation of MtRDR1 protein in these extracts. A polyclonal anti-MtRDR1 rabbit serum was prepared using a maltose binding protein (MBP) MtRDR1 fusion protein made in Escherichia coli (Additional file 5). This was used for Western blot analysis of RDR-enriched extracts (Fig. 5). As the anti-MtRDR1 serum detected a number of background plant protein bands, western blot analysis with preimmune serum was also carried out (Fig. 5). The only polypeptide specifically detected by the anti-MtRDR1 serum was of c.131 kDa, the mass predicted for MtRDR1 (Additional file 6), and this was present only in RDR-enriched extracts from SA-treated, MtRDR1-transgenic plants (Fig. 5). Ubiquitination and multiple phosphorylation sites are predicted to occur in the MtRDR1 protein sequence (Additional file 7). This raises the possibility that SA treatment stabilizes the protein (through phosphorylation or inhibition of ubiquitin-mediated breakdown) or, alternatively, that SA triggers recruitment of MtRDR1 into an active complex.

Fig. 5
figure 5

Western blot analysis for MtRDR1 in RDR1-enriched protein extracts. Leaves of empty vector control transgenic or MtRDR1-transgenic N. benthamiana plants were infiltrated with a control solution of water containing 0.05 % (v/v) ethanol (-) or 2.5 mM SA (+). RDR1-enriched protein extracts were prepared 48 h later and subjected to western blot analysis with either preimmune rabbit serum (left panel) or polyclonal rabbit anti-MtRDR1 serum (right panel). The positions of protein molecular mass markers are indicated (kDa). A protein (indicated with a star) corresponding to the predicted size of 131 kDa for MtRDR1 was specifically detected in SA-treated MtRDR1-transgenic plants with the anti-MtRDR1 serum but not with the preimmune serum. Equal loading of gel lanes with protein, based on the Bradford [59] assay was confirmed by Ponceau S staining of the western blot membrane (lower panels)

Discussion

We studied the effects of RDR1 on the long-distance movement and spread of TMV. Expression of MtRDR1 in N. benthamiana did not prevent TMV.GFP movement into non-inoculated leaf tissue or enhance the inhibitory effect of SA on virus movement into these tissues. However, MtRDR1 expression inhibited the extent of spread into the region proximal to the apical meristem and this movement restriction became more pronounced following treatment of MtRDR1-transgenic plants with SA. Schwach and colleagues [28] showed that knock-down of NbRDR6 gene expression in the same host (RDR6i-transgenic N. benthamiana) allowed entry of PVX into the meristem and that this resulted in an exacerbation of PVX-induced disease symptoms throughout the plant. In line with previous work [49] we found that in N. benthamiana plants, TMV.GFP enters tissue adjacent to the meristem, although it does not appear to enter the meristem itself. In MtRDR1-transgenic plants, however, the ability of TMV.GFP to approach the meristem is drastically curtailed. Furthermore, in the MtRDR1-transgenic plants the normally lethal effect of TMVU1 is ameliorated and recovery from disease occurs. Our results with constitutive expression of MtRDR1 complement and extend the results and conclusions of Schwach and colleagues [28]; indicating that RDR1, as well as RDR6, can function to inhibit virus entry to tissue adjacent to the meristem. Both studies point to a relationship between the effectiveness of RNA silencing-mediated exclusion of virus from meristematic and adjacent tissues with the severity of virus-induced disease symptoms. This suggests that, with respect to the elaboration of disease symptoms, that there is a critical developmental stage for plant cells and tissues, which corresponds physically to a certain point along the sub-apical zone. If viral entry into these tissues is delayed, for example by host defense mechanisms, until after this critical developmental stage has been completed, plants will to some extent be able to recover from virus infection.

It has been postulated that the hyper-susceptibility of the laboratory accession of N. benthamiana to a number of tobamoviruses including TMV can be attributed to the absence of an active RDR1 in this species [42]. Ying and colleagues [43] proposed an alternative hypothesis, suggesting that RDR1 can enhance virus spread and accumulation and that the loss of RDR1 functionality in N. benthamiana may be due to selective pressure to maintain high levels of RDR6-dependent antiviral defense. This hypothesis was formulated mainly to explain results after challenge with non-tobamoviruses, as challenge with TMV did not alter symptoms and at best only modestly decreased virus accumulation in systemically-infected tissue of N. benthamiana plants expressing NtRDR1 [43]. We have seen no increase in susceptibility to TMV in our study of N. benthamiana plants expressing the transgene derived from the RDR1 of M. truncatula. Furthermore, wild accessions of N. benthamiana that contain a non-truncated, functional NbRDR1 are naturally protected against severe symptoms from tobamovirus infection and are not more susceptible to poty- or cucumovirus infection than N. benthamiana carrying NbRDR1m (43) indicating that RDR1 does indeed have a particular role in ameliorating tobamovirus infection in N. benthamiana.

A difference in the effects of RDR1 and RDR6 on virus spread is apparent in the pattern of spread of PVX.GFP into upper uninoculated leaves previously observed in RDR6i plants [28] and in the pattern of systemic TMV.GFP movement in MtRDR1-transgenic plants in the present study (Fig. 1). Specifically, knock-down of NbRDR6 accelerated entry of virus into well-developed upper leaves, whereas constitutive expression of MtRDR1 did not slow the appearance of TMV.GFP in these leaves. These differences suggest that the roles of RDRs 1 and 6 do not overlap completely and that each may be effective in protection against different viruses with differing infection and movement strategies, as suggested previously [42].

These results suggest that RDR1 may play a role similar to RDR6 in protecting the meristem from viral invasion. The strong AtRDR1 promoter activity in Arabidopsis phloem tissue observed by Xu and colleagues [41] may also help restrict the spread of virus into the vasculature and subsequently the meristem, although curiously the AtRDR1 promoter:β-glucuronidase reporter transgene was found to be poorly active in younger tissues. However, since RDR1 gene expression is induced by chemical signals associated with SAR (SA, nitric oxide and hydrogen peroxide) [39, 40] as well as by many other phytohormones (jasmonic acid, abscisic acid, auxin, ethylene) and wounding [18, 39, 41], it may be that the multiplicity of factors affecting RDR1 expression, and its therefore complex regulation, allows plants to reinforce meristem protection against invasion in times of stress or perhaps control the accessibility of this tissue to endogenous factors during development.

Previous studies showed that plants expressing a functional RDR1 are less susceptible to TMV-induced disease, yet RDR1 does not seem to be required for successful SA-induced resistance to viruses in directly inoculated leaves [18, 19, 37]. Interestingly, we found here not only that expression of MtRDR1 enabled N. benthamiana to recover from severe disease induced by TMV strain U1, but also that this recovery was enhanced by SA treatment. Whilst this could have been due to the activity of one or more of the RDR1-independent SA-inducible antiviral defense systems that are known to exist [37], SA treatment of control-transgenic N. benthamiana plants only delayed TMV-induced plant death by a few days suggesting that the protective activity of other SA-induced antiviral defense systems was insufficient in this case to account for the marked enhancement of recovery after SA treatment in the MtRDR1-transgenic N. benthamiana plants.

Although SA and MtRDR1 worked synergistically in protection against severe TMV-induced disease, it is not possible to say with certainty whether or not these factors were working together or separately. However, the hypothesis that SA enhanced RDR1 activity was favored by the observation that MtRDR1-transgenic plants contained more extractable RDR activity following SA treatment, despite there being no concomitant SA-induced increase in MtRDR1 transgene expression. There was no direct effect of SA on in vitro RDR activity in plant extracts (Additional file 4), which indicates that SA indirectly influences RDR activity in planta. Detection of a band cross-reacting with an antibody raised against MtRDR1 in RDR1-enriched preparations from SA-treated but not untreated MtRDR1-transgenic plants was consistent with the detection of greater RDR1 activity. It also suggested that RDR1 can be regulated by SA through post-translational process(es) either resulting in a more stable MtRDR1 or involving recruitment of MtRDR1 protein to a complex. An in silico analysis of the MtRDR1 protein sequence revealed a number of candidate ubiquitination and phosphorylation sites (Additional file 7). Post-translational protein modification plays an important role in host immunity [50, 51], for example: ubiquitin-dependent proteolysis negatively regulates defense responses in tobacco [52]; NPR3 and 4 are E3 ligase adapters that control NPR1 activity by proteolysis in an SA-dependent manner in Arabidopsis [53, 54]; and SUMO proteases negatively regulate SA biosynthesis [55]. It may therefore be that SA induces, in some manner, stabilization of the MtRDR1 protein against small modifier protein-mediated degradation and/or activation/deactivation of the protein by phosphorylation. The idea of assembly of a complex, which includes RDR1 and forms in response to SA, is made more plausible by recent findings that several RNA silencing factors, including RDR6, form complexes with each other and with other cellular components such as membranes [56, 57]. Another possibility is that SA enhances translation of the MtRDR1 transcript leading to increased synthesis of MtRDR1 protein but it is difficult to envisage a likely mechanism for this.

Conclusions

In conclusion, our results support an important role for the phytohormone-inducible factor RDR1 in resistance by limiting viral access to the region adjacent to the meristem and thereby ameliorating the severity of virus-induced disease. Its ability to inhibit access to the meristematic region correlates with amelioration of TMV-induced disease symptoms and is enhanced by SA. Our data suggests that SA may enhance RDR1 activity at a post-transcriptional level in addition to the previously documented effects of this phytohormone on transcription of the RDR1 gene.

Methods

Plant growth conditions

Seeds of the laboratory accession of Nicotiana benthamiana Domin. [42] non-transgenic, control transgenic (Line V19-1) and MtRDR1-transgenic (Line R15-1) plant lines [43] and of tobacco (N. tabacum L. cv. Xanthi) were germinated in soil and plants cultivated in a controlled environment room (Conviron Ltd., Winnipeg, Manitoba, Canada) with a 16 h photoperiod (200 μE.m2.s−1 of photosynthetically active radiation) at 22 °C and 60 % relative humidity.

Virus strains, chemical treatments and detection of infection

The viral strains used in this study were the common (U1) strain of TMV [58], and TMV.GFP (TMV30B.GFP constructed by Shivprasad et al. [45]). Capped TMV.GFP infectious RNA was re-generated by in vitro transcription as previously described and inoculated onto 3- to 5-week old N. benthamiana plants to prepare aliquots of infectious sap as described by Murphy et al. [46]. For whole-plant treatments with SA, 5-to-6 week old N. benthamiana plants were sprayed for four consecutive days with either a control solution [0.05 % (w/v) ethanol] or 1 mM SA dissolved in 0.05 % (w/v) ethanol before mechanical inoculation with TMV U1 or TMV.GFP on one or two lower leaves [37]. Treatment of leaf tissue by infiltration was carried out as described previously using 2.5 mM SA dissolved in 0.05 % (w/v) ethanol [37]. Western blot detection of TMV using anti-TMV coat protein serum was carried out using a previously described method [46]. Observations of GFP fluorescence utilized either a hand-held UV lamp and Nikon Cool-Pix digital camera or epifluorescence microscopy using a Nikon Optiphot 2 (Nikon Ltd., Kingston-upon-Thames, UK) [59].

RNA extraction, cDNA synthesis and PCR

Total RNA was prepared [60] and treated with DNase I using the TURBO DNase-free kit (Ambion, Austin, TX). First-strand cDNA synthesis was carried out with 1 μg treated RNA using Superscript III (Invitrogen, Paisley, UK) with random pentadecamer primers according to the manufacturer’s instructions. Primers for semi-quantitative or quantitative PCR are listed as a table in Additional file 8. For semi-quantitative RT-PCR analysis of plant gene expression cDNA was amplified with RedTaq DNA polymerase (Sigma-Aldrich, Poole, UK) using the following conditions: 1 cycle of 94 °C for 2 min; 27 to 40 cycles of 94 °C for 15 s, 55 °C for 30 s, 72 °C for 40 s; and 1 cycle of 72 °C for 5 min. PCR products were analysed by electrophoresis on 1.5 % agarose gels. For RT-qPCR analysis of MtRDR1 transcript levels, the cDNA template was diluted 1:5 and qPCR performed using SYBR Green JumpStart Taq ReadyMix (Sigma-Aldrich, St Louis, MO) according to the manufacturer’s instructions. Reactions were conducted in duplicate (40 cycles: 94 °C for 15 s, 57 °C for 30 s, 72 °C for 40 s) on a Chromo4 PCR system (Bio-Rad, Hemel-Hempsted, UK) and analysed using the LinRegPCR program [61]. Relative transcript levels of MtRDR1 were calculated using the 2-ΔΔC(t) method [62] using NbEF1α as an internal reference (previously authenticated for stability in the presence of SA [63]).

Isolation and assay of RDR activity

The isolation of RDR1-enriched extracts and assays for RDR activity were carried out as described by Xie and colleagues [18]. Leaf tissue (2 g fresh weight) from 21 days-old N. tabacum or N. benthamiana plants was harvested 48 h post-infiltration with 1 mM SA in 0.05 % (w/v) ethanol or control solution [0.05 % (w/v) ethanol]. Tissue was homogenised in 4 ml of buffer A (50 mM Tris-acetate pH 7.4, 10 mM potassium acetate, 1 mM EDTA, 10 mM 2-mercaptoethanol and 0.5 mM phenylmethysulfonylfluoride), centrifuged at 1000 × g for 12 min and centrifuged again at 14,000 × g for 10 min after adding glycerol to a final proportion of 20 % (v/v). An equal volume of 4 M (NH4)2SO4 was added to each supernatant and the proteins precipitated over 2 h with gentle agitation (50 rpm) on an orbital shaker. Precipitated protein was collected by centrifugation at 10,000 × g for 20 min. Pellets were washed briefly with 1 ml buffer B [25 mM Tris-acetate pH 8.2, 1 mM EDTA, 20 % (v/v) glycerol and 3 mM 2-mercaptoethanol] twice. The RDR1-enriched protein extracts were re-suspended in 300–500 μl buffer B and dialysed overnight against buffer B. All preparation steps were carried out at 4 °C. Protein concentrations were assayed using the method of Bradford [64] with a Bio-Rad protein quantification kit.

In vitro RDR assays were performed in a 50 μl final volume containing 25 μg of RDR1-enriched protein preparation with 50 mM Tris–HCl (pH 8.0), 10 mM MgCl2, 1 mM DTT, 1 mM each of ATP, GTP, UTP, and 2 μM CTP supplemented with α-[32P]CTP (Perkin-Elmer, Little Chalfont, UK) and 4 μg total RNA from TMV-infected N. tabacum leaves [18]. Reactions were initiated by adding the RDR1-enriched protein preparation and incubated for 2 h at 30 °C. Reactions were terminated by incubation at 95 °C for 5 min and placed on ice. Incorporation of α-[32P]CTP into RDR reaction products was quantified by scintillation counting of radioactivity incorporated into trichloroacetic acid-precipitable material [65] using OptiPhase Hisafe 3 scintillation cocktail (Perkin-Elmer) and a 2000 Tri-CARB liquid scintillation analyser (Packard, Illinois, USA). RDR assays were carried out using at least three biological replicates.

Expression of recombinant MtRDR1 in Escherichia coli and production of rabbit polyclonal anti-MtRDR1

The plasmid pMAL-MtRDR1 was obtained by cloning the MtRDR1 in pMAL-c2E (New England Biolabs, MA, USA) in frame to the carboxyl terminus of the MBP coding sequence using the KpnI and HindIII restriction enzyme sites. Cells of E. coli TB1 strain (NEB) were co-transformed with pMAL-MtRDR1 and the CodonPlus plasmid (Stratagene). CodonPlus plasmids contain genes for tRNAs that recognize the arginine codons AGA and AGG, the isoleucine codon AUA, and the leucine codon CUA, respectively, which improves the availability of tRNAs that most frequently restrict translation of AT-rich eukaryotic genes in E. coli. The recombinant protein expressed by the pMAL-RDR1 clone was induced and purified by amylose resin column chromatography  as described in the manufacturer’s instructions (NEB) (Additional file 5). The amylose column fractions were further purified on a HiLoad Superdex200 column attached to an AKTA purification system (GE Healthcare Life Sciences, USA). Purified fractions were run on a SDS-PAGE gel and stained with Coomassie Brilliant Blue R-250 (Additional file 5). Putative MBP-MtRDR1 fusion protein was extracted from the SDS page gel and verified by MALDI-TOF MS [66] (data not shown). Gel slices containing 1 mg of MtRDR1 protein were sent to Covance Inc. (Denver, PA, USA) for antibody production, where two rabbits (OK127 and OK129) were injected with 125 μg each of MtRDR1 protein at 14, 35, 49 and 70 days following a sampling for pre-immune serum. Test bleeds were collected on the 45th day and production bleeds were collected on the 59th and 77th days. The test bleed and production bleeds were purified by using NAb Protein A/G Spin Columns (Pierce, USA). Antibody obtained from rabbit OK129 exhibited better activity in the preliminary experiments (data not shown). Thus, anti-MtRDR1 and the corresponding pre-immune serum from rabbit OK129 were used in the current study.

Western blotting for MtRDR1

RDR1-enriched protein extracts were prepared as above and equal amounts of protein analyzed by SDS-PAGE on 15 % (w/v) acrylamide gels [67] before electrophoretic transfer to nitrocellulose [68]. Equal loading of protein samples was confirmed using Ponceau S staining. Western blot analysis of MtRDR1 accumulation was carried out using a rabbit polyclonal anti-MtRDR1 serum (1:10,000 dilution) or control preimmune rabbit serum (1:10,000 dilution) as primary antibodies followed by anti-rabbit IgG conjugated to horseradish peroxidase (1:15,000 dilution) secondary antibody. Bound antibody was detected using the Western Lightning Chemiluminescence Reagent Plus (PerkinElmer) and exposure to Konica Minolta AX film (Konica Minolta Medical and Graphic, Japan).

In silico predictions of phosphorylation and ubiquitination sites in the MtRDR1 sequence

In silico translation of the MtRDR1 cDNA sequence was carried out using the ExPASy Translate tool (http://web.expasy.org/translate/) from the Swiss Institute of Bioinformatics ExPASy Bioinformatics Resource Portal. Potential phosphorylation sites in the MtRDR1 protein sequence were predicted using NetPhos 2.0 software (www.cbs.dtu.dk/services/NetPhos/) [69]. NetPhos 2.0 generates a score in the range 0.000 to 1.000 to indicate the likelihood of a serine, threonine or tyrosine being a phosphorylation site, with residues with a score >0.500 being assigned as potential phosphorylation sites. UbPred software (http://omictools.com/ubiquitination-sites-category) was used to predict potential ubiquitination sites [70]. Residues with ubiquitination scores >0.62 are assigned as potential ubiquitination sites, with scores in the range 0.62–0.69 being low confidence predictions and those in the range 0.69–0.84 being high confidence predictions.

Availability of supporting data

The data sets supporting the results of this article are included within the article and its additional files.

Abbreviations

AGO:

Argonaute

CMV:

cucumber mosaic virus

DCL:

Dicer-like nucleases

EV:

empty vector

GFP:

green fluorescent protein

HR:

hypersensitive response

MBP:

maltose binding protein

NPR1:

Non-Expressor of PR proteins 1

PR:

pathogenesis-related proteins

PVX:

potato virus X

RDR:

RNA-dependent RNA polymerases

SA:

Salicylic acid

SAR:

systemic acquired resistance

TMV:

tobacco mosaic virus

VSR:

viral suppressors of RNA silencing

References

  1. Lewsey MG, Palukaitis P, Carr JP. Plant-virus interactions: defence and counter-defence. In: Parker J, editor. Annual plant reviews volume 34: molecular aspects of plant disease resistance. Oxford: Blackwell; 2009. p. 134–76.

    Google Scholar 

  2. Carr JP, Lewsey MG, Palukaitis P. Signaling in induced resistance. Adv Virus Res. 2010;76:57–121.

    Article  PubMed  CAS  Google Scholar 

  3. Maleck K, Levine A, Eulgem T, Morgan A, Schmid J, Lawton KA, et al. The transcriptome of Arabidopsis thaliana during systemic acquired resistance. Nat Genet. 2000;26:403–10.

    Article  PubMed  CAS  Google Scholar 

  4. Lewsey MG, Murphy AM, MacLean D, Dalchau N, Westwood JH, Macaulay K, et al. Disruption of two defensive signaling pathways by a viral RNA silencing suppressor. Mol Plant-Microbe Interact. 2010;23:835–45.

    Article  PubMed  CAS  Google Scholar 

  5. Durrant WE, Dong X. Systemic acquired resistance. Annu Rev Phytopathol. 2004;42:185–209.

    Article  PubMed  CAS  Google Scholar 

  6. Carr JP, Beachy RN, Klessig DF. Are the PR1 proteins of tobacco involved in genetically engineered resistance to TMV? Virology. 1989;169:470–3.

    Article  PubMed  CAS  Google Scholar 

  7. Cutt JR, Harpster MH, Dixon DC, Carr JP, Dunsmuir P, Klessig DF. Disease response to tobacco mosaic virus in transgenic tobacco plants that constitutively express the pathogenesis-related protein PR1b. Virology. 1989;173:89–97.

    Article  PubMed  CAS  Google Scholar 

  8. Linthorst HJM, Meuwissen RLJ, Kauffmann S, Bol JF. Constitutive expression of pathogenesis-related proteins PR-1, GRP, and PR-S in tobacco has no effect on virus-infection. Plant Cell. 1989;1:285–91.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  9. Palukaitis P, Groen SC, Carr JP. The Rumsfeld paradox: some of the things we know that we don’t know about plant virus infection. Curr Opin Plant Biol. 2013;16:513–9.

    Article  PubMed  CAS  Google Scholar 

  10. Kachroo P, Yoshioka K, Shah J, Dooner HK, Klessig DF. Resistance to turnip crinkle virus in Arabidopsis is regulated by two host genes and is salicylic acid dependent but NPR1, ethylene, and jasmonate independent. Plant Cell. 2000;12:677–90.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  11. Wong CE, Carson RA, Carr JP. Chemically induced virus resistance in Arabidopsis thaliana is independent of pathogenesis-related protein expression and the NPR1 gene. Mol Plant-Microbe Interact. 2002;15:75–81.

    Article  PubMed  CAS  Google Scholar 

  12. Goto K, Kobori T, Kosaka Y, Natsuaki T, Masuta C. Characterisation of silencing suppressor 2b of Cucumber mosaic virus based on examination of its small RNA-binding abilities. Plant Cell Physiol. 2007;48:1050–60.

    Article  PubMed  CAS  Google Scholar 

  13. Gonzalez I, Martinez L, Rakitina DV, Lewsey MG, Atencio FA, Llave C, et al. Cucumber mosaic virus 2b protein subcellular targets and interactions: their significance to RNA silencing suppressor activity. Mol Plant-Microbe Interact. 2010;23:294–303.

    Article  PubMed  CAS  Google Scholar 

  14. Gonzalez I, Rakitina D, Semashko M, Taliansky M, Praveen S, Palukaitis P, et al. RNA binding is more critical to the suppression of silencing function of Cucumber mosaic virus 2b protein than nuclear localization. RNA. 2012;18:771–82.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  15. Ji LH, Ding SW. The suppressor of transgene RNA silencing encoded by Cucumber mosaic virus interferes with salicylic acid-mediated virus resistance. Mol Plant-Microbe Interact. 2001;14:715–24.

    Article  PubMed  CAS  Google Scholar 

  16. Alamillo JM, Saénz P, García JA. Salicylic acid mediated and RNA-silencing defence mechanisms cooperate in the restriction of systemic spread of plum pox virus in tobacco. Plant J. 2006;48:217–27.

    Article  PubMed  CAS  Google Scholar 

  17. Jovel J, Walker M, Sanfaçon H. Salicylic acid-dependent restriction of Tomato ringspot virus spread in tobacco is accompanied by a hypersensitive response, local RNA silencing, and moderate systemic resistance. Mol Plant-Microbe Interact. 2011;24:706–18.

    Article  PubMed  CAS  Google Scholar 

  18. Xie Z, Fan B, Chen C, Chen Z. An important role of an inducible RNA-dependent RNA polymerase in plant antiviral defense. Proc Natl Acad Sci U S A. 2001;98:6516–21.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  19. Yu D, Fan B, MacFarlane SA, Chen Z. Analysis of the involvement of an inducible Arabidopsis RNA-dependent RNA polymerase in antiviral defence. Mol Plant-Microbe Interact. 2003;16:206–16.

    Article  PubMed  CAS  Google Scholar 

  20. Blevins T, Rajeswaran R, Shivaprasad PV, Beknazariants D, Si-Ammour A, Vazquez F, et al. Four plant dicers mediate viral small RNA biogenesis and DNA virus induced silencing. Nucleic Acids Res. 2006;34:6233–46.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  21. Garcia-Ruiz H, Takeda A, Chapman EJ, Sullivan CM, Fahlgren N, Brempelis KJ, et al. Arabidopsis RNA-dependent RNA polymerases and dicer-like proteins in antiviral defense and small interfering RNA biogenesis during Turnip Mosaic Virus infection. Plant Cell. 2010;22:481–96.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  22. Fusaro AF, Matthew L, Smith NA, Curtin CJ, Dedic-Hagen J, Ellacott GA, et al. RNA interference-inducing hairpin RNAs in plants act through the viral defence pathway. EMBO Rep. 2006;2006(7):1168–75.

    Article  Google Scholar 

  23. Morel JB, Godon C, Mourrain P, Beclin C, Boutet S, Feuerbach F, et al. Fertile hypomorphic ARGONAUTE (ago1) mutants impaired in post-transcriptional gene silencing and virus resistance. Plant Cell. 2002;14:629–39.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  24. Iwakawa HO, Tomari Y. Molecular insights into microRNA-mediated translational repression in plants. Mol Cell. 2013;52:591–601.

    Article  PubMed  CAS  Google Scholar 

  25. Wassenegger M, Krczal G. Nomenclature and functions of RNA-directed RNA polymerases. Trends Plant Sci. 2006;11:142–51.

    Article  PubMed  CAS  Google Scholar 

  26. Mourrain P, Beclin C, Elmayan T, Feuerbach F, Godon C, Morel J-B, et al. Arabidopsis SGS2 and SGS3 genes are required for posttranscriptional gene silencing and natural virus resistance. Cell. 2000;101:533–42.

    Article  PubMed  CAS  Google Scholar 

  27. Dalmay TD, Horsefield R, Braunstein TH, Baulcombe DC. SDE3 encodes an RNA helicase required for post-transcriptional gene silencing in Arabidopsis. EMBO J. 2001;20:2069–78.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  28. Schwach F, Vaistij FE, Jones L, Baulcombe DC. An RNA-dependent RNA polymerase prevents meristem invasion by potato virus X and is required for the activity but not the production of a systemic silencing signal. Plant Physiol. 2005;138:1842–52.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  29. Rakhshandehroo F, Takeshita M, Squires J, Palukaitis P. The influence of RNA-dependent RNA polymerase 1 on Potato virus Y infection and on other antiviral response genes. Mol Plant-Microbe Interact. 2009;10:1312–8.

    Article  Google Scholar 

  30. Qin C, Shi N, Gu M, Zhang H, Li B, Shen J, et al. Involvement of RDR6 in short-range intercellular RNA silencing in Nicotiana benthamiana. Sci Rep. 2012;2:467.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Harvey JJW, Lewsey MG, Patel K, Westwood JH, Heimstädt S, Carr JP, et al. An antiviral defense role of AGO2 in plants. PLoS One. 2011;6, e14639.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  32. Jaubert M, Bhattacharjee S, Mello AFS, Perry KL, Moffett P. ARGONAUTE2 mediates RNA silencing antiviral defenses and Potato virus X in Arabidopsis. Plant Physiol. 2011;156:1556–64.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  33. Scholthof HB, Alvarado VY, Vega-Arreguin JC, Ciomperlik J, Odokonyero D, Brosseau C, et al. Identification of an ARGONAUTE for antiviral RNA silencing in Nicotiana benthamiana. Plant Physiol. 2011;156:1548–55.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  34. Zhang X, Zhao H, Gao S, Wang WC, Katiyar-Agarwal S, Huang HD, et al. Arabidopsis Argonaute 2 regulates innate immunity via miRNA393*-mediated silencing of a Golgi-localized SNARE gene, MEMB12. Mol Cell. 2012;42:356–66.

    Article  Google Scholar 

  35. Lewsey MG, Carr JP. Effects of DICER-like proteins 2, 3 and 4 on cucumber mosaic virus and tobacco mosaic virus infections in salicylic acid treated plants. J Gen Virol. 2009;90:3010–4.

    Article  PubMed  CAS  Google Scholar 

  36. Gilliland A, Singh DP, Hayward JM, Moore CA, Murphy AM, York CJ, et al. Genetic modification of alternative respiration has differential effects on antimycin A-induced versus salicylic acid-induced resistance to Tobacco mosaic virus. Plant Physiol. 2003;132:1518–28.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  37. Lee WS, Fu SF, Verchot-Lubicz J, Carr JP. Genetic modification of alternative respiration in Nicotiana benthamiana affects basal and salicylic acid-induced resistance to potato virus X. BMC Plant Biol. 2011;11:41–50.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  38. Pandey SP, Baldwin IT. RNA-directed RNA polymerase 1 (RdR1) mediates the resistance of Nicotiana attenuata to herbivore attack in nature. Plant J. 2007;50:40–53.

    Article  PubMed  CAS  Google Scholar 

  39. Hunter LJR, Westwood JH, Heath G, Macaulay K, Smith AG, MacFarlane SA, et al. Regulation of RNA-dependent RNA polymerase 1 and isochorismate synthase gene expression in Arabidopsis. PLoS One. 2013;8, e66530.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  40. Liao YWK, Sun ZH, Zhou YH, Shi K, Li X, Zhang GQ, et al. The role of hydrogen peroxide and nitric oxide in the induction of plant-encoded RNA-dependent RNA polymerase 1 in the basal defense against tobacco mosaic virus. PLoS One. 2013;8, e76090.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  41. Xu T, Zhang L, Zhen J, Fan Y, Zhang C, Wang L. Expressional and regulatory characterization of Arabidopsis RNA-dependent RNA polymerase 1. Planta. 2013;237:1561–9.

    Article  PubMed  CAS  Google Scholar 

  42. Yang SJ, Carter SA, Cole AB, Cheng NH, Nelson RS. A natural variant of a host RNA-dependent RNA polymerase is associated with increased susceptibility to viruses by Nicotiana benthamiana. Proc Natl Acad Sci U S A. 2004;101:6297–302.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  43. Wylie SJ, Zhang C, Long V, Roossinck MJ, Koh SH, Jones MGK, et al. Differential responses to virus challenge of laboratory and wild accessions of Australian species of Nicotiana, and comparative analysis of RDR1 gene sequences. PLoS One. 2015;10, e0121787.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Ying X-B, Dong L, Zhu H, Duan C-G, Du Q-S, Lv D-Q, et al. RNA-dependent RNA polymerase 1 from Nicotiana tabacum suppresses RNA silencing and enhances viral infection in Nicotiana benthamiana. Plant Cell. 2010;22:1358–72.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  45. Shivprasad S, Pogue GP, Lewandowski DJ, Hidalgo J, Donson J, Grill LK, et al. Heterologous sequences greatly affect foreign gene expression in tobacco mosaic virus-based vectors. Virology. 1999;255:312–23.

    Article  PubMed  CAS  Google Scholar 

  46. Murphy AM, Gilliland A, York CJ, Hyman B, Carr JP. High-level expression of alternative oxidase protein sequences enhances the spread of viral vectors in resistant and susceptible plants. J Gen Virol. 2004;85:3777–86.

    Article  PubMed  CAS  Google Scholar 

  47. Jupin I, Chua NH. Activation of the CaMV as-1 cis-element by salicylic acid: differential DNA-binding of a factor related to TGA1a. EMBO J. 1996;15:5679–89.

    PubMed  CAS  PubMed Central  Google Scholar 

  48. Butterbrodt T, Thurow C, Gatz C. Chromatin immunoprecipitation analysis of the tobacco PR-1a and the truncated CaMV 35S promoter reveals differences in salicylic acid-dependent TGA factor binding and histone acetylation. Plant Mol Biol. 2006;61:665–74.

    Article  PubMed  CAS  Google Scholar 

  49. Cheng NH, Su CL, Carter SA, Nelson RS. Vascular invasion routes and systemic accumulation patterns of tobacco mosaic virus in Nicotiana benthamiana. Plant J. 2000;23:349–62.

    Article  PubMed  CAS  Google Scholar 

  50. Shirsekar G, Dai L, Hu Y, Wang X, Zeng L, Wang G-L. Role of ubiquitination in plant innate immunity and pathogen virulence. J Plant Biol. 2010;53:1–10.

    Article  Google Scholar 

  51. Lee J, Nam J, Park HC, Na G, Miura K, Jin JB, et al. Salicylic acid-mediated innate immunity in Arabidopsis is regulated by SIZ1 SUMO E3 ligase. Plant J. 2007;49:79–90.

    Article  PubMed  CAS  Google Scholar 

  52. Conrath U, Klessig DF, Bachmair A. Tobacco plants perturbed in the ubiquitin-dependent protein degradation system accumulate callose, salicylic acid, and pathogenesis-related protein 1. Plant Cell Rep. 1998;17:876–80.

    Article  CAS  Google Scholar 

  53. Spoel SH, Mou Z, Tada Y, Spivey NW, Genschik P, Dong X. Proteasome-mediated turnover of the transcription coactivator NPR1 plays dual roles in regulating plant immunity. Cell. 2009;137:860–72.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  54. Fu ZQ, Yan S, Saleh A, Wang W, Ruble J, Oka N, et al. NPR3 and NPR4 are receptors for the immune signal salicylic acid in plants. Nature. 2012;486:228–32.

    PubMed  CAS  PubMed Central  Google Scholar 

  55. Bailey M, Srivastava A, Conti L, Nelis S, Zhang C, Florance H, et al. Stability of small ubiquitin-like modifier (SUMO) proteases OVERLY TOLERANT TO SALT1 and -2 modulates salicylic acid signalling and SUMO1/2 conjugation in Arabidopsis thaliana. J Exp Bot. 2016;67:353–63.

  56. Kumakura N, Takeda A, Fujioka Y, Motose H, Takano R. SGS3 and RDR6 interact and colocalize in cytoplasmic SGS3/RDR6-bodies. FEBS Lett. 2009;583:1261–6.

    Article  PubMed  CAS  Google Scholar 

  57. Pontes O, Vitins A, Ream TS, Hong E, Pikaard CS, Costa-Nunes P. Intersection of small RNA pathways in Arabidopsis thaliana sub-nuclear domains. PLoS One. 2013;8, e65652.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  58. Carr JP. Tobacco mosaic virus. In: Talbot NJ, editor. Annual plant reviews volume 11: plant-pathogen interactions. Oxford: Blackwell; 2004. p. 27–67.

    Google Scholar 

  59. Zhou T, Murphy AM, Lewsey MG, Westwood JH, Zhang H, González I, et al. Domains of the cucumber mosaic virus 2b silencing suppressor protein affecting inhibition of salicylic acid-induced resistance and priming of salicylic acid accumulation during infection. J Gen Virol. 2014;95:1408–13.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  60. Berry JO, Nikolau BJ, Carr JP, Klessig DF. Transcriptional and post-transcriptional regulation of ribulose 1,5-bisphosphate carboxylase gene expression in light- and dark-grown amaranth cotyledons. Mol Cell Biol. 1985;5:2238–46.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  61. Ramakers C, Ruijter JM, Deprez RH, Moorman AF. Assumption-free analysis of quantitative real-time PCR data. Neurosci Lett. 2003;339:62–6.

    Article  PubMed  CAS  Google Scholar 

  62. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-ΔΔC(t)) method. Methods. 2001;25:402–8.

    Article  PubMed  CAS  Google Scholar 

  63. Westwood JH, Lewsey MG, Murphy AM, Tungadi T, Bates A, Gilligan CA, et al. Interference with jasmonic acid-regulated gene expression is a general property of viral suppressors of RNA silencing but only partly explains virus-induced changes in plant-aphid interactions. J Gen Virol. 2014;95:733–9.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  64. Bradford MM. A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein dye-binding. Anal Biochem. 1976;72:248–54.

    Article  PubMed  CAS  Google Scholar 

  65. Pfeiffer P, Laquel P, Hohn T. Cauliflower mosaic virus replication complexes: characterisation of the associated enzymes and of the polarity of the DNA synthesised in vitro. Plant Mol Biol. 1984;3:261–70.

    Article  PubMed  CAS  Google Scholar 

  66. Lei Z, Anand A, Mysore KS, Sumner LW. Electroelution of intact proteins from SDS-PAGE gels and their subsequent MALDI-TOF MS analysis. Methods Mol Biol. 2007;355:353–63.

    PubMed  CAS  Google Scholar 

  67. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature. 1970;227:680–5.

    Article  PubMed  CAS  Google Scholar 

  68. Towbin H, Staehelin T, Gordon J. Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc Natl Acad Sci U S A. 1979;76:4350–4.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  69. Blom N, Gammeltoft S, Brunak S. Sequence and structure-based prediction of eukaryotic protein phosphorylation sites. J Mol Biol. 1999;294:1351–62.

    Article  PubMed  CAS  Google Scholar 

  70. Radivojac P, Vacic V, Haynes C, Cocklin RR, Mohan A, Heyen JW, et al. Identification, analysis, and prediction of protein ubiquitination sites. Proteins: Struct Funct Bioinf. 2010;78:365–80.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank Zhixiang Chen for his advice on the RDR assay protocol. Xiaoqiang Wang and Zhentian Lei at the Samuel Roberts Noble Foundation for the AKTA purification protocol and analysis of recombinant MBP:MtRDR1 fusion protein, respectively. David Baulcombe, Peter Palukaitis, Joel Milner and Lydia Hunter are thanked for stimulating discussions and useful advice and Adrienne Pate for expert technical assistance. FSF was funded by grants from the Cambridge Overseas Trust and the Ministry of Education of Taiwan, and WSL was funded by a studentship from the Biotechnology and Biological Sciences Research Council (BBSRC) and work in the Carr lab was funded by BBSRC grants (BB/D008204/1, BB/D014376/1, BB/J011762/1), The Leverhulme Trust (F/09 741/F, RPG-2012-667), and Cambridge University Isaac Newton Trust [Grant 12.07(1)]. RSN and SRC were funded by the Samuel Roberts Noble Foundation, Inc.

Funders had no role in design, in the collection, analysis, and interpretation of data; in the writing of the manuscript; and in the decision to submit the manuscript for publication.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John P. Carr.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

WSL, FSF, ZL and AMM carried out the experimental procedures with additional experiments by EAD, LG and MGL. SRC and RSN prepared and characterized the anti-MtRDR1 and supplied seed of N. benthamiana transformed with empty vector or vector containing MtRDR1. JPC initiated and directed the research. JPC, WSL, AMM and FSF wrote the manuscript with all authors providing additional information. All authors read and approved the final manuscript.

Additional files

Additional file 1:

Symptom severity in TMV-GFP infected Nicotiana tabacum . (PDF 10005 kb)

Additional file 2:

Extent of TMV.GFP spread into Nicotiana tabacum stem tissue adjacent to the shoot apical meristem. (PDF 2233 kb)

Additional file 3:

Relative MtRDR1 transcript level in MtRDR1 transgenic Nicotiana tabacum after SA treatment. (PDF 48 kb)

Additional file 4:

RDR activity in extracts from Nicotiana tabacum and Nicotiana benthamiana plants is not increased by in vitro treatment with SA. (PDF 484 kb)

Additional file 5:

Expression of a MBP-MtRDR1 fusion protein in E. coli and its purification by amylose affinity chromatography and gel filtration. (PDF 635 kb)

Additional file 6:

Calibration curve to determine the molecular weight (log 10 MW) of protein detected by anti-MtRDR1 serum on SDS-PAGE gel. (PDF 72 kb)

Additional file 7:

Predicted ubiquitination and phosphorylation sites in the MtRDR1 amino acid sequence. (PDF 153 kb)

Additional file 8:

Table of primers used for semi-quantitative and quantitative PCR. (PDF 76 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, WS., Fu, SF., Li, Z. et al. Salicylic acid treatment and expression of an RNA-dependent RNA polymerase 1 transgene inhibit lethal symptoms and meristem invasion during tobacco mosaic virus infection in Nicotiana benthamiana . BMC Plant Biol 16, 15 (2016). https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-016-0705-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-016-0705-8

Keywords